Sunteți pe pagina 1din 521

I. J. Dunn, E. Heinzle, J. Ingham, J. E.

Pfenosil
Biological Reaction Engineering
Dynamic Modelling Fundamentals
with Simulation Examples

Second, Completely Revised Edition

Biological Reaction Engineering. Second Edition. \. J. Dunn. E. Heinzle, J. Ingham, J- E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheitn
ISBN: 3-527-30759-1
Also of Interest

Ingham, X, Dunn, I. J., Heinzle, E., Pfenosil, J. E.


Chemical Engineering Dynamics
An Introduction to Modelling and Computer Simulation
Second, Completely Revised Edition
2000, ISBN 3-527-29776-6
Irving J. Dunn, Elmar Heinzle, John Ingham, Jifi E. Pf enosil

Biological
Reaction Engineering
Dynamic Modelling Fundamentals
with Simulation Examples

Second, Completely Revised Edition

WILEY-
VCH
WILEY-VCH GmbH & Co. KGaA
Dr. Irving J. Dunn This book was carefully produced. Nevertheless,
ETH Zurich authors and publisher do not warrant the informa-
Department of Chemical Engineering tion contained therein to be free of errors. Rea-
CH-8092 Zurich ders are advised to keep in mind that statements,
Switzerland data, illustrations, procedural details or other
items may inadvertently be inaccurate.
Professor Dr. Elmar Heinzle
University of Saarland
Department of Technical Biochemistry
P.O. Box 15 11 50 First Edition 1992
D-66041 Saarbrucken Second, Completely Revised Edition 2003
Germany

Dr. John Ingham


University of Bradford Library of Congress Card No.: Applied for.
Department of Chemical Engeering British Library Cataloguing-in-Publication Data:
Bradford BD7 1DP A catalogue record for this book is available from
United Kingdom the British Library.

Dr.JiriE.Prenosil
ETH Zurich
Department of Chemical Engineering Bibliographic information published by Die Deut-
CH-8092 Zurich sche Bibliothek. Die Deutsche Bibliothek lists this
Switzerland publication in the Deutsche Nationalbibliografie;
detailed bibliographic data is available in the
Internet at <http://dnb.ddb.de>.

2003 WILE Y-VCH Verlag


GmbH & Co. KGaA, Weinheim

All rights reserved (including those of translation


into other languages). No part of this book may be
reproduced in any form - by photoprinting, micro-
film, or any other means - nor transmitted or
translated into a machine language without writ-
ten permission from the publishers. Registered
names, trademarks, etc. used in this book, even
when not specifically marked as such, are not to
be considered unprotected by law.
Printing: Strauss Offsetdruck, Morlenbach
Bookbinding: GroBbuchbinderei J. Schaffer
GmbH & Co. KG, Griinstadt
Printed in the Federal Republic of Germany.
Printed on acid-free paper.

ISBN 3-527-30759-1
Table of Contents

TABLE OF CONTENTS V
PREFACE XI

PART I PRINCIPLES OF BIOREACTOR MODELLING 1

NOMENCLATURE FOR PART I 3

1 MODELLING PRINCIPLES 9
1.1 FUNDAMENTALS OF MODELLING 9
7.7.7 Use of Models for Understanding, Design and Optimization of Bioreactors 9
1.1.2 General Aspects of the Modelling Approach 10
1.1.3 General Modelling Procedure..... 72
1.1.4 Simulation Tools 75
7.7.5 Teaching Applications 75
1.2 DEVELOPMENT AND MEANING OF DYNAMC DIFFEREOTTAL BALANCES 16
1.3 FORMULATION OF BALANCE EQUATIONS ..21
7.5.7 Types of Mass Balance Equations 27
1.3.2 Balancing Procedure 23
1.3.2.1 Case A. Continuous Stirred Tank Bioreactor 24
1.3.2.2 CaseB. Tubular Reactor 24
1.3.2.3 Case C. River with Eddy Current 25
1.3.3 Total Mass Balances 33
1.3.4 Component Balances for Reacting Systems 34
1.3.4.1 Case A. Constant Volume Continuous Stirred Tank Reactor 35
1.3.4.2 Case B. Semi-continuous Reactor with Volume Change 37
1.3.4.3 Case C. Steady-State Oxygen Balancing in Fermentation 38
1.3.4.4 Case D. Inert Gas Balance to Calculate Flow Rates 39
7.5.5 Stoichiometry, Elemental Balancing and the Yield Coefficient Concept.. 40
1.3.5.1 Simple Stoichiometry 40
1.3.5.2 Elemental Balancing 42
L3.5.3 Mass Yield Coefficients 44
VI Table of Contents

1.3.5.4 Energy Yield Coefficients 45


1.3.6 Equilibrium Relationships 46
1.3.6.1 General Considerations 46
1.3.6.2 Case A. Calculation of pH with an Ion Charge Balance 47
1.3.7 Energy Balancing for Bioreactors..... 49
1.3.6.3 Case B. Determining Heat Transfer Area or Cooling Water Temperature 52
2 BASIC BIOREACTOR CONCEPTS 55
2.1 INFORMATION FOR BIOREACTOR MODELLING... .....55
2.2 BIOREACTOR OPERATION .....56
2.2.7 Batch Operation 57
2.2.2 Semicontinuous or Fed Batch Operation..... ....58
2.2.3 Continuous Operation , 60
2.2.4 Summary and Comparison 63
3 BIOLOGICAL KINETICS 67
3.1 ENZYME KINETICS 68
3.1.1 Michaelis-Menten Equation 68
3.1.2 Other Enzyme Kinetic Models 73
3.1.3 Deactivation 76
3.1.4 Sterilization 76
3.2 SIMPLE MICROBIAL KINETICS 77
3.2.1 Basic Growth Kinetics 77
3.2.2 Substrate Inhibition of Growth 80
3.2.3 Product Inhibition 81
3.2.4 Other Expressions for Specific Growth Rate 81
3.2.5 Substrate Uptake Kinetics 83
3.2.6 Product Formation 85
3.2.7 Interacting Microorganisms ....86
3.2.7.1 Case A. Modelling of Mutualism Kinetics..... 88
3.2.7.2 Case B. Kinetics of Anaerobic Degradation 89
3.3 STRUCTURED KINETIC MODELS ..........91
3.3.1 Case Studies 93
3.3.1.1 Case C. Modelling Synthesis of Poly-B-hydroxybutyric Acid (PHB) 93
3.3.1.2 Case D. Modelling of Sustained Oscillations in Continuous Culture 94
3.3.1.3 Case E. Growth and Product Formation of an Oxygen-Sensitive Bacillus-
subtilis Culture 97
4 BIOREACTOR MODELLING 101
4.1 GENERAL BALANCES FOR TANK-TYPE BIOLOGICAL REACTORS 101
4.1.1 The Batch Fermenter. 103
4.1.2 The Chemostat 104
4.1.3 The Fed Batch Fermenter 1 06
4.1.4 Biomass Productivity 109
4.1.5 Case Studies 109
Table of Contents VII

4.1.5.1 Case A. Continuous Fermentation with Biomass Recycle 110


4.1.5.2 Case B. Enzymatic Tanks-in-series Bioreactor System 112
4.2 MODELLING TUBULAR PLUG FLOW BIOREACTORS 113
4.2.1 Steady-State Balancing 113
4.2.2 Unsteady-State Balancing for Tubular Bioreactors 775
5 MASS TRANSFER 117
5.1 MASS TRANSFER IN BIOLOGICAL REACTORS 117
5.7.7 Gas Absorption with Bioreaction in the Liquid Phase 777
5.1.2 Liquid-Liquid Extraction with Bioreaction in One Phase 778
5.1.3 Surface Biocatalysis 778
5.7.4 Diffusion and Reaction in Porous Biocatalyst 779
5.2 INTERPHASEGAS-LIQUID MASS TRANSFER 119
5.3 GENERAL OXYGEN BALANCES FOR GAS-LIQUID TRANSFER 123
5.3.1 Application of Oxygen Balances 725
5.3.1.1 Case A. Steady-State Gas Balance for the Biological Uptake Rate 125
5.3.1.2 Case B. Determination of KLa Using the Sulfite Oxidation Reaction 126
5.3.1.3 Case C. Determination of Kj^a by a Dynamic Method 126
5.3.1.4 Case D. Determination of Oxygen Uptake Rates by a Dynamic Method 128
5.3.1.5 Case E. Steady-State Liquid Balancing to Determine Oxygen Uptake Rate.. 129
5.3.1.6 Case F. Steady-State Deoxygenated Feed Method for KJJI 130
5.3.1.7 Case G. Biological Oxidation in an Aerated Tank 131
5.3.1.8 Case H. Modelling Nitrification in a Fluidized Bed Biofilm Reactor 133
5.4 MODELS FOR OXYGEN TRANSFER IN LARGE SCALE BIOREACTORS 137
5.4.1 Case Studies for Large Scale Bioreactors 7 39
5.4.1.1 Case A.Model for Oxygen Gradients in a Bubble Column Bioreactor 139
5.4.1.2 Case B.Model for a Multiple Impeller Fermenter 140
6 DIFFUSION AND BIOLOGICAL REACTION IN IMMOBILIZED
BIOCATALYST SYSTEMS 145
6.1 EXTERNAL MASS TRANSFER 146
6.2 INTERNAL DIFFUSION AND REACTION WITHIN BIOCATALYSTS ..... 149
6.2.1 Derivation of Finite Difference Model for Diffusion-Reaction Systems. 151
6.2.2 Dimensionless Parameters from Diffusion-Reaction Models 754
6.2.5 The Effectiveness Factor Concept. 755
6.2.4 Case Studies for Diffusion with Biological Reaction 757
6.2.4.1 Case A. Estimation of Oxygen Diffusion Effects in a Biofilm 157
6.2.4.2 Case B. Complex Diffusion-Reaction Processes (Biofilm Nitrification).... 157
7 AUTOMATIC BIOPROCESS CONTROL FUNDAMENTALS 161
7.1 ELEMENTS OF FEEDBACK CONTROL 161
7.2 TYPES OF CONTROLLER ACTION 163
7.2.7 On-OffControl 163
7.2.2 Proportional (P) Controller 764
7.2.3 Proportional-Integral (PI) Controller 765
VIII Table of Contents

7.2.4 Proportional-Derivative (PD) Controller 166


7.2.5 Proportional-Integral-Derivative (PID) Controller 167
7.3 CONTROLLER TUNING 169
7.3.1 Trial and Error Method 769
7.3:2 Ziegler-Nichols Method. 769
7.3.3 Cohen-Coon Controller Settings 170
7.3.4 Ultimate Gain Method 777
7.4 ADVANCED CONTROL STRATEGIES 172
7.4.1 Cascade Control 772
7.4.2 Feed Forward Control 173
7.4.3 Adaptive Control 774
7.4.4 Sampled-Data Control Systems 774
7.5 CONCEPTS FOR BIOPROCESS CONTROL 175
7.5.7 Selection of a Control Strategy 776
7.5.2 Methods of Designing and Testing the Strategy 7 78
REFERENCES 181
REFERENCES CITED IN PART I 181
RECOMMENDED TEXTBOOKS AND REFERENCES FOR FURTHER READING 184
PART II DYNAMIC BIOPROCESS SIMULATION EXAMPLES AND
THE BERKELEY MADONNA SIMULATION LANGUAGE. 191

8 SIMULATION EXAMPLES OF BIOLOGICAL REACTION


PROCESSES USING BERKELEY MADONNA 193
8.1 INTRODUCTORY EXAMPLES 193
8.7.7 Batch Fermentation (BATFERM) 793
8.7.2 ChemostatFermentation (CHEMO) 799
8.1.3 Fed Batch Fermentation (FEDBAT) 204
8.2 BATCH REACTORS 209
8.2.7 Kinetics of Enzyme Action (MMKINET) 209
8.2.2 Lineweaver-Burk Plot (LINEWEAV) .....272
8.2.3 Oligosaccharide Production in Enzymatic Lactose Hydrolysis (OLIGO) 215
8.2.4 Structured Model for PHB Production (PHB) ....279
8.3 FED BATCH REACTORS 224
8.3.1 Variable Volume Fermentation (VARVOL and VARVOLD) 224
8.3.2 Penicillin Fermentation Using Elemental Balancing (PENFERM) 230
8.3.3 Ethanol Fed Batch Diauxic Fermentation (ETHFERM) 240
8.3.4 Repeated Fed Batch Culture (REPFED) 245
8.3.5 Repeated Medium Replacement Culture (REPLCUL) 249
8.3.6 Penicillin Production in a Fed Batch Fermenter (PENOXY) 253
8.4 CONTINUOUS REACTORS 257
8.4.7 Steady-State Chemostat (CHEMOSTA) 257
8.4.2 Continuous Culture with Inhibitory Substrate (CONINHIB) 267
8.4.3 Nitrification in Activated Sludge Process (ACTNITR) 267
Table of Contents IX

8.4.4 Tubular Enzyme Reactor (ENZTUBE) 272


8.4.5 Dual Substrate Limitation (DUAL) 275
8.4.6 Dichloromethane in a Biofllm Fluidized Sand Bed (DCMDEG) 280
8.4.7 Two-Stage Chemostat with Additional Stream (TWOSTAGE) 286
8.4.8 Two Stage Culture with Product Inhibition (STAGED) 290
8.4.9 Fluidized Bed Recycle Reactor (FBR) 295
8.4.10 Nitrification in a Fluidized Bed Reactor (NITBED)... 299
8.4.11 Continuous Enzymatic Reactor (ENZCON) 305
8.4.12 Reactor Cascade with Deactivating Enzyme (DEACTENZ) 308
8.4.13 Production ofPHB in a Two-Tank Reactor Process (PHBTWO) 314
8.5 OXYGEN UPTAKE SYSTEMS 318
8.5.1 Aeration of a Tank Reactor for Enzymatic Oxidation (OXENZ) 318
8.5.2 Gas and Liquid Oxygen Dynamics in a Continuous Fermenter (INHIB) 321
8.5.3 Batch Nitrification with Oxygen Transfer (NITRIF) 327
8.5.4 Oxygen Uptake and Aeration Dynamics (OXDYN) 331
8.5.5 Oxygen Electrode for Kia (KLADYN, KLAFIT and ELECTFIT) 335
8.5.6 Biofiltration Column with Two Inhibitory Substrates (BIOFILTDYN). 342
8.5.7 Optical Sensing in Microtiter Plates (TITERDYN and T1TERB1O) 349
8.6 CONTROLLED REACTORS 354
8.6.1 Feedback Control of a Water Heater (TEMPCONT) 354
8.6.2 Temperature Control of Fermentation (FERMTEMP) 358
8.6.3 Turbidostat Response (TURBCON) 363
8.6.4 Control of a Continuous Bioreactor, Inhibitory Substrate (CONTCON)367
8.7 DIFFUSION SYSTEMS ....371
8.7.1 Double Substrate Biofilm Reaction (BIOFILM) 377
8.7.2 Steady-State Split Boundary Solution (ENZSPLIT).... 377
8.7.3 Dynamic Porous Diffusion and Reaction (ENZDYN).... 383
8.7.4 Oxygen Diffusion in Animal Cells (CELLDIFF) 388
8.7.5 Biofilm in a Nitrification Column System (NITBEDFILM) 393
8.8 MULTI-ORGANISM SYSTEMS ..400
8.8.1 Two Bacteria with Opposite Substrate Preferences (COMMENSA) 400
8.8.2 Competitive Assimilation and Commensalism (COMPASM) 406
8.8.3 Stability of Recombinant Microorganisms (PLASMID) 411
8.8.4 Predator-Prey Population Dynamics (MIXPOP) 417
8.8.5 Competition Between Organisms (TWOONE) 422
8.8.6 Competition between Two Microorganisms in a Biofilm (FILMPOP). 425
8.8.7 Model for Anaerobic Reactor Activity Measurement (ANAEMEAS).... 433
8.8.8 Oscillations in Continuous Yeast Culture (YEASTOSC) 441
8.8.9 Mammalian Cell Cycle Control (MAMMCELLCYCLE) 445
8.9 MEMBRANE AND CELL RETENTION REACTORS 451
8.9.1 Cell Retention Membrane Reactor (MEMINH) 451
8.9.2 Fermentation with Pervaporation (SUBTILIS) 455
8.9.3 Two Stage Fermentor With Cell Recycle (LACMEMRECYC) 464
8.9.4 Hollow Fiber Enzyme Reactor for Lactose Hydrolysis (LACREACT). 470
8.9.5 Animal Cells in a Fluidized Bed Reactor (ANIMALIMMOB) 477
X Table of Contents

9 APPENDIX: USING THE BERKELEY MADONNA LANGUAGE.. 483


9.1 A SHORT GUIDE TO BERKELEY MADONNA 483
9.2 SCREENSHOT GUIDE TO BERKELEY MADONNA 488
10 ALPHABETICAL LIST OF EXAMPLES 497

11 INDEX 499
Preface

Our goal in this textbook is to teach, through modelling and simulation, the
quantitative description of bioreaction processes to scientists and engineers. In
working through the many simulation examples, you, the reader, will learn to
apply mass and energy balances to describe a variety of dynamic bioreactor
systems. For your efforts, you will be rewarded with a greater understanding of
biological rate processes. The many example applications will help you to gain
confidence in modelling, and you will find that the simulation language used,
Berkeley Madonna, is a powerful tool for developing your own simulation
models. Your new abilities will be valuable for designing experiments, for
extracting kinetic data from experiments, in designing and optimizing
biological reaction systems, and for developing bioreactor control strategies.
This book is based on part of our successful course, "Biological Reaction
Engineering", which has been held annually in the Swiss mountain resort of
Braunwald for the past twenty five years and which is now known, throughout
European biotechnology circles as the "Braunwald Course". More details can
be found at our website www.braunwaldcourse.ch. Modelling is often
unfamiliar to biologists and chemists, who nevertheless need modelling
techniques in their work. The general field of biochemical reaction engineering
is one that requires a very close interdisciplinary interaction between applied
microbiologists, biochemists, biochemical engineers, engineers and managers; a
large degree of collaboration and mutual understanding is therefore important.
Professional microbiologists and biochemists often lack the formal training
needed to analyze laboratory kinetic data in its most meaningful sense, and
they may sometimes experience difficulty in participating in engineering
design decisions and in communicating with engineers. These are just the very
types of activity required in the multi-disciplinary field of biotechnology.
Chemical engineering's greatest strength is its well-developed modelling
concepts, based on mass and energy balances, combined with rate processes.
Biochemical engineering is a discipline closely related to conventional
chemical engineering, in that it attempts to apply physical principles to the
solution of biological problems. This approach may be applied to the
measurement and interpretation of laboratory kinetic data or as well to the
design of large-scale fermentation, enzymatic or waste treatment processes. The
necessary interdisciplinary cooperation requires the biological scientists and
chemical engineers involved to have at least a partial understanding of each
other's field. The purpose of this book is to provide the mathematical tools
necessary for the quantitative analysis of biological kinetics and other
biological process phenomena. More generally, the mathematical modelling
XII Preface

methods presented here are intended to lead to a greater understanding of how


the biological reaction systems are influenced by process situations.
Engineering science depends heavily on the use of applied theory,
quantitative correlations and mathematics, and it is often difficult for all of us
(not only the biological scientist) to surmount the mathematical barrier, which
is posed by engineering. A mistake, often made, is to confuse "mathematics"
with the engineering modelling approach. In modelling an attempt is made to
analyze a real and possibly very complex situation into a simplified and
understandable physical analog. This physical model may contain many
subsystems, all of which still make physical sense, but which now can be
formulated as mathematical equations. These equations can be handled
automatically by the computer. Thus the engineer and the biologist are freed
from the difficulties of mathematical solution and can tackle complex problems
that were impossible before. Models, however, still have to be formulated and
one of the most important tools of the biochemical engineer, in this operation,
is the use of material balance equations. Though it may not be easy for the
microbiologist to fully appreciate the importance of differential equations, mass
balance equations are not so difficult to understand, since the first law of
conservation, namely that matter can neither be created nor destroyed, is
fundamental to all science. Mass balances, when combined with kinetic rate
equations, to form simple mathematical models, can be used with very great
effect as a means of planning, conducting and analyzing experiments. Models
are especially important as a means of obtaining a better understanding of
process phenomena. A rational approach to experimentation and design
requires a considerable knowledge of the system, which can really only be
achieved by means of a mathematical model. This book attempts to
demonstrate this by way of the many detailed examples.
The contribution made to biotechnology by the biochemical engineering
modelling approach is especially important because the basic procedure can be
developed from a few fundamental principles. An aim of this book is to
demonstrate that you do not have to be an engineer to learn modelling and
simulation. The basic concepts of the material balance, combined with
biological and enzyme kinetics, are easily applied to describe the behavior of
well-stirred tank and tubular fermenters, mixed culture dynamics, interphase
gas-liquid mass transfer and internal biofilm diffusional limitations, as
demonstrated in the computer examples supplied with this book. Such models,
when solved interactively by computer simulation, become much more
understandable to non-engineers.
The Berkeley Madonna simulation language, used for the examples in this
book, is especially suitable because of its sophisticated computing power,
interactive facility and ease of programming. The use of this digital simulation
programming language makes it possible for the reader, student and teacher to
experiment directly with the model, in the classroom or at the desk. In this way
it is possible to immediately determine the influence of changing various
Preface XIII

operating parameters on the bioreactor performance - a real learning


experience.
The simulation examples serve to enforce the learning process in a very
effective manner and also provide hands-on confidence in the use of a
simulation language. The readers can program their own examples, by
formulating new mass balance equations or by modifying an existing example
to a new set of circumstances. Thus by working directly at the computer, the
no-longer-passive reader is able to experiment directly on the bioreaction
system in a very interactive way by changing parameters and learning about
their probable influence in a real situation. Because of the speed of solution, a
true degree of interaction is possible with Berkeley Madonna, allowing
parameters to be changed easily. Plotting the variables in any configuration is
easy during a run, and the results from multiple runs can be plotted together
for comparison. Other useful features include data fitting and optimization.
In our experience, digital simulation has proven itself to be absolutely the
most effective way of introducing and reinforcing new concepts that involve
multiple interactions. The thinking process is ultimately stimulated to the point
of solid understanding.

Organization of the Book

The book is divided into two parts: a presentation of the background theory in
Part I and the computer simulation exercises in Part II. The function of the text
in Part I is to provide the basic theory required to fully understand and to make
full use of the computer examples and simulation exercises. Numerous case
studies provide illustration to the theory. Part II constitutes the main part of this
book, where the simulation examples provide an excellent instructional and
self-learning tool. Each of the more than fifty examples is self-contained,
including a model description, model equations, exercises, computer program
listing, nomenclature and references. The exercises range from simple
parameter-changing investigations to suggestions for writing a new program.
The combined book thus represents a synthesis of basic theory and computer-
based simulation examples.
Quite apart from the educational value of the text, the introduction and use
of the Berkeley Madonna software provides the reader with the considerable
practical advantage of a differential equation solution package. In the appendix
a screenshot guide is found concerning the use of the software.
Part I: "Principles of Bioreactor Modelling" covers the basic theory
necessary for understanding the computer simulation examples. This section
presents the basic concepts of mass balancing, and their combination with
kinetic relationships, to establish simple biological reactor models, carefully
presented in a way that should be understandable to biologists. In fact,
engineers may also find this rigorous presentation of balancing to be valuable.
XIV Preface

In order to achieve this aim, the main emphasis of the text is placed on an
understanding of the physical meaning and significance of each term in the
model equations. The aim in presenting the relevant theory is thus not to be
exhaustive, but simply to provide a basic introduction to the theory required for
a proper understanding of the modelling methodology.
Chapter 1 deals with the basic concepts of modelling, the basic principles,
development and significance of differential balances and the formulation of
mass and energy balance relationships. Emphasis is given to physical
understanding. The text is accompanied by example cases to illustrate the
application of the material.
Chapter 2 serves to introduce the varied operational characteristics of the
various types of bioreactors and their differing modes of operation, with the
aim of giving a qualitative insight into the quantitative behavior of the
computer simulation examples.
Chapter 3 provides an introduction to enzyme and microbial kinetics. A
particular feature of the kinetic treatment is the emphasis on the use of more
complex structured models. Such models require much more consideration to
be given to the biology of the system during the modelling procedure, but
despite their added complexity can nevertheless also be solved with relative
ease. They serve as a reminder that biological reactions are really infinitely
complex.
Chapter 4 is used to derive general mass balance equations, covering all types
of fermentation tank reactors. These generalized equations are then simplified
to show their application to the differing modes of stirred tank bioreactor
operation, discussed previously and which are illustrated by the simulation
examples.
Chapter 5 explains the basic theory of interfacial mass transfer as applied to
fermentation systems and shows how equations for rates of mass transfer can be
combined with mass balances, for both liquid and gas phases. A particular
extension of this approach is the combination of transfer rate and material
balance equations to models of increased geometrical complexity, as
represented by large-scale air-lift and multiple-impeller fermenters.
Chapter 6 treats the cases of external diffusion to a solid surface and internal
diffusion combined with biochemical reaction, with practical application to
immobilized biocatalyst and biofilm systems. Emphasized here is the
conceptual ease of handling a complex reaction in a solid biocatalyst matrix.
The resulting sets of tractable differential-difference equations are solved by
simulation techniques in several examples.
Chapter 7 describes the importance of control and summarizes control
strategies used for bioreaction processes. Here the fundamentals of feedback
control systems and their characteristic responses are discussed. This material
forms the basis for performing the many recommended control exercises in the
simulation examples. It also will allow the reader-simulator to develop his or
her own control models and simulation programs.
Preface XV

Part II, "Dynamic Bioprocess Simulation Examples and the Berkeley Madonna
simulation language" comprises Chapter 8, with the computer simulation
examples, and Chapter 9, which gives the instructions for using Madonna, Each
example in Chapter 8 includes a description of its physical system, the model
equations, that were developed in Part I, and a list of suggested exercises. The
programs are found on the CD-ROM. These example exercises can be carried
out in order to explore the model system in detail, and it is suggested that work
on the computer exercises be done in close reference to the model equations
and their physical meaning, as described in the text. The exercises, however, are
provided simply as an idea for what might be done and are by no means
mandatory or restrictive. Working through a particular example will often
suggest an interesting variation, such as a control loop, which can then be
programmed and inserted. The examples cover a wide range of application and
can easily be extended by reference to the literature. They are robust and are
well tested by a variety of undergraduate and graduate students and by also the
350 participants, or so, who have previously attended the Braunwald course. In
tackling the exercises, we hope you will soon come to share our conviction that,
besides being very useful, computer simulation is also fun to do.
For the second edition, the text was thoroughly revised and some of our
earlier, less relevant material was omitted. On the other hand, a number of new
examples resulting mainly from the authors' latest research and teaching work
were added. There was also an opportunity in this new edition to eliminate
most of the past errors and to avoid new ones as much as possible. Most
importantly, the examples have been rewritten in Berkeley Madonna, which all
of our reader-simulators will greatly appreciate.
Our book has a number of special characteristics. It will be obvious, in
reading it through, that we concentrate only on those topics of biological
reaction engineering that lend themselves to modelling and simulation and do
not attempt to cover the area completely. Our own research work is used to
illustrate theoretical points and from it many simulation examples are drawn. A
list of suggested books for supplementary reading is found at the end of
Chapter 6, together with the list of cited references. The diversity of the
simulation examples made it necessary to use separate nomenclature for each.
The symbols used in Chapters 1 - 6 are defined at the end of Part I. The
authors' four nationalities and three mother tongues, made it difficult to settle
on American or British spelling. Somehow we like "modelling" better than
"modeling".
We are confident that the book will be useful to all life scientists wishing to
obtain an understanding of biochemical engineering and also to those chemical
and biochemical engineers wanting to sharpen their modelling skills and
wishing to gain a better understanding of biochemical process phenomena. We
hope that teachers with an interest in modelling will find this to be a useful
textbook for undergraduate and graduate biochemical engineering and
biotechnological courses.
XVI Preface

Acknowledgements

A major acknowledgement should be made to the excellent pioneering texts of


R. G. E. Franks (1967 and 1972) and also of W. L. Luyben (1973), for
inspiring our interest in digital simulation.
We are especially grateful to our students and to the past-participants of the
Braunwald course, for their assistance in the continuing development of the
course and of the material presented in this book. Continual stimulus and
assistance has also been given by our doctoral candidates, especially at the
Chemical Engineering Department, ETH-Zurich, as noted throughout the
references.
We are grateful to and have great respect for the developers of Berkeley
Madonna and hope that this new version of the book will be useful in drawing
attention to this wonderful simulation language.
Part I Principles of
Bioreactor Modelling

Biological Reaction Engineering, Second Edition, I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
Nomenclature for Part I

Symbols Units

A Area m2
A Magnitude of controller input signal various
a Specific area m2/m3
a Constant in Logistic Equation 1/h
b Constant in Luedeking-Piret relation 1/h
b Constant in Logistic Equation m3/kg h
B Magnitude of controller output signal various
c Fraction carbon converted to biomass
C Concentration kg/m3, kmol/m3
CP Heat capacity kJ/kg K, kJ/mol K
CPR Carbon dioxide production rate mol/h
D Diffusivity m2/h
D Dilution rate 1/h
d Differential operator and diameter -, m
d Fraction carbon converted to product
DO Dissolved oxygen g/m3, 9 air sat.
E Enzyme concentration g/m3
E Ethanol kg/m3
ES Enzyme-substrate concentration g/m3
f Fraction carbon converted to CO2
f Frequency in the ultimate gain method 1/h
F Flow rate m3/h and m3/s
G Gas flow rate m3
G Intracellular storage product kg/m3
hi Partial molar enthalpy kJ/mol
H Henry's Law constant bar m3/kg
AH Enthalpy change kJ/mol or kJ/kg
I Inhibiting component concentration kg/m3
I Cell compartment masses kg/m3
j Mass flux kg/m2h, mol/m2h
K Mass transfer coefficient 1/h
K Constant in Cohen-Coon method various
KD Acid-base dissociation constant

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
Nomenclature

k Constant various
Kca Gas-liquid mass transfer coefficient 1/h
kGa Gas film mass transfer coefficient 1/h
KI Inhibition constant kg/m3, kmol/m3
KLa Gas-liquid mass transfer coefficient 1/h
kLa Liquid film mass transfer coefficient 1/h
KM Michaelis-Menten constant kg/m3, kmol/m3
Kp Proportional controller gain constant various
KW Dissociation constant of water
KS Monod saturation coefficient kg/m3
L Length m
M Mass kg or mol
m Maintenance coefficient 1/h
N Mass flux kg/m2 h
N Molar flow rate mol/h
n Number of mols
n Reaction order _
OTR Oxygen transfer rate mol/h and kg/h
OUR Oxygen uptake rate mol/h and kg/h
P Pressure bar
P Product concentration kg/m3 and g/m3
P Output control signal various
Q Total transfer rate kg/h and mol/h
q Specific rate kg/kg biomass h
R Ideal gas constant bar m3/ K mol
R Recycle flow rate m3/h
R Residual active biomass kg/m3
r Reaction rate kg/m3h, kmol/m3h
fi Reaction rate of component i kg i/m3h
r
i/j Reaction rate of component i to j kg /m3h, kmol/m3h
RQ Respiration quotient mol CO2/mol 2
rx Growth rate kg biomass/m3 h
S Concentration of substrate kg/m3, kmol/m3
S Slope of process reaction curve various
s Stoichiometric coefficient
T Temperature CorK
T Enzyme activity kg/m3
T Time lag h, min. or s
L
t Time h, min and s
Tr Transfer rate mol/m3 h
U Heat transfer coefficient kJ/m2 C h
V Volume m3
V Flow velocity m/h
v
max Maximum reaction rate kmol/m3 h
w Wastage stream flow rate m3/h
Nomenclature

w Mass fraction
X Biomass concentration kg/m3
Y Yield coefficient kg/kg
Yi Yield of i from j kg i/kg j,mol i/mol j
y Mol fraction in gas
Z Length variable m

Greek

8 Controller error various


a Partial differential operator
8 Concentration difference quantity kg/m3
A Difference operator
O Thiele Modulus
Effectiveness factor
Specific growth rate 1/h
Maximum growth rate 1/h
V Stoichiometric coefficient
p Density kg/m3
Z Summation operator
T Residence time h and s
T Controller time constant s
T Electrode time constant s

Indices
* Refers to equilibrium concentration
0 Refers to initial, inlet, external, and zero order
1 Refers to time ti, outlet, component 1, tank 1, and first order
2 Refers to tank 2, time t2 and component 2
1,2,..., n Refers to stream, volume elements and stages
A Refers to component A, anions and bulk
A- Refers to anions
a Refers to ambient
Ac Refers to acetoin and acetoin formation
aer Refers to aerobic
agit Refers to agitation
anaer Refers to anaerobic
app Refers to apparent
ATP/S,Ac Refers to ATP yield from reaction glucose --> Ac
ATP/S,CO2 Refers to ATP yield from glucose oxidation
ATP/NADH Refers to ATP produced from NADH
ATP/X Refers to consumption rate ATP > biomass
Nomenclature

avg Refers to average


B Refers to component B, base, backmixing, and surface position
Bu Refers to butanediol
CO2 Refers to carbon dioxide
d Refers to deactivation and death
D Refers to derivative control
D Refers to D-value in sterilization
E Refers to electrode
E Refers to energy by complete oxidation
f Refers to final
G Refers to gas and to cellular compartment
H+ Refers to hydrogen ions
i Refers to component i and to interface
I Refers to inhibitor
I Refers to integral control
inert Refers to inert component
K Refers to cellular compartment
K+ Refers to cations
L Refers to liquid
m Refers to maximum
m Refers to metabolite
max Refers to maximum
n Refers to tank number
NH4 Refers to ammonium
NC>2 Refers to nitrite
NOs Refers to nitrate
O and O2 Refer to oxygen
P Refers to product
PA Refers to product A
PB Refers to product B
Q Refers to heat
Q/O2 Refers to heat-oxygen ratio
Q/S Refers to heat-substrate ratio
R Refers to recycle stream
r Refers to reactor
r,S Refers to reaction of substrate
s Refers to settler
S Refers to substrate and surface
SL Refers to liquid film at solid interface
Sn Refers to substrate n
tot Refers to total
X Refers to biomass
X/i Refers to biomass-component i ratio
X/S Refers to biomass-substrate ratio
Nomenclature

Refers to difference between cations and ions


Bar above symbol refers to dimensionless variable
1 Modelling Principles

1.1 Fundamentals of Modelling

1.1.1 Use of Models for Understanding, Design and


Optimization of Bioreactors

An investigation of bioreactor performance might conventionally be carried


out in an almost entirely empirical manner. In this approach, the bioreactor
behavior would be studied under practically all combinations of possible
conditions of operation and the results then expressed as a series of
correlations, from which the resulting performance might hopefully be
estimated for any given set of new operating conditions. This empirical
procedure can be carried out in a very routine way and requires relatively little
thought concerning the actual detail of the process. While this might seem to
be rather convenient, the procedure has actually many disadvantages, since very
little real understanding of the process would be obtained. Also very many
experiments would be required in order to obtain correlations that would cover
every process eventuality.
Compared to this, the modelling approach attempts to describe both actual
and probable bioreactor performance, by means of well-established theory,
which when described in mathematical terms, represents a working model for
the process. In carrying out a modelling exercise, the modeller is forced to
consider the nature of all the important parameters of the process, their effect
on the process and how each parameter can be defined in quantitative terms,
i.e., the modeller must identify the important variables and their separate
effects, which, in practice, may have a very highly interactive combined effect
on the overall process. Thus the very act of modelling is one that forces a
better understanding of the process, since all the relevant theory must be
critically assessed. In addition, the task of formulating theory into terms of
mathematical equations is also a very positive factor that forces a clear
formulation of basic concepts.
Once formulated, the model can be solved and the behavior predicted by the
model compared with experimental data. Any differences in performance may

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
10 1 Modelling Principles

then be used to further redefine or refine the model until good agreement is
obtained. Once the model is established it can then be used, with reasonable
confidence, to predict performance under differing process conditions, and it
can also be used for such purposes as process design, optimization and control.
An input of plant or experimental data is, of course, required in order to
establish or validate the model, but the quantity of experimental data required,
as compared to that of the empirical approach is considerably reduced. Apart
from this, the major advantage obtained, however, is the increased
understanding of the process that one obtains simply by carrying out the
modelling exercise.
These ideas are summarized below.
Empirical Approach: Measure productivity for all combinations of reactor
operating conditions, and make correlations.
- Advantage: Little thought is necessary
- Disadvantage: Many experiments are required.

Modelling Approach: Establish a model, and design experiments to determine


the model parameters. Compare the model behavior with the experimental
measurements. Use the model for rational design, control and optimization.
- Advantage: Fewer experiments are required, and greater understanding is
obtained.
- Disadvantage: Some strenuous thinking may be necessary.

1.1.2 General Aspects of the Modelling Approach

An essential stage in the development of any model, is the formulation of the


appropriate mass and energy balance equations (Russell and Denn, 1972). To
these must be added appropriate kinetic equations for rates of cell growth,
substrate consumption and product formation, equations representing rates of
heat and mass transfer and equations representing system property changes,
equilibrium relationships, and process control (Blanch and Dunn, 1973). The
combination of these relationships provides a basis for the quantitative
description of the process and comprises the basic mathematical model. The
resulting model can range from a very simple case of relatively few equations
to models of very great complexity.
Simple models are often very useful, since they can be used to determine the
numerical values for many important process parameters. For example, a
model based on a simple Monod kinetics can be used to determine basic
parameter values such as the specific growth rate (JLI), saturation constant (Ks),
biomass yield coefficient (Yx/s) and maintenance coefficient (m). This basic
kinetic data can be supplemented by additional kinetic factors, such as oxygen
transfer rate (OTR), carbon dioxide production rate (CPR), respiration quotient
(RQ) based on off-gas analysis and related quantities, such as specific oxygen
1.1 Fundamentals of Modelling 11

uptake rate (qo2)> specific carbon dioxide production rate (qco2)> may also be
derived and used to provide a complete kinetic description of, say, a simple
batch fermentation.
For complex fermentations, involving product formation, the specific
product production rate (qp) is often correlated as a complex function of
fermentation conditions, e.g., stirrer speed, air flow rate, pH, dissolved oxygen
content and substrate concentration. In other cases, simple kinetic models can
also be used to describe the functional dependence of productivity on cell
density and cell growth rate.
A more detailed "structured kinetic model" may be required to give an
adequate description of the process, since cell composition may change in
response to changes in the local environment within the bioreactor. The greater
the complexity of the model, however, the greater is then the difficulty in
identifying the numerical values for the increased number of model parameters,
and one of the skills of modelling is to derive the simplest possible model that
is capable of a realistic representation of the process.
A basic use of a process model is thus to analyze experimental data and to
use this to characterize the process, by assigning numerical values to the
important process variables. The model can then also be solved with
appropriate numerical data values and the model predictions compared with
actual practical results. This procedure is known as simulation and may be
used to confirm that the model and the appropriate parameter values are
"correct". Simulations, however, can also be used in a predictive manner to test
probable behavior under varying conditions; this leads on to the use of models
for process optimization and their use in advanced control strategies.
The application of a combined modelling and simulation approach leads to
the following advantages:

1. Modelling improves understanding, and it is through understanding that


progress is made. In formulating a mathematical model, the modeller is
forced to consider the complex cause-and-effect sequences of the
process in detail, together with all the complex inter-relationships that
may be involved in the process. The comparison of a model prediction
with actual behavior usually leads to an increased understanding of the
process, simply by having to consider the ways in which the model might
be in error. The results of a simulation can also often suggest reasons as
to why certain observed, and apparently inexplicable, phenomena occur
in practice.

2. Models help in experimental design. It is important that experiments be


designed in such a way that the model can be properly tested. Often the
model itself will suggest the need for data for certain parameters, which
might otherwise be neglected, and hence the need for a particular type of
experiment to provide the required data. Conversely, sensitivity tests on
the model may indicate that certain parameters may have a negligible
12 1 Modelling Principles

effect and hence that these effects therefore can be neglected both from
the model and from the experimental program.

3. Models may be used predictively for design and control. Once the
model has been established, it should be capable of predicting
performance under differing sets of process conditions. Mathematical
models can also be used for the design of relatively sophisticated control
algorithms, and the model, itself, can often form an integral part of the
control algorithm. Both mathematical and knowledge based models can
be used in designing and optimizing new processes.

4. Models may be used in training and education. Many important aspects


of bioreactor operation can be simulated by the use of very simple
models. These include such concepts as linear growth, double substrate
limitation, changeover from batch to fed-batch operation dynamics, fed-
batch feeding strategies, aeration dynamics, measurement probe
dynamics, cell retention systems, microbial interactions, biofilm diffusion
and bioreactor control. Such effects are very easily demonstrated by
computer, as shown in the accompanying simulation examples, but are
often difficult and expensive to demonstrate in practice.

5. Models may be used for process optimization. Optimization usually


involves considering the influence of two or more variables, often one
directly related to profits and one related to costs. For example, the
objective might be to run a reactor to produce product at a maximum
rate, while leaving a minimum amount of unreacted substrate.

1.1.3 General Modelling Procedure

One of the more important features of modelling is the frequent need to


reassess both the basic theory (physical model) and the mathematical equations,
representing the physical model (mathematical model) in order to achieve the
required degree of agreement, between the model prediction and actual plant
performance (experimental data).
As shown in Fig. 1.1, the following stages in the modelling procedure can be
identified:

(i) The first stage involves the proper definition of the problem and hence the
goals and objectives of the study. These may include process analysis,
improvement, optimization, design and control, and it is important that the aims
of the modelling procedure are properly defined. All the relevant theory must
then be assessed in combination with any practical experience with the process,
1.1 Fundamentals of Modelling 13

Physical Model
New

LJt
Mathematical Model
Revise ideas
and equations

NO
experiments

Experimental Data

Comparison
Solution: C = f(t)
OK?
YES

Use for design,


optimization and control

Figure 1.1. Information flow diagram for model building.

and perhaps alternative physical models for the process need to be developed
and examined. At this stage, it is often helpful to start with the simplest possible
conception of the process and to introduce complexities as the development
proceeds, rather than trying to formulate the full model with all its complexities
at the beginning of the modelling procedure.

(ii) The available theory must then be formulated in mathematical terms.


Most bioreactor operations involve quite a large number of variables (cell,
substrate and product concentrations, rates of growth, consumption and
production) and many of these vary as functions of time (batch, fed-batch
operation). For these reasons the resulting mathematical relationships often
consist of quite large sets of differential equations. The thick arrow in Fig. 1.1
designates both the importance and the difficulty of this mathematical
formulation.

(iii) Having developed a model, the model equations must then be solved.
Mathematical models of biological systems are usually quite complex and
highly non-linear and are such that the mathematical complexity of the
14 1 Modelling Principles

equations is usually sufficient to prohibit the use of an analytical means of


solution. Numerical methods of solution must therefore be employed, with the
method preferred in this text being that of digital simulation. With this
method, the solution of very complex models is accomplished with relative ease,
since digital simulation provides a very easy and a very direct method of
solution.
Digital simulation languages are designed specially for the solution of sets
of simultaneous differential equations using numerical integration. Many fast
and efficient numerical integration routines are now available and are
implemented within the structure of the languages, such that many digital
simulation languages are able to offer a choice of integration routine. Sorting
algorithms within the structure of the language enable very simple programs to
be written, having an almost one-to-one correspondence with the way in which
the basic model equations were originally formulated. The resulting simulation
programs are therefore very easy to understand and also to write. A further
major advantage is a convenient output of results, in both tabulated and
graphical form, that can be obtained via very simple program commands.

(iv) The validity of the computer prediction must be checked and steps (i) to
(iii) will often need to be revised at frequent intervals during the modelling
procedure. The validity of the model depends on the correct choice of the
available theory (physical and mathematical model), the ability to identify the
model parameters correctly and the accuracy of the numerical solution method.
In many cases, owing to the complexity and very interactive nature of
biological processes, the system will not be fully understood, thus leaving large
areas of uncertainty in the model. Also the relevant theory may be very
difficult to apply. In such cases, it is then often very necessary to make rather
gross simplifying assumptions, which may subsequently be eliminated or
improved as a better understanding is subsequently obtained. Care and
judgement must also be used such that the model does not become over
complex and so that it is not defined in terms of too many immeasurable
parameters. Often a lack of agreement between the model and practice can be
caused by an incorrect choice of parameter values. This can even lead to quite
different trends being observed in the variation of particular parameters during
the simulation.
It should be noted, however, that often the results of a simulation model do
not have to give an exact fit to the experimental data, and often it is sufficient to
simply have a qualitative agreement. Thus a very useful qualitative
understanding of the process and its natural cause-and-effect relationships is
obtained.
1.1 Fundamentals of Modelling 15

1.1.4 Simulation Tools

Many different digital simulation software packages are available on the market
for PC and Mac application. Modern tools are numerically powerful, highly
interactive and allow sophisticated types of graphical and numerical output.
Most packages also allow optimisation and parameter estimation. BERKELEY
MADONNA is very user-friendly and very fast. We have chosen it for use in
this book, and details can be found in the Appendix. With it data fitting and
optimisation can be done very easily. MODELMAKER is also a more recent,
powerful and easy to use program, which also allows optimisation and
parameter estimation. ACSL-OPTIMIZE has quite a long history of
application in the control field, and also for chemical reaction engineering.
MATLAB-SIMULINK is a popular and powerful software for dynamic
simulation and includes many powerful algorithms for non-linear optimisation,
which can also be applied for parameter estimation.

1.1.5 Teaching Applications

For effective teaching, the introduction of computer simulation methods into


modelling courses can be achieved in various ways, and the method chosen will
depend largely on how much time can be devoted, both inside and outside the
classroom. The most time-consuming method for the student is to assign
modelling problems to be solved outside the classroom on any available
computer. If scheduling time allows, computer laboratory sessions are
effective, with the student working either alone or in groups of up to three on
each monitor or computer. This requires the availability of many computers,
but has the advantage that pre-programmed examples, as found in this text, can
be used to emphasize particular points related to a previous theoretical
presentation. This method has been found to be particularly effective when
used for short, continuing-education, professional courses. By use of the
computer examples the student may vary parameters interactively and make
program alterations, as well as working through the suggested exercises at his or
her own pace. Demonstration of a particular simulation problem via a single
personal computer and video projector is also an effective way of conveying
the basic ideas in a short period of time, since students can still be very active in
suggesting parametric changes and in anticipating the results. The best
approach is probably to combine all three methods.
16 1 Modelling Principles

1.2 Development and Meaning of Dynamic


Differential Balances

As indicated in Section 1.1, many models for biological systems are expressed
in terms of sets of differential equations, which arise mainly as a result of the
predominantly time-dependent nature of the process phenomena concerned.
For many people and especially for many students in the life sciences, the
mention of differential equations can cause substantial difficulty. This section
is therefore intended, hopefully, to bring the question of differential equations
into perspective. The differential equations arise in the model formulation,
simply by having to express rates of change of material, due to flow effects or
chemical and biological reaction effects. The method for solution of the
differential equations will be handled automatically by the computer. It is
hoped that much of the difficulty can be overcome by considering the
following case. In this section a simple example, based on the filling of a tank
of water, is used to develop the derivation of a mass balance equation from the
basic physical model and thereby to give meaning to the terms in the equations.
Following the detailed derivation, a short-cut method based on rates is given to
derive the dynamic balance equations.
Consider a tank into which water is flowing at a constant rate F (m3/s), as
shown in Fig. 1.2. At any time t, the volume of water in the tank is V (m3) and
the density of water is p (kg/m3).

Figure 1.2. Tank of water being filled by stream with flow rate F.

During the time interval At (s), a mass of water p F At (kg) flows into the tank.
As long as no water leaves the tank, the mass of water in the tank will increase
by a quantity p F At, causing a corresponding increase in volume, AV.
Equating the accumulation of mass in the tank to the mass that entered the tank
during the time interval A t gives,
pAV = p F A t
Since p is constant,
-Fh
At -
1.2 Development and Meaning of Dynamic Differential Equations 17

Applying this to very small differential time intervals (At > dt) and replacing
the A signs by the differential operator "d", gives the following simple first
order differential equation, to describe the tank filling operation,

dV
dT = F

What do we know about the solution of this equation? That is, how does the
volume change with time or in model terms, how does the dependent variable,
V, change with respect to the independent variable, t? To answer this, we can
rearrange the equation and integrate it between appropriate limits to give,

t
o
or for constant F,
= F Jtf l l dt= F(ti-to)
o

Integration is equivalent to summing all the contributions, such that the total
change of volume is equal to the total volume of water added to the tank,

IV = IF At

For the case of constant F, it is clear that the analytical solution to the
differential equation is,
V = F t + constant

In this case, as shown in Fig. 1.3, the constant of integration is the initial volume
of water in the tank, VQ, at time t = 0.

dt
Vo

Figure 1.3. Volume change with time for constant flow rate.
18 1 Modelling Principles

Note that the slope in the variation of V with respect to t, dV/dt, is constant, and
that from the differential equation it can be seen that the slope is equal to F.
Suppose F is not constant but varies linearly with time.

F = Fo-kt

The above model equation applies also to this situation.


Solving the model equation to obtain the functional dependence of V with
respect to t,
JdV = |F dt = J(F0 - k t)dt = FO Jdt - k Jt dt

Integrating analytically,
kt"
V = FO t - + constant
The solution is,

v = FO t - kt'
+ V0

."t
Figure 1.4. Variation of F and V for the tank-filling problem.

Note that the dependent variable starts at the initial condition, (Vo), and that the
slope is always F. When F becomes zero, the slope of the curve relating V and t
also becomes zero. In other words, the volume in the tank remains constant
and does not change any further as long as the value of F remains zero.

Derivation of a Balance Equation Using Rates

A differential balance can best be derived directly in terms of rates of change.


For the above example, the balance can then be expressed as:

/The rate of accumulation^ =


/The flow rate of mass^
V of mass within the tank ) \ entering the tank )
1.2 Development and Meaning of Dynamic Differential Equations 19

Thus, the rate of accumulation of mass within the tank can be written directly as
dM/dt where the mass M is equal to p V. The rate of mass entering the tank is
given by p F, where both sides of the equation have units of kg/h.

=
PF
and
d(oV)
= pF

Thus this approach leads directly to a differential equation model, which is the
desired form for dynamic simulation. Note that both terms in the above
relationship are expressed in mass quantities per unit time or kg/h.

At constant density, the equation again reduces to,

dV
dT = F

which is to be solved for the initial condition, that at time t = 0, V=Vo and for a
variation in flow rate, given by,

F = F0-kt
which is valid until F = 0.

These two equations, plus the initial condition, form the mathematical
representation or the mathematical model of the physical model, represented
by the tank filling with an entering flow of water. Thus this approach leads
directly to a differential equation model, which is the desired form for
simulation. This approach can be applied not only to the total mass but also to
the mass of any component.
We have seen an analytical solution to this model, but it is also interesting to
consider how a computer solution can be obtained by a numerical integration
of the model equations. This is important since analytical integration is seldom
possible in the case of real complex problems.

Computer Solution

The numerical integration can in principle be performed using the relations:

dV
20 1 Modelling Principles

where t - to represents a very small time interval and V - VQ is the resulting


change in volume of the water in the tank. As before, the flow is assumed to
decrease with time according to F = FQ - k t.
This integration procedure is equivalent to the following steps:
1) Setting the integration time interval.
2) Assigning a value to the inlet water flow rate at the initial value, time
t = to-
3) The term involving the water flow rate, F-kt, is equal to the derivative
value, (dV/dt), at time t = to.
4) Knowing the initial value of V and the slope dV/dt, enables a new value of
V to be calculated over the small interval of time, equivalent to the
integration time interval or integration step length.
5) At the end of the integration time interval, the value of V will have
changed to a new value, representing the change of V with respect to time
from its original value. The new value of V can thus be calculated.
6) Using the new value of V, a new value for the rate of change of V with
respect to time, (dV/dt), at the end of the integration time interval can
now be calculated.
7) Knowing the value of V and the value of dV/dt at the end of the
integration time interval, a new value of V can be estimated over a further
step forward in time or integration time interval.
8) The entire procedure, as represented by steps (2) to (7) in Fig. 1.5 below,
is then repeated with the calculation moving forward with respect to time,
until the value of F reaches zero. At this point the volume no longer
increases, and the resulting steady-state value of V is obtained, including
all the intermediate values of V and F, which were determined during the
course of the calculation.

Figure 1.5. Graphical portrayal of numerical integration, showing slopes and approximated
values of V at each time interval.
1.3 Formulation of Balance Equations 21

Using such a numerical integration procedure, the computer can thus be used
to generate data concerning the time variations of both F and V. In practice,
more complex numerical procedures are employed in digital simulation
languages to give improved accuracy and speed of solution than illustrated by
the above simplified integration technique.

1.3 Formulation of Balance Equations

1.3.1 Types of Mass Balance Equations

Steady-State Balances

One of the basic principles of modelling is that of the conservation of mass,


which for a steady-state flow process can be expressed by the statement,

( Rate of mass flow^


into the system J
f Rate of mass flow^j
^ out of the system J

Dynamic Total Mass Balances

Many bioreactor applications are, however, such that conditions are in fact
changing with respect to time. Under these circumstances, a steady-state mass
balance is inappropriate and must be replaced by a dynamic or unsteady-state
mass balance, which can be expressed as:

( Rate of accumulation of ^
mass in the system J
( Rate of ^
^mass flow inj
( Rate of ^
^mass flow out J

Here the rate of accumulation term represents the rate of change in the total
mass of the system, with respect to time, and at steady-state this is equal to zero.
Thus the steady-state mass balance represented earlier is seen to be a
simplification of the more general dynamic balance, involving the rate of
accumulation.
At steady-state:
22 1 Modelling Principles

( Rate of ^
= 0 = (Mass flow in) - (Mass flow out)
I accumulation of mass .

hence, when a steady-state is reached:

(Mass flow in) = (Mass flow out)

Component Balances

The previous discussion has been in terms of the total mass of the system, but
most fluid streams, encountered in practice, contain more than one chemical or
biological species. Provided no chemical change occurs, the generalized
dynamic equation for the conservation of mass can also be applied to each
component. Thus for any particular component:

Rate of
Mass flow of A Mass flow of >
accumulation of mass
of component
in the system
(the component _
into the system J (
the component out
of the system ,

Component Balances with Reaction

Where chemical or biological reactions occur, this can be taken into account by
the addition of a further reaction rate term into the generalized component
balance. Thus in the case of material produced by the reaction:

' Rate of ^ ' Mass flow ^ ^ Mass flow ^ Rate of


accumulation of the of the production
of mass = component - component + of the
of component into out of component
^ in the system, ^the system, ^the system, by the reaction>

and in the case of material consumed by the reaction:

Rate of ' Mass flow ' Mass flow Rate of


accumulation of the of the consumption
of mass = component - component - of the
of component into out of component
^in the system, v the system, ^the system^ by the reaction
1.3 Formulation of Balance Equations 23

Elemental Balances

The principle of the mass balance can also be extended to the atomic level and
applied to particular elements. Thus in the case of bioreactor operation, the
general mass balance equation can also be applied to the four main elements,
carbon, hydrogen, oxygen and nitrogen and also to other elements if relevant
to the particular problem. Thus for the case of carbon:

' Rate of accumulation^! (Mass flowrate of "\ ( Mass flow rate of \


of carbon in = carbon into - the carbon out
the system J ^ the system J V of the system )

Note the elemental balances do not involve reaction terms since the elements do
not change by reaction.
The computer example PENFERM, is based on the use of elemental mass
balance equations for C, H, O and N which, when combined with other
empirical rate data, provide a working model for a penicillin production
process.
While the principle of the mass balance is very simple, its application can
often be quite difficult. It is important therefore to have a clear understanding
of both the nature of the system (physical model), which is to be modelled by
means of the mass balance equations, and also of the methodology of
modelling.

1.3.2 Balancing Procedure

The methodology described below outlines six steps, I through VI, to establish
the model balances. The first task is to define the system by choosing the
balance or control region. This is done using the following procedure:

I. Choose the balance region such that the variables are


constant or change little within the system. Draw
boundaries around the balance region

The balance region may be a reactor, a reactor region, a single phase within a
reactor, a single cell, or a region within a cell, but will always be based on a
region of assumed constant composition. Generally the modelling exercises will
involve some prior simplification. Often the system being modelled is usually
considered to be composed of either systems of tanks (stagewise or lumped
24 1 Modelling Principles

parameter systems) or systems of tubes (differential systems), or even


combinations of tanks and tubes, as used in Case C, Sec. 1.3.2.3.

1.3.2.1 Case A. Continuous Stirred Tank Bioreactor

AO

Total mass = pV

Mass of A = C VA

Balance region

Figure 1.6. The balance region around the continuous reactor.

If the tank is well-mixed, the concentrations and density of the tank contents
are uniform throughout. This means that the outlet stream properties are
identical with the tank properties, in this case CA and p. The balance region can
therefore be taken around the whole tank.
The total mass in the system is given by the product of the volume of the
tank contents V (m3) multiplied by the density p (kg/m3), thus Vp (kg). The
mass of any component A in the tank is given as the product of V times the
concentration of A, CA (kg of A/m3 or kmol of A /m3), thus V CA (kg or kmol).

1.3.2.2 Case B. Tubular Reactor

Balance region

'AO A1

Figure 1.7. The tubular reactor concentration gradients.

In the case of tubular reactors, the concentrations of the products and reactants
will vary continuously along the length of the reactor, even when the reactor is
operating at steady-state. This variation can be regarded as being equivalent to
1.3 Formulation of Balance Equations 25

that of the time of passage of material as it flows along the reactor and is
equivalent to the time available for reaction to occur. Under steady-state
conditions the concentration at any position along the reactor will be constant
with respect to time, though not with position. This type of behavior can be
approximated by choosing the balance regions sufficiently small so that the
concentration of any component within a region can be assumed to be
approximately uniform. Thus in this case, many uniform property subsystems
(well-stirred tanks or increments of different volume but of uniform
concentration) comprise the total reactor volume.

13.2.3 Case C. River with Eddy Current

For this example, the combined principles of both the stirred tank and
differential tubular modelling approaches need to be applied. As shown in Fig.
1.8 the main flow along the river is very analogous to that of a column or
tubular process, whereas the eddy region can be approximated by the behavior
of a well-mixed tank. The interaction between the main flow of the river and
the eddy, with flow into the eddy from the river and flow out from the eddy
back into the river's main flow, must be included in any realistic model.
The real-life and rather complex behavior of the eddying flow of the river,
might thus be represented, by a series of many well-mixed subsystems (or
tanks) representing the main flow of the river. This interacts at some particular
stage of the river with a single well-mixed tank, representing the turbulent eddy.
In modelling this system by means of mass balance equations, it would be
necessary to draw boundary regions around each of the individual subsystems
representing the main river flow, sections 1 to 8 in Fig. 1.9, and also around the
tank system representing the eddy. This would lead to a very minimum of nine
River

Eddy

Figure 1.8. A complex river flow system.


26 1 Modelling Principles

component balance equations being required. The resulting model could be


used, for example, to describe the flow of a pollutant down the river in rather
simple terms.
1 1 1 1 I I 1

>
* 1 ',
i
2
2 ',
i
3
3
1

i
4
4
1

i
, 5
5
1

i
, 66
1

i
, 7
7
1

i
, 8
8 > fc C
r

>k
Flow interaction
V
f

fitSliM^B

:i
iK&^'ffK<&y^

Figure 1.9. A multi-tank model for the complex river flow system.

//. Identify the transport streams which flow across the


system boundaries

Having defined the balance regions, the next task is to identify all the relevant
inputs and outputs to the system. These may be well-defined physical flow
rates (convective streams), diffusive fluxes, and also interphase transfer rates.
It is important to assume a direction of transfer and to specify this by means of
an arrow. This direction might reverse itself, but will be accomodated by a
reversal in sign.

Out by diffusion
Convective
flow out
Convective flow
in

In by diffusion

Figure 1.10. Balance region showing convective and diffusive flows in and out.

///. Write the mass balance in word form

This is an important step because it helps to ensure that the resulting


mathematical equation will have an understandable physical meaning. Just
1.3 Formulation of Balance Equations 27

starting off by writing down equations is often liable to lead to fundamental


errors, at least on the part of the beginner. All balance equations have a basic
logic as expressed by the generalized statement of the component balance
given below, and it is very important that the mathematical equations should
retain this. Thus:

' Rate of f Mass flow ^\ f Mass flow ^ / Rate of


accumulation of the of the production
of mass component component of the
of component into out of component by
the system ) \the system^ ^the system^ \ the reaction /

This can be abbreviated as,

(Accumulation) = (In) - (Out) + (Production)

IV. Express each balance term in mathematical form with


measurable variables

A. Rate of Accumulation Term

This is given by the derivative of the mass of the system, or the mass of some
component within the system, with respect to time. Hence:

dMi
(Rate of accumulation of mass of component i within the system) =

where M is in kg or mol and time is in h, min or s.


Volume, concentration and, in the case of gaseous systems, partial pressure
are usually the measured variables. Thus for any component i

dMj _ d(CjV)
=
dt dt

where, Q is the concentration of i (kmol/m3 or kg/m3), and pi is the partial


pressure of i within the gas phase system. In the case of gases, the Ideal Gas
Law can be used to relate concentration to partial pressure and mol fraction.
Thus,
piV= niRT

where R is in units compatible with p, V, n and T.


In terms of concentration,
28 1 Modelling Principles

_ = ni
c = Pi yiP
i T RT = "RT
where yi is the mol fraction of the component in the gas phase and p is the total
pressure.
The accumulation term for the gas phase can be written as,
/piV
dMj _ d(CjV) _ d(QV) _ . d _

For the total mass of the system:

dM _ d(p V)
dt = dt
with units

m3 s

B. Convective Flow Terms

Total mass flow rates are given by the product of volumetric flow multiplied by
density. Component mass flows are given by the product of volumetric flow
rates times concentration.

Mass \
(VolumeJ
kg m 3 kg
s - s m3

Total mass flow = F p

Component mass flow MI = F Q

A stream leaving a well-mixed region, such as a well stirred tank, has the same
properties as the system volume as a whole, since for perfect mixing the
contents of the tank will have uniform properties, identical to the properties of
the fluid leaving at the outlet. Thus, the concentrations of component i both
within the tank and in the tank effluent are equal to Qi, as shown in Fig. 1.11.
1.3 Formulation of Balance Equations 29

lilf

Figure 1.11. Convective flow terms for a well-mixed tank bioreactor.

C. Diffusion of Components

As shown in Fig. 1.12, diffusional flow contributions can be expressed by


analogy to Pick's Law for molecular diffusion

Ji = -i dZ

where jt is the flux of any component i flowing across an interface (kmol/m2 h


or kg/m2 h) and dQ/dZ (kmol/m) is the concentration gradient as shown in
Fig. 1.12.

Figure 1.12. Diffusion flux j j driven by concentration gradient (Qo - Cji) / AZ through
surface area A.

In accordance with Pick's Law, diffusive flow always occurs in the direction of
decreasing concentration and at a rate proportional to the concentration
gradient. Under true conditions of molecular diffusion, the constant of
proportionality is equal to the molecular diffusivity for the system, Dj (m2/h).
For other cases, such as diffusion in porous matrices and turbulent diffusion, an
30 1 Modelling Principles

effective diffusivity value is used, which must be determined experimentally.


The concentration gradient may have to be approximated in finite difference
terms (Finite differencing techniques are described in more detail in Sec. 6.2).
Calculating the mass rate requires the area, through which diffusive transfer
occurs.

( Massrate ( Diffusivity YConcentrationY Area ^ (


of of gradient perpendicular =-DJ
(^component i (^component ij\^ ofi J^ to transport J

kg 2 _ m 2 kg 2 _ kg
sm 2 m
" s m4 m " T

D. Interphase Transport

Interphase mass transport also represents a possible flow into or out of the
system. In bioreactor modelling applications, this is most frequently
represented by the case of oxygen transfer from air to the liquid medium,
followed by oxygen taken up by the cells during respiration. In this case, the
transfer of oxygen occurs across the gas liquid interface, which exists between
the surface of the air bubbles and the surrounding liquid medium, as shown in
Fig. 1.13.

Figure 1.13. Transfer of oxygen across a gas-liquid interface of specific area "a" into a liquid
phase of volume V.

Other applications may involve the supply of oxygen to the bioreactor by


transfer from the air, across a membrane and then into the bulk liquid. Where
there is interfacial transfer from one phase to another, the component balance
equations will need appropriate modification to take this into account. Thus, an
oxygen balance for the well-mixed gas phase, with transfer from the gas to the
liquid, can be written as,
1.3 Formulation of Balance Equations 31

/ Rate of \
accumulation /Mass flow\ /Massflow\ / of interfacial
Rate \
of the of the of the mass transfer
mass of oxygen = oxygen - oxygen - from the gas
in the gas into the from the phase into
V phase system J Vgas phase ) ^ gas phase> V the liquid s

This form of transfer rate equation will be examined in much more detail in
Chapter 5. Suffice it to say here that the rate of transfer can be expressed in the
form shown below:

f Rate Of \ ( Mass ^ ^Area peA /Concentration^ /SystemA


Uass transferJ= tra"sPrtf ^ volume ) ^driving force ) VvolumeJ
V coefficient.

= KaACV

where, a is a specific area for mass transfer, A/V (m2/m3), A is the total
interfacial area for mass transfer (m2), V is the liquid phase volume (m3), AC is
the concentration driving force (kmol/m3 or kg/m3) and, K is the overall mass
transfer coefficient (1/s). Mass transfer rate expressions are usually expressed
in terms of kmol/s, and can be converted to mass flows (kg/s), if desired.
The units of the terms in the equation (with appropriate mass quantity units)
are:
kg 1 kg
mj
Production Rate

This term in the component balance equation allows for the production or
consumption of material bv
by reaction and is incorporated into the component
balance equation. Thus,

Rate of >\ / Massflow\ / Massflow\ / Rate of


accumulation of the of the production
of mass component component of the
of component into out of component by
the system ) \the system / the system / v the reaction /

Chemical production rates are often expressed on a molar basis and, as in the
case of the interfacial mass transfer rate expressions, can be easily converted to
mass flow quantities (kg/s). The production rate can then be expressed as
32 1 Modelling Principles

f Mass rate \ /^Reaction rate\


production of = rA V = ^ per volume ) (Volume of system)
^component Ay

kg __ kg
m3
s m3

Equivalent molar quantities may also be used. The quantity r^ is positive when
A is formed as product, and TA is negative when a reactant A is consumed.
The growth rate for cells can be expressed in the same manner, using the
symbol rx- Thus,

/ Mass rate of ^ /^Growth rate^


Vbiomass production^ = rx V = ^per vo lume) (Volume of system)

kg _ kg
mj
s m3

The consumption rate of substrate, r$, is often directly related to the cell growth
rate by means of a constant yield coefficient YX/S, which has the units of kg
biomass produced per kg substrate consumed. Thus,

( Ma< rate \ growth rate V 1 \


= er
U>nsumptionJ M> volume ABiomass-substrate yield) (Volume)

kg ~ kg biomass kg substrate
s m3 m =
s m3 kg biomass

V. Introduce other relationships and balances such that the


number of equations equals the number of dependent
variables

The system mass balance equations are often the most important elements of
any modelling exercise, but are themselves rarely sufficient to completely
formulate the model. Other relationships are therefore needed to supplement
the material balance relations, both to complete the model in terms of other
important aspects of behavior and to satisfy the mathematical rigor of the
modelling, such that the number of unknown variables must be equal to the
number of defining equations.
1.3 Formulation of Balance Equations 33

Examples of this type of relationships which are not based on balances, but
which nevertheless form an important part of any model are:

Reaction rates as functions of concentration, temperature, pH


Stoichiometric or yield relationships for reaction rates
Ideal gas law behavior
Physical property correlations as functions of concentration
Pressure variations as a function of flow rate
- Dynamics of measurement instruments as a function of the instrument
response time
- Equilibrium relationships (e.g., Henry's law)
- Controller equations
- Correlations of mass transfer coefficient, gas holdup volume, and
interfacial area, as functions of system physical properties and degree of
agitation or flow velocity

How these and other relationships are incorporated within the development of
particular modelling instances are shown later in the cases given throughout the
text and in the simulation examples.

VI. For additional insight with complex problems, draw


an information flow diagram

Information flow diagrams can be useful in understanding complex


interactions (Franks, 1966). They help to identify missing relationships and
provide a graphical aid to a full understanding of the interactive nature of
system. An example is given in the simulation example BATFERM.

1.3.3 Total Mass Balances

In this section the application of the total mass balance principles will be
presented. Consider some arbitrary balance region, as shown in Fig. 1.14 by
the shaded area. Mass accumulates within the system at a rate dM/dt, owing to
the competing effects of a convective flow input (mass flow rate in) and an
output stream (mass flow rate out).
34 1 Modelling Principles

Mass flow rate out

Mass flow rate In

Figure 1.14. Balancing the total mass of an arbitrary system.

The total mass balance is expressed by,

( Mass flow \ fMass flow out\


=
U .he system) - Uhe system J

dM
= Mass rate in - Mass rate out

or in terms of volumetric flow rates, F, densities, (p), and volume, V

d(p V) system = F
3t oPo-FiPi
When densities are equal, as in the case of water flowing in and out of a tank,
dV
dT = F O-FI
The steady-state condition of constant volume in the tank (dV/dt = 0) occurs
when the volumetric flow in, FQ, is exactly balanced by the volumetric flow out,
FI. Total mass balances therefore are mostly important for those bioreactor
modelling situations in which volumes are subject to change.

1.3.4 Component Balances for Reacting Systems

Each chemical species can be described with a component balance around an


arbitrary, well-mixed, balance region, as shown in Fig. 1.15.
1.3 Formulation of Balance Equations 35

Species i
Species i outflow
inflow

Figure 1.15. Component balancing for species i.

Thus for any species i, involved in the system, the component mass balance is
given by:

/ Rate of \ 'Mass flow of^ /^MassflowoA f Rate of N


accumulation component i component i production of
of mass =
into out of component!
of component i ^ the system ) \ by reaction
i in tnp cvct<=m i v the system ^ }

Expressed in terms of volume, volumetric flow rate and concentration, this is


equivalent to:
~ = (F 0 C i0 )-(F 1 C il )+(r i V)

with units of mass/time:

,3*6.
m~
_ m^ m3
ill 3 _

m-

1.3.4.1 Case A. Constant Volume Continuous Stirred Tank


Reactor

A constant volume, continuous, tank reactor with reaction A > 2B is


considered here, as shown in Fig. 1.16.
36 1 Modelling Principles

F
FQ C AO C BO 1 C A 1 C B1

>f

'.; '.??: 5: ' ' :; f'.<$ 'XI ' ;:?-'?:;':- ''M

I! i ll
ilili
;|i

Figure 1.16. Continuous stirred tank reactor with reaction A > 2B.

Component A is converted to component B in a 1 to 2 molar ratio.


The component balances for A and B are:
d(VC A1 )
dt = F0 CAO -
d(VC B i)
dt = F0CBo -

Here it is convenient to use molar masses, such that each term has the units of
kmol/h.
Under constant volume conditions:

d(VCA) = VdC A

d(VCB) = VdC B

and in addition FQ = FI. Thus the two model equations, then simplify to give:

dCAi F
= - CAI) +
V
and
dCBi =
F
~dT~ V ( c BO~C B i)

In these two balances there are four unknowns CAI, Q*i, rAl an(^ rB l 8
kinetics are assumed to be first order, as often found in biological systems at
low concentration. Then:
r
Al = -
According to the molar stoichiometry,
1.3 Formulation of Balance Equations 37

rfil = -2r M = + 2 k C A !

Together with the kinetic relations there are 4 equations and 4 unknowns, thus
satisfying the conditions necessary for the model solution. With the initial
conditions, CAI and CBI at time t = 0, specified, the solution to these two
simultaneous equations, combined with the two kinetic relations, will give the
resulting changes of concentrations CAI and CBI as functions of time. The
simulation example ENZCON, is similar to the situation of Case A.

1.3.4.2 Case B. Semi-continuous Reactor with Volume


Change

The chemical reaction and reaction rate data are the same as in the preceding
example, but now the reactor has no effluent stream. The operation of the
reactor is therefore semi-continuous.

AO

*
2B

Figure 1.17. A semi-continuous reactor example.

The kinetics are as before:


t ^ moles
rA = -kC A T
m s
In terms of moles the stoichiometry gives,

rB = - 2 rA = + 2 k CA

The component balances with no flow of material leaving the reactor are now:

= FC AO + r A V

d(V C )
atB
= IB v
38 1 Modelling Principles

The number of unknowns is now five and the number of equations is four, so
that an additional defining relationship is required for solution. Note that V
must remain within the differential, because the volume of the reactor contents
is now also a variable and must be determined by a total mass balance.
Assuming constant density p, this gives the defining equation as:

= FF
dt

With initial conditions for the initial molar quantities of A and B, (VGA,
and the initial volume of the contents, V, at time t = 0 specified, the resulting
system of equations can be solved to obtain the time varying quantities VCA(t),
VCs(t), V(t) and hence also concentrations CA and CB as functions of time.
Similar variable volume situations are found in examples FEDBAT, and
VARVOL.

1.3.4.3 Case C. Steady-State Oxygen Balancing in


Fermentation

Calculation of the oxygen uptake rate, OUR, by means of a steady-state oxygen


balance is an important application of component balancing for fermentation.
In the reactor of Fig. 1.1.8, the entering air stream flow rate, oxygen
concentration, temperature and pressure conditions are shown by the subscript
0 and the exit conditions by the subscript 1.
F
Gas > i>yi> T i> PI

Air
F
0'yO'T0'P0

Figure 1.18. Entering air and exit gas during the continuous aeration of a bioreactor.

Writing a balance around the combined gas and liquid phases in the reactor
gives,

f Rate of accum-^j_ f Flowrate^i ( Flowrate"\ /^Rate of O2 uptake A


( ulationofO 2 J~ [ofO 2 in J~(ofO 2 out J~ I by the cells J
1.3 Formulation of Balance Equations 39

At steady-state, the accumulation terms for both phases are zero and

Flow of O2 in - Flow of C>2 out = Rate of C>2 uptake.

For gaseous systems, the quantities are often expressed in terms of molar
quantities.
Often only the inlet air flow rate FQ and the mol fraction of 62 in the outlet
gas, yi9 are measured. It is often assumed that the total molar flow rate of gas is
constant. This is a valid assumption as long as the number of carbon dioxide
mols produced is nearly equal to the number of oxygen mols consumed or if
the amounts of oxygen consumed are very small, relative to the total flow of
gas.
Converting to molar quantities, using the Ideal Gas Law,
pV = nRT
or in flow terms:
pF = N R T

where N is the molar flow rate, R is the gas constant and F is the volumetric flow
rate. Thus, for the inlet gas flow:
P

where NO is molar flow rate of the oxygen entering. Note that the pressure, po,
and temperature, TO, are measured at the point of flow measurement.
Assuming NO = NI, then measurement of NO gives enough information to
calculate oxygen uptake rate, OUR, from the steady-state balance. Thus,

0 = yo NO - yi NI - ro2 VL

OUR = ro2 VL = yo NO - yi NI

If NO is not equal to NI, then this equation will give large errors in oxygen
uptake rate, and NI must be measured, or determined indirectly by an inert
balance. This is explained in the Sec. 1.3.4.4 below.

1.3.4.4 Case D. Inert Gas Balance to Calculate Flow Rates

Differences in the inlet and outlet gas flow rates of a tank fermenter can be
calculated by measuring one gas flow rate and the mol fraction of an inert gas
in the gas stream. Since inert gases, such as nitrogen or argon, are not
consumed or produced within the system (rinert = 0), their mass rates must
40 1 Modelling Principles

therefore be equal at the inlet and outlet streams of the reactor, assuming
steady-state conditions apply. Then for nitrogen

/Molar flow of\ /Molar flow of\


=
V nitrogen in ) ^ nitrogen out /

and in terms of mol fractions,

NO YO inert = NI yi inert

From this balance, calculation of NI can be made on the basis of a combination


of measurements of NO and the inert gas partial pressures (yinertX at both inlet
and outlet conditions.
N,1 = y i inert

Since the inlet mol fraction for nitrogen in air is known, the outlet mol fraction,
yi inert' must be measured. This is often done by difference, having measured
the mol fraction of oxygen and carbon dioxide concentration in the exit gas.

1.3.5 Stoichiometry, Elemental Balancing and the


Yield Coefficient Concept

Stoichiometry is the basis for any quantitative treatment of chemical and


biochemical reactions. In biochemical processes it is a necessary basis for
building kinetic models.

1.3.5.1 Simple Stoichiometry

The Stoichiometry of chemical reactions is used to relate the relative quantities


of the different materials which react with one another and also the relative
quantities of product that are formed. Most chemical and biochemical reactions
are relatively simple in terms of their molar relationship or Stoichiometry. For
single reactions stoichiometric coefficients are clearly defined and may usually
easily be determined. Some examples are given below:

C3H4O3 + NADH + H+ < C3H6O3 + NAD+


Pyruvic Acid Lactic Acid
1.3 Formulation of Balance Equations 41

This relation indicates that 1 mol of pyruvic acid reacts with 1 mol of NADH to
produce 1 mol of lactic acid.
Another example of stoichiometry is that of the oxidative decarboxylation of
pyruvic acid to yield acetyl-CoA
C3H4O3 + CoA-SH + NAD+ -> CH3CO-S-CoA + CO2 + NADH + H+
Pyruvic Acid Acetyl-CoA

Stoichiometry relations also describe more complex pathways and can be


written with exact molar relationships, like the pentose-phosphate pathway
below.

Glucose + 12 NADP+ + ATP + 7 H2O -> 6 CO2 + 12 (NADPH + H+) +


+ ADP + Pi

where 1 mol of glucose reacted, consumes 7 mol of water and produces 6 mol
of carbon dioxide. Here the molar quantities of NADPH and ATP produced
and consumed, respectively, are shown.
For many complex biological reactions, however, not all the elementary
reactions and their contributions to the overall observed reaction stoichiometry
are known (Roels, 1983; Bailey and Ollis, 1986; Moser, 1988).
Thus the case of a general fermentation is usually approximated by an
overall reaction equation, where
Substrate + Nitrogen source + O2 -> Product + CO2 + H2O

v N H3(t)NH 3 + v02(t)O2 >

VC02(t) CO2 + VH2o(0 H2O

where the i-th product, such as metabolites or biomass, is given by a general


formula.
In the case above, the generalized elemental formulae are used for substrate,
biomass and products, but the nitrogen source is given simply as ammonia. The
stoichiometric coefficients, v, for each component are taken relative to that of
substrate and their coefficients may vary as a function of time as a result of
changing fermentation conditions. Some indication as to the relative
magnitudes of the stoichiometric coefficients can be obtained from elemental
balancing techniques, but in general the problem is so complex that other
concepts, such as the more approximate yield coefficient concept, are used to
relate the relative proportions of materials undergoing conversion during the
fermentation.
42 1 Modelling Principles

1.3.5.2 Elemental Balancing

The technique of elemental balancing can be represented as follows:


Taking the general case of

CHmOi + a NH3 + b 02 > c CHpOnNq + d CHrOsNt + e H2O + f CO2


[substrate] [biomass] [product]

where c, d and f are the fractions of carbon converted to biomass, product and
CO2, respectively.
Elemental balances for C, H, O and N give

C 1 = c +d +f
H m+3 a = c p + dr + 2e
O l+2b =cn + ds + e + 2f
N a =cq+dt

In this general problem there are too many unknowns for the solution method
to be taken further, since the elemental balances provide only four equations
and hence can be solved for only four unknowns. Assuming that the elemental
formulae for substrate, biomass and product and hence 1, m, n, p, q, r, s and t are
defined, there still remain six unknown stoichiometric coefficients a, b, c, d, e
and f and only four elemental balance equations. Thus the elemental balances
need supplementation by other measurable quantities such as substrate, oxygen
and ammonia consumption rates (assuming controlled pH conditions), and
carbon dioxide or biomass production rates, such that the condition is satisfied
that the number of unknowns is equal to the number of defining equations. In
principle the problem then becomes solvable. In practice, there can be
considerable difficulties and inaccuracies involved, although the technique of
elemental balancing can still provide useful data. The application of so-called
macroscopic principles (Roels, 1980, 1982 and 1983; Heijnen and Roels, 1981)
introduces a more strict systematic system of analysis. This is depicted in Fig.
1.19.
1.3 Formulation of Balance Equations 43

<P2 Substrate

H
C32 b2 Oc2 Nd2

04 N Source

C34 Hb4 Oc4 Nd4

Figure 1.19. Flow inputs into a system.

The system is represented here in terms of the various flow inputs, where f is the
corresponding flow vector

() = <]> <1> O O O <E> <5

The steady state balance for the system is then represented by: <|) E = 0
where E is the elemental composition matrix

E = a 4 b4 c 4 d 4 O4
0 0 2 0 O5
1 0 2 0 O6
0 2 1 0 <D7

(C) (H) (O) (N)

The combination of 7 unknown quantities and 4 elemental balance equations)


leave 3 quantities are independent. Thus assuming fluxes <E>1 (biomass), O2
(substrate) and <J>3 (product) are known, the unknown fluxes 04, 05, Og and
<I>7 can be obtained by methods of linear algebra and which are detailed by
Roels (1983).
44 1 Modelling Principles

In more complex cases with growth and product formation, more


information is needed. The introduction of the concept of the degree of
reduction is useful. For organic compounds this is defined as the number of
equivalent available electrons per gram atom C, that would be transferred to
CC>2, H2O and NH3 upon oxidation. Taking charge numbers: C = 4, H = 1 , O =
-2, and N = -3, reductance degrees (y) can be defined for

substrate (S) ys = 4 + m - 2 1
biomass (X) yx = 4 + p - 2 n - 3 q
product (P) yp = 4 + r - 2 s - 3 t

The reductances for NHs, f^O and CC>2 are of course zero.
Often the elemental composition of the substrate is not known and then the
reductance method may be supplemented by the following regularities, which
apply to a wide variety of organic molecules.

Qo2 = 27 J per g equivalent of available electrons transferred to oxygen


Yx = 4.29 g equivalent of available electrons per equivalent 1 g atom C in
biomass
GX = 0.462 g carbon / g dry biomass

1.3.5.3 Mass Yield Coefficients

Yield coefficients are biological variables, which are used to relate the ratio
between various consumption and production rates of mass and energy. They
are typically assumed to be time-independent and are calculated on an overall
basis. This concept should not be confused with the overall yield of a reaction
or a process. The biomass yield coefficient on substrate (Yx/s) is defined as:

v
Y
x/S = rs

In batch systems, reaction rates are equal to accumulation rates, and therefore
/dX\
Y
IdTj dX
X/S = - TdST = - dS"
IdTj

After integration from time 0 to time t the integral value is obtained:


1.3 Formulation of Balance Equations 45

.. = amount of biomass produced


* x/s total amount of substrate consumed

X(t)-X(t=0)
S(t=0)-S(t)

For a steady-state continuous system the mass balances give

rs
SQ-S!
where index 0 and 1 indicate feed and effluent values, respectively.
In the literature, yield coefficients for biomass with respect to nutrients are
most often used (e.g. Dekkers, 1983; Mou and Cooney, 1983; Roels, 1983;
Moser, 1988). In many cases this is very useful because the biomass
composition is quite uniform, and often product selectivity does not change
very much during an experiment involving exponential growth and associated
production. Some useful typical values are given in Table 1.1.

1.3.5.4 Energy Yield Coefficients

Energy yield coefficients may be defined similarly to mass yield coefficients.


In terms of oxygen uptake,

TO amount of heat released


Yq/02 = = 7
TQ2 amount or oxygen consumed

In terms of carbon substrate consumed,

v - rQ - amount of heat released


rs amount of carbon source consumed
46 1 Modelling Principles

Table 1.1. Typical mass and energy yield values (Roels, 1983; Atkinson and
Mavituna, 1991).

Type of yield coefficient Dimension Value

Y
X/S,aer c-mol / c-mol 0.4-0.7
Yx/S,anaer c-mol / c-mol 0.1-0.2
Yx/02 (Glucose) c-mol / mol 1-2
YX/ATP c-mol / mol 0.35
Y
Q/O2 kJ / mol 380-490
YQ/C02 kJ / mol 460
YQ/x,aer (Glucose) kJ / c-mol 325-500
Yq/x,anaer kJ / c-mol 120-190

Note: The molecular weight of biomass is taken here as 24.6 g/C-mol


The yield coefficients are usually determined as a result of a large number of
elementary biochemical reactions and it can easily be understood that their
values might vary depending on environmental and operating conditions.

A detailed description of some of these dependencies is given in the literature.


Despite this inconsistency, measured yield coefficients are often very useful for
practical purposes of process description and modelling.

1.3.6 Equilibrium Relationships

1.3.6.1 General Considerations

In many biological systems, processes with large ranges of time constants have
to be described. Usually it is important to start with a simplification of a system,
focusing on the most important time constant or rate. For example, if the
growth of an organism is to be modelled with a time constant of the order of
hours, it is very useful to ignore all aspects of biological evolution with time
constants of years. Also fast equilibrium reactions or conformational changes
of proteins having time constants below milliseconds should be ignored. Fast
reactions can, however, be very important when considering allosteric activation
or deactivation of proteins or simply pH-changes during biochemical reactions.
1.3 Formulation of Balance Equations 47

pH changes can have dramatic effects on the enzyme and microbial activity but
can also strongly influence absorption and desorption of carbon dioxide.
A typical equilibrium reactions is the dissociation of a receptor-protein ligand
complex, RL, into the free protein, L, and the receptor protein, P

This reaction is characterized by the corresponding dissociation equilibrium


constant KD

CLP k_j
In most cases such relationships can be used to express the concentration of all
concentrations in explicit form using, e.g. a protein balance.

C Ptot r1 _i_ c*
~ ^P ~r ^LP

^ ^ KD
- ^ptot Cr + K
D

The total concentration, Cptot, is then included in a material balance equation


and the concentrations of the free receptor and the receptor-ligand complex are
determined by the equilibrium relationship. This is also true for a simple acid-
base equilibrium relationship.
In more complex cases with interactions of various receptors or with a buffer
system containing several components, it is not possible to express the
concentrations in explicit forms and a non-linear algebraic equation has to be
solved during the simulation. The implementation of such problems into
BerkeleyMadonna is shown below with the example of pH calculation

1.3.6.2 Case A. Calculation of pH with an Ion Charge


Balance.

Modelling systems with variable pH requires modelling of acid-base equilibria,


whose reactions are almost instantaneous. Production of acids or bases causes a
variation of pH, which depends on the buffer capacity of the system. pH also
influences the biological kinetics. It has been shown that only the undissociated
acid forms are kinetically important substrates in anaerobic systems. The
48 1 Modelling Principles

concentration of these species is a function of the pH as can be seen in the


equilibrium equation

Acid Base' + H+

with dissociation constant


CBase- H+
CAcid

where CAcid is the concentration of the undissociated acid and CBase" is the
concentration of the corresponding base (salt).
An ion charge balance can be written

(cations * charge) = (anions * charge)

In the pH range of interest (usually around pH = 7) all strong acids and strong
bases are completely dissociated. Moderately strong acids and bases exist in
both the dissociated and non-dissociated forms,
In the usual pH range the sum of the cations are much larger than the H+
ions.
ICK+CH+
where ]CK+ is the total cation concentration.
Negative ions originate mainly from strong acids (e.g. Cl% SO42') but also
arise from weak acids (Ac", Pr, Bu~, HCO3'). The concentration of CO32' is
always much smaller than that of
The ion balance reduces to

KBj C C
V1 KAi
KW Btot,i+ K+ =

where KAI are the acid dissociation constants (e.g. KAC); KBI are the base
dissociation constants (e.g. KNHS); KW is the dissociation constant of water;
Cfitot,i are the total concentrations of base i; CAtot,i are the total concentrations
of acid i and EC An" is the sum of the anions.
The pH can be estimated from the above equation for any situation by
solving the resulting non-linear implicit algebraic equation, provided the total
concentrations of the weak acids, CAtot,i> weak bases, CBtot,i cations of strong
bases, CK+, and anions of strong acids, CAIT> are known.
1.3 Formulation of Balance Equations 49

It is convenient to use only the difference between cations and anions

After neglecting any ammonia buffering effect, it is useful to rearrange the


above equations in the form,

The example ANAMEAS, Sec. 8.8.6 includes this ion balance for pH
calculation. This equation represents an algebraic loop in a dynamic simulation
which is solved by iteration at each time interval until 8 approaches zero. This is
accomplished with the root-finding feature of Berkeley Madonna.
If there is pH control, then strong base or acid would be usually added. The
addition of strong alkali for pH control would cause an increase in CK+
which in accordance with the above equation would result in a decrease of CH+.
An alternative approach, which avoids an algebraic loop, is to treat the
instantaneous equilibrium reactions as reactions with finite forward and
backward rates. These rates must be adjusted with their kinetic constants to
maintain the equilibrium for the particular system; that is, these rates must be
very fast compared with the other rates of the model. This approach replaces
the algebraic loop iteration with a stiff er and larger set of differential equations.
This could be an advantage in some cases.

1.3.7 Energy Balancing for Bioreactors

Energy balances are needed whenever temperature changes are important, as


caused by reaction heating effects or by cooling and heating for temperature
control. For example, such a balance is needed when the heat of fermentation
causes a variation in bioreactor temperature. Energy balances are written
following the same set of rules as given above for mass balances in Sec. 1.3.
Thus the general form is as follows:

Accumu-^ Rate of^ 'Rate of ^ 'Rate of" 'Rate of > 'Rate of >
lation energy energy energy energy energy
rate of in by out by out by generated added by
^Energy , ^flow , flow <transfer> ^by reactiony ^agitation ^
50 1 Modelling Principles

The above balance in word form is now applied to the measurable energy
quantities of the continuous reactor shown in Fig. 1.20.
AL H
agit

\
M)' PO P1
' '

U,A,T S

l!

Figure 1.20. A continuous tank fermenter showing only the energy-related variables.

An exact derivation of the energy balance was given by Aris (1989) as,

S
"dT = - h n ) ) + U A (Ta - TI) + rQ V + AHagit
agi

where ni is the number of moles of component i, cpi are the partial molar heat
capacities and hi are the partial molar enthalpies. In this equation the rate of
heat production, TQ, takes place at temperature TI. If the heat capacities, cpi, are
independent of temperature, the enthalpies at TI can be expressed in terms of
heat capacities as
hn = hio + Cpi (TI -TO)
and with
S S
2>i0 - 2X = vp
1=1 1=1
Thus with these simplifications,

Vpc p L = F0 p cp (T0 - TI) + U A (Ta - TI) +rQ V + AHagi

The units of each term of the equation are energy per time (kJ/h or kcal/h).
1.3 Formulation of Balance Equations 51

Accumulation Term
Densities and heat capacities of liquids can be taken as essentially constant.

dT
VpcPdF
has units:
m3 (kg/m3) (J/kg K) K _ kJ
s ~ s

Here (p cp T) is an energy "concentration" term and has the units,

/jnass\ / energy \ _ /energy \


Vvolumey Vmass degree^ VaegreeJ - Vyolume,/

Thus the accumulation term has the units of energy/time (e.g. J/s)

Flow Terms
The flow term is F p CP (T0 - TI)

with the units, /energy


i^nn^J\ /volume\
\ctisr) = /energy \
This term actually describes heating of the stream entering the system with TO
to the reaction temperature TI. It is important to note here that this term is
exactly the same for a continuous reactor as for a fed-batch system.

Heat Transfer Term


The important quantities in this term are the heat transfer area A, the
temperature driving force or difference (Ta-Ti), where Ta is the temperature of
the heating or cooling source, and the overall heat transfer coefficient, U. The
heat transfer coefficient, U, has units of energy/time area degree, e.g. J/s m2 C.
The units for U A AT are thus,

(heat transfer rate) = U A (Ta - TI)

energy =
energy
~B55~ area time degree (area) (degree)

The sign of the temperature difference determines the direction of heat flow.
Here if T a > TI heat flows into the reactor.

Reaction Heat Term


The term rq V gives the rate of heat released by the bioreaction and has the
units of
52 1 Modelling Principles

energy _ energy
volume time ( volume ) - time

The rate term TQ can alternatively be written in various ways as follows:


In terms of substrate uptake and a substrate-related heat yield,

rq = rs YQ/S

In terms of oxygen uptake and an oxygen-related heat yield,

rq = ro2YQ/Q2

In terms of a heat of reaction per mol of substrate and a substrate uptake rate,

rQ = AHr,s rs

Here rs is the substrate uptake rate and AHr?s is the heat of reaction for the
substrate, for example J/mol or kcal/kg. The rs AHr>s term therefore has
dimensions of (energy/time volume) and is equal to TQ.

Other Heat Terms


The heat of agitation may be the most important heat effect for slow growing
cultures, particularly with viscous cultures. Other terms, such as heat losses
from the reactor due to evaporation, can also be important.

1.3.6.3 Case B. Determining Heat Transfer Area or Cooling


Water Temperature

For aerobic fermentation, the heats of reaction per unit volume of reactor are
usually directly related to the oxygen uptake rate, ro2-
Thus for a constant-volume batch reaction with no agitation heat effects, the
general energy balance is

/Accumulation rate^ /Energy out^ +


/Energy generated\
V of energy ) ~ ~ \ by transfer J V by reaction )

where YQ/Q2 often has a value near 460 kJ/mol 2, as given in Table 1.1.
1.3 Formulation of Balance Equations 53

If T is constant (dT/dt = 0):

UA(Ti-Ta) = r 0 2YQ/o 2 V

(heat transfer rate) = (rate of heat release)

Using this steady-state energy balance, it is possible to calculate the cooling


water temperature (Ta) for a given oxygen uptake rate and cooling device.
Thus,
_-ro 2
UA

Alternatively this same relation can be used in other ways:

1) To calculate the additional heat transfer area required for a known


increase in cooling water temperature.
2) To calculate the biomass concentration allowable for a given cooling
system, knowing the specific oxygen uptake rate (kg O2 / kg biomass h).
3) To calculate the cooling area required for a continuous fermenter with
known volume inlet, temperature, flow rate and biomass production rate.
2 Basic Bioreactor Concepts

2.1 Information for Bioreactor Modelling

Both physical and biological information are required in the design and
interpretation of biological reactor performance, as indicated in Fig. 2.1.
Physical factors that affect the general hydrodynamic environment of the
bioreactor include such parameters as liquid flow pattern and circulation time,
air distribution efficiency and gas holdup volume, oxygen mass transfer rates,
intensity of mixing and the effects of shear. These factors are affected by the
bioreactor geometry and that of the agitator (agitator speed, effect of baffles)
and by physical property effects, such as liquid viscosity and interfacial tension.
Both can have a large effect on gas bubble size and a corresponding effect on
both liquid and gas phase hydrodynamics. The biokinetic input involves such
factors as cell growth rate, cell productivity and substrate uptake rate. Often this
information may come from laboratory data, obtained under conditions which
are often far removed from those actually existing in the large scale bioreactor.
Although shown as separate inputs in Fig. 2.1, there are, in fact, considerable
interactions between the bioreactor hydrodynamic conditions and the cell
biokinetics, morphology and physiology, and one of the arts of modelling is to
make proper allowance for such effects. Thus in the large scale bioreactor,
some cells may suffer local starvation of essential nutrients owing to a
combination of long liquid circulation time and an inadequate rate of nutrient
supply, caused by inadequate mixing or inefficient mass transfer. Agitation and
shear effects can affect cell morphology and hence liquid viscosity, which will
also vary with cell density. This means that the processes of cell growth affect
the bioreactor hydrodynamics in a very complex and interactive manner.
Changes in the cell physiology, such that the cell processes are switched from
production of further biomass to that of a secondary metabolite or product, can
also be affected by selective limitation on the quantity and rate of supply of
some essential nutrient in the medium. This can in turn be influenced by the
bioreactor hydrodynamics and also by the mode of the operation of the
bioreactor.
The overall problem is therefore very complex, but as seen in Figure 2.1,
when all the information is combined successfully in a realistic and well
founded Bioreactor Model, the results obtained can be quite impressive and

Biological Reaction Engineering, Second Edition, I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
56 2 Basic Bioreactor Concepts

may enable such factors as cell and product production rates, product
selectivities, optimum process control and process optimization to be
determined with some considerable degree of confidence.

Physical Aspects Biokinetics


(flow patterns, residence (order, inhibition,pH,
time, mass transfer) temperature)

Production rate
Selectivity
Control

Figure 2.1. Information for bioreactor modelling.

2.2 Bioreactor Operation

The rates of cell growth and product formation are, in the main, dependent on
the concentration levels of nutrients and products within the bioreactor. The
concentration dependencies of the reaction or production rate are often quite
simple, but may also be very complex. The magnitude of the rates, however,
depend upon the level of concentrations, and it will be seen that concentration
levels within the bioreactor depend very much on its type and mode of
operation. Differing modes of operation for the bioreactor can therefore lead
to differing rates of cell growth, to differing rates of product formation and
hence to substantially differing productivities.
Generally, the various types of bioreactor can be classified as either stirred
tank or tubular and column devices and according to the mode of operation as
batch, semi-continuous or continuous operation.
2.2 Bioreactor Operation 57

2.2.1 Batch Operation

Most industrial bioreactors are operated under batch conditions. In this, the
bioreactor is first charged with medium, inoculated with cells, and the cells are
allowed to grow for a sufficient time, such that the cells achieve the required
cell density or optimum product concentrations. The bioreactor contents are
discharged, and the bioreactor is prepared for a fresh charge of medium.
Operation is thus characterized by three periods of time: the filling period, the
cell growth and cell production period and the final emptying period as
depicted in Fig. 2.2. It is only during the reacting period, that the bioreactor is
productive. During the period of cell growth, strictly speaking, no additional
material is either added to or removed from the bioreactor, apart from minor
adjustments needed for control of pH or foam, small additions of essential
precursors, the removal of samples and, of course, a continuous supply of air
needed for aerobic fermentation. Concentrations of biomass, cell nutrients and
cell products thus change continuously with respect to time, as the various
constituents are either produced or consumed during the time course of the
fermentation, as seen in Fig. 2.3.

Filling Reacting Emptying Cleaning

Figure 2.2. Periods of operation for batch reactors.

concentration
A biomass
ubstrate

product

time

Figure 2.3. Concentration-time profiles during batchwise operation.


58 2 Basic Bioreactor Concepts

During the reaction period, there are changes in substrate and product
concentration with time, and the other time periods are effectively lost as
regards production.
Since there is no flow in or out of the bioreactor, during normal operation,
the biomass and substrate balances both take the form,

(Rate of accumulation within the reactor) = (Rate of production)

This will be expressed in more quantitative terms in Ch. 4.

Batch reactors thus have the following characteristics:

1) Time-variant reaction conditions


2) Discontinuous production
3) Downtime for cleaning and filling

2.2.2 Semicontinuous or Fed Batch Operation

In semi-continuous or fed batch operation, additional substrate is fed into the


bioreactor, thus prolonging operation by providing an additional continuous
supply of nutrients to the cells. No material is removed from the reactor, apart
from normal sampling, and therefore the total quantity of material within the
reactor will increase as a function of time. However if the feed is highly
concentrated, then the reactor volume will not change much and can be
regarded as essentially constant.

Figure 2.4. Fed batch bioreactor configuration.


2.2 Bioreactor Operation 59

Semi-continuous operation shares the same characteristics as pure batch


operation, in that concentration levels generally change with time and that some
downtime occurs during the initial charging and final discharge period at the
end of the process.
The ability to manipulate concentration levels within the bioreactor by an
appropriate controlled feeding strategy confers a high degree of flexibility to
fed batch or semi-continuous operation, since differing concentration levels can
be utilized to manipulate the rates of reaction. In Fig. 2.4, both the volumetric
feeding rate, F, and the feed substrate concentration SQ, may be constant or may
vary with time, giving the possibility of such feeding strategies as:

1. Slow constant feeding, which can be shown to result in linear growth


of the total cell biomass.

2. Exponential feeding to maintain constant substrate concentration and,


resulting in unlimited, exponential cell growth.

3. Feedback control of the feed rate, based on monitoring some key


component concentration.

The important characteristics of fed batch operation are therefore as follows:

1. Extension of batch growth or product production by additional


substrate feeding.

2. Possibility of operating with separate conditions for growth and


production phases.

3. Control possibilities on feeding policies.

4. Development of high biomass and product concentration.

For fed-batch operation, the cell balance follows the same form as for batch
operation, but since additional substrate feeding to the reactor now occurs, the
substrate balance takes the form:
Rate
( <* "| ( Substrate \ ( Substrate >|
accumulation = (f^d [n) _ consumption
V of substrate J \ rate )

Under controlled conditions, in which the substrate concentration is maintained


constant or kept small, the accumulation term in the above equation will also be
small, with the result that the feed rate of substrate into the reactor will balance
the rate of consumption by reaction.
60 2 Basic Bioreactor Concepts

One other balance equation, however, is also necessary, i.e. the total mass
balance,

f Rate of accumulation of ^ ( Mass flow rate of feed ^


=
V mass in the reactor / V to the reactor )

which for constant density conditions reduces to

(Rate of change of volume) = (Volumetric rate of feeding)

Further extensions of fed batch operation are possible, such as the cyclic or
repeated fed batch, which involves changing volume with a filling and
emptying period. The changing reactor concentrations repeat themselves with
each cycle. This operation has similarities with continuous operation and
approaches most closely to continuous operation, when the amount withdrawn
is small and the cycle time is short. The simulation examples FEDBAT, Sec.
8.1.3 and in Sec. 8.3 (VARVOL, PENFERM, PENOXY, ETHFERM, REPFED)
allow detailed investigations of fed batch performance to be made on the
computer.

2.2.3 Continuous Operation

In continuous operation fresh medium is added continuously to the bioreactor,


while at the same time depleted medium is continuously removed. The rates of
addition and removal are such that the volume of the reactor contents is
maintained constant. The depleted material, of course, contains any products
that have been excreted by the cells and, in the case of suspended-cell culture,
also contains effluent cells from the bioreactor.
Continuous reactors are of two main types, as indicated in Fig. 2.5, and these
may be considered either as discrete stages, as in the continuous, stirred-tank
bioreactor, or as differential devices, as represented by the continuous tubular
or column reactor.

Continuous tank bioreactor Continuous tubular bioreactor

Figure 2.5. The two main types of continuous reactors.


2.2 Bioreactor Operation 61

As shown later, these two differing forms of continuous reactor operation have
quite different operational characteristics. Both however are characterized by
the fact that after a short transient period, during which conditions within the
bioreactor change with time, the bioreactor will then achieve a steady state. This
means that operating conditions, both within the bioreactor and at the
bioreactor outlet, then remain constant, as shown in Fig. 2.6.

Startup Steady state


Concentration period

time

Figure 2.6. Startup of a continuous reactor.

Continuous reactors, however, have found little use as biological reactors on a


production scale, although there are a few important examples (Id's single-cell
protein air lift process, wastewater treatment and the isomerization of corn sugar
to fructose syrup). Frequent use is made of continuous reactors in the
laboratory for studying the kinetics of organism growth and for enzyme
reaction kinetics. This is because the resulting form of the balance equation,
leads to an easy method for the determination of reaction rate, as discussed in
Ch. 4.
The behavior of the two differing forms of continuous reactor, are best
characterized by their typical concentration profiles, as shown in Fig. 2.7. In
this case, S is the concentration of any given reactant consumed, and P is the
concentration of any given product.

Tank Tube
So So

Cone. Cone.

distance distance

Figure 2.7. Profiles of substrate and product in steady state continuous tank and tubular
reactors.

As seen, the concentrations in a perfectly mixed tank are uniform, throughout


the whole of the reaction vessel contents and are therefore identical to the
concentration of the effluent stream. In a tubular reactor the reactant
concentration varies continuously, falling from a high value at the inlet to the
62 2 Basic Bioreactor Concepts

lowest concentration at the reactor outlet. The product concentration rises from
inlet to outlet. These differences arise because in the tank reactor the entering
feed is continuously being mixed with the reactor bulk contents and therefore
being diluted by the tank contents. The feed to the tubular reactor, however, is
not subject to mixing and is transformed only by reaction, as material moves
down the reactor.
No real situation will exactly correspond to the above idealized cases of
perfect mixing or zero mixing (plug flow), although the actual behavior of
tanks and tubes tends in the limit towards the corresponding idealized model.
The characteristics of continuous operation are as follows:

1. Steady state after an initial start-up period (usually)


2. No variation of concentrations with time
3. Constant reaction rate
4. Ease of balancing to determine kinetics
5. No down-time for cleaning, filling, etc.

The balance equations at steady state for a well-mixed tank reactor have the
form
0 = (Input) - (Output) + (Production)

since at steady-state the rate of accumulation and therefore the rate of change is
zero.
This equation predicts that the reaction rate causes a depletion of substrate
from the feed condition to the outlet, (the product will increase) and that the
rate of production can be obtained from this simple balance:

(Rate of production) = (Rate of output) - (Rate of input)

For a non well-mixed reactor such as a tubular or column reactor, steady-state


implies the same non-transient conditions, but now concentrations also vary
with position. The same situation also applies to the case of a series of well-
mixed tanks.
The balance form is then:
0 = (Rate of input) - (Rate of output) + (Overall Rate of Production)

Here the overall rate of reaction is obtained by summing or integrating over


every part of the reactor volume.
The concentration characteristics of a tubular reactor, as shown in Fig. 2.7,
are well approximated by a series of tank reactors. Referring to Fig. 2.8, and
moving downstream along the reactor cascade, the substrate concentration
decreases stepwise from tank to tank, while the product concentration increases
in a similar stepwise manner. As the number of tanks in the cascade increases,
so the performance becomes more and more similar to that of a tubular reactor.
In the case of a reaction, whose rate of reaction increases with increasing
2.2 Bioreactor Operation 63

substrate concentration S, the multiple tank configuration or a tubular reactor


would thus have a kinetic advantage over that of a single tank. The same is true,
in the case of product inhibition kinetics, in which the rate would be lowered by
high product concentration, P. Substrate inhibition systems would be run
preferably in single tanks, however, since then the substrate concentration is
always at its lowest value.

Cone.

distance

Figure 2.8. Stirred tanks in series and their concentration profiles.

A calculation of the tank volume or residence time requirement involves the


formulation of the tank balance equations, as before and then the application of
the equations, successively from tank to tank such that the effluent from the
preceding tank is the feed of the next and so on. Tanks-in-series bioreactor
operations are illustrated by the simulation examples TWOSTAGE, STAGED
and DEACTENZ in Sec. 8.4.

2.2.4 Summary and Comparison

The operating characteristics of the various reactor modes are summarized in


Table 2.1.
The important bioreactor operating parameters will depend on the mode of
operation. In batch operation, concentration levels can be varied by adjustment
of the initial values, whereas in continuous and semi-continuous operation, the
concentration levels depend on the feed rate and feed concentration. As
indicated previously, the manner in which the bioreactor is operated can
therefore give rise to different concentration levels and therefore differing
productivities. The consequent concentration profiles depend, of course, on the
reaction kinetics, which express the rate of reaction as a function of the
concentrations of reactants and products.
64 2 Basic Bioreactor Concepts

Table 2.1. Summary of reactor modes.


Mode of operation Advantages Disadvantages

Batch Equipment simple. Suitable Downtime for loading and


for small production. cleaning. Reaction
conditions change with
time.

Continuous Provides high production. Requires flow control.


Better product quality due Culture may be unstable
to constant conditions. over long periods.
Good for kinetic studies.

Fed batch Control of environmental Requires feeding strategy to


conditions, e.g. substrate obtain desired
concentration. concentrations.

Table 2.2 lists the main operating parameters for the three differing modes of
bioreactor operation.

Table 2.2. Operating variables for batch and continuous bioreactors.


Batch Continuous Semicontinuous

Initial medium composition Inlet medium Feed and initial substrate


and inoculum composition composition

Temperature, pressure Temperature, pressure Temperature, pressure

pH if controlled pH if controlled pH if controlled

Reaction time Liquid flow rate Liquid flow rate


(residence time) (residence time)
Aeration rate
Aeration rate Feeding rate and control
Stirring rate program
Stirring rate
Aeration rate

Stirring rate
2.2 Bioreactor Operation 65

The foregoing discussion of the varying characteristics of the different reactor


types and their concentration profiles allows a qualitative comparison of the
volume requirements for the different types of reaction, according to the
particular kinetics. For this it is first necessary to consider the qualitative nature
of the basic forms of kinetic relationship: zero order, first order, product and
substrate inhibition. The detailed quantitative treatment of these kinetic forms is
dealt with in Ch. 3.
The rate of a zero order reaction is independent of concentration. Many
bioreactions at high substrate concentrations follow zero order kinetics and are
therefore insensitive to concentration and to the effects of concentration
gradients. From the kinetic viewpoint, therefore, any reactor type would be
equally suitable.
First order reaction rates are directly proportional to concentration.
Bioreactions at low concentration are generally first order, and this would favor
operation in either a batch or a tubular/column type reactor. This is because
reactant concentrations in such reactors are generally high overall and hence
the overall rates of reaction are also consequently high. Hence the reactor
volume required for a given duty would generally be small. (In the case of a
batch reactor, this of course neglects the time lost for filling, emptying and
cleaning.)
A reaction with substrate inhibition would be best run in a tank at low
substrate concentration, since the concentration would be low throughout the
whole of the tank contents. Conversely, product inhibition would be more
pronounced in tank reactors, since product concentration would be at its
highest. In this case, a tubular type reactor or batch reactor would be preferred.

Table 2.3. Kinetic considerations for reactor choice.


Continuous
Reaction Batch Tank Tanks-in- Continuous Fed Batch
Kinetics Series or Single Tank
Tubular
Zero order OK OK OK Low con-
version
First order Best Best Low con- OK
version only
Substrate Low initial Low tank Best Best
inhibition concentration concentrations

Product Best Best Low con- Low con-


inhibition version only version only

Production OK for temp- Possible Not suitable Best for con-


triggered by erature-shift centration-
shift in shift
environment
66 2 Basic Bioreactor Concepts

Table 2.3 compares the performance of batch tanks, continuous tubular or


tanks-in-series reactors and single continuous tank reactors. As discussed in
Sec. 4.2.1, batch tank concentration-time profiles are exactly analogous to the
steady state concentration-distance profiles obtained in continuous tubular
reactors. In terms of performance, therefore, the batch reactor would be the
same as a tube, when compared on the basis of equal batch time in the tank and
time of passage through the tube. Tanks-in-series reactors, as shown in Fig. 2.8,
involve step wise gradients, which in the limit are very similar to those of
continuous tubular reactors, hence, making their performance similar to that of
a tubular reactor. Owing to the high degree of mixing which leads to a uniform
concentration, the performance of the single continuous stirred tank reactor is
very different to that of the other reactor types. An exact quantitative
comparison can be made using the mass balance equations developed in Ch. 4
for each reactor type.
Biological Kinetics

As explained in Sec. 2.1, a realistic bioprocess model will usually require the
input of kinetic rate data. In the case of even simple chemical reactions, this
data has to be obtained by laboratory experiment. Since biochemical reactions
are controlled by enzymes, it is appropriate to start with a consideration of
simple enzyme kinetics (Sec. 3.1), In the case of modeling the behavior of
enzyme reactors, knowledge of the enzyme reaction kinetics is most important.
The sheer complexity of the biological reactions, occurring in a living cell,
seem to imply an almost impossible task in obtaining meaningful rate data for
biological modelling applications. Fortunately this is not the case and, as
shown in Section 3.2, a quite reasonable overall description of cell growth rate
data is possible, based on an overall empirical relationship, the Monod
Equation, which has been found to give a good fit to many general
observations of cell growth. This overall view, based on the net result of many
simultaneously occurring and highly interacting biochemical reactions, of
course represents an incredible oversimplification of the actual situation.
Fortunately it seems to work in many instances and can also be easily modified
to allow for the uptake of substrate by the cells and to include such additional
effects, as substrate limitation, multiple substrate limitation and product
inhibition. It is interesting, that the basic enzyme rate equation, or Michaelis-
Menten equation, based on the theory indicated in Sec. 3.1, is of the same basic
form as the empirically-based Monod equation for the growth of micro-
organisms.
When used in this manner, the cell kinetics are completely devoid of any
mechanistic interpretation and constitute what is known as an unstructured
kinetic model (Fig.3.1 A). In other cases, it may be necessary to look in some
detail at individual cell processes and reactions, in order to obtain a more
realistic description, thus leading on to the use of structured kinetic models
(Fig. 3.IB) as described in Sec. 3.3. In a most simple case, the cells are
composed of a catalytic part comprising proteins, RNA, DNA and other cellular
compounds and of a storage part, e.g. poly-hydroxy-alkanoic acids (PHAs) or
inclusion bodies of recombinant proteins. A most simple segregated model
considers different stages of cells and therefore a distribution of cell stages in a
culture without structuring the cell composition (Fig. 3.1C). In the most
realistic, but most complex situation the model is structured and segregated

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
68 3 Biological Kinetics

(Fig. 3.ID). For the purpose of this book, the differences of these models can
be best described by their different balance regions.

Non-structured Structured

B
3CO

t
o

Figure 3.1. Types of kinetic models for cells. Balance regions: A Total cell biomass, B Cell
parts, C biomass parts, D Biomass and cell parts.

3.1 Enzyme Kinetics

3.1.1 Michaelis-Menten Equation

The rate of reaction catalyzed by a soluble enzyme can usually be described by


the Michaelis-Menten kinetic equation. This equation can be derived from the
accepted Briggs-Haldane mechanism for a simple enzyme reaction, which is
very similar to that for conventional chemical heterogeneous catalysis.
Thus,
3.1 Enzyme Kinetics 69

>
S 4- E < ES

ES - P + E

where E is enzyme, S is substrate, P is product and ES is an enzyme-substrate


complex.
For a batch reaction, the balances for S and ES are written in terms of the
mechanism as,
JQ

= - ki S E + k_i (ES)
dt

= ki S E - (k_i + ki) (ES)


dt
with initial conditions at t = 0:

S = S0 (ES) = 0

The concentration changes for a batch reactor are shown qualitatively in Fig.
3.2. While the enzyme concentration is usually much lower than that of the
substrate, most of the enzyme is present during the reaction in the form of the
enzyme-substrate complex, ES.
Analytical solution is then possible by assuming a quasi-steady state for the
enzyme-substrate complex, ES,
d(ES)
-= 0
dt
This assumption is valid for E SQ.
Using the total enzyme mass balance,

EO = E + (ES)

the above equations can be solved for the unknown concentrations E and ES to
give,

and
ki S E k.i + k2
E = E = E
0 ~ k_i .+ k2 k_i + k2
70 3 Biological Kinetics

o
1
Q)
O

-^ time

Figure 3.2. Concentration changes of the reaction species for a simple enzymatic reaction
taking place in a batch reactor.

Substituting for E and ES the substrate balance becomes,

dS _ k2SEo
dt ~

giving the Michaelis-Menten equation,

where the parameters in terms of the mechanistic model are for the maximum
reaction rate (kmol/m3min):

v max = k 2 E 0

and for the Michaelis-Menten constant (kmol/m3):

k-i + k2
KM - ki

The Michaelis-Menten equation exhibits three distinct regions for the reaction
rate. At very low and very high substrate concentrations the rs versus S curve is
essentially linear, as seen in Fig. 3.3.
3.1 Enzyme Kinetics 71

vm

rS

Michaehs-Menten region

vm/2""

0 KM 5 KM 10 KM 15 KM

Figure 3.3. Reaction rate versus substrate concentration for the Michaelis-Menten equation.

The low concentration region can be approximated by first-order kinetics. The


Michaelis-Menten equation becomes for S KM,

_ VmaxS

For high substrate concentration (S KM) the relation approaches zero-order,


r
S = v max

and the rate of reaction is thus independent of substrate concentration and is


constant at the maximum value.
In the intermediate substrate concentration range, 0.1 KM < S < 10 KM, the
full Michaelis-Menten equation must be used to guarantee an accuracy for rs
greater than 10 %. The parameters vmax and KM can be determined from
experimental data, either graphically following a linearization of the Michaelis-
Menten equation or, better, numerically by using nonlinear parameter
estimation techniques.
Graphical determination of KM and vmax is based on rearrangement of the
Michaelis-Menten equation into a linear form,

Inversion and rearrangement give,


72 3 Biological Kinetics

1 KM I
rs v
max S max

A graphical representation of this equation is called the Lineweaver-Burk


diagram (Fig. 3.3) from which the kinetic parameters vmax and KM may be
determined.

Figure 3.3. Lineweaver-Burk double-reciprocal plot.

Typical values of the enzyme kinetic constants are given in Table 3.1. It is
interesting to note that the rate of formation of the enzyme-substrate complex
can be extremely fast, with the constant ki approaching 1 x 1010 L/mol s. This
is the maximum value for a rate constant of a reaction that is limited by
diffusion of a small substrate molecule in aqueous solution.

Table 3.1. Typical values of the constants of the Michaelis-Menten equation.

Constant Value range

105 - 109 L/mol s


10- 104 1/s
1 - 106 1/s
KM 10-6 10-l mol/L
3.1 Enzyme Kinetics 73

The simulation example MMKINET enables a computer study of the basic


characteristics of the Michaelis-Menten equation to be carried out, and
LINEWEAV simulates the study of the Lineweaver-Burk plot for a batch
enzyme reaction.

3.1.2 Other Enzyme Kinetic Models

The reaction mechanism of enzyme catalysis can be very complex, resulting in


complicated kinetic equations that are treated in specialized textbooks, as given
in the reference section. Some of the more readily used forms of the modified
Michaelis-Menten kinetics are presented here.

Double Michaelis-Menten Kinetics


This refers to the case when two substrates are involved in the reaction:
v
max Si 82 _
r
s - (KMi + Si) (KM2 + S2)
Inhibition
Inhibition occurs when a substance, inhibitor (I), reduces the rate of an
enzyme-catalyzed reaction, usually by the inhibitor binding to the enzyme
active site. Three simple types of reversible inhibition kinetics are given in Tab.
3.2.

Table 3.2. Enzymatic inhibition kinetics.

Mechanism Inhibition Rate equation, rs


v
I maxS
competitive KM (1 + I/KI) + S

E+SES-P
c c non-competitive (i + I/KI) (KM + S)
El ESI

ES +1 => ESI uncompetitive vmax


74 3 Biological Kinetics

Usually, the substance I is the substrate or the product, and the reaction kinetics
are known as substrate or product inhibition, respectively.

Allosteric Kinetics
A simple model to describe allosteric inhibition is given, in which the enzyme
can bind to more than one substrate molecule. Thus:

nS + E

z
ESn ) nP + E

when n is the number of substrate molecules.


The resulting kinetic expression is referred to as Hill kinetics,

Vmax S n
s
" K M n + Sn

As shown in Fig. 3.5, for values of n > 1 an S-shaped function results. KM is


the substrate concentration with r$ = vmax/2. The simulation example PHB
employs this kinetic form.

Temperature and pH Influence


Rates of biological reactions, including growth rates, exhibit a maximum when
plotted versus temperature or pH. The maximum point is referred to as the
temperature optimum or pH optimum for the system. The term temperature
optimum must be used with caution because the curve is a result of two
temperature dependent processes, the enzyme catalysed reaction and the
enzyme deactivation reaction, respectively.
3.1 Enzyme Kinetics 75

Figure 3.5. Michaelis-Menten and Hill kinetics: v max = 1; Km = 5; n = 1, 2, 3, 5.

At temperatures well below the optimum the enzyme deactivation may be


neglected and the temperature influence on the reaction rate described by the
Arrhenius equation. At higher temperatures both enzymatic reaction and
enzyme deactivation rate equations must be solved together with their respective
kinetic constants expressed in terms of the Arrhenius equation.
Most enzymes exhibit a distinct pH optimum. This can be explained by
dissociation of acidic and basic groups of the enzyme, especially of its active
center. The following equation is a useful description of this.

r =
K
S,H+CH+CH/ K I,H
76 3 Biological Kinetics

3.1.3 Deactivation

Biocatalysts in reactors usually undergo irreversible conformational changes


generally known as denaturation or deactivation. This often causes an
exponential decrease of activity with time and can be described by a first-order
reaction rate process:
rd = -kdE

Considering that for a batch reactor, dE/dt = r^, the integrated form can be
written as
E = E 0 e- kdt

Substitution in the Michaelis-Menten equation yields

= MOJL kdt
KM+S
This equation suggests an exponential decrease of reaction rate regardless of
substrate concentration. The simulation example DEACTENZ, Sec. 8.4.12
illustrates this.
Engineering models for the kinetics of deactivation are given by Prenosil et
al. (1987).

3.1.4 Sterilization
Similar to enzyme deactivation, sterilization kinetics can be viewed as a process
of inactivation or the removal of viable organisms or cells from the system.
Inactivation can be achieved by using heat, radiation or chemicals. It is a
statistical process, with the rate of killing being usually proportional to the
number of the organisms at any time. Therefore it can be described again by
first-order kinetics:
rd = - k d X

where

For a batch reactor,


dX
= - k d X =-k d =k 0 e- E a / R T X

which upon integration gives,


3.1 Enzyme Kinetics 77

x
_ ,,-kdt
where XQ is the initial live biomass concentration, X is the viable biomass after
the treatment time t, and kd is the specific deactivation constant (1/s).
The sterilization time will depend on the initial level of contamination. For
this purpose the D-value is defined as the treatment time required to reduce the
population by a factor of ten. This time is related to the rate constant by

2.3

3.2 Simple Microbial Kinetics

3.2.1 Basic Growth Kinetics

Under ideal conditions for growth, when a batch fermentation is carried out, it
can be observed experimentally that the quantity of biomass, and therefore also
the concentration, increases exponentially with respect to time. This phenomena
can be explained by the fact that all cells have the same probability to multiply.
Thus the overall rate of biomass formation is proportional to the biomass itself
This leads to an autocatalytic reaction, which is described by a first order rate
expression as

where rx is the rate of cell growth (kg cell/m3 s), X is the cell concentration
(kg cell/m3) and k is a kinetic growth constant (1/s). For a batch system, this is
equivalent to,

where dX/dt is the rate of change of cell concentration with respect to time
(kg cell/m3 s). The analytical solution of this simple, first-order differential
equation is of the form
In X = k t + In X0
or,
= ekt

where XQ is the initial cell concentration at time t = 0.


78 3 Biological Kinetics

Plotting experimental growth data in the form of the natural logarithm of cell
concentration versus time will often yield a straight line over a large portion of
the curve, as shown below in Fig. 3.6.

InS
Limitation \Stationary
InX
X Death
Exponential

Lag

time

Figure 3.6. Biomass and substrate concentrations during batch growth.

In the range from ti to t2 the logarithmic curve is linear, and this is the region
of exponential growth. Three other regions can be identified: between t = 0
and ti, there exists a period of cell adaptation or lag phase, and before t2 there
is a region where the growth is limited by the lack of a particular substance,
which is known as the limiting substrate.
The slope of the linear part of the curve between ti and t2 is the growth rate
per unit mass of cells or specific growth rate and is given the symbol \i.

1 dX
=
X" "dT = = specific growth rate = |i

In many processes cells begin to die (after ts), because of lack of nutrients,
toxic effects or cell aging. This process can typically be described by a first
order decay,
rd = - kd X

where rd is the death rate and k^ is the specific death rate, with the same
dimensions as the specific growth rate. This expression is identical with
sterilization kinetics, Sec. 3.1.4.
The exponential and limiting regions can be described by a single relation,
that sets JLI equal to a function of substrate concentration. It is observed
experimentally that |a is at a maximum when the particular limiting substrate
concentration S is large, and for low concentration ja is proportional to S. Over
the whole range from low to high S, |i is described by the following Monod
equation.
3.2 Simple Microbial Kinetics 79

Thus |i varies with S in the same fashion as does the enzymatic rate of
Michaelis-Menten kinetics.
Again, this is a two-parameter equation involving two constants, the
maximum specific growth rate |im and the saturation constant KS. It is best
considered to be an empirical relation, but since it has the same form as the
Michaelis-Menten enzyme kinetics equation, it is sometimes taken to be related
to a limiting enzymatic step. Although very simple, it often describes
experimental data for growth rates very well. The form of this relation is shown
in Fig. 3.7.
M
Monod Relation

Figure 3.7. Specific growth rate versus limiting substrate concentration according to the
Monod relation.

The important properties of this relationship are as follows:

S - 0,

Urn
S = KS, Jl = ~

The first introductory simulations in Sec. 8.1 are based on Monod kinetics.
When two substrates can be limiting, it is often the case that a double Monod
type relationship can be used, as given in Sec. 3.2.4 and as shown by the
simulation examples NITRIF, Sec. 8.5.3, and BIOFILM, Sec. 8.7.1.
80 3 Biological Kinetics

3.2.2 Substrate Inhibition of Growth

Many substrates can be utilized by organisms at low concentration, but at high


concentrations they can also act as toxic growth inhibitors. The |i versus S
curve may then appear in the form shown in Fig. 3.7, and can be described by
the relation:
_ |imS
^ " (KS + s + s2/KO
whose shape depends on the values for KS and KI.
This is a modified Monod relation to allow for the inhibitory effects of high
substrate concentration. As shown in Fig. 3.8, the inhibition term (S2/Ki),
which is small in magnitude at low values of S, increases dramatically at high
values of S and causes a decrease in \i. Note that high values of KI correspond
to a decreasing effect of substrate inhibition. It is seen that larger values of KS
shift the left side of the curves towards the right, while increasing values of KI
raise the right side of the curves. Thus a wide range of shapes can be achieved
by varying the three parameters, but a maximum value of (I is always obtained
at some intermediate value of S.

1.0 -,

0.8 -
K
0.6 -
I

0.0

Figure 3.7. Substrate inhibition kinetics for various values of KS and Kj. The parameters used
are as follows: For all curves |im = 1.0 1/h. Curve A: KS = 1 and KI = 10, Curve B: KS = 0.1 and
KI = 10; Curve C: KS = 1 and KI = 20; Curve D: KS = 0.1 and KI = 20. The units of KS, KI and S
are g/m^.
3.2 Simple Microbial Kinetics 81

The substrate inhibition kinetic curve has the following characteristics, which
result from the kinetic equation:

1) When S = Ks
_
- - 2 + Ks/Kj
2) When S = KI

- 2 + Ks/Ki

3) The maximum occurs at S = (Ks Ki)-5 and

Mm
M- =
2 (Ks/Ki)-5 + 1

3.2.3 Product Inhibition

When the formation of product inhibits the rate of cell growth, the basic Monod
equation can be modified, by the addition of a product inhibition term P/Kj.
Thus,

3.2.4 Other Expressions for Specific Growth Rate

The Modified Monod Form


MmS
M- - KS S0 + S

shows the influence of initial concentration, which is sometimes observed if


other components are limiting.

The Teisser Equation


ILL = |Li m (l-e-S/k)

relates |Li to S exponentially.


82 3 Biological Kinetics

The Contois Equation

- KX + S

expresses the effective saturation constant as being proportional to the biomass


concentration X. At high X, |i is inversely proportional to X. This is
sometimes used to represent a diffusion limitation in flocculating or
immobilized biomass.

The Logistic Equation


M, = ( a - b X )

encompasses exponential growth and the levelling off to zero growth rate at
high X. For a batch fermentation the biomass balance is,

= aX-bX2

Thus when X is small, growth is exponential and given by

dX

When X is large,

At steady state or zero growth rate,

0 = a X - b X2
and thus
X = a/b

Multiple-Substrate Monod kinetics


can be used to describe the influence of many substrates, which for two
substrates takes the form,
Si ^ f S2 A
l + SiJ ^2 + 827

In this way either substrate may be limiting under conditions when the other
substrate is in excess. Note that the multiplicative effect gives for S\ = K\ and
3.2 Simple Microbial Kinetics 83

82 = K2, ILL = |LLm/4. An example of such kinetics is the simultaneous requirement


of glucose and oxygen by aerobically growing organisms.

Double-Monod kinetics
can also be written for two substrates as parallel reactions, according to

lSi k2S2 V 1

This form gives an additive, fractional contribution for each substrate. Thus for
Si = KI and 82 = K2, the result is |Li = |Lim/2. For the case Si = 0 and 82 large,
then ji = |iim k2/(ki+k2). Each substrate thus allows a different maximal growth
rate. If both Si and 82 are large then |ii = |iim. Note that the flexibility of this
kinetic form requires twice as many kinetic parameters as the simpler double
Monod kinetics. An example of this kinetics is the parallel use of alternative
substrates, such as various types of sugars.

Diauxic Monod Growth can be modelled for two substrates by the relation

K 2 + S2 + S / K !
in this way the consumption of substrate 82 will be inhibited until Si is
exhausted, for suitably low values of Kj. Diauxic growth can be observed in
many organisms. An example is E. Coli, where the uptake of lactose is
repressed in the presence of glucose. The simulation example SUBTILIS, Sec.
8.9.2 uses this kinetic form.

3.2.5 Substrate Uptake Kinetics

The rate of uptake of substrate by micro-organisms is generally considered to


be either related to that of growth or to that required for cell maintenance. This
can be expressed as:
rs =
~rx v
?5 " m X

where rs is the rate of substrate uptake by the cells (kg substrate/m3 s). As
explained in Section 1.3.5, YX/S is the stoichiometric factor or yield coefficient
and relates mass cells/mass substrate.
84 3 Biological Kinetics

The maintenance factor, m, gives the (mass substrate/mass cells time)


required for non-growth functions. The total substrate utilization for cell
maintenance is, of course, taken to be proportional to the total quantity of cells,
and therefore for a batch reactor it is proportional to cell concentration, X.
Often the uptake and production rates of substances are expressed in terms
of the particular quantities related to unit mass of cells and are then known as
specific cell quantities. Thus:

For the specific growth rate (1/h)


rx
= X

For the specific substrate uptake rate (kg S/kg biomass h),

rs
qs = x

For the specific oxygen uptake rate (kg C>2/kg biomass h),
r
O2
qo2 =
For the specific carbon dioxide uptake rate (kg CCVkg biomass h),

qco2 = "IT"
For the specific product production rate (kg P/kg biomass h),

qp =

Note that qx = rx/X = |JL is the specific biomass production rate.

Specific rate quantities may take simple or complicated forms, for example, for
the case of substrate:
rs = y^" - m X
then,

qs =
where |i is also a function of S.
3.2 Simple Microbial Kinetics 85

By necessity, in wastewater treatment systems the substrate concentration, S, is


taken often as total dissolved organic carbon, rather than considering a specific
substance, such as glucose. The biomass concentration, X, also must be related
to the total of all microbial species present. Naturally a gross simplification of
such a complex system results.
In wastewater treatment systems, biomass growth is immeasurably slow,
whereas the substrate uptake can usually be measured fairly accurately. Under
such circumstances it is then more useful to base the kinetics on the more
measurable rate and to express r$ as a separate rate equation, which is
independent of rx-
For example, this can be done using an expression, analogous in form to that
of the Monod equation,
VmS

where the constant vm is proportional to the quantity of biomass in the system


and is the maximum rate of substrate consumption, observed at high S.

3.2.6 Product Formation

The kinetics of product formation can be exceedingly complex. Product is


sometimes formed during growth and sometimes after all growth has stopped.
A useful equation for the rate of product formation, flexible enough to find
frequent application, is that of the Luedeking-Piret model:

where rp is the rate of product formation (kg product/m3 s). YX/P is a yield
factor (kg cell produced/kg product produced), relating the growth and product
stoichiometry in the "growth associated" term of the Luedeking-Piret equation,
and b is a non-growth related term and is important for cultures which produce
product independent of growth. Often both coefficients of the above equation
(YX/P and b) are not constant but are functions of substrate or product
concentration.
When little is known about the detailed kinetics of product formation, a more
general expression rp = qp X is used, where qp will usually vary with culture
conditions and concentrations.
86 3 Biological Kinetics

3.2.7 Interacting Microorganisms

Multiple-organism populations will involve interactions in which one species of


organisms will exert some influence on another. Such interactions and their
models are described below. Considering two microbial species, A and B, three
types of interactions on each other are possible; a positive beneficial effect (+),
a negative detrimental effect (-), or a neutral effect (0). The resulting
interaction possibilities are given in Table 3.3.
The different types of interactions can also be described by a graphical
representation. Thus the growth kinetics can be described by a solid arrow
connecting the substrate to the product, where the organism involved is shown
above the arrow. A solid arrow from one substrate symbol to the same symbol
in another growth path indicates that the product from one organism acts as a
substrate for another. Substrate or product inhibition is indicated by a dashed
arrow connecting the component to the inhibited organism. The symbols +, -
or 0 at the right hand side of the diagram indicate whether or not the organism
involved has benefited by the interaction. This is made clear by the examples
below:

Table 3.3. Definition of microbial interaction types.

Interaction-type Organisms
A B
Neutralism 0 0
Commensalism 0 +
Mutualism
Predator-prey
Predator-prey
Amensalism 0
Amensalism 0
Competition

Predator-Prey Kinetics
Organism A consumes substrate S, and organism B consumes organism A.

The simulation example MIXPOP demonstrates this type of system.


3.2 Simple Microbial Kinetics 87

Commensalism
Organism A uses substrate 82 to produce product P; organism B uses substrate
Si to produce product 82, which benefits organism A since product 82 acts as
its substrate.
A

81

The following processes with the compound 82, shown in brackets, involve this
form of commensalism:
- nitrification (NC>2~ )
- anaerobic digestion (organic acids)
- methanogenation (H2, 62) as found in simulation example ANAMEAS,
Sec. 8.8.7.

Commensalism with Product Removed


Organism A utilizes a substrate 82, which inhibits the growth of B on substrate
Si.
s 2 - - -> p 2 0

This effect may be found in the removal of toxic wastes in mixed cultures with
multiple carbon sources. An example is found in ANAMEAS in which the
hydrogen substrate of the methanogens (A) inhibits the acetogenic organisms
(B).

Mutualism with Product Removed


Organism A utilizes substrates 82 to produce product P. Organism B utilizes
substrate S\ to produce 82, which inhibits organism B.

An example of this would be found in anaerobic digestion for hydrogen gas


(see ANAMEAS, Sec. 8.8.7)
88 3 Biological Kinetics

Mutualism with Products Used Mutually as Substrates


Both organisms benefit from each other's products.

3.2.7.1 Case A. Modelling of Mutualism Kinetics

Organism A utilizes the product from organism B, which also helps B because
PB inhibits its growth.

An example of kinetic modelling is presented for this case, in which the growth
of the two organisms, A and B, takes place in a batch reactor with initial
substrate concentrations SIQ and 820- The growth rate is expressed by Monod-
type kinetics and constant yield factors are used to express the substrate uptake
and product formation rates.

Substrate Si balance,
_
dt -

Substrate 82 balance.
_
dt -
BS2

Product PB balance,
dPB
IT A PB
Species A balance,
dXA
"3f" =

Species B balance,
3.2 Simple Microbial Kinetics 89

dXB =
~3T ^B XB
The kinetics are given by Monod-type relations, with a double form of Monod
equation employed for species A and a product inhibition term employed for
B,
Si PB
HA - MmA KI + Si

B = K2 + S2 + (PB/KI)
Other examples of interacting microorganism effects are given in the
simulation examples ACTNITR (neutralism), Sec. 8.4.3, COMPASM
(competitive assimilation and commensalism), Sec. 8.8.2, MIXPOP (predator-
prey population dynamics), Sec. 8.8.4 and TWOONE (competition between
organisms), Sec. 8.8.5.

3.2.7.2 Case B. Kinetics of Anaerobic Degradation

Anaerobic degradation is a very complex multi-substrate, multi-organism


process, as is depicted in Fig. 3.9. It is shown here how its modelling can be
approached using Monod-type kinetics. This problem is of interest because
overloading of an anaerobic reactor causes accumulation of intermediates
(organic acids, hydrogen) and consequent inhibition of the final methanogenic
step (Gujer and Zehnder, 1983).
In a first step, polymer materials (carbohydrates, proteins or lipids) are
hydrolyzed to yield the monomer compounds (amino acids, sugars and fatty
acids). In a second acidogenic step, these compounds are fermented to organic
acids (mainly acetic, propionic and butyric acid) and hydrogen. In the third
acetogenic step, organic acids with more than three atoms of carbon per
molecule are converted to acetic acid and hydrogen. In a last methanogenic
step, the intermediates, acetic acid and hydrogen and carbon dioxide are
converted to methane. Five different groups of organisms accomplish these
reactions as shown in Fig. 3.9.
According to thermodynamic calculations (Archer, 1983) the oxidation of
propionic to acetic acid should only be possible at a hydrogen partial pressure
f PH2 < 10~4 bar. From a series of observations it seems evident that
disturbances in the methanogenic step, which is generally considered to be the
most sensitive one, lead to the accumulation of H2- Additionally the state of an
anaerobic reactor may be characterized by its volatile fatty acid levels based on
the CH4/CO2 ratio and the total gas production rate.
90 3 Biological Kinetics

(0
'55

CO
g0)
TJ
"5
o>

W
'(0
0)
0)
D)
O

I
O

Figure 3.9. Reaction scheme of anaerobic degradation. The symbols are: Poly - polymer
material (proteins, fats, hydrocarbons, etc.); XJJY - Biomass hydrolyzing Poly; Mono -
monomeric materials from hydrolysis of Poly; XAG ~acid generating biomass; HPr - propionic
acid; Pr" - propionate; Xpr - biomass growing on propionate; HBu - butyric acid; Bu" - butyrate;
XBU - biomass growing on butyrate; HAc - acetic acid; Ac" - acetate; XAC - biomass growing on
acetate; XH - biomass growing on hydrogen and carbon dioxide. Dashed arrows indicate gaseous
compounds transfer to the liquid phase. T - gaseous compounds transferred to liquid gas phase.

The respective reaction rates rj for the production of biomass Xi, for the
consumption of substrate Si and for the formation of product Pi in each step
are:
rxi = Hi Xi - kdi Xi
3.3 Structured Kinetic Models 91

Hi Xj
rpi =

where k^ is the specific death rate, including maintenance metabolism, and the
specific growth rates take the Monod form,

Mimax Si
W - KSi + Si

or its modified form in the case of substrate inhibition by acetate,

Mimax Si

These kinetics can then be combined with the mass balances as discussed in Ch.
4 for each component, Si and Pi, and for the biomass balances for each
organism type, Xi
Following this approach a model was established (Denac et al., 1988) and
combined with particular control algorithms to simulate a continuous anaerobic
digestor with feed rate control. This included a gas phase balance,
thermodynamic equilibrium constraints and acid-base equilibria using an ion
charge balance (Sec. 1.3.6.2). The simulations were used to adjust the control
parameters, which were employed on laboratory reactors (Heinzle et al., 1992).
The simulation example ANAEMEAS, Sec. 8.8.7, gives details concerning this
model.

33 Structured Kinetic Models


In many cases the characterization of biological activity by simply the total
biomass concentration is insufficient for a true model representation
(Fredrickson et al., 1970; Roels, 1983; Moser, 1988). A variation in the
biomass activity per unit biomass concentration may be caused by:
- Loss of plasmids (Imanaka and Aiba, 1981). See also the simulation
example PLASMID.
- Induction and repression of genes.
- Variation of RNA content of the cells (Harder and Roels, 1982;
Furukawaet al., 1983).
Variation of enzyme content of the cells.
Post-translational modification of proteins, e.g. phosporylation.
Signaling networks.
92 3 Biological Kinetics

Membrane transport.
- Accumulation of storage materials (Heinzle and Lafferty, 1980; Heinzle
et al., 1983). See example PHB, Sec. 8.2.4.
- Morphological changes, e.g., branching of filamentous organisms,
volume to surface ratio of yeast cells (Fig. 3.10), Furukawa et al., 1983.

Such variations in biomass activity and composition require a more complex


description of the cellular metabolism and a more structured approach to the
modelling of cell kinetics. Structured models are based on a compartmental
description of the cell mass as shown in Fig. 3.1.
In general it is very difficult experimentally to obtain sufficient mechanistic
knowledge about the cell metabolism for the development of a "realistic"
structured model. Parameter estimation may be very difficult, and the
application of complex numerical methods may easily lead to physically
meaningless results. Often the verification of even simple unstructured models
is not possible owing to experimental difficulties. This problem becomes much
more significant with increasing complexity of the model. For this reason,
structured models are seldom used for design or control. Structured models
may be useful to model transient behavior of a biological system. Such models
also may be required if a wide range of changes of environmental conditions
have to be described with one model and one set of parameters (Moes et al.,
1985 and 1986).
Changes of cellular composition as functions of steady-state growth rate are
well known. For example, in long-term experiments the cellular composition
of lipids, carbohydrates, protein, RNA and DNA in Baker's Yeast were found to
change as a function of dissolved oxygen and dilution rate (Furukawa et al.,
1983). In this work the yeast shape and specific area also changed with
dissolved oxygen concentration (Fig. 3.10).

L/d; S/V x 10 [1/m]

2.0 -
L/d

1.5

S/V
1.0

0.01 0.1 1.0 D0[g/m3]

Figure 3.10. Dissolved oxygen concentration, DO, influenced the shape of yeast in
continuous culture as given by the ratios of length/diameter (L/d) and surface/volume (S/V).
3.3 Structured Kinetic Models 93

In what follows three cases involving structured kinetics models will be


discussed briefly. Case C describes a simple, two-compartment model with
storage material. Case D gives an example of a complex, structured three-
compartment model. Here the biomass contains storage material and an
enzyme that degrades the storage material. In this example the problem of
model discrimination is discussed briefly. Case E describes the application of
ATP balancing, a method of linking stoichiometry of various pathways in
complex models.

3.3.1 Case Studies

3.3.1.1 Case C. Modelling Growth and Synthesis of Poly-B-


hydroxybutyric Acid (PHB)

Fig. 3.11 represents the process of cell growth and the synthesis of intracellular
product PHB. Residual biomass (R) is the difference between total cell dry
mass (X) and product PHB (P). Synthesis of PHB occurs with a single limiting
substrate NH4+ (S) and constant dissolved gas concentrations of H2, C>2 and
CO2 (So). Mass flows are indicated by solid lines, and regulatory mechanisms
are symbolized by dashed lines.

inhibiting

inhibiting
Figure 3.11. Schematic diagram of growth and synthesis of the intracellular product PHB (P)
with constant concentrations of dissolved gases H2, O2, and CO2 (SG) X is the total biomass
(X=R+P).
94 3 Biological Kinetics

It can be seen that the catalytically active biomass, R, is produced from both S
and SQ. During exponential growth the PHB content is constant, and thus the
rate of intracellular product formation is proportional to the rate of formation
of the residual biomass. On this basis, the basic mass balance equations for a
batch process can be formulated as shown in the simulation example PHB, Sec.
8.2.4. This model was used successfully in describing experimental batch
growth and the PHB product formation (Heinzle and Lafferty, 1980), as shown
in Fig. 3.12.

S [g/L] X,P,R [g/L] P/X [-]

3 -

2 -

1 -

Figure 3.12. Comparison of simulation results from the structured PHB model with
experimental data (Heinzle and Lafferty, 1980).

3.3.1.2 Case D. Modelling of Sustained Oscillations in


Continuous Baker's Yeast Culture

Oscillations of continuous culture of baker's yeast have been observed by a


number of researchers. An example of such sustained oscillations is shown in
Fig. 3.13 (Heinzle et al., 1983).
3.3 Structured Kinetic Models 95

15 10
10 0.5
X [g/L]
'1 DO [mg/L]
0.05
0.0
0.5

0
E [g/L] S [mg/L]
2
0

100
2.0
50 RQ
1.5
q X [mmol/h L]
1.0
0 10 20
t[h]
Figure 3.13. Oscillating profiles from a continuous culture of S. cerevisiae. (Heinzle et al.,
1983). Symbols used are X (total biomass), DO (dissolved oxygen), E (ethanol), S (glucose),
QCO2 and QO2 (specific gas reaction rates), and RQ (respiratory quotient).

One possible mechanism to explain the observed results is that a storage


material (G) having the same oxidation state as glucose (S) must cyclically be
produced and consumed. This storage material was identified as trehalose and
glycogen. Under conditions of high glucose uptake rate and high degradation

Figure 3.14. Structured model to describe Baker's Yeast oscillations.


96 3 Biological Kinetics

rate of G, ethanol, E, is accumulated in the medium and can be later oxidized


to yield biomass, R. From this and additional information on the activity of the
enzyme T, which catalyzes the degradation of G, the reaction scheme shown in
Fig. 3.14 was developed and used as a basis for model formulation.
Here R is the biomass, not including the intracellular storage material, G. The
enzyme, T, is not considered to contribute significantly to the total biomass and
was therefore neglected in the total biomass balance. The detailed kinetic model
for continuous culture is given in the simulation example YEASTOSC, Sec.
8.8.8.
Many parameters could be determined from independent experiments. Some
were taken from the literature, and some, especially those describing the
enzyme T activity had to be based on simulation results. The model leads to
oscillations (Fig. 3.15), whose existence and dependency on operating
conditions agree qualitatively with experimental observations.

(g/L)

r
10 15
Time (h) Time (h)
D=0.05 h'1

I ' ' ' ' I


10 15
Time (h)

Figure 3.15. Simulation of the Baker's yeast model (simulation example YEASTOSC, Sec.
8.8.8, showing oscillations of all the components, Q.
3.3 Structured Kinetic Models 97

3.3.1.3 Case E. Growth and Product Formation of an Oxygen-


Sensitive Bacillus subtilis Culture

This example shows how knowledge of the biochemical pathways, when


combined with experimental data, can lead to model development. In this
research an oxygen-sensitive culture was to be used for mixing studies, and it
was important to establish the kinetic model (Moes et al, 1985, 1986) in order
to describe the batch profiles as shown in Fig. 3.17.
Since it was not possible to describe the growth behavior by simple Monod-
type models, an ATP balance was used to establish the available energy for
biomass synthesis. This was possible because the biochemical pathways (Fig.
3.18) for the fermentation and their associated chemical energy production and
consumption steps were known.

Gl Ac, Bu (g/L) X (9/L)

- 3
10

- 2

- 1

Figure 3.17. Growth and product formation of Bacillus subtilis at constant DO. Gl - glucose,
X - biomass, Ac - acetoin, Bu - butanediol.

The formulation starts with a steady state ATP balance, which assumes that all
energy-producing steps are balanced by those that consume energy. The form
of this balance is as fpllows:
i A ' I' U

dt = 0 = [ qs/co2 YATF/s,co2 + qs/Ac YATP/S,AC +


u - qBu/Ac) YATP/BU - qATp/x ]X
98 3 Biological Kinetics

In this equation q$/GO2 is the rate at which sugar S is converted to CC>2 by


respirative growth. The rate of the sugar conversion to acetoin is qs/Ac- The
rates of conversion of acetoin to butanediol and the reverse reaction are given
by qAc/Bu, and QBu/Ac- Energy is required for growth, and the ATP
consumption rate is qATP/X- The corresponding ATP yields Y convert these
rates to ATP equivalents. The nomenclature at the end of this example defines
all symbols in detail.
Here knowledge of the yields of ATP for each step is important. Rate
expressions for the reaction pathways, as given in Fig. 3.18, were thus
established for each step in terms of the participating reactants. Batch mass
balances for all species were then written in terms of the individual rates of
production and consumption of each relevant component.

NADH

Cells
6CO 2

Glucose >* 2Pyruvate

NADH
ATP Acetoin ^ > Butanediol

NADH ^ NAD+ r
NAD"1"

ADP ^ ATP

Figure 3.18. Biochemical pathways for the production of acetoin and butanediol.

Using the steady-state approximation that the ATP level does not vary
significantly, allows setting the condition that dATP/dt = 0. The steady state
ATP balance is then solved for qATP/X, which is the rate of ATP available for
growth. The required yields can be calculated from the reactions given in Fig.
3.17. In these calculations it was assumed that at high oxygen concentration 1
mol NADH was converted to 3 mol ATP in the respiratory chain. At low
oxygen concentrations, the conversion equivalent was assumed to be a function
of oxygen as determined by parameter estimation, based on the experimental
data.
The glucose substrate balance can be written in terms of the rates at which
substrate was consumed for complete oxidation (qs/GO2) fr biomass
(qATP/X YX/ATP / YX/S) and for product formation (qs/Ac)'

dS ,. qATP/x YX/ATP x
X
3.3 Structured Kinetic Models 99

It was observed experimentally that the reaction stoichiometry for the


formation of metabolites was almost constant and was given by Yp/S = 0.57
mol (Ac + Bu)/mol glucose. Thus, another balance can be written for the
substrate:
_ qS/Ac X
dt Y
P/S
These two equations were used to determine dS/dt and qs/GO2-

The biomass balance is,


dX =
qATP/x YX/ATP X

The metabolite balances are,

dAc
"dt" = ( qS/Ac ~ qAc/Bu + qBu/Ac )
and,
dBu
=
~dt~ ( ^Ac/flu ~ qBu/Ac ) X

In the above balances, all specific rate terms, q, are in the units
(mol/g biomass h). All concentrations (ATP, S, Ac, Bu) are in mol/L units
except X (g/L). All yield coefficients Y are in mol/mol units except when
involving biomass, e.g. YX/S is in units of g/mol.
Empirical Monod-type kinetic relationships, not given here, were established
to calculate the rate of glucose to acetoin, qs/Ac an(i the reversible acetoin to
butanediol rates, qAc/Bu and qBu/Ac as a function of reactant concentrations for
glucose, S, for acetoin, Ac, for butanediol, Bu, and for dissolved oxygen, DO.
Additional empirical kinetic terms were needed to fit the following
experimental observations:

1) Growth stopped when S approached zero.


2) qs/Ac was limited by C>2
3) qBu/Ac was promoted by high C>2
4) qAc/Bu was promoted by low C>2
5) The influence of DO with S = 0 was negligible.

The many kinetic parameters were determined partly by direct experiments and
partly by fitting the data using a parameter estimation computer program. The
influence of oxygen was determined using data from experiments at controlled
oxygen conditions and determining the best values of the oxygen sensitive rates
by parameter estimation procedures. A simple graphical procedure then
allowed determination of the appropriate constants.
100 3 Biological Kinetics

The quantities YATP/NADH and YX/ATP are linearly dependent on each other and
could therefore not be determined from experimental data. The maximum
value of YATP/NADH was arbitrarily fixed at the maximum theoretical value of
3.0, which has a direct influence on the estimation of YX/ATP (here 5.7 g/mol).
Good agreement of the batch curves with the model at constant DO was
achieved as shown in Fig. 3.18. From these results it is seen that the model
predicts the metabolite and biomass profiles. The model was quite versatile and
reasonably accurate, considering the large differences in biomass formation at
high DO (X = 3.4 g/L) and low DO (X = 2.5 g/L), as well as the variation of the
butanediol formation at high DO (Bu = 0.2 g/L) and low DO (Bu = 2 g/L),

X(g/L) GI(g/L)

t(h)

Figure 3.19. Comparison of simulation results with the Bacillus subtilis fermentation (Moes
et al., 1986). X=biomass, Ac=acetoin, Bu=butanediol, Gl=glucose.

The simulation example SUBTILIS, Sec. 8.9.2, employs a more conventional


approach to the kinetics but also makes use of the biochemical pathways.
4 Bioreactor Modelling

4.1 General Balances for Tank-Type Biological


Reactors

Fermentation systems obey the same fundamental mass and energy balance
relationships as do chemical reaction systems, but special difficulties arise in
biological reactor modelling, owing to uncertainties in the kinetic rate
expression and the reaction stoichiometry. In what follows, material balance
equations are derived for the total mass, the mass of substrate and the cell mass
for the case of the stirred tank bioreactor system as shown in Fig. 4.1.

' X0! F0 1f F1

Figure 4.1. The variables for a tank fermenter.

In this generalized case, feed enters the reactor at a volumetric flow rate FQ, with
cell concentration, XQ, and substrate concentration, SQ. The vessel contents,
which are well-mixed are defined by volume V, substrate concentration Si and
cell concentration X\. These concentrations are identical to those of the outlet
stream, which has a volumetric flow rate FI.

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
102 4 Bioreactor Modelling

General Balance Form


As shown previously, the general balance form can be derived by setting:

(Rate of accumulation) = (Input rate) - (Output rate) + (Production rate)

and can be applied to the whole volume of the tank contents.


Expressing the balance in equation form, gives:

Total mass balance:


d(Vp)
= P(FO-FI)
Substrate balance:
d(VSi)
gj = F o S o - F i S i + r s V

Organism balance:
d(V Xi)
J J t = FO XQ - FI X j + rx V

where the units are: V (m3), p (kg/m3), F (m3/s), S (kg/m3), X (kg/m3) with rs
and rx (kg/m3 s).
The rate expressions can be simply:

= KS + S!

and using a constant yield coefficient,

-*x
fs =
Yxl

but other forms of rate equation may equally apply.

The above generalized forms of equations can be simplified to fit particular


cases of bioreactor operation.
4.1 General Balances for Tank-Type Biological Reactors 103

4.1.1 The Batch Fermenter

A batch fermenter is shown schematically in Fig. 4.2.

Jiiilii

Figure 4.2. The batch fermenter and variables.

Starting from an inoculum (X at t = 0) and an initial quantity of limiting


substrate, S at t = 0, the biomass will grow, perhaps after a short lag phase, and
consume substrate. As the substrate becomes exhausted, the growth rate will
slow and become zero when substrate is completely depleted. The above
general balances describe the particular case of a batch fermentation if V is
constant and F = 0. Thus,

Total balance:

Substrate balance:
-
dSi
V^r = r s V
Organism balance:
V
T = rxV
Suitable rate expressions for r$ and rx and the specification of the initial
conditions would complete the batch fermenter model, which describes the
exponential and limiting growth phases but not the lag phase. The simulation
example BATFERM, Sec. 8.1.1, demonstrates use of this model.
104 4 Bioreactor Modelling

4.1.2 The Chemostat

A chemostat, as shown in Fig. 4.3, normally operates with sterile feed (Xo = 0)
at constant volume steady state conditions, meaning that dV/dt = 0,
d(VSi)/dt = 0, d(VXi)/dt = 0. In addition constant density conditions can be
taken to apply

UN

Figure 4.3. The chemostat and its variables.

The total mass balance simplifies to,

o = FQ-
The dynamic component balances are then,

Substrate balance
VdSi
= F(S 0 -Si)
Cell balance
VdXi
= F(X0-Xi)

where F is the volumetric flow through the system.


These dynamic equations are used in the simulation example CHEMO, Sec.
8.1.2.
At steady state, dSi/dt = 0 and dXi/dt = 0, Hence for the substrate balance:

0 = F (S0 - Si) + rs V
4.1 General Balances for Tank-Type Biological Reactors 105

Chemostats normally operate with sterile feed, XQ = 0, and hence for the cell
balance,
0 = - F X! + rx V

Inserting the Monod-type rate expressions gives:

For the cell balance,


FXi
= rx = Ji X!
hence
F
D
H = v =

Here D is the dilution rate and is equal to 1/T, where i = V/F and is equal to the
tank mean residence time.

For the substrate balance,


]
from which:
Xi = YX/S (So-Si)

Thus the specific growth rate in a chemostat is controlled by the feed flowrate,
since [I is equal to D at steady state conditions.
Since |Li, the specific growth rate, is a function of the substrate concentration,
and |Li is also determined by dilution rate, the flow rate F then also determines
the outlet substrate concentration Si. The last equation is, of course, also
simply a statement that the quantity of cells produced is proportional to the
quantity of substrate consumed, as related by the yield factor YX/S-
The curves in Fig. 4.4 represent solutions to the steady-state chemostat
model as obtained from the simulation example CHEMOSTA, Sec. 8.4.1, with
KS = 1.0.
The variables Xi, Si, as well as the productivity DXi are plotted versus D.
Thus as flow rate is increased, D also increases and causes the steady state value
of Si to increase and the corresponding value of Xi to decrease. It is seen
when D nears |im, Xi becomes zero and Si rises to the inlet value SQ. This
corresponds to a complete removal of the cells by flow out of the tank; a
phenomenon known as "washout". Fig. 4.4 shows washout occurring for D well
below |jm, which is caused by the large value of KS. . When D is nearly zero
(low flow rates) then Si approaches zero, and Xi approaches YSo The
productivity of biomass, DXi, (kg X/m3 h) passes through a maximum rather
close to the washout region.
106 4 Bioreactor Modelling

X1,S1,DX1
10.0 T

5.0 -

0.25 1.0 D (1/h)

Figure 4.4. Variation of the steady state variables in a chemostat with Monod kinetics as a
function of dilution rate.

Chemostat applications are largely found in research laboratories. Microbial


physiology studies can be made conveniently, since the cells can be controlled
by the flow rate to grow at a particular value of specific growth rate \i. Kinetic
studies can be made by measuring the concentration of the limiting substrate
for a range of |u (=D) values, permitting the kinetics, |j = f(Si), to be
determined. The yield coefficient can be determined by steady state
measurements of substrate, biomass and product. The influence of any
substrate in the culture can be tested by adding it to the medium at various
concentrations.

4.1.3 The Fed Batch Fermenter

As shown in Fig 4.5 the outlet is zero for a fed batch fermenter, and the inlet
flow, FO, may be variable. As a result the reactor volume will change with time.
4.1 General Balances for Tank-Type Biological Reactors 107

S.F,
o1 ' o

Figure 4.5. The fed batch fermenter and its variables.

The balance equations then become for constant density,

T7
F
dT = 0

d(VSi)
dt = F0 S0 + rs V

d(V XQ
dt = rxV

Expanding the differential terms, which are products of the two variables V
and Si and V and Xi, respectively, and substituting for dV/dt = FQ gives:

VdSi
= F 0 (S 0 -Si)

The above equations are identical to those for a constant volume chemostat
reactor. It can be shown by simulation that a quasi-steady state can be reached
where dXi/dt = 0 and fi = F/V (Dunn and Mor, 1975) as seen in the Fig. 4.7.
Since V increases, therefore n must decrease, and thus the reactor moves
through a series of changing steady states for which |a = D, during which Si
and p decrease, and Xi remains constant. This is analogous to a constant
volume reactor with slowly decreasing F. These phenomena are demonstrated
by the simulation examples FEDBAT, Sec. 8.1.3, and VARVOL, Sec. 8.3.1.
108 4 Bioreactor Modelling

A fed batch fermenter, in which the inlet feed rate is very low, will not exhibit
a large increase in volume and will not reach a quasi-steady state, unless X is
very high. Assuming V to be approximately constant, the general equations
can be integrated analytically for the case of |j = constant, giving an
exponential increase in X. The constant |u condition is maintained by constant
Si, which can be obtained via exponential feeding. Another phenomenon can
be proven from these equations for the case of constant feed rate and
essentially constant V; this is the linear growth situation, where X increases
linearly with time. As shown in Fig. 4.6, the slope of the curve is related to the
feed rate and the yield coefficient. If V changes as a consequence of dilute
feed, then the total quantity of biomass (VX) will increase linearly.

FS
0 Y X/S

Figure 4.6. Linear growth under conditions of feed limitation with constant volume.

Figure 4.7. Repeated fed batch operation in terms of dimensionless variables for substrate
inhibition kinetics. Two cycles of operation are shown. The dimensionless variables are
defined in the simulation example VARVOL, Sec. 8.3.1.
4.1 General Balances for Tank-Type Biological Reactors 109

Other important types of operation described by the general balances are


cyclic fed batch (Keller and Dunn, 1978). An example of cyclic fed batch
operation in which the quasi-steady state can be observed is shown in Fig. 4.7.
Similar results can be obtained from the simulation example REPFED.

4.1.4 Biomass Productivity

The specific biomass production rate for a chemostat, DXi, (kg biomass/m3 h)
can be calculated by applying the above model equations. Thus,

D X i = DY x /s(So-Si) = DYx/s

The conditions for the maximum value of DXi as shown in Fig. 4.4, can be
obtained by setting

d(DXi)
=
dD

or by running the simulation example CHEMOSTA, Sec. 8.4.1.


The production rate for a batch culture can be calculated by dividing the
biomass concentration by the time for the culture. Thus, the batch biomass
production rate is equal to Xf/tf.
Comparing the continuous biomass production rate to the batch biomass
production rate, it is found that the continuous fermenter will have a 2 - 3 times
larger maximum biomass production rate. This is because the batch
fermentation starts at the inoculum value of X and has correspondingly low
initial growth rates.
Biomass production rates for fed batch fermenters are calculated by taking
the total mass produced over the time of operation, or (VX)f - (VX)o/tf . For
cyclic operation the biomass production has been shown to depend on the
starting and final volume ratio (Keller and Dunn, 1978). This and more
complex questions regarding product productivity for particular kinetics can be
answered by making suitable changes in the simulation example REPFED, Sec.
8.3.4. The productivity of a repeated fed batch as compared to chemostat
operation will depend on the operating variables, as well as the kinetics (Dunn
et al., 1979).
110 4 Bioreactor Modelling

4.1.5 Case Studies

4.1.5.1 Case A. Continuous Fermentation with Biomass


Recycle

The retention of active biomass is a means of improving the reactor


productivity or efficiency for substrate uptake. The biomass separation could
be performed by any suitable process (flotation, sedimentation, membrane
filtration, or centrifugation). The cell recycle stream has a volumetric flow rate
R and a biomass recycle concentration XR, where X R > X I and Xi is the
biomass concentration in the stream leaving the fermenter.
Consider the operation of a biological reactor with cell recycle as shown in
Fig. 4.8.

S^X^F+R

Cell recycle

Figure 4.8. A bioreactor with cell separation and recycle.

The mass balances around the entire system are as follows: Biomass
accumulates both within the reactor of volume Vr and also within the separation
unit with volume Vs. Assuming that biomass leaves only in the wastage stream
and that growth occurs only in the reactor, the balance is then
dXi
=0-WXR

where W is the wastage flow rate.


At steady state,
0 = -
4.1 General Balances for Tank-Type Biological Reactors 111

Thus, the wastage rate of biomass must equal its production rate, otherwise Xi
will change. Wastage rate is an important control parameter in wastewater
treatment, where the separator is usually a sedimentation tank.
For the substrate, which is consumed only in the reactor section,

dSi dSR

At steady state,
0 = F So FI Si W Sj + r$ Vr

Here it is seen that the uptake rate is equal to the difference between the inlet
and outlet mass flows. The efficiency of the continuous biomass separation
determines XI/XR, which controls the recycle ratio, R/(F+R).
Considering the fermentation tank only, the balances are as follows:
For the biomass,

At steady state,
0 =RXR-(F + R)X!+rxVr

Cell separation and recycle lead to high cell concentrations in the reactor,
which, when neglecting the contribution by growth, would be XR/(F+R). Since
the rates are proportional to Xi, an increase in reactor efficiency is obtained.
Assuming, rx = |n Xi gives,

MX,
where, XR > Xi , and D = F/V is the nominal dilution rate. This equation means
that the specific growth rate is decreased from the chemostat value, D. This is
due to the reduction in substrate concentration, Si, which is caused by the
higher biomass concentration, resulting from the cell recycle. Washout is
impossible due to the complete biomass retention, and for this reason flow rates
greater than in a chemostat are possible.
The substrate balance gives
JQ

Vr == F S0 + R Si - (F + R) Si + rs V
dt
which shows that at steady state
112 4 Bioreactor Modelling

This would also be the case without cell recycle, since the substrate is assumed
to pass unchanged in concentration through the separator.
The above equation can be written as

where the right hand side of this equation is the F/M (food/biomass) ratio. This
gives a theoretical basis to the F/M concept, which is well known for the control
of waste treatment plants. The simulation example ACTNITR, Sec. 8.4.3,
enables the main operating characteristics of cell recycle systems to be studied.
A related simulation example MEMINH, Sec. 8.9.1, considers the retention of
enzyme using a membrane, and SUBTILIS, Sec. 8.9.2, involves the retention of
biomass.

4.1.5.2 Case B. Enzymatic Tanks-in-series Bioreactor


System

A three tanks-in-series system is used for an enzymatic reaction, as shown in


Fig. 4.9.

:
: :';: y~: : w> Vv

illi 81 S2
> . 1
111

>
11 S;^;:-);'::;HaKS illi
Figure 4.9. Tanks-in-series bioreactor.

For the first tank,


dSi
jp = F(S 0 -Si)
dividing by
SQ-SI
=

where i\ = Vi/F and is the mean residence time of the liquid in tank 1.
The balances for tanks 2 and 3 have the same form except for the subscripts:
4.2 Modelling Tubular Plug Flow Bioreacrors 113

For known flow rate, F, and known tank volumes, there are six unknowns in
these three equations. Note that different tank sizes may be accounted for by
differing values for the tank residence times TI, 12 and 13.
If the kinetic terms r$ are only dependent on S, then the above equations can
be solved without any further balances. It is often the case that enzymatic rate
equations of the form given below can be used for each tank n = 1, 2 and 3:

This gives now six equations and six unknowns, and the problem is solvable by
simulation methods.
If the situation is more complex, such that r$ depends on other components,
for example P, in the case of product inhibition or biomass X, then additional
balance equations for these components must be included in the model. When
combined with equations for the complete kinetics description (rates as a
function of all the influential concentrations), the model can be solved to obtain
the dynamics of the system and also the final steady state values. It can be
shown that a three tanks-in-series reactor system will provide a good
approximation to the performance of a corresponding tubular reactor, except
for very high conversions.

4.2 Modelling Tubular Plug Flow Bioreactors

4.2.1 Steady-State Balancing


The tubular reactor can be modelled for steady state conditions by considering
the flow as a series of fluid elements or disks of liquid, which behave as a batch
reactor during their time of passage through the reactor. This can be
understood by considering a pulse of unreacting tracer in Fig. 4.10 that passes
from entrance to exit unchanged without mixing.
114 4 Bioreactor Modelling

tracer pulse response

Figure 4.10. Plug flow idealization of the tubular reactor with no axial mixing.

A reaction will cause a steady state axial concentration profile as shown in Fig.
4.11. Thus at steady state, the concentrations vary with distance in a manner
which is analogous to the time-varying concentrations that occur in a batch
reactor.

Concentration

Figure 4.11. Axial profiles of steady-state concentrations in a tubular reactor.

This means that steady-state tubular reactor behavior can be modelled by direct
analogy to that of a simple batch reactor. Thus using the batch reactor
substrate balance (p = constant),
dS
dF = fS

The flow velocity, v, for the liquid is defined as,


dZ
v
= dF
or,
dZ
dt =

where v = F/A and F is the volumetric flow rate through the tube with cross-
sectional area A. Thus substituting for dt,

dS_ rs_
=
dZ v
4.2 Modelling Tubular Plug Flow Bioreacrors 115

This is the steady state tubular reactor design equation.


With the kinetics model, rs = f(S), the equation can be integrated from the
inlet, at position Z = 0, to the outlet, at Z = L, to obtain the steady state
concentration profile of S. Additional component balances would be required
for more complex kinetics. This is demonstrated by the simulation example
for a tubular enzyme reactor ENZTUBE, Sec. 8.4.4.

4.2.2 Unsteady-State Balancing for Tubular


Bioreactors

If dynamic information is needed for tubular or column systems, then changes


with respect to both length and time must be considered. In order to achieve
this, the reactor can be considered by dividing the volume of the reactor into N
finite-differenced axial segments (Fig. 4.12), and treating each segment
effectively as a separate stirred tank.

)))))>
Figure 4.12. Finite-differencing the tubular reactor.

Figure 4.13. Balancing the difference segment n for the tubular reactor.

Formulating the substrate balance for S over a single segment n of volume AV


= A AZ:

f Accumulation^ ( input \ { output \ ^production rate\


V rate of S ) = Vrate of S) ~ \ rate of S) + ^ of S by growth)

The balances have the same form. Thus for segment n,


dSn
AV
"dT = S n-l F n-l - Sn Fn + rSn AV

Constant density gives F n _i = Fn = F. Dividing by AV,


116 4 Bioreactor Modelling

dSn = A(SF)
+ rsn
dt AV

Setting AV AdZ and AS > dS gives the partial differential equation, which
describes changes in time and distance, as,

38
- =
1 d(SF)
-- x J_
at A az
When the volumetric flow, F, is constant,

as _F as + rs
at ~ ~ A az
At steady state,

and
V + fS
dZ
which is the steady state equation derived earlier in Section 4.2.1.
To model the dynamics by simulation methods, the partial differential
equation must be written in difference form as,

AV ^- = F (Sn.j - Sn) + rsn AV


or

dt AV/F

where AV/F = x , the residence time of the liquid in a single segment.


Mass Transfer

5.1 Mass Transfer in Biological Reactors

Multiphase reaction systems usually involve the transport of material between


two or more phases. Usually one of the reactants is transferred from one phase
into a second phase, in which the reaction takes place. The following cases are
examples of biological systems.

5.1.1 Gas Absorption with Bioreaction in the Liquid


Phase

The gas phase is dispersed as gas bubbles within the liquid phase. Mass transfer
occurs across the gas-liquid interface, out of the gas into the liquid, where the
reaction occurs. The typical example is aeration of the bioreactor broth and
the supply of oxygen to the cells as shown in Fig. 5.1.

Figure 5.1. Absorption of oxygen from an air bubble to the liquid medium.

Biological Reaction Engineering, Second Edition, I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
118 5 Mass Transfer

5.1.2 Liquid-Liquid Extraction with Bioreaction in


One Phase

An immiscible liquid phase is dispersed in a continuous liquid phase. Mass


transfer of a reactant takes place across the liquid-liquid interface, shown here
(Fig. 5.2) from the continuous phase into the dispersed phase, where reaction
occurs. An example might be the transfer of a substrate in an oil phase to an
enzyme in the droplet aqueous phase.

Figure 5.2. Liquid-liquid extraction plus reaction.

5.1.3 Surface Biocatalysis

In this case, a liquid phase is in contact with solid biocatalyst. Substrates A and
B diffuse from the liquid to the reaction sites on the surface of the solid, where
reaction occurs. The product C must similarly be transferred away from the
solid reaction surface, as shown in Fig. 5.3. Examples are found with
immobilized enzyme and cell systems. In Sec. 6.1 the modelling aspects of this
type of system are considered in detail.
5.2 Interface Gas-Liquid Mass Transfer 119

Figure 5.3. Reaction of two substrates on a solid biocatalyst surface.

5.1.4 Diffusion and Reaction in Porous Biocatalyst

Here a porous biocatalyst sphere is suspended in a liquid medium. Substrates


diffuse into the porous internal structure of the biocatalyst support and react.
Similarly, the products must diffuse away from the reaction sites within the
solid to the outer surface, where they are then transported into the liquid.
Detailed modelling of this process is treated in Ch. 6.

Figure 5.4. Reaction within a solid biocatalyst.

5.2 Interphase Gas-Liquid Mass Transfer

Concentration gradients are the driving forces for mass transfer. Actual
concentration gradients (Fig. 5.5) in the very near vicinity of the gas-liquid
120 5 Mass Transfer

interface, under mass transfer conditions, are very complex. They result from
an interaction between the mass transfer process and the local fluid
hydrodynamics, which change gradually from stagnant flow, close to the
interface, to perhaps fully-developed turbulence within each of the bulk phases.
According to the Two-Film Theory, the actual concentration profiles, as
represented in Fig. 5.5 can be approximated by linear gradients, as shown in
Fig. 5.6.
A thin film of fluid is assumed to exist at either side of the interface. Away
from these films, each fluid is assumed to be in fully developed turbulent flow.
There is therefore no resistance to mass transfer within the bulk phases, and the
concentrations, CG and CL, are uniform throughout each relevant phase. At the
phase interface itself, it is assumed there is no resistance to mass transfer, and
the interfacial concentrations, CGI and CLI, are therefore in local equilibrium
with each other. All the resistance to mass transfer must, therefore, occur within
the films. In each film, the flow of fluid is assumed to be stagnant, and mass
transfer is assumed to occur only by molecular diffusion and therefore to be

Interface

Gas

Figure 5.5. Concentration gradients at a gas-liquid interface.

described by Pick's law, which says that the flux JA (mol/s m2) for the molecular
diffusion of some component A is given by,

dZ
5.2 Interface Gas-Liquid Mass Transfer 121

Interface
Gas Liquid

Figure 5.6. Concentration gradients according to the Two-Film Theory.

where D is the molecular diffusion coefficient (m2/s) and dC/dZ is the steady
state concentration gradient (mol/m3). Thus applying the same concept to mass
transfer across the two films,
JA = DG Z Z
G L

where DG and DL are the effective diffusivities of each film, and ZG and ZL are
the respective thicknesses of the two films.
The above equations can be expressed in terms of mass transfer coefficients
kc and kL (m2/s) for the gas and liquid films,

JA = k G (C G -C G i) = k L (C Li -C L )

The total rate of mass transfer, Q (mol/s), is given by,

Q = JAA = jA(aV)

where "A" is the total interfacial area available for mass transfer, and "a" is
defined as the specific area for mass transfer or interfacial area per unit liquid
volume (m2/m3). Thus for the total rate of mass transfer:

In terms of the total interfacial area A,

Q = k G A(C G -C G i ) = k L A(C L i -C L )

In terms of a and VL,

Q = ko a (Co - CGi) VL = k L a(C L i -C L )V L


122 5 Mass Transfer

Since the mass transfer coefficient, k, and the specific interfacial area, a, depend
on the same hydrodynamic conditions and system physical properties, they are
frequently combined and referred to as a "ka value" or more properly a mass
transfer capacity coefficient.
In the above theory, the interfacial concentrations CGI and CLI cannot be
measured, and are therefore of relatively little use, even if the values of the film
coefficients are known. For this reason, by analogy to the film equations,
overall mass transfer rate equations are defined, based on overall coefficients of
mass transfer, KG and KL, and overall concentration driving force terms, where:

Q = K G A(C G -C G *) = K L A(C L *-C L )

Here, CG* and CL* are the respective equilibrium concentrations, corresponding
to the bulk phase concentrations, CL and CG, respectively, as shown in Fig. 5.6.
Equilibrium relationships for gas-liquid systems, at low concentrations of
component A usually obey Henry's law, which is a linear relation between gas
partial pressure, PA, and equilibrium liquid phase concentration, CLA*:

PA=
where HA (bar m3/kg) is the Henry's law constant for component A in the
medium. Henry's law is generally accurate for gases with low solubility, such as
the solubility of oxygen in water or in fermentation media. Thus from this
relation, as shown in Fig. 5.7, the corresponding equilibrium concentrations can
be easily established.

CLC*

Figure 5.7. Equilibrium concentrations based on Henry's law.

For gases of low solubility, e.g., oxygen and carbon dioxide in water, the
concentration gradient through the gas film is very small, as compared to that
within the liquid film, as illustrated in Fig. 5.6. This results from the relatively
5.3 General Oxygen Balances for Gas-Liquid Transfer 123

low resistance to mass transfer in the gas film, as compared to the much greater
resistance to mass transfer in the liquid film. The main resistance to mass
transfer is predominantly within the liquid film. This causes a large change in
concentration (Cy - CL), since the resistance is almost entirely on the liquid
side of the interface.
At the interface, the liquid concentration, Cy, is in equilibrium with that of
the gas, CGI, and since CGI is very close in magnitude to the bulk gas
concentration, CLI must then be very nearly in equilibrium with the bulk gas
phase concentration, CG- This is known as liquid film control and corresponds
to the situation where the overall resistance to mass transfer resides almost
entirely within the liquid phase. The overall mass transfer capacity coefficient
is KLa. Hence the overall mass transfer rate equation used for slightly soluble
gases in terms of the specific area is

Q = KLa (C L *-C L )V L

where CL* is in equilibrium with CG, as given by Henry's law,

C G = HCL*,

Mass transfer coefficients in fermentation are therefore generally spoken of as


KL values or K^a values for the case of mass transfer capacity coefficients.

5.3 General Oxygen Balances for Gas-Liquid


Transfer
In order to characterize aeration efficiency, to predict dissolved oxygen
concentration, or to follow the biological activity it is necessary to develop
models, which include expressions for the rate of oxygen transfer and the rate
of oxygen uptake by the cells. Well-mixed phase regions, in which the oxygen
concentration can be assumed uniform, can be described by simple balancing
methods. Situations in which spatial variations occur require more complex
models, as described in Sec. 5.4. The following generalized oxygen balance
equations are derived for well-mixed phases, using the well-mixed tank
concept. In the situation in Fig. 5.8, both the liquid and gas phases are defined
by distinct well-mixed regions and by the total volumes of each phase, VL and
VG.
124 5 Mass Transfer

Gas CQO GO

Figure 5.8. The balance regions for well-mixed gas and liquid phases in a continuous reactor.

For the gas phase the oxygen balance can be developed as follows:

Rate of > ( Rate of ^


accumulation Flow of f Flow of \ transfer
of oxygen oxygen in _ oxygen out _ of oxygen
in gas ; inlet gas streamy Vin exit stream/
v from gas ,

Thus, for the gas phase,

dCGi
K L a(C L i*- C L i)V L

where VQ represents the volume of gas in the dispersed phase, or the gas
holdup.
For the liquid phase,

Rate of ^ /Flow of ( Flow of Rate of


accumulation oxygen oxygen transfer
of oxygen in inlet out in of oxygen
^ in liquid liquid exit ^ from gas
V stream J
' Rate of
consumption
of oxygen
in liquid

Rate of oxygen consumption = -rO2 = -qo2


5.3 General Oxygen Balances for Gas-Liquid Transfer 125

Thus for the liquid phase,

dCLi
- LiC L i + K L a(C L 1 *-C L i)V L -

The above equations include accumulation, convective flow, interphase transfer


and biological oxygen uptake terms. Here CLI* is the equilibrium solubility of
oxygen corresponding to the gas phase concentration, CGI , and is calculated
by Henry's law, according to the relationship:

Typical units are as follows: CG and CL (kg/m3); G and L (m3/s); K^a (1/s);
VG and VL (m3); qO2 (kg/kg s); X (kg/m3).
In the next sections, the general equations, given above, will be applied to
important special situations.

5.3.1 Application of Oxygen Balances

5.3.1.1 Case A. Steady-State Gas Balance to Determine the


Biological Uptake Rate

The convective terms in the generalized liquid balance equation can usually be
neglected, owing to the low solubilities of oxygen in water (about 8 g/m3). This
gives the steady state liquid balance, dCL/dt=0, relation as:

K L a(C L i*- CLI) = qo2Xi

Thus at steady-state, the oxygen transfer rate is effectively equal to the oxygen
uptake rate. Even during batch fermentations this is approximately true.
Substituting this relationship into the steady state gas balance gives,

0 = GO CGO - GI CGI - qo2 X VL


The above equation can also be derived from a steady state balance around the
entire two-phase system. It shows that the biological oxygen uptake rate can be
calculated from knowledge of the gas flow rates and the gas concentrations.
126 5 Mass Transfer

This application is very important in fermentation technology, since it permits


the on-line monitoring of the rate of fermentation, by gas balancing methods
(Heinzle and Dunn, 1991, Ingham and Dunn, 1991).

5.3.1.2 Case B. Determination of Ki,a Using the Sulfite


Oxidation Reaction

If a chemical reaction, classically the oxidation of sodium sulfite, is used to take


up the oxygen from solution, then the term qo2 X VL in the liquid phase
balance may be replaced by the chemical reaction term, ro2 VL- At steady-
state,
K L a(C L i*-C L i) = r02
Usually ro2 is obtained by taking samples and titrating for the fractional
conversion of sulfite, which can be related by stoichiometry to the oxygen
consumption. Since the chemical reaction causes the liquid dissolved
concentration CLI to fall essentially to zero and with CLI* calculated from the
oxygen concentration in the exit gas, the value of the overall mass transfer
capacity coefficient, K^a, can be estimated. An improved method uses the gas
balance instead of titration to obtain ro2 in the manner outlined above for
qO2 X and also provides a check on the sulfite measurements. The sulfite
method is useful for comparing aeration systems, but the values are difficult to
relate to actual fermentation conditions owing to the very different physical
conditions (coalescence, aeration rates) (Ruchti et al., 1985).

5.3.1.3 Case C. Determination of Ki,a by a Dynamic Method

If water is initially deoxygenated and is then re-aerated, the concentration of


the dissolved oxygen will increase with time, from zero to effectively 100% air
saturation at the end of the experiment. The exact form of the response curve
obtained depends on the values of KLa, the driving force, (CLI* - CLI), and the
measurement dynamics of the dissolved oxygen electrode. The liquid balance,
for the unsteady state batch aeration condition, gives:

=
K L a(CLi*-C L i)V L
5.3 General Oxygen Balances for Gas-Liquid Transfer 127

The classical dynamic KLa method assumes that K^a and CLI* are constant.
Under these conditions, the differential equation can be integrated analytically
to give the relationship:

r *(
CL* ^
CL - rCL\}
= K L at

Plotting the natural logarithmic concentration function on the left side of the
equation versus time, should, in principle, give a straight-line relationship, with
^a as the slope.
Usually deoxygenation is accomplished with nitrogen, so that initially the gas
phase consists of nitrogen, which is gradually displaced and mixed with air.
Under these conditions, CLI* is nt constant, and a gas balance must be
employed to calculate the variation in CGI versus t. Since the liquid phase
concentration, CLI, is measured by means of a membrane covered oxygen
electrode, the dynamics of measurement method usually cannot be neglected.
The dynamics of the measurement electrode can be described, approximately,
by a first-order lag equation,
dCE

where TE represents the electrode time constant, and CE is the measurement


signal.
The fractional response of the electrode for a step change in CL would
appear as shown in Fig. 5.9.

time

Figure 5.9. Response of electrode for a step change in CE from zero to 100 % saturation
according to a first-order lag model.

Note that TE corresponds to the time for the electrode to reach 63 % of the final
response. The overall process dynamics involves thus the gas phase, the liquid
128 5 Mass Transfer

phase and the electrode response. The three responses might appear as shown
below:

time

Figure 5.10. Response of the gas, the liquid and the electrode measurement during a dynamic
KLa experiment.

The values of three individual time constants determine the process response.
These are TG = VG/G, (representing the dynamics of the gas phase), 1/KLa
(representing the dynamics of the liquid phase mass transfer process), and IE
(representing the measurement dynamics). This is illustrated in the simulation
example KLADYN, Sec. 8.5.5.

5.3.1.4 Case D. Determination of Oxygen Uptake Rates by a


Dynamic Method

Low oxygen uptake rates, as exist in slow growing systems (plant and animal
cell cultures, aerobic sewage treatment processes, etc.), cannot easily be
measured by a gas balance method, since the measured difference between inlet
and outlet oxygen gas phase concentrations is usually very small. Due to the
low solubility of oxygen in the liquid media, quite small oxygen uptake rates
will cause measurably large changes in the dissolved oxygen concentration.
Thus it is possible to measure qo2 X either by taking a sample and placing it in
a small chamber or by turning off the reactor air supply, according to the
liquid balance equation
dCLi
5.3 General Oxygen Balances for Gas-Liquid Transfer 129

Dissolved oxygen concentration decreases linearly and is equal to qo2 X as


shown in Fig. 5.11.

Figure 5.11. Oxygen uptake rate determined by a dynamic method.

When the time required for an appreciable decrease in dissolved oxygen is


large, as compared to the electrode time constant, the method is quite accurate
and no correction for the electrode measurement dynamics is required (Mona
et al., 1979). If the response is too fast the sample can be diluted. This method
is illustrated by the simulation example OXDYN, Sec. 8.5.4. A similar
simulation example, ANAMEAS, Sec. 8.8.7, illustrates dynamic measurements
in anaerobic systems.

5.3.1.5 Case E. Steady-State Liquid Balancing to Determine


Oxygen Uptake Rate

If the biomass is immobilized or retained by membranes within the reactor,


oxygen can be supplied to the cells by means of a circulating liquid supply,
which is aerated in a separate unit, external to the reactor.
130 5 Mass Transfer

CL1

CLO

Figure 5.12 Oxygen uptake rate determined by a steady-state liquid balance.

It then becomes possible to determine the oxygen uptake rate, simply by


measuring the liquid flow rate and the difference in dissolved oxygen in the
liquid inlet and outlet flow streams, according to the following steady-state
liquid phase balance equation:

0 = L(CLo - CLI) - qo2 Xi VL

Thus the rate of oxygen supply via the liquid is equal to the rate of oxygen
uptake by the cells. This method provides a very sensitive way of measuring
low oxygen uptake rates (Keller et al., 1992, Tanaka et al., 1982). The case-
study H in Sec. 5.3.1.8 is an example of this use for an experimental reactor.
The simulation example FBR, Sec. 8.4.9, also demonstrates this method.

5.3.1.6 Case F. Steady-State Deoxygenated Feed Method for


KLa

Feeding a deoxygenated liquid continuously to an aerated tank (Fig. 5.11)


allows the oxygen transfer rate to be determined by difference measurement.
Thus the liquid phase balance becomes

0 = L (CLO - CLI) + KLa (CLi* - C L i)V L

Knowing the flow rate L, the oxygen liquid concentrations CLO and CLI and the
outlet oxygen in the gas phase (to determine CLI*) permits the calculation of
5.3 General Oxygen Balances for Gas-Liquid Transfer 131

KLa. Another variation of this would be to gas with oxygen-enriched air or


with nitrogen, which would avoid the difficulty of producing a continuous
source of deoxygenated liquid. A similar steady state method has been
employed to obtain steady oxygen concentration profiles in column (Meister et
al., 1980), and tubular bioreactors (Ziegler et al., 1977). A suitable steady state
model for the tubular reactor then allows calculating the unknown K^a by
parameter estimation (Shioya et al., 1978).
Deoxygenated
C
liquid LO

t
/"VX "N^

nM&^/m^M:WM.
-^^^ r-^^-~\_/~\^^x

Xiiiiiiiiiiji ^^^SRSSSS
^^^iO^wiBii
O

Wiiiiilmiiiiiii
Air

Figure 5.12. Steady-state dissolved oxygen difference measurement for Kj^a.

5.3.1.7 Case G. Biological Oxidation in an Aerated Tank

A batch reactor liquid is aerated with a continuous flow of air to support a


biological reaction, as shown in Fig. 5.13.
air

air

Figure 5.13. A batch bioreactor with continuous aeration.


132 5 Mass Transfer

The biological reaction in the liquid phase is first-order in oxygen


concentration. Since oxygen is relatively insoluble (approximately 8 g/m3
saturation for air-water) the transfer rate is important to maintain a high
dissolved oxygen concentration CL. The batch oxygen balance for the liquid
phase is then:

f Rate of \ ,^ f
accumulation of (Transfer rate of f N\ / Uptake rate of
=
V 02 in liquid ) l02 mto the hqmdj ^ by the cells

dCL
= K L a(CL*-C L )VL - k C L V L

A steady-state can be reached for which the mass transfer rate is equal to the
oxygen uptake rate by reaction:

0 = KLa (CL* - CL) - k CL


giving for CL
KLaCL*
CL = K L a

Using this equation, the reaction rate constant, k, can be determined if CL is


measured and K^a is known or measured. The equilibrium value, CL*, can be
calculated from the gas phase concentration, and if there is little oxygen
depletion it can be calculated from the inlet gas conditions.
If the oxygen depletion in the gas phase is appreciable, then the mole
fraction of oxygen in the exit may not be the same as in the inlet, and a gas
phase balance must be applied to determine CL*:

Figure 5.14. Inlet and outlet oxygen mole fractions and total gas molar flow rates.
5.3 General Oxygen Balances for Gas-Liquid Transfer 133

From the ideal gas law as shown before, assuming a well-mixed gas phase in
steady state, N = (p / RT) F, where NO is the molar flow rate of air and F is the
air volumetric flow rate.
/ Rate of O2 \ / Rate of O2 \ / Transfer rate of \
0 =
V in by flow ) ~ Vout byflow) ~ \ C>2 to the liquid )

Using the nomenclature in Fig 5.13,

0 = yoN0-yiNi-KLa(CL*-CL)VL
where

Assuming NO = NI, these equations can be solved to obtain yi and CL-


Solving for CL gives,
CL = k~" CL*

or for the apparent reaction rate,


k ,
re = - k~ CL

Thus it is possible to distinguish between two different regimes for this system,
transfer control and reaction control:
1) Reaction rate control applies for low values of k/KLa, when re approaches
k CL*, and CL approaches CL*
2) Diffusion control applies for high values of k/KLa, when re approaches
KL& CL* and CL approaches 0.
3) If KLa = k, then rc = - (k/2) CL*, and CL approaches (1/2) CL*.

5.3.1.8 Case H. Modelling Nitrification in a Fluidized Bed


Biofilm Reactor

Nitrification is a two-step microbiological process, in which the ammonium ion


is oxidized to nitrite ion and further to nitrate ion as shown:

NH4+ - NO2- - NO3-


134 5 Mass Transfer

This reaction is important in waste water treatment because of the toxicity of


ammonia and its large oxygen demand. Several known organisms can gain
energy from either of the two oxidation steps, but most commonly
Nitrosomonas and Nitrobacter are responsible for steps (1) and (2),
respectively. These organisms grow very slowly, obtaining their carbon from
dissolved carbonate. Due to the very slow grow rates, it is of interest to retain
the biomass within the reactor. One possibility considered here is to
immobilize the biomass as a natural biofilm on a fluidized bed of sand (Tanaka
and Dunn, 1982).
The stoichiometric relations for the reaction steps (1) and (2) are:
, 3 ,
NH4+ + j O2 -> NO2" + H2O + 2 H+

O2 -> NO3"

Summing the above steps (1) and (2) gives

NH4+ + 2O 2 -

The reactor of volume, Vr, consisted of a conical sand bed column, which was
fluidized by the liquid recycle stream flowing up through the bed. The recycle
stream was oxygenated in a separate, baffled, tank contactor of volume VT, with
turbine impeller and air or oxygen sparging. The reactor and oxygenator were
thus separate parts of a recycle loop configuration. This could be operated
batchwise or with a continuous feed and effluent stream flow to and from the
system. When operating at high recycle rates, the whole system acted
effectively as one well-mixed tank system.
The reactor-oxygenator recycle loop can be analyzed as a total system or
broken down into its individual components as shown in Fig. 5.14. These
include liquid phase balance over the reactor and combined phase, liquid phase
and gas phase balances over the oxygenator and over the total system.
5.3 General Oxygen Balances for Gas-Liquid Transfer 135

Figure 5.15. Mass balancing regions for the fluidized bed reactor nitrification system.

The mass balances to be considered are those for oxygen and the nitrogen-
containing reactants and products. The oxygen balance taken over the total
system can be simplified by neglecting the accumulation terms and the liquid
flow terms, that will be small compared to the gas rates and the consumption by
reaction, owing to the relatively low solubility of oxygen in the liquid medium.
Thus the oxygen balance becomes,

0 =

Here Vr is the volume of the reactor column.


The nitrogen (N) components, NH4+, NCV, and NOs', in the liquid phase
can be balanced around the total system by considering the accumulation, flow,
and reaction terms for each of the N-containing components. For the total
system each component equation has the form,

VdCp
= F(C N i-C N 2 )
dt

When the reactor is operated as a batch system, F = 0, and when used as a


continuous steady state reactor, dCN2/dt = 0. This equation can be used in
column systems for very low single-pass conversion, when the differences in
local reaction rate at the reactor inlet and outlet are not large. Although the
reactions actually occur in the solid phase, because of the high solid-liquid
interfacial area, the system is treated here as being quasi-homogeneous.
The gas-liquid interfacial mass transfer area will often be small enough to be
important for the overall process, and it is therefore useful to consider the gas
136 5 Mass Transfer

and liquid phases as separate balance regions. The absorption tank can be
described by the oxygen balances for the liquid phase:

0 = F R (C L4 -C L3 ) + K L a(C L *-C L3 )V T

and for the gas phase:

0 = G(CGi - CG2) - KLa (CL* - CL3) VT

The liquid phase oxygen balance for the total system is

0 = KL

where ro2 is the oxygen uptake rate by the reaction. These equations, which
assume ideally mixed phases, are useful in designing the gas absorber
according to the required oxygen transfer coefficient.
Balancing the oxygen around the reactor gives

0 =
Since CL4 at the reactor outlet is usually very low, then,

FR CL3 = - ro2 Vr

which says that the oxygen uptake rate by reaction must be equal to the supply
rate from the oxygenation tank. This is the condition of reaction-rate
limitation by the oxygen transfer in the absorber.
From the stoichiometry, the relationships between the molar reaction rates
(rNH4> rO2 rH* r2,NO2 an^ fNO3) can be found. Thus, for example, the first
nitrification step gives
2
TNH4 = T r l , O 2 = ~ri,NO2

and the total rate for 02 is given by the sum of the rates for steps (1) and (2).
r
O2 = r l,O2 + r2,O2

From the measured concentration dependency of these rates, the reaction


kinetics of the individual steps can be determined. The dependency of these
rates on the individual concentrations can then be used to establish the reaction
kinetic model. This model is the basis of the simulation example NITRIF, Sec.
8.5.3. A similar type of recycle, fluidized-bed reactor is the theme of
simulation examples FBR, Sec. 8.4.9 and DCMDEG, Sec. 8.4.6.
5.4 Models for Oxygen Transfer in Large Scale Bioreactors 137

5.4 Models for Oxygen Transfer in Large Scale


Bioreactors

Large-scale industrial fermenters can generally be expected to exhibit


deviations from the two idealized flow conditions of perfect mixing or perfect
plug flow. Thus the assumption of completely mixed gas or liquid phases may
not be valid. Little experimental information is available on concentration
inhomogeneities or concentration gradients within large bioreactors. Residence
time distribution information, from which a physical and mathematical model
could be established, is also generally not available.
Convection currents within the liquid phase of a bioreactor are usually caused
by the mechanical energy inputs of agitation and aeration. It is often
reasonable to assume that slowly changing quantities, such as biomass
concentration, substrate concentration, pH and temperature are uniform within
the whole mass of bioreactor liquid. Oxygen must be considered, however, as a
rapidly changing substrate, owing to its low solubility in fermentation media. It
is therefore necessary to consider that differences in oxygen transfer and
uptake rates will create oxygen concentration gradients throughout the reactor.
Buoyancy forces carry the gas from the lower gas inlet point up to the top
liquid surface. In the absence of mechanical agitation, the gas phase might
move from the bottom to the top of the reactor in an approximate plug flow
manner, with very little backmixing. If the stirring power supplied to the
fermenter, however, is sufficient to create liquid velocities, that are greater than
the free rise velocity of a bubble (about 26 cm/sec) then the bubbles will
circulate around the fermenter, before eventually escaping. Very high power
inputs can cause the smaller bubbles to circulate many times within the vessel
and spend an appreciable time before reaching the surface. Under such
conditions, if no bubble coalescence occurs, the gas phase would contain a
fraction of small bubbles, depleted of oxygen but with a large surface area.
Obviously any well-mixed phase assumption becomes difficult to justify.
The gas phase flow conditions in large scale industrial fermenters usually lie
somewhere between the extreme cases of idealized plug flow and perfect
mixing. Experimental residence-time distribution information, obtained by
helium tracer techniques under actual operating conditions, are then necessary
to characterize the gas phase flow. Unfortunately very little experimentation on
industrial scale equipment has been reported.
Hydrostatic pressure gradients in tall fermenters will cause large differences
in the oxygen solubility, CL*, with regard to the depth position in the tank. In a
10m tall reactor, the oxygen solubility for a given gas composition will be twice
that at the bottom of the tank as compared to the top surface, since the total
pressure is effectively doubled. This is seen by Henry's law which can be written
as:
138 5 Mass Transfer

d* = yH2p
where yo2 is the mole fraction of oxygen in the air and p is the total pressure at
some point in the tank.
The possibility that oxygen gas compositions, dissolved oxygen
concentrations, oxygen solubilities, gas holdup volumes, bubble sizes and other
transfer parameters can vary with depth in a tall fermenter introduces a much
greater degree of complexity to the problem of modelling the reactor. This
makes it difficult to obtain data on oxygen mass transfer coefficients.
Although it is impossible to give specific recommendations that apply to any
particular situation, a further discussion of possible models and their
underlying assumptions may help to define the problem. Incorporated into the
more complex models, discussed below, are such factors as gas and liquid phase
flow pattern, gas composition gradients and the effects of hydrostatic pressure.
Great caution and wisdom must be exercised to avoid creating a model that is
too complex to verify by experimentation. Experienced engineers will say
"Keep it simple!" and "Avoid too much model!".
All large scale reactors, whether multi-impeller tanks or column fermenters,
will display some axial dissolved oxygen concentration gradients. The most
general method for modelling is to represent the reactor using balances in a
series of sections or stages. Mass balances in multi-stage process are easy to
formulate, since both the liquid and gas phases may be assumed to be well-
mixed, for any given stage of the cascade.

Figure 5.16. A single gas-liquid stage with backmixing of the liquid phase.

The formulation of the mass balances for a single stage, as shown in Fig.
5.16, follows closely that described previously, except that now the reactor is
made up of many stages which are interconnected by the flows of gas and
liquid between stages and by diffusive mass transfer mechanisms.
5.4 Models for Oxygen Transfer in Large Scale Bioreactors 139

5.4.1. Case Studies

5.4.1.1 Case A. Model for Oxygen Gradients in a Bubble


Column Bioreactor

The application of the stagewise modelling approach is shown below, where a


bubble column reactor is modelled as a five-stage reactor system. The reactor
will be assumed to operate cocurrently, as would be also the case for the riser of
an airlift bioreactor.

Exit Gas Exit Liquid


A CQB A CLS

Gas
ill

Gas Feed Liquid Feed

GO L' LO

Figure 5.17. Stagewise model of a bubble column bioreactor.


140 5 Mass Transfer

The oxygen balance equations for the gas and liquid phases of each stage are
as follows:

if = FG(C G n-l-C Gn ) - K L a(C Ln *-C Ln )V L


f\(~^

VL-dT = F L(C L n-i-C Ln ) + K L a(C L n*-C L n)VL - rn VL

where,
P
r<~Ln * - 1
H rC r c * -
Gn or CLn
02
-77-
rl
and
rn = Qo2m |

For simplification only oxygen is assumed limiting and growth is not


considered; however the biomass concentration is contained in the maximum
oxygen uptake Qo2m- The dynamics of the oxygen transfer and uptake
processes are obtained by solving these differential equations simultaneously
for each stage. The resulting solution then gives CLH and CGn > for each stage
as functions of time and also yields the resulting final steady state values.
Note that the biomass concentrations Xn are assumed constant, otherwise
biomass balance and growth kinetics equations would have to be added to the
model. Using simulation methods, other effects, such as the effect of
hydrostatic pressure on CG or on bubble size could be included. The
simulation example DCMDEG, Sec. 8.4.6, demonstrates some aspects of the
stagewise modelling approach.

5.4.1.2 Case B. Model for a Multiple Impeller Fermenter

Mixing in a tank reactor is complex, and it would be necessary to consider


liquid flow in both directions. It is generally assumed, however, that the
intensity of mixing is such that no radial variations occur. Fig. 5.16 represents
a multiple impeller reactor with well-mixed liquid zones in the region of each
impeller. The reactor can be described approximately by means of a three-
stage model. Mixing of the liquid in a direction which is directly opposite to
that of the main flow liquid (here upwards) can be incorporated into the model,
by the assumption of a backmixing stream, with flow rate FB- This backmixing
stream accounts for a flow interaction between the mixing zones and for
deviations from ideal stage mixing. To determine FB, a tracer experiment
5.4 Models for Oxygen Transfer in Large Scale Bioreactors 141

would need to be performed to obtain the necessary information regarding the


degree of backmixing actually existing in the reactor.

Exit Gas Exit Liquid c


FG3 A A C L3
d \_^
Gas

L3 L2
G2

Gas

'LI
G1
FL+FB

Gas

F
I G- C GO \\>cu>
Inlet Gas Liquid Feed

Figure 5.18. Stagewise approximation for stirrer regions in multi-stirrer tank.

To model this system, the liquid-phase impeller zones are assumed to be well-
mixed, and the plug-flow gas is described by a series of well-mixed phases,
together with an arithmetic-mean, concentration-driving-force approximation.
Here the flow rates and mass transfer coefficients are assumed constant.

Stage 1:

dCLi
VL -ar = FLCLO + FBCL2 - (FL + FB) CLI +
+ K L a(CLi*-C L i)VL + riV L

V dCGl
G
"""' xx-<
-CGI) -KLa(CLi -CLI)VL
f\ \ -rr- //~1 ^ /"I \A 7

where the plug flow nature of the gas is partially accounted for by
142 5 Mass Transfer

and
CLI
= - Qo2r

Stage 2:

dCL2
VL -gf = (FL+ FB) CLI + FBCL3 - (FL+ FB) CL2 - FB CL2

K L a(C L 2*-C L 2)V L

= FG (CGI - CG2) - KLa (CL2* - CL2)VL


where
CL2. .

CL2
and n
r2 = -Qo2mK 0 + CL2

Stage 3:

dCL3

K L a(C L 3*-CL3)V L + r 3 V L

V
G - CG3) - KLa (CL3*- CL3) VL
where
CL3. .
CL3
and = -QO2n K0 + CL3

The above equations describe the dynamic oxygen concentrations in the multi-
impeller continuous bioreactor. Note that the liquid phase balances for the two
end stages 1 and 3 differ from that of the intermediate stage 2, owing to the
5.4 Models for Oxygen Transfer in Large Scale Bioreactors 143

absence of any backmixing flow contribution exterior to the column. A batch


reactor would be described by setting the liquid flow, FL, equal to zero. Since
the biomass balance and growth kinetics are not included here, the solution
would be valid at only one time during the fermentation, corresponding to the
assumed value of Qo2m> which is proportional to the value of X existing at that
time. Variations in X are, however, easily incorporated into the model by
adding cell and substrate balance equations.
Diffusion and Biological Reaction in
Immobilized Biocatalyst Systems

The retention and immobilization of enzymes and cells usually requires the
presence of an additional solid carrier phase or flocculant cell mass. As
illustrated in Fig 6.1, in order to reach a reaction site, substrate S must first be
transported by convection from the bulk liquid to the exterior stagnant film
(point A). Then transport by diffusion must occur through the film (from A to
B) to the surface of the carrier (point B), where surface reaction can take place.
If further reaction sites are available within the carrier matrix, an additional
internal diffusion path (from B to C) is then also required. Similarly product P,
formed within the carrier matrix, must diffuse out of the matrix towards the
surface, and then away from the surface via the external mass transfer laminar
film to the bulk liquid.

Concentration
Diffusion film

i
Bulk Solid
liquid carrier

A B B

Figure 6.1. Concentration profiles for a biocatalyst immobilized on a solid carrier.

The stagnant film and the immobilization matrix constitute mass transfer
resistances which may slow the overall reaction rate, since reaction cannot
proceed at a rate greater than the rate at which substrate is supplied by the
mechanism of diffusion. The diffusional mass transfer process via the external
film is referred to as external mass transfer. Since the reaction site may often
be located within a gel, a porous solid, biofilm or biofloc, the transfer of
substrate or substrates from the exterior surface of the biocatalyst to reaction
sites, located within the internal structure of the carrier, is also usually
necessary. This process is therefore referred to as internal mass transfer or
intraparticle transfer. In what follows, external transport and internal transport

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
146 6 Diffusion and Biological Reaction in Immobilized Biocatalyst Systems

are considered separately, although, of course, the two effects can exert a
combined effect in reducing the effective reaction rate, compared to that which
would be obtained if there were no diffusional limitations.

6.1 External Mass Transfer

Fig. 6.2 illustrates the substrate concentration profile, existing in the very near
region of an immobilized biocatalyst surface, supported on a non-porous
carrier. Also shown is the idealized concentration profile, as represented by the
film theory. As previously discussed in Sec. 5.2, the "film theory" assumes the
presence of a stagnant layer of liquid to exist at the solid-liquid interface. This
stagnant region is termed the diffusion film or Nernst-diffusion film and
constitutes the external resistance to mass transfer. It thus determines the rate of
supply of substrate to the surface, for subsequent reaction.

Substrate cone.

SA A

Figure 6.2. External diffusion model of substrate transport to a reactive enzyme immobilized
on a solid surface.

The rate of supply of substrate to the surface is defined by mass transfer


considerations, such that the mass flux to the catalyst surface is given by,

JS = ks L (SA-S B )
6.1 External Mass Transfer 147

where, js is the mass flux (mol/m2 s), ksL is the mass transfer coefficient (m/s)
and SA, SB are the substrate concentrations for the bulk and surface conditions
[mol/m3], respectively.
The steady-state balance can be written for the transport-reaction process,

(Rate of supply by diffusion) = (Surface reaction rate)


k
SL (SA - SB) = ks SB = rapp

In the following treatment, the surface reaction is assumed to be first-order,


such as found for a biocatalytic reaction with Michaelis-Menten kinetics and
S KM- The apparent reaction rate per unit surface area, r app (mol/m2 s), is
equal to the rate of both processes.
Solving the equation, for the surface concentration, SB,

SB = -SA
and hence

= ksSB=U^
l k S+ k SL

Two extreme conditions can be identified:

1) For ks/ksL 1, SB approaches zero, and the reaction is completely mass


transfer controlled, with rapp =

2) For ks/ksL .! SB approaches SA, and the reaction is kinetically


controlled, with an apparent rate equal to that defined by the reaction
kinetics, with rapp = ks SA-

The intermediate situation is given by the full equation, for which the apparent
reaction rate is influenced by both the true kinetic rate constant ks and by the
diffusional mass transfer coefficient ksL-
For a zero-order reaction:
k S L(S A -S B ) = ks
where ks is now a zero-order kinetic rate constant. The concentration at the
reaction surface SB is thus,
SB = S A -ks/k S L
148 6 Diffusion and Biological Reaction in Immobilized Biocatalyst Systems

which indicates that the ratio of the magnitudes of the kinetic rate constant to
that of the mass transfer coefficient determines SB- If the reaction is zero-
order, the overall order of reaction rate is not influenced by diffusional
considerations, but the effective rate will still be reduced, owing to the lowered
concentration SB-
For Michaelis-Menten kinetics, which encompass the range between effective
zero and first order reaction kinetics, the relation between rate of supply and
the rate of reaction becomes,
kSL (SA - SB) = : > - = r app

After rearrangement, the resulting quadratic equation can be solved for SB, with
the solution indicating that, in general, the magnitudes of all the coefficients
can influence the overall reaction rate and also that the external transfer can
change the overall observed reaction kinetics. Thus they no longer follow the
Michaelis-Menten form with respect to bulk concentration, and the apparent
kinetics can differ substantially from the intrinsic true reaction kinetics. Under
these conditions, it is no longer correct to equate the Michaelis-Menten
constant, KM, to the substrate concentration at which the observed reaction rate
is equal to the half of the maximum observed rate. This can be most easily
seen from the above equation; when SB is low, the effective surface rate reduces
to the form, rapp = vm SB/KM- The overall reaction rate then becomes,

= (v m /K M )S A
app
(v m /K M k S L ) + l

showing that the apparent rate of reaction depends on the magnitude of the
mass transfer coefficient
It is only possible to measure true reaction kinetics, by operating
experiments in a truly kinetic regime, such that any influence of the external
diffusional mass transfer is negligible. This can be achieved by ensuring that
the ratio of vm/ksL is sufficiently low. Under these conditions,

which are the intrinsic Michaelis-Menten kinetics.


The ratio can be made low by increasing the mass transfer coefficient, k$L,
and by increasing the mass transfer rate enhancing parameters, such as flow
velocity and stirring speed. Conversely those factors affecting the maximum
reaction rate, vm, should be decreased, for example enzyme loading and
temperature.
The regimes of possible external mass transfer influence on the observed
kinetics are summarized in Table 6.1, together with the important parameters.
6.1 External Mass Transfer 149

Table 6.1. Characteristics of overall reaction rate influenced by external


transfer.

Regime of operation Reaction parameter having an influence


on overall rate

Transfer control Temperature (slight influence due to viscosity and


diffusion rate). Stirring speed in tank. Flow in packed
and fluidized beds. S in bulk liquid.

Kinetics control S in bulk liquid. Enzyme loading on surface.


Temperature.

Intermediate regime All of above.

6.2 Internal Diffusion and Reaction within


Biocatalysts
Reactions with enzymes and whole cells entrapped or immobilized in a porous
solid matrix will be subjected to a mass transfer influence. Example systems
are whole cells immobilized in alginate, enzymes adsorbed on ion-exchange
resins, or naturally occurring biological films on surfaces or flocculated
biomass. In the case of a biological film attached to an impermeable solid, the
substrate can enter from only one surface, as shown in Fig. 6.1, through the
diffusion layer A-B and into the biocatalyst B-C.
In the case of an alginate bead, a biofloc or its two-dimensional
approximation, substrate can enter from opposing directions, as shown below in
Fig 6.3. In this case, the diffusion will result in a symmetrical, steady state
concentration profile. The case of complete penetration of substrate through
the biofloc is shown by the solid line. Whereas an incomplete penetration, as
shown by the dashed line, results in the center of the film being completely
ineffective, in terms of reaction capability. Note that in this case, the effects of
external diffusion are neglected.
The uptake of substrates within solid material requires transport by a
diffusional process. The driving force for diffusion is a gradient in
concentration, and the diffusional flux is given by Pick's law,
150 6 Diffusion and Biological Reaction in Immobilized Biocatalyst Systems

dCA
jA = -D A
dZ

with j having units of kg/m2 s.

'AO
'AO
Diffusion
Diffusion

'AO

Figure 6.3. Internal concentration profiles in a symmetrical rectangular biocatalyst matrix.

If a reaction occurs within the matrix, a concentration gradient will be


established as a result of the simultaneous diffusion and reaction processes.
The reaction rate at each position, being usually a function of concentration,
will vary, and the overall or apparent reaction rate per unit volume of matrix
(kg/s m3), rapp, will be determined by the transfer rate at the surface (kg/s),

/Apparent rate\ =
/Rate of substrate^ /Net rate of reaction^
Vin bulk liquid) \entering matrix ) = v within matrix )

Vr a p p = (j|z=o)A = r a v g A L

where the units of each term are kg/s. Here ravg represents an average value in
the matrix, which will increase with higher internal substrate concentrations.
Regarding the influence of diffusion for a particular situation, it is possible
to arrive at some quantitative guidelines without considering any mathematical
details. Obviously the concentration profiles are caused by a competition
between reaction and diffusion. The ratio of the maximum intrinsic reaction
rate (not influenced by transfer) to maximum diffusion rate provides a useful
dimensionless parameter,
6.2 Internal Diffusion and Reaction within Biocatalyst 151

r
max A L r(C 0 ) A L maximum reaction rate
JA D (Co/L) A ~ maximum diffusion rate

For first order reaction, r = k CQ, this dimensionless group becomes k L2/D,
and for zero order reaction, r = k, it is k L2/D CQ.
Therefore for any kinetic form of equation, the distance coordinate or length
of diffusion path, L, plays an important role since the ratio of maximum
diffusion rate to maximum reaction rate varies according to L2. The higher the
value of this ratio, the greater in magnitude are the substrate gradients. With
this qualitative feeling for diffusion-reaction phenomena, more quantitative
aspects can be considered.

6.2.1 Derivation of Finite Difference Model for


Diffusion-Reaction Systems

Diffusion with biological reaction can be treated by mathematical modelling,


and from this it is possible to develop equations describing changes of
concentration, with respect to both time and position. The same technique of
finite differencing is used as in the modelling of the dynamic behavior of
tubular bioreactors, Sec. 4.2.2.
Consider the case where the substrate varies from a concentration SQ in the
bulk liquid to some concentration, at the position L (a wall or the center of a
symmetrical particle). At the center by symmetry or at a wall, owing to the
absence of diffusion into the wall, the concentration gradient must be zero.
The actual continuous concentration profile, through the slab, may be
approximated by a series of increments, as indicated in Fig. 6.4 and by a series
of biocatalyst matrix elements as shown in Fig. 6.5.

Liquid

Figure 6.4. Finite differencing a solid, showing concentration gradient approximation.


152 6 Diffusion and Biological Reaction in Immobilized Biocatalyst Systems

jn-1 In I n+1
n-1 > ,
n n+1

Figure 6.5. Series of volume elements connected with diffusion fluxes.

A magnified view of element n is shown in Fig. 6.6, where the flux, jn, depends
on the local concentration gradient, and the reaction rate, rsn, depends on the
local concentration in element n.

Jn-1

AZ

Figure 6.6. A single element n of volume AV and thickness AZ, showing the diffusion fluxes.

A component mass balance is written for each segment and for each
component as

/Accumulation^ /Diffusion^ /Diffusion\ / Production \


=
V rate ) \ rate in ) ~ \ rate out ) + \rate by reaction/

A AZ
dSn
~dT = Jn-1 A - jn A + rSn A AZ

Using Pick's law in the difference form,

(Sn-i ~ Sn)
s
and similarly for Jn gives,

n-l ~ Sn)
+rSnAAZ
6.2 Internal Diffusion and Reaction within Biocatalyst 153

Dividing by A AZ,

dSn (S n ,j -2S n + S n+ i)
s 5 + rsn
T

The equivalent partial differential equation is,

as a2s + rs

The finite-differenced forms of the model equations, however, are especially


suitable for simulation programming. Thus, N equations are obtained, one
substrate balance equation for each element, and these are solved
simultaneously. Note that the boundary conditions, for elements 1 and N, must
be described separately.
For the above case, the boundary conditions are dS/dZ = 0 at Z = L and
S = SQ at Z = 0. Thus the equations for the first and last elements must be
written accordingly, as shown in simulation example BIOFILM, Sec. 8.7.1.
Note also that it would be also possible, in principle, to include external
diffusion effects, by formulating a boundary condition, balance for the first
element as:

/Accumulation^ =
/External trans- A /Diffusion^ f Production \
V rate / V port rate in ) ~ \ rate out ) + Vrate by reaction/

where the external transport rate through area A is,

Q = k S L (So-Si)A

where SR is the bulk reactor concentration.

Coupling the Biocatalyst Matrix to the Reactor Liquid.


The biocatalyst diffusion model can be combined with a well-mixed tank
model, as shown in Fig. 6.7. The bulk liquid-phase component balances take
the form:
^= |(S F -SO)-Jsa| z = 0
where a=A/V,
, A dS ,
Js a| z=o = - DS v dZ I z=o

where,
154 6 Diffusion and Biological Reaction in Immobilized Biocatalyst Systems

dS -
dZ I z=o =

The direction of the positive diffusion is into the biofilm, since dS/dZ is
negative and (SR - Si) /AZ is positive. Here SR corresponds to So in Fig. 6.4,
and Si is the concentration in the first element. The boundary condition is
S = SR at Z = 0.

F,SF F,SO

Figure 6.7. Coupling the biofilm model to the continuous tank model.

In this way it is possible to simulate immobilized biocatalyst performance in a


single tank or in a column by using a tanks-in-series model, The simulation
example BIOFILM, Sec. 8.7.1, demonstrates this approach.

6.2.2 Dimensionless Parameters from Diffusion-


Reaction Models

There are several advantages of formulating model equations in dimensionless


form. The number of variables in the model is reduced, thus reducing the
number of experiments or simulations required to investigate all combinations.
It is also possible, on the basis of the numerical values of the parameters, to
access the relative importance of certain terms. Finally, the dimensionless form
makes the solution much more generalized because the units of the individual
quantities are no longer important.
The governing dimensionless parameters can be obtained by re-examining
the defining model equations and arranging them such that the variables range
6.2 Internal Diffusion and Reaction within Biocatalyst 155

only between the values of zero and unity. Thus new dimensionless variables,
S=S/S 0 , Z = Z/L and dimensionless time, f = t / ( L I D ) , can be defined.
Substituting these new variables into the diffusion-reaction, partial differential
equation, for the case of a first order biochemical reaction, gives,

as
S 2 =
(L /D8)at
or
as _ _ _
at - az2 " DS s

Thus the solution depends only on the value of [lq L2/DsL which is a
dimensionless diffusion-reaction parameter. For zero-order reaction the
equation becomes,

as a2s2
at " az
where ko L2/Ds SQ is the governing parameter.
It is seen that the dimensionless parameters in the model have the same form
and significance as was derived from the qualitative reasoning presented earlier.
For heterogeneous reaction systems this dimensionless group is known as the
Damkohler Number, and its square root is called the Thiele Modulus.
In the above equations, all the terms, excepting that of the reaction term, have
dimensionless parameters of unity. On this basis, it can be said that if the
reaction parameter for a first order reaction, [ki L2/DsL has a value of 1.0 or
greater, then the reaction will have a large effect on the solution, that is, on the
concentration gradients. Similarly, ko L2/D$ SQ will govern a zero order
reaction. Such "order of magnitude analysis" is important for physical
understanding and also to obtain information from differential equations
without having to actually develop an analytical solution.
Dimensionless formulation of equations is also explained in the simulation
example VARVOL, Sec. 8.3.1, and KLADYN, Sec. 8.5.5.

6.2.3 The Effectiveness Factor Concept

The relative influence of diffusion on biochemical reaction rate, can be


expressed by means of an effectiveness factor, T|, where,
156 6 Diffusion and Biological Reaction in Immobilized Biocatalyst Systems

actual apparent rate


T| = rate at bulk liquid concentration

Solutions of the above diffusion reaction model are available in the literature
for simple reaction orders (Satterfield and Sherwood, 1963). In Fig. 6.8 the
values of t| for zero, first and second reaction order have been plotted against
the Thiele Modulus ,<f>, where,
O = '\/L 2 kSo n - 1 /D

This figure shows that a zero order reaction is not influenced by concentration
gradients until the substrate falls to zero in the matrix, corresponding to O > \2
. The other reaction-types are influenced by low concentrations, as the curves
for T| indicate. It is seen, for example, that for a first order reaction a value of
O = l corresponds to t| = 0.8.

1.0 - Incomplete penetration at O >V~2~

0.8 - A
A = zero-order
0.6 - B = first-order
C = second-order
0.4 -

0.2 -

0 = L(kSo n ' 1 /D) 172

Figure 6.8. Effectiveness factor TJ versus the dimensionless reaction/diffusion parameter


(Thiele Modulus O ) for reactions in a flat film with diffusion from one side (after Satterfield and
Sherwood, 1963).
6.2 Internal Diffusion and Reaction within Biocatalyst 157

6.2.4 Case Studies for Diffusion with Biological


Reaction

6.2.4.1 Case A. Estimation of Oxygen Diffusion Effects in a


Biofilm

For a biofilm or floe, whose oxygen uptake might be taken as a constant (zero
order), the corresponding group would be [L2 qo2 Xbi0fiim/D Co2L where qo2
Xbiofiim corresponds to the rate constant k and the oxygen concentration in the
outside liquid phase is Co2- Note that Xbiofiim is the biomass per unit of
biofilm volume and is not easy to measure. Substituting values obtained from
an aerobic biofilm nitrification experiment gives,

2 L2 qQ2 X (0.01 mm2) (80 mg O2/L min)


0 =
= D C02 (0.1 mm2/min) (8 mg O2/L) = l

For this order-of-magnitude analysis, the value of 1.0 can be used to separate
the regions of reaction and diffusion dominance. Thus it is seen if L = 0.1 mm,
then the dimensionless group will have a value of 1.0, and it could therefore be
expected that a film or floe thickness greater than 0.1 mm would be oxygen
limited. From the exact solution, as seen in Fig. 6.8, gradients would appear for
a zero order reaction at a value of <&2 = 2.0, instead of 1.0. This example
shows how the Thiele Modulus can be used to make useful estimates for
diffusion reaction problems, providing rate and diffusion data are available.

6.2.4.2 Case B. Complex Diffusion-Reaction Processes


(Biofilm Nitrification)

Nitrification reactions, considering only the substrate conversion reactions and


ignoring the slow organism growth processes, the reactions can be written as,

NH4+ + 3/2O 2 ->

N02- + 1/2 02 -> N03


158 6 Diffusion and Biological Reaction in Immobilized Biocatalyst Systems

The oxygen requirements for the first and second steps can be related to the
nitrogen content of NH4+ and NC>2~. These values are si = 3.5 mg 62 / mg
NH4+ - N and 82 = 1.1 mg 62 / mg NCV - N. The low yields and low growth
rates make it unnecessary to consider growth requirements and kinetics. In
previous work (Tanaka and Dunn, 1982) the intrinsic substrate uptake kinetics
for the two steps were shown to have a double Monod form for the first step,

rNH4 = v m i
. K NH4
and for the second step,

rN02 = vm2

where v m i and vm2 represent the maximum rates for a particular biomass
concentration and the chemical symbols represent concentrations.
Considering the diffusion phenomena in the biofilm to be represented by
one-dimensional diffusion with quasi-homogeneous reaction, differential
balance equations can be written for all reactants and products to describe the
concentration profile in the film. Proceeding as described in Sec. 6.2.1, a
component mass balance is written for segment n and for each component:

(Accumulation^ _ (Diffusion^ _ (Diffusion^ ( Production ^


^ rate ) ~ \ rate *n ) ~ \ rate out J Vrate ^ reactin J

and the equivalent partial differential equation is obtained by letting AZ


approach zero as
3S 32S

Applying this to each component gives the following balances:

For NH4+,
3NH4+ a2NH4+
= DNH4 -wo - TNH4

For NO2",
aNQ2- = D
~3r NO2 ^2 + rNH4 - TNO2
6.2 Internal Diffusion and Reaction within Biocatalyst 159

For NO3%
3N03 32N03

For O2,
3202
" s l rNH4 ~

The stoichiometric oxygen requirements for the first and second reaction steps
are given by si and s2. The boundary conditions used represent the bulk liquid
phase or reactor concentrations and the zero gradient at the biofilm- solid
interface, as discussed earlier.
These equations can be written as differential-difference equations using the
finite-differencing technique (Sec. 6.2.1). Thus for each of N increments, four
component balances will be needed. Three simulation examples, BIOFILM,
ENZDYN, CELLDIFF in Sec. 8.7, demonstrate this approach.
This system was also analyzed in terms of dimensionless variables. A
comparison of the resulting dimensionless NH4+ and O2 balances reveals that,
when the second reaction is neglected, the equations are identical if
DNH4 = Do2 and if,
Q2R

where the subscript R refers to the concentrations in the bulk reactor liquid.
Under these conditions, to a good approximation, the penetration distances
of O2 and NH4+ would be the same. The ratio O2R/NH4+R, which can be varied
according to the reactor operating conditions, can thus be used as a criterion to
evaluate whether NtLj."1" or O2 might be penetration-limiting. The O2R/NH4+R
criterion indicates which component can be limiting, O2 if the ratio is less than
3.5 or NH4+ if the ratio is greater than 3.5. These conditions are not sufficient
for limitation, but indicate which component would be limiting. Simulation
results from a model that was developed using finite-differencing demonstrates
this phenomenon. The profiles in Fig. 6.9, are for the case O2R/NH4+R = 0.07.
160 6 Diffusion and Biological Reaction in Immobilized Biocatalyst Systems

NH 4 ,NO3,N0 2 (mg/L) 0 2 (mg/L)

100
H 20

80

60
NO
10
40

20

Figure 6.9. Steady state profiles for constant bulk concentrations showing incomplete
oxygen penetration.

Coupling the liquid and biofilm for a batch nitrification reactor as explained in
Fig. 6.7, gave the results in Fig. 6.10. Here the influence of oxygen limitation
caused the oxygen in the midpoint of the film to rise as the nitrogen substrates
were successively consumed.

NH4 NO 3, NO2" (mg/L) 0 2 [mg/L]

100 80
12
80 64
10
60 48 8

40 32 6
4
20 16
2
0 0 0
120
t (min)

Figure 6.10. Simulated biofilm nitrification profiles in a batch reactor. The N-component
concentrations are in the bulk liquid. O2 is in the midpoint of the biofilm and indicates
limitation during the first 60 minutes.
7 Automatic Bioprocess Control
Fundamentals

The purpose of automatic process control is to maintain time-dependent


changes of the relevant process variables (deviations, errors), within prescribed
limits and without a direct action of an operator. Process control may be
considered as a corrective action involving three steps:

1. Measuring the variable to be controlled (controlled variable)


2. Comparing the measurement with the desired value (set point)
3. Adjusting some other variable (manipulated variable) that has a direct
effect on the controlled variable, until the set point is reached.

A number of advantages or reasons for process control may be listed, which


include uniform and higher quality products, safety, increase of productivity,
minimization of waste, optimization, freeing the labor force from drudgery and
danger, and decrease of labor costs.
Obviously, process control is highly dynamic in nature, and therefore its
modelling requires the solution of sets of differential equations, and it is
therefore highly suited to solution by digital simulation. A brief introduction to
the basic principles of process control required for solution of simple
simulation examples is given here.

7.1 Elements of Feedback Control


The simple temperature control of a fermenter shown in Fig. 7.1 illustrates the
essential idea of any automatic control system that the process and the
controller form a closed loop, which usually functions in a feedback fashion.

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
162 7 Automatic Bioprocess Control Fundamentals

Controller
mechanism
Desired
value

I Measured
I value
I
Thermo-
couple

Fermenter

Figure 7.1. Simple feedback control system.

The components of such a control system can be best understood using a


generalized block diagram (Fig. 7.2). They are the process itself, the
measuring element (thermometer), the controller (including a comparator),
the final control element (automatic control valve) and the transmission lines.
The information on the measured variable, temperature, taken from the system
is used to manipulate the flow rate of the cooling water in order to keep the
temperature at the desired constant value, or set point.

Controller Load
mechanism
1 Com)jarator Controller Actuator Process
_ i' Controlled
Desired+c
value | *A/
1
1
J
$~ 1
K
1
1
-*U variable

Measuring
element
Measured k A A A A
A A A A A

value
^A^A^AAA^A

Figure 7.2. Block diagram of the feedback control system in Fig. 7.1.
7.2 Types of Controller Action 163

Similar temperature control systems are given for a simple water heater,
TEMPCONT, in Sec. 8.6.1 and for a batch fermenter, FERMTEMP, in Sec.
8.6.2

7.2 Types of Controller Action

7.2.1 On-Off Control

The most common and simplest type of control is an on-off or two position
action, sometimes called discontinuous control (Fig. 7.3). An example is a
contact thermometer, which closes or opens the heater circuit. The controller
changes the value of the controller output, or the manipulated variable, from
one extreme to the other, when the controlled variable moves above or below
the set point. This leads to oscillations that could become very fast, depending
on the speed of response of the process. The real on-off controller has
therefore a built in feature called a differential gap or a dead zone. It is a small
interval on either side of the set point, within which the controller does not
respond. When the controlled variable moves outside the dead zone, the
manipulated variable goes on or off. This is illustrated in Fig. 7.3. Such shifts
from the set point are known as offset. Such a controller is simple and
inexpensive, but the oscillatory nature of the action and the offset make it very
imperfect.
The usefulness of this type of control was demonstrated for a biological,
sequential batch process by Hediger and Prenosil, (1985). More sophisticated
function control modes consider the magnitude and time behavior of the
control error. Three principal functional modes of control generally employed
for process control are proportional (P), integral (I) and derivative (D) control.
164 7 Automatic Bioprocess Control Fundamentals

100

Manipulated
variable
0

Control Xl___4/_
variable .Set point

Differential gap

Figure 7.3. On-Off controller with differential gap or dead zone.

7.2.2 Proportional (P) Controller

The produced output signal P is proportional to the detected error, e, according


to
P = PO + Kp 8

where Kp is the proportional gain, and P0 is the controller output for zero error.
An example of a level control is shown in Fig. 7.4. The action of this type of
controller is shown in Fig. 7.5 and 7.6.

Figure 7.4. Response of proportional-mode controller to sinusoidal error input.


7.2 Types of Controller Action 165

Systems with proportional control often exhibit pronounced oscillations and


for sustained changes in load the controlled variable attains a new equilibrium,
steady-state position, or control point. The difference between this point and
the set point is the offset (Fig. 7.5) Integral and derivative modes are used
mostly in combination with the basic proportional control mode. The
simulation Example INHIB, Sec. 8.5.2 includes the application of this simple
control mode in the recommended exercises.

F + AF

Setpoint

Figure 7.5. An example of proportional-mode level controller illustrating offset.

7.2.3 Proportional-Integral (PI) Controller

Pi-controller, sometimes called automatic reset, produces an output signal


related to the error by

l
K
P = Kpe+ ?
TI

or for e = constant
dP

for t=Ti
dP
dF = K P
166 7 Automatic Bioprocess Control Fundamentals

where i\ is the integral time constant or reset time. This is the time required to
enable the controller to repeat the initial proportional output action (Fig. 7.6).
The integral part of the control mode eliminates the offset and it is especially
useful for correction of very small errors because the controller output P will
continue to change as long as an error persists. This can be understood by
considering a constant error, which would cause P to increase linearly (Fig. 7.6)
at a rate proportional to the error. This type of controller is found most often
and the simulation examples TEMPCONT, FERMTEMP, TURBCON and
CONTCON in Sec. 8.6 demonstrate the use of this control mode. The
examples also show the ease by which the programming of the PI controller
equations is made using the simulation language.

2Kp
Controlled
output PI action
Kp

P action

Error
signal

Figure 7.6. Response of a proportional-integral controller to a unit step change in error.

7.2.4 Proportional-Derivative (PD) Controller

A controller with derivative function projects the error in the immediate future
and the controller output is proportional to the current rate of the error change.
The output signal varies only if the error is changing.
7.2 Types of Controller Action 167

d
P = P0 + Kp e + Kp TD gj-

where TD is the derivative time constant. A PD-controller output is compared


with pure proportional mode in Fig. 7.7.

P (alone)

Figure 7.7. Response of a PD-Controller to a constant rate of decrease in error. Comparison


with P and PID modes.

The drawbacks of the derivative control mode standing alone are that a constant
error (e ^ 0) gives no response at all, since de/dt = 0, and therefore an
unnecessarily large response might occur as a result of small but fast error
changes.

7.2.5 Proportional-Integral-Derivative (PID)


Controller

In industrial practice it is common to combine all three modes, sometimes


termed as Proportional-Reset-Rate-Control. The action is proportional to the
error (P) and its change (D) and continues if residual error is present (I):
168 7 Automatic Bioprocess Control Fundamentals

KD r de
P = P0 + Kp e + ^ J dt+ Kp TD -gj-
TI
o
This combination gives the best control using conventional feedback
equipment. It retains the specific advantages of all three modes: proportional
correction (P), offset elimination (I) and stabilizing, quick-acting character,
which is especially suitable to overcome lag presence (D). The action of a PID-
controller as a response to a ramp function is shown in Fig, 7.7. The
performance of the different feedback control modes can also be seen in Fig.
7.8.

Controlled
variable

/ Uncontrolled
response

Figure 7.7. Response of controlled variable to a step change in error using different control
modes.

The selection of the best mode of control depends largely on the process
characteristics. Further information can be found in the recommended texts
listed in the reference section. Simulation methods are often used for testing
control methods.
7.3 Controller Tuning 169

7.3 Controller Tuning

The purpose of controller tuning is to choose the controller constants, such that
the desired performance is obtained. This usually means that the control
variables should be restored in an optimal way, following either a change in the
set point or as a result of an input disturbance to the system. Thus the
controller constants can be set by experimentation. A rational basis for such
experimental tuning is given in what follows. Other methods for tuning
combine process dynamic experimentation with theoretically-based control
methods; some of the standard methods are also described below. The
simulation example TEMPCONT, Sec. 8.6.1, provides exercises for controller
tuning using the methods explained below.

7.3.1 Trial and Error Method

Controllers can be adjusted by changing the values of gain Kp, reset time i\ and
derivative time ID- By experimentation, either on the real system or by
simulation, the controller can be set by trial and error. Each time a disturbance
is made the response is noted. The following procedure might be used to test
the control with small set point or load changes:
1. Starting with a small value, Kp can be increased until the response is
unstable and oscillatory. This value is called the ultimate gain KPQ.
2. K p is then reduced by about 1/2.
3. Integral action is brought in with high i\ values; they are reduced by
factors of 2 until the response is oscillatory, and tj is set at 2 times this
value.
4. Include derivative action, increase ID until noise develops and set ID at
1/2 this value.
5. Increase Kp in small steps to achieve the best results.

7.3.2 Ziegler-Nichols Method

This method is an empirical open-loop tuning technique, obtained by


uncoupling the controller. It is based on the characteristic curve of the process
response to a step change in manipulated variable, equal to A. This response is
called a process reaction curve, whose magnitude is B. The two parameters
170 7 Automatic Bioprocess Control Fundamentals

important for this method are given by the normalized slope of the tangent
through the inflection point, S = slope/A and by its intersection with the time
axis (lag time TL), as determined graphically in Fig. 7.9. The actual tuning
relations which are based on empirical criteria for the "best" closed-loop
response are given in Table 7.1.

Manipulated
Variable
X

Slope

Time

Figure 7.9. Process reaction curve as a response to a step change in manipulated variable.

7.3.3 Cohen-Coon Controller Settings

Cohen and Coon observed that the response of most uncontrolled (controller
disconnected) processes to a step change in the manipulated variable was a
sigmoidal shape curve. This can be modelled approximately by a first order
system with time lag TL, as given by the intersection of the tangent through the
inflection point with the time axis. The theoretical values of the controller
settings obtained by the analysis of this system are summarized in Table 7.1
The model parameters for a step change A to be used with Table 7.1 are
calculated as follows:
K = B/A T = B/S

where B is from Fig. 7.9 and S is the slope at the inflection point/A.
7.3 Controller Tuning 171

Table 7.1. Controller Settings Based on Process Responses

Controller Kp

Ziegler-Nichols

P 1/TLS
PI 0.9/TL S 3.33TL
PID 1.2/TLS 2TL TL/2

Co/ten-Coon

P T

TL \ 30 + 3TL/T
PI

4 TL \ 32 + 6TL/T

Ultimate Gain

P 0.5 Kpo
PI 0.45 Kp0 l/1.2fp0
PID 0.6 Kpo l/2fpo l/8fpO

7.3.4 Ultimate Gain Method

The previous tuning transient response methods are sensitive to disturbances


because they rely on open-loop experiments. Several closed loop methods
were developed to eliminate this drawback. One of them is the empirical tuning
method, ultimate gain or continuous-cycling method. The ultimate gain, Kpo,
is the gain which brings the system with the proportional control mode to
sustained oscillations (stability limits) of the frequency fpo, where l/f p o is called
the ultimate period. It is determined experimentally by increasing Kp from low
172 7 Automatic Bioprocess Control Fundamentals

values in small increments until continuous cycling begins. The controller


settings are then calculated from Kpo and f p o according to the tuning rules
given in Table 7.1.
While this method is very simple it can be quite time consuming in terms of
number of trials required and if the process dynamics are slow. In addition, it
may be hazardous to experimentally force the system into unstable operation.
Simulation methods can be very useful if a suitable model is available as was
shown by Heinzle et al. (1992) for a one and two stage anaerobic system, using
a kinetic model similar to the simulation example, ANAMEAS, Sec. 8.8.7.

7.4 Advanced Control Strategies

7.4.1 Cascade Control

In control situations with more then one measured variable but only one
manipulated variable, it is advantageous to use control loops for each measured
variable in a master - slave relationship. In this, the output of the primary
controller is usually used as a set point for the slave or secondary loop.
This may be relevant for some wastewater treatment plants where the high
concentration of some substrate may be toxic for the microorganisms. For
example, the simulation Example TURBCON, Sec. 8.6.3, could be easily
adapted to this situation if the substrate concentration were subject to significant
changes, as shown schematically in Fig. 7.10.
7.4 Advanced Control Strategies 173

m |
1
^ V^
lilllll

XI . .. N
! :
';S:'|||::||:|- illlll

/6oncentrationV/Concentriilonl
Controller )H f errand
x. ,s \ transmitter /
^
ill lit:

c Biomass
controller
J_
[ Turbidometer

Figure 7.10. Cascade control of a fermenter with toxic substrate.

The simulation example TURBCON, Sec. 8.6.3. could be modified similarly by


adding biomass as a measured variable. An interested reader may try to
implement the cascade control strategy in these simulation programs.

7.4.2 Feed Forward Control

Feedback control may never be perfect as it only reacts to the disturbances


which are measured in the system output. The feedforward method tries to
eliminate this drawback by another approach. Rather than using the process
output as the measured variable, this is taken as the measured inlet disturbance
and its effect on the process is anticipated by means of a process model. Thus
action is taken on the manipulated variable by the model, which relates the
measured variable at the inlet, the manipulated variable and the process output.
The success of this control strategy depends largely on the accuracy of the
model prediction. For this reason sometimes an additional feedback loop is
used.
Many of the continuous process simulation examples in Ch. 8 could be
altered in this fashion. It would be interesting to program an example using
simple kinetics for the feedforward control and to describe the "actual" system
with more complex kinetics. The discrepancies between the "simple" model
174 7 Automatic Bioprocess Control Fundamentals

prediction and the more complex "actual" process kinetics could then be taken
care of by a feedback control loop.

7.4.3 Adaptive Control

This control system can automatically modify its behavior according to the
changes in the system dynamics and disturbances. Especially systems with
nonlinear and unsteady characteristics call for use of this control strategy.
There are a number of actual adaptive control systems. Programmed or
scheduled adaptive control uses an auxiliary measured variable to identify
different process phases for which the control parameters can be either
programmed or scheduled. The "best" values of these parameters for each
process state must be known a priori. Sometimes adaptive controllers are used
to optimize two or more process outputs, by measuring these and fitting the
data with empirical functions, as employed on anaerobic treatment process, by
Ryhiner, et al. (1992).

7.4.4 Sampled-Data Control Systems

When discontinuous measurements are involved the control system is referred


to as sampled-data. Concentration measurements by chromatography would
represent such a case.

Controlled
variable

Figure 7.11. Sampled control strategy.

Here a special consideration must be given to the sampling interval T (Fig.


7.11). In general the sampling time will be short enough if the sampling
7.5 Concepts for Bioprocess Control 175

frequency is equal to 2 times the highest frequency of interest or T is equal to


0.5 times the minimum period of oscillation. When the sampling time satisfies
the above criteria, the system will behave as if it were continuous. Details of this
and other advanced control topics are given in specialized process control
textbooks, some of which are listed in the reference list.

7.5 Concepts for Bioprocess Control

Bioprocess control consists of establishing a strategy for the management of the


biocatalyst environment. In a natural environment microorganisms and cells of
higher organisms very rarely produce large amounts of products. In
biotechnological processes organisms are usually kept in a completely
"unnatural" environment. Control is therefore often necessary to induce them
to produce substances in economically important amounts.
Process optimization is closely linked with control. The objectives of
optimization and control may be to maximize productivity, final concentration,
yield or to minimize effluent concentration and energy costs. Although the
criteria for optimal processes differ widely, all bioprocesses need control and
automation to run under optimal conditions. The selection of control variables
strongly depends on the process and the final goal to be achieved.
Information about the dependency of biological rates, yields and selectivities
on environmental conditions is usually required, as given in Ch. 3. All
biological reactions have distinct temperature and pH optima, and all respond
to substrate concentrations. Therefore it is common to control these variables.
Heat is produced in all biological reactions and therefore temperature
control is necessary. In large-scale production, heat removal capacity may be
the limiting factor. It is important to maintain the temperature at an optimum
level. This is the theme of the simulation example, FERMTEMP, Sec. 8.6.2.
As discussed in Ch. 3, bioreaction rates usually follow the Arrhenius' Law below
the optimal temperature, which means that the growth rates can be expected to
increase exponentially with increasing temperature. Above the optimum
temperature, further temperature increases usually cause a dramatic decrease in
activity, mainly due to inactivation of enzymes. Also, temperature shocks may
be important since enzyme formation may often be induced by a shock at the
end of the exponential growth phase.
The variable pH has certain similarities with temperature because there
usually exists an optimum for biological activity; it can be relatively easily
measured and is often controlled. The biological rates also exhibit a maximum
at the optimal pH value, which is usually in the neutral pH 7 region. Again
control is often required since in almost all biological reactions acids (e.g.
lactic, pyruvic acid) or bases (e.g. NH3) are either produced or consumed.
176 7 Automatic Bioprocess Control Fundamentals

Biological rates usually depend on substrate concentration (e.g.: sugar,


mineral salts, oxygen, precursor, etc.) though in many cases kinetics are of
zero-order type above certain concentration levels. In the latter case control is
not important, since high concentrations will guarantee maximum rates. The
situation is more complex if process selectivity changes with substrate
concentration. The most well-studied process of this kind is Baker's Yeast
production, where high glucose concentration (> 100 mg L"1) and oxygen
limitation causes undesired ethanol formation. Substrates (e.g. mineral salts,
components of wastewater), precursors (e.g. in antibiotica production or
transformation processes) or products (e.g. ethanol) may be inhibiting or even
toxic at higher concentration levels. In such processes it is necessary to control
the concentration within certain limits.

7.5.1 Selection of a Control Strategy

The first step in controlling a process is to choose a control strategy. Simple


examples are the set-point control of constant temperature, pH, substrate and
precursor concentration. Table 7.2 gives examples of methods and strategies
for control of biological reactors. Much of the difficulty in control lies in
finding a suitable sensor. Calculated values using indirect measurements can be
very useful, e.g. measuring oxygen uptake to control substrate level. Often
variables such as pH or dissolved oxygen (DO) control can be used to
indirectly keep substrate concentration constant

Table 7.2. Examples of methods and strategies for the control of bioprocesses
(Heinzle and Saner, 1991).
Process Method Controlled Manipulated
and strategy variable(s) variable(s)

Baker's yeast Discrete RQ Glucose


production control (Gas analysis) feed rate

Baker's yeast Feedback RQ Glucose


production control (Gas analysis) feed rate

Baker's yeast Two point DO Feed rate,


production control (electrode) agitation speed
aeration rate
7.5 Concepts for Bioprocess Control 177

Table 7.2. (Continued).


Process Method Controlled Manipulated
and strategy variable(s) variable(s)
Ethanol PID Sucrose cone, by Feed rate
production enzyme thermister

Wastewater Adaptive DO Aeration rate


treatment control (electrode)

Diverse Various DO Stirring speed,


fermentations (electrode) gas composition,
aeration rate,
pressure

Bacillus subtilis Cascade DO Gas flow,


at low DO (electrode) valve setting

Penicillin Set-points Growth rate Feed rate


production (growth and (CO2 rate)
production)

a-Amylase by Feed profile Feed rate Feed rate


Bacillus amylo- (off-line)
liquefaciens

Fed-batch Feed profile Substrate and Feed rate


penicillin biomass cone.

Baker's yeast Feed profile RQ Feed rate

Cephalosporin C Profiled pH pH, temp. Alkali feed rate,


production and temperature (electrode, cooling water rate
control thermister)

Recombinant Conventional Growth rate Glucose and alkali


E. coli with added (pH) rate
glucose

Phenol oxidation Adaptive- Phenol uptake Flow rate


questing (O2 uptake)

For constant value or set-point control usually constant control parameters are
used. Because of non-linearities or varying process dynamics (e.g. exponential
178 7 Automatic Bioprocess Control Fundamentals

growth phase followed by production phase) control parameters of a linear


controller may be inadequate to control the process. It is therefore necessary to
adapt control parameters (e.g. proportional gain) according to the process
requirements. Minimization of an objective function can be used to guide the
adaptive tuning of the controller. A rather simple method uses a somewhat
empirical adjustment mechanism, which is driven by a secondary measurement.
This was used for DO control by Heinzle et al. (1986), in which the oxygen
uptake rate measurement was used to adjust the controller gain.
In addition to constant value control, optimal profile control may be applied.
The predefined optimal profile is then followed, which may be calculated from
off-line simulation and optimization. An example is found in the exponential
feeding profiles that can be calculated from the models for fed batch
fermenters in Ch 4.
If no suitable dynamic model is available and the process changes in
unpredictable ways, then on-line adaptive optimizing control may be useful.
This would however require measurements of the key inputs and outputs of the
process. An example is the optimization of a continuous anaerobic process by
Ryhiner et al. (1989) in which the methane and organic acids output rates were
correlated with the input flow rate. The optimization involved a compromise
between high methane rates and low organic acid concentration.

7.5.2 Methods of Designing and Testing the


Strategy

Selection of a control strategy and its parameters (e.g. for a PID controller)
may be difficult, since the process and controller dynamics are often not well
understood. If possible, it is useful to use dynamic models to select a control
strategy, and to use it for testing and tuning. An example with anaerobic
digestion is given by Heinzle et al. (1992). In Fig. 7.12 are shown the results of
a simulation and a corresponding experiment for the control of the propionic
acid concentration by manipulation of the feed flow rate.
7.5 Concepts for Bioprocess Control 179

Propionic acid [mg/l]


Feed flow [ml/min]

1
I

0 1 2 3 4

6(
Propionic acid [mg/l]
Feed flow [ml/min]
5(

4(

o
CD
O
"E
3( -

V2
i
I
20-

1(
2 3
Time [h]

Figure 7.12. Control of anaerobic digestion of whey wastewater. Simulation (A) and
experiment (B) of control after step change (Heinzle et al. (1992). Here the controlled variable
was the propionic acid concentration, and the manipulated variable was the feed flow rate.
References

References Cited in Part I

Archer, D.B. (1983) The Microbiological Basis of Process Control in


Methanogenic Fermentation of Soluble Wastes. Enzyme Microb. Technol. 5,
570-577.

Aris, R. (1989) Elementary Chemical Reactor Analysis. Butterworths,


Boston.

Atkinson, B. and Mavituna, F. (1991) Biochemical Engineering and


Biotechnology Handbook. 2nd. Ed., Stockton Press, New York.

Bailey, J.E. and Ollis, D.F. (1986) Biochemical Engineering Fundamentals.


2nd. Ed., McGraw-Hill, N.Y.

Blanch, H.W. and Dunn, I.J. (1973) Modelling and Simulation in Biochemical
Engineering. Adv. Biochem. Eng, 3, 127-165.

Dekkers, R.M. (1983) State Estimation of a Fed-batch Baker's Yeast


Fermentation., in: Modelling and Control of Biotechnological Processes, (Ed.:
A.Halme) Pergamon Press, Oxford, 73.

Denac, M., Miguel, A., Dunn, I.J. (1988) Modeling Dynamic Experiments on
the Anaerobic Degradation of Molasses Wastewater. Biotechnol. Bioeng. 31, 1-
10.

Dunn, I.J. and Mor, J.R. (1975) Variable Volume Continuous Cultivation.
Biotechnol. Bioeng. 17, 1805-1822.

Dunn, I.J., Shioya, S. and Keller, R. (1979) Analysis of Fed Batch


Fermentation Processes. Annals N.Y. Acad. ScL 326, 127-139.

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
References 181

Franks, R.G.E. (1966) Mathematical Modeling in Chemical Engineering. Wiley,


New York.

Franks, R.G.E. (1973) Modeling and Simulation in Chemical Engineering.


Wiley, New York.

Fredrickson, A.G., Megee, R.D., Tsuchita, HM. (1970) Mathematical Models


for Fermentation Processes. Adv. Appl. Microbiol. 13, 419-465.

Furukawa, K., Heinzle, E., and Dunn, IJ. (1983) Influence of Oxygen on the
Growth of Saccharomyces cerevisiae in Continuous Culture. Biotechnol.
Bioeng. 25, 2293-2317.

Gujer, W., Zehnder, A.J.B. (1983) Conversion Processes in Anaerobic


Digestion. Water Sci. Technol. 15, 127-167.

Harder, A., Roels, J.A. (1982) Application of Simple Structured Models in


Bioengineering. Adv. Biochem. Eng./Biotechnol. 21, 56-107.

Hediger, T. and Prenosil, I.E. (1985) Microprocessor Automated Sequential


Batch Process, Biotechnol. Progr.l, 216-225.

Heinzle, E. and Lafferty, R.M. (1980) Continuous Mass Spectrometric


Measurement of Dissolved H2, O2, and CO2 during Chemolitho- autotrophic
Growth of Alcaligenes eutrophus strain H 16. Eur. J. Appl. Microbiol.
Biotechnol. 11, 8-16.

Heinzle, E., Furukawa, K., Dunn, I.J., and Bourne, J.R. (1983) Experimental
Methods for On-line Mass Spectrometry in Fermentation Technology.
Bio/Technology 1, 14-16.

Heinzle E. and Dunn I. J. (1991) Methods and Instruments in Fermentation


Gas Analysis, in Biotechnology, Vol 4, (Ed.: H.-J.Rehm and R.Reed). VCH,
Weinheim, 27-74.

Heinzle, E. and Saner, U. (1991) Methodology for Process Control in Research


and Development, Ed. Pons, M.-N., in Bioprocess Monitoring and Control.,
Hanser, Munich. 223-304.

Heinzle, E., Dunn, IJ. and Ryhiner, G. (1992) Modelling and Control for
Anaerobic Wastewater Treatment, Adv. Biochem. Eng., Springer Verlag.

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
182 References

Imanaka, T., Aiba, S. (1981) A Perspective on the Application of Genetic


Engineering: Stability of Recombinant Plasmid. Ann. N. Y. Acad. Sci. 369, 1-
14.

Ingham, J., and Dunn, I.J. (1991) "Bioreactor Off-Gas Analysis", in "Bioreactors
in Biotechnology", Ed. A. Scragg, Ellis-Horwood, Chichester, 195-220.

Keller, R. and Dunn, I.J. (1978), Computer Simulation of the Biomass


Production Rate of Cyclic Fed Batch Continuous Culture, J. Appl. Chem.
Biotechnol. 28, 508-514.

Keller, R. and Dunn, I.J. (1978) Fed Batch Microbial Culture: Models, Errors,
and Applications. J. Appl. Chem. Biotechnol. 28, 508-514.

Keller, J., Dunn, I. J. and Heinzle, E. (1991). A Fluidized Bed Reactor for
Animal Cell Culture, in preparation, Biotechnol. Bioeng.

Luyben, W.L. (1973) Process Modeling, Simulation, and Control for Chemical
Engineers. McGraw-Hill.

Meister, D., Post, T., Dunn, I.J. and Bourne, J.R. (1979) Design and
Characterization of a Multistage, Mechanically Stirred Column Absorber.
Chem. Eng. Sci., 34, 1376.

Moes, J., Griot, M., Keller, J., Heinzle, E., Dunn, I.J., and Bourne, J.R. (1985) A
Microbial Culture with Oxygen-sensitive Product Distribution as a Tool for
Characterizing Bioreactor Oxygen Transport. Biotechnol. Bioeng. 27, 482-
489.

Moes, J., Griot, M., Heinzle, E., Dunn, I.J., and Bourne, J.R. (1986) A Microbial
Culture as an Oxygen Sensor for Reactor Mixing Effects. Ann. N. Y. Acad. Sci.
469, 482-489.

Mona, R., Dunn, I.J. and Bourne, J.R. (1979) Activated Sludge Process
Dynamics with Continuous TOC and Oxygen Uptake Measurements.
Biotechnol. Bioeng. 21, 1561-1577.

Moser, A. (1988) Bioprocess Technology, Springer, N.Y.

Mou, D.G. and Cooney, C.L. (1983) Growth Monitoring and Control through
Computer-aided On-line Mass Balancing in a Fed-batch Penicillin
Fermentation. Biotechnol. Bioeng. 25, 225-255.
References 183

Prenosil, J., Dunn, I.J., and Heinzle, E. (1987) Biocatalyst Reaction Engineering,
in: Biotechnology Vol.7a, (Ed.: HJ.Rehm and R.Reed) VCH, Weinheim, 489-
545.

Roels, J.A. (1983) Energetics and Kinetics in Biotechnology. Elsevier


Biomedical Press, Amsterdam.

Ruchti, G., Dunn, I.J., Bourne, J.R. and v. Stockar, U. (1985) Practical
Guidelines for Determining Oxygen Transfer Coefficients with the Sulfite
Oxidation Method. Chem. Eng. J. 30, 29-38.

Russell, T.W.F., Denn, M.M. (1972). Introduction to Chemical Engineering


Analysis. Wiley, New York.

Ryhiner, G., Dunn, IJ, Heinzle, E, Rohani S., (1992) Adaptive On-line Optimal
Control of Bioreactors: Application to Anaerobic Degradation, J. BiotechnoL.
22, 89-106.

Ryhiner, G., Heinzle, E, Dunn, IJ. (1991) Modelling of Anaerobic Degradation


and Its Application to Control Design: Case Whey, in: Dechema Biotechnology
Conferences Vol.3, (Ed.: Behrens, D. and Driesel, A.J.), 469-474.

Saner, U., Bonvin, D., Heinzle, E. (1990) Application of Factor Analysis for
Elaboration of Stoichiometry and its On-line Application in Complex Medium
Fermentation of B. subtilis, in: Dechema Biotechnology Conferences Vol.3,
(Ed.: Behrens, D. and Driesel, A.J.), 775-778.

Satterfield, C.N. and Sherwood, T.K. (1963) The Role of Diffusion in Catalysis.
Addison-Wesley, N.Y.

Shioya, S., Dang, N.D.P. and Dunn, IJ. (1978). Bubble Column Fermenter
Modeling: A Comparison for Pressure Effects. Chem. Eng. Sci., 33, 1025 -
1030.

Tanaka, H. and Dunn, IJ. (1982). Kinetics of Biofilm Nitrification. BiotechnoL


Bioeng., 24, 669 - 689.

Tanaka, H., Uzman, S., Dunn, IJ. (1981). Kinetics of Nitrification Using a
Fluidized Sand Bed Bioreactor with Attached Growth. BiotechnoL Bioeng., 23,
1683 - 1702.

Ziegler, H., Meister, D., Dunn, IJ., Blanch, H.W., Russell, T.W.F. (1977). The
Tubular Loop Fermenter: Oxygen Transfer, Growth Kinetics and Design.
BiotechnoL Bioeng. 19, 507.
184 References

Recommended Textbooks and References for Further Reading

Biochemical Engineering

Aiba, S., Humphrey, A.E. and Millis, N.F. (1973) Biochemical Engineering.
Academic Press, N.Y.

Atkinson, B. and Mavituna, F. (1991) Biochemical Engineering and


Biotechnology Handbook., 2nd. Ed., Stockton Press, New York.

Bailey, I.E. and Ollis, D.F. (1986) Biochemical Engineering Fundamentals.


2nd. Ed., McGraw-Hill, N.Y.
Blanch, H.W., Clark, D.S. (1996) Biochemical Engineering; Marcel Dekker,
N.Y.

Bu'lock, J. and Kristiansen Eds (1987) Basic Biotechnology Acad. Press,


London 1987.
Doran, P. M. (1995) Bioprocess Engineering Principles; Academic Press
Limited: London.

Glick, B.R., Pasternak, J.J. (1995) Molekulare Biotechnolgie; Spektrum,


Heidelberg.

Grady, C. P. L. and Lim H. C. (1980) Biological Wastewater Treatment, Marcel


Dekker.
Hastings, A. (1997) Population biology. Concepts and models; Springer, N.Y.

Heinrich, R., Schuster, S. (1996) The Regulation of Cellular Systems; Chapman


& Hall, New York.

Klefenz, H. (2002) Industrial Pharmaceutical Biotechnology; Wiley-VCH.

Ladisch, M.R. (2001) Bioseparations Engineering: Principles, Practice and


Economics. Wiley, New York.

Lee, J. M. (1992) Biochemical Engineering. Prentice Hall.

Moo-Young, M., Ed. (1985) Comprehensive Biotechnology, Vols. 1-4


Pergamon Press, Oxford.
References 185

Moser, A. (1988) Bioprocess Technology.Springer, N.Y.

Nielsen, J., Villadsen, J. (1994) Bioreaction Engineering Principles. Plenum


Press.

Pirt, S. John (1975) Microbe and Cell Cultivation. Blackwell Scientific Publ.,
Oxford.

Rehm H.J. and Reed R. Eds (1988) Fundamentals of Biochemical Engineering,


in Biotechnology, Vol. 2, VCH, Weinheim,

Roels, J.A. (1983) Energetics and Kinetics in Biotechnology. Elsevier


Biomedical Press, Amsterdam.

Schiigerl, K., Bellgardt, H. (Eds.) (2000) Bioreaction Engineering. Springer,


Berlin, Heidelberg.

Schiigerl, K. (1987) Bioreaction Engineering. Wiley, New York.

Schiigerl, K. (1994) Solvent Extraction in Biotechnology : Recovery of


Primary and Secondary Metabolites. Springer, Berlin.
Shuler, M. L., Kargi, F. (2002) Bioprocess Engineering. Basic Concepts;
Prentice-Hall.
Spier, R.E., Griffiths, J.B. (1985) Animal Cell Biotechnology. Vol. 1-3,
Academic Press.

Wang, D.I.C., Cooney, Ch.L., Demain, A.L., Dunnill, P., Humphrey, A.E., Lilly,
M.D. (1979) Fermentation and Enzyme Technology. Wiley, New York.

Wingard, L.B., Jr., Katschalski-Katzir and L. Goldstein Eds (1976-1983)


Applied Biochemistry and Bioengineering, Vols. 1-4 Acad. Press, London.

Bioreactor Design and Modelling

Asenjo, J.A., Merchuk, J.C. (1995) Bioreactor system design; Marcel Dekker,
N.Y.
Hannon, B., Ruth, B. (1997) Modeling Dynamic Biological Systems; Springer-
Verlag, New York.
186 References

Scragg, A.H. (1991). Bioreactors in Biotechnology. Ellis Horwood.

Sinclair, C.G., Kristiansen, B., Bu'Lock, J.D. (1987) Fermentation Kinetics and
Modelling. Open University Press, Milton Keynes.

Schugerl, K. (1987) Bioreaction Engineering.Vol.1, John Wiley, Chichester.


Subramanian, G. (1998) Bioseparation and Bioprocessing. A Handbook
Volume II: Processing, Quality and Characterization, Economics, Safety and
Hygiene, Wiley-VCH, Weinheim.

van't Riet, K., Tramper, J. (1991) Basic Bioreactor Design. M. Dekker, New
York.

Vieth, W.R. (1994) Bioprocess Engineering, J. Wiley & Sons, N.Y.

Webb, C., Black, G.M., and Atkinson, B. (1986) Process Engineering Aspects of
Immobilized Cell Systems.Pergamon Press Ltd., Oxford.

Enzyme Engineering and Kinetics

Bisswanger, H. (2002) Enzyme Kinetics. Principles and Methods; Wiley-VCH.


Buchholz, K., Kasche, V. (1996) Biokatalysatoren und Enzymtechnologie,
VCH.Weinheim.

Cornish-Bowden, A. (1979) Fundamentals of enzyme kinetics, Butterworth,


London.

Drauz, K., Waldmann, H. (1995) Enzyme Catalysis in Organic Synthesis.


Volume I. VCH Weinheim.
Drauz, K., Waldmann, H. (1995) Enzyme Catalysis in Organic Synthesis.
Volume II. VCH, Weinheim.

Fessner, W.-D. (1999) Biocatalysis - From Discovery to Application, Springer,


Berlin.

Godfrey, T., West, S. (1996) Industrial enzymology, Macmillan Press, London.


Hayashi, K.and Sakamoto, N. (1986) Dynamic Analysis of Enzyme Systems,
Japan Sci. Soc. Press, Tokyo, Springer Verlag, Berlin.
References 187

Kennedy J. F., Ed. (1987) Enzyme Technology, in Biotechnology Vol. 7a,


VCH, Weinheim,

Liese, A., Seelbach, K., Wandrey, C. (2000) Industrial Biocatalysis, Wiley-VCH.

Scheper, T. (1997) Advances in Biochemical Engineering Biotechnology. New


Enzymes for Organic Synthesis, Springer.

Segel, I.H. (1975) Enzyme Kinetics: Behavior and Analysis of Rapid


Equilibrium and Steady-state Enzyme Systems. Wiley, New York.

Metabolic Engineering

Lee, Papoutsakis (1999) Metabolic Engineering, Marcel Dekker: New York,


Basel.
Stephanopoulos, G. N., Aristidou, A. A., Nielsen, J. (1998) Metabolic
Engineering. Principles and Methodologies, Academic Press: USA.

Chemical Reaction Engineering

Aris, R. (1989) Elementary Chemical Reactor Analysis. Butterworth Publ.,


Stoneham.

Fogler, H. S. (1992) Elements of Chemical Reaction Engineering, Prentice-


Hall.
Hagen, J. (1993) Chemische Reaktionstechnik, VCH: Weinheim.
Ingham, J., Dunn, I.J., Heinzle, E., Prenosil, I.E. (2000) Chemical Engineering
Dynamics: An Introduction to Modelling and Computer Simulation, VCH
Verlagsgesellschaft mbH: Weinheim, Germany,.

Levenspiel, O. (1999) Chemical Reaction Engineering. John Wiley & Sons,


New York.
Massart, D. L., Vandeginste, B. G. M., Buydens, L. M. C., de Jong, S., Lewi, P.
J., Smeyers-Verbeke, J. (1997) Handbook of Chemometrics and Qualimetrics:
Part A, Elsevier: Amsterdam.
188 References

Richardson, J. F., Peacock, D. G. (1994) Coulson & Richardson's Chemical


Engineering. Volume 3: Chemical & Biochemical Reactors & Process Control,
Pergamon, Trowbridge.

Satterfield, C.N. and Sherwood, T.K. (1963) The Role of Diffusion in Catalysis,
Addison-Wesley, New York.

Modelling and Simulation

Basmadjian, D. (1999) The Art of Modeling in Science and Engineering,


Chapman & Hall/CRC: Boca Raton.
Deaton, M. L., Winebrake, J. J. (1999) Dynamic Modelling of Environmental
Systems, Springer: New York.
Dodson, C. T. J., Gonzalez, E. A. (1995) Experiments in Mathematics using
Maple, Springer-Verlag: Berlin,
Franks, R.G.E. (1966) Mathematical Modeling in Chemical Engineering. Wiley,
New York.

Franks, R.G.E. (1972) Modeling and Simulation in Chemical Engineering.


Wiley, New York.

Russell, T.W.F., Denn, M.M. (1972). Introduction to Chemical Engineering


Analysis. Wiley, New York.
Ruth, M., Hannon, B. (1997) Modeling Dynamic Economic Systems, Springer
Verlag, New York.

Dynamics and Control

Astrom, K.J. and Wittenmark, B. (1989). Adaptive Control. Addison-Wesley,


Reading.

Coughanowr D. R. and Koppel L. B. (1965) Process System Analysis and


Control. McGraw-Hill, New York.
References 189

Fish, N. M., Fox, R.L, and Thornhill, N.F. (1989) Computer applications in
fermentation technology: Modelling and control of biotechnological processes.
Elsevier, London.

Halme, A. (1983) Modelling and Control of Biotechnical Processes. Pergamon


Press, Oxford.

Johnson, A. (1986) Modelling and Control of Biotechnological Processes.


Pergamon Press, Oxford.

Luyben W. L. (1973) Process Modeling, Simulation, and Control for Chemical


Engineers, McGraw Hill, New York.

Pons, M.-N., Ed.(1991) Bioprocess Monitoring and Control. Hanser, Munich.

Snape, J. B., Dunn, I. J., Ingham, J., Prenosil, J. E. (1995) Dynamics of


Environmentel Bioprocesses, VCH Verlagsgesellschaft mbH, Weinheim,.

Stephanopoulos, G. (1984) Chemical Process Control: An Introduction to


Theory and Practice, Prentice Hall.

Weber. W. J., Jr., DiGiano, F. A. (1996) Process Dynamics in Environmental


Systems, Wiley.
Part II Dynamic Bioprocess
Simulation Examples
and the Berkeley
Madonna Simulation
Language

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
8 Simulation Examples of Biological
Reaction Processes Using Berkeley
Madonna

8.1 Introductory Examples

8.1.1 Batch Fermentation (BATFERM)

System

The system is represented in Fig. 1, and the important variables are biological
dry mass or cell concentration, X, substrate concentration, S, and product
concentration, P. The reactor volume V is well-mixed, and growth is assumed
to follow kinetics described by the Monod equation, based on one limiting
substrate. Substrate consumption is related to cell growth by a constant yield
factor YX/S- Product formation is the result of both growth and non-growth
associated rates of production, where either term may be set to zero as required.
The lag and decline phases of cell growth are not included in the model.

Figure 1. Stirred batch fermenter with model variables.

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
194 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Model

Mass Balances:

(Rate of accumulation) = (Rate of production)


For cells
VTT = r xv
or
dX

For substrate
V
dS = r
dF s
or
dS
dF = rS
For product
dP
V-3T =
or
dP
dF = r?
Kinetics:
rx = f i X

with the Monod relation, constant yield relation, and product formation
kinetics:

tx/s

rP = (ki + k2 |^) X

where ki is the non-growth associated coefficient, and k2 is the coefficient


associated with growth.
If the number of equations is equal to the number of unknowns, the model is
complete and the solution can be obtained. The easiest way to demonstrate this
is via an information flow diagram, as shown below in Fig. 2.
8.1 Introductory Examples 195

x
o Biomass X
^
Balance
r
T*
4_ x Growth
Rate
M
Monod
Kinetics
So Substrate 1 A s
^
Balance f'x
r
4_ s Substrate
Rate
PO Product p
Balance
r -^M
A
4
p Product
Rate ,-_

Figure 2. Information flow diagram of the batch fermenter model equations,

It is seen in that all the variables required for the solution of any one equation
block are obtained as the products of other blocks. The information flow
diagram thus emphasizes the complex inter-relationship involved in even this
very simple problem. Solution begins with the initial conditions XQ, SQ and PQ
at time t=0. The specific growth rate |i is calculated, enabling rs, rx and rp to
be calculated, and hence the initial gradients dX/dt, dS/dt and dP/dt. At this time
the integration routine takes over to estimate revised values of X, S and P over
the first integration step length. The procedure is repeated for succeeding step
lengths until the entire X, S and P concentration time profiles have been
calculated up to the required final time.

Program

The following Berkeley Madonna program solves the above fermentation


problem:

{BATPERM}
{Batch growth with product formation}

{Constants}
UM=0 . 3
KS = 0 . 1 ;kg/m3
Kl=0.03 ;kgP/kgX h
K2=0.08 ;kgP/kgX h
Y=0. 8 ;kg X/kg S
196 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

X0=0.01 /Initial biomass inoculum, kg/m3


SO = 10 ;Initial substrate cone., kg/m3
P0=0 ;Initial product conc.,kg/m3

{Initial Conditions}
INIT X=XO
INIT S=SO
INIT P=PO

(Mass Balances}
X'= RX ;BIOMASS BALANCE
S'= RS ;SUBSTRATE BALANCE
P' = RP ;PRODUCT BALANCE

{Kinetics}
RX = U*X ; BIOMASS RATE EQUATION, kg/m3 h
U = UM*S/(KS + S) ;MONOD EQUATION, 1/h
RS = -RX/Y ; SUBSTRATE RATE EQUATION, kg/m3 h
RP= (K1 + K2*U) *X /PRODUCT RATE EQUATION, g/m3 h

Limit S>=0.0

The semicolon or curly brackets are used for comments.


INIT specifies the initial conditions. XQ, SQ and PQ are used here for the initial
conditions, or the values at time=0. The form X' designates the time derivative
or d/dt(X) can be used. Most models are conveniently structured in terms of
mass balances and kinetics. Any result quantity on the left of the equal sign is
stored for further calculations or for use in graphing. Usually concentration
versus time is of interest, but rates versus concentrations make very useful plots
for understanding the kinetics. The five integration methods require specifying
time intervals, such as DT, DTMIN and DTMAX. This requires a bit of
experience. Care must be taken to see that the same results are obtained by two
different methods or for at least two different DT values.

As is seen in the Appendix, Berkeley Madonna provides many possibilities to


change the parameters and graph new runs. These include the following:
changing parameters with the parameter window and making overlay plots;
changing parameters with sliders; using the Batch Runs facility.
8.1 Introductory Examples 197

Nomenclature

Symbols

k ] and k2 Product formation constants 1/h and kg/kg


KS Saturation constant kg/m 3
P Product concentration mg/m3
r Reaction rate kg/m 3 h and kg/m 3 h
S Substrate concentration kg/m 3
V Reactor volume m3
X Biomass concentration kg/m 3
Y Yield coefficient kg/kg
H Specific growth rate 1/h

Indices

Refers to non-growth association rate


2 Refers to growth-association rate
m Refers to maximum
P Refers to product
S Refers to substrate
X Refers to biomass

Exercises

1. Vary KS, Mm separately and observe the effects in the graphs. It is


useful to zoom in on regions of importance by using the zoom tool
in the tool bar.

2. Vary the product kinetics constants (Kj and K2>, and observe the
effects. Observe the P versus time curve when S reaches zero.

3. Plot the rates versus the concentrations.


198 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

The plots of X, S and P versus T in Fig. 3 show that when substrate is depleted,
the growth stops, and the product continues to increase, but only linearly. The
results of Fig. 4 are obtained by varying the product formation rate constants,
ki in three runs using a slider, which is defined in the Parameter Menu.

Run 1:1500 steps in 0 seconds


.10

Figure 3. Plots of X, S and P versus time during batch growth and production.

Run 3: 1500 steps in 0.0333 seconds


^.......^ -10

"l"""'"*"'.^_ / 9

-8
** /
\ /* 7

-S:2
P:2
X / /' .5 cn
-S:3 -4
P:3 \ /s' \,"~'" 3

-'7^'^ 2

.**&'*' \ 1
-rp^ fm
vtf*EV \ .. 0
0 5 1 0 15 20 25 3()
TIME

Figure 4. Plots of P and S versus time created by varying the product formation rate constant
8.1 Introductory Examples 199

8.1.2 Chemostat Fermentation (CHEMO)

System

A continuous fermenter, as shown in Fig. 1, is referred to as a chemostat. At


steady state the specific growth rate becomes equal to the dilution rate, |a = D.
Operation is possible at flow rates (F) which give dilution rates (D = F/V) below
the maximum specific growth rate (|um). Washout of the organisms will occur
when D > (a. The start-up, steady state and washout phenomena can be
investigated by dynamic simulation.

D,S F S,X

Figure 1. Chemostat with model variables.

Model

The program BATFERM may be easily modified to allow for chemostat


operation with sterile feed by modifying the mass balance relationships to
include the inlet and exit flow terms. The corresponding equations are then:

For cells
dX
.= - D X + rx
200 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

For substrate
dS
3f = D (SF - S) + rs
For product
dP
df = -DP + rp

where D is the dilution rate and Sp the concentration of the limiting substrate in
the feed.

The same kinetic expressions as in BATFERM will be applied here.

Program

Note the conditional statement for D which allows a batch startup.

{CHEMO}
(Chemostat startup and steady state. Startup as
batch reactor until time=tstart}

{Constants}
UM=0.3 ; 1/h
KS = 0.1 ; kg/m3
Kl = 0.03 ; kgP/kgX h
K2=0.08 ; kgP/kgX
Y = 0.8 ; kg X/kg S
X0 = 0.01 ; Initial biomass inoculum, kg/m3
S0=10 ; Initial substrate cone., kg/m3
P0=0 ; Initial product conc.,kg/m3
SF = 10 ; Feed cone. ,kg/m3
Dl = 0.25 ; Dilution rate, 1/h
t start = 5 ; Start time for the feed

(Initial Conditions}
Init X=XO
Init S=SO
Init P=PO

{Mass Balances}
X'=-D*X+RX ; BIOMASS BALANCE EQUATION
S =D* (SF-S) +RS ; SUBSTRATE BALANCE EQUATION
P'=-D*P+RP ; PRODUCT BALANCE EQUATION
8.1 Introductory Examples 201

{Kinetics}
RX = U*X ; BIOMASS RATE EQUATION, kg/m3 h
U = U M * S / ( K S + S) ; MONOD EQUATION, 1/h
RS=-RX/Y ; SUBSTRATE RATE EQUATION, kg/m3 h
RP= (K1 + K2*U) *X ;PRODUCT RATE EQUATION, kg/m3 h

{Conditional equation for D}


D=if time>=tstart then Dl else 0

Prod=D*X /Productivity for biomass, kg/m3 h

Nomenclature

Symbols

D Dilution rate 1/h


ki and Product formation constants 1/h and kg/kg
KS Saturation constant kg/m3
P Product concentration mg/m3
r Reaction rate kg/m3h and
kg/m3
S Substrate concentration kg/m3
X Biomass concentration kg/m3
Y Yield coefficient kg/kg
ILL Specific growth rate 1/h
1 Time lag constant

Indices

F Refers to feed
MONOD Refers to Monod kinetics
P Refers to product
S Refers to substrate
X Refers to biomass
202 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Exercises

1. Increase D interactively to obtain washout.

2. Note the steady state values of X and S; calculate Y from these.

3. Change SF. Does this alter S at steady state? Why?

4. Calculate S at steady state from D. Verify by simulation.

5. Change the program to account for biomass in the feed.

6. Operate initially as a batch reactor with D = 0, and switch to


chemostat operation with D < |jm. Does this reduce the time to reach
steady state? Is the exact time of switchover important?

7. Include maintenance requirements to the substrate uptake kinetics


using RS = - ( U / Y + M ) * X . Remember to add a value of the
maintenance coefficient M to the constants. Investigate the influence of
the value of M on the steady state biomass concentration.

8. Using a Parameter Plot, obtain steady state values of X and S for a


range of Dl.

9. Rapidly-changing dynamic fermentations do not follow


instantaneous Monod kinetics. Modify the model and the program with
a dynamic lag on jo, such that d|j /dt= (nMonod - l-O/t- Compare the
response to step changes in D for suitable values of the time lag constant
t.

Results

The graphical output in Fig. 2 shows three startups of the fermenter under
initially batch growth conditions, using three values for D l . The break in the
concentration-time dependency as feeding starts is quite apparent, and the new
transient then continues up to the eventual steady state chemostat operating
condition or washout in the case of one run. For the results of Fig. 3 the
program was changed by adding the line PROD = X*D, and the final, steady
state value of production rate was plotted versus Dl for twenty runs, using the
Parameter Plot feature of Madonna.
8.1 Introductory Examples 203

Run 3: 4000 steps in 0.0333 seconds


10

Figure 2. Startups of the chemostat after initial batch growth for 3 values of Dl.

Run 8: 200000 steps in 1.38 seconds


2

Figure 3. Productivity in a chemostat. Steady states are shown for 20 runs using the Parameter
Plot.
204 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

8.1.3 Fed Batch Fermentation (FEDBAT)

System

In this case the model equations allow for the continuous feeding of sterile
substrate, the absence of outflow from the fermenter and the increase in volume
(accumulation of total mass) in the fermenter, schematically as shown in Fig. 1.
Simulation of fed batch fermenters can be used to demonstrate the important
characteristics of quasi-steady state, linear growth, and use of alternative feed
strategies.

F,SF

V
X
s
p

Figure 1. Fed batch fermenter with model variables.

Model

For fed batch operation, the equations become as follows:

Total balance
dV
F
dT =
For cells

For substrate
8.1 Introductory Examples 205

For product

where F is the volumetric feed rate, Sp is the feed concentration and V is the
volume of the fermenter contents at time t. Thus the mass quantities, VX, VS,
and VP are calculated and are divided by the volume at each time interval to
obtain the concentration terms required for the kinetic relationships. The
kinetics are taken to be the same as in BATFERM.

Program

The "IF" statement in the program causes the continuous feed to start when time
reaches tfeed, at which point batch operation stops and the fedbatch starts.

(FEDBAT)
{Fermentation with batch start up}

{Flow rate is initially zero and is turned on at


time=tfeed.}

{ Constants}
UM=0.3 ; 1/h
KS = 0 . 1 ; kg/m3
Kl = 0.03 ; kgP/kgX h
K2 = 0.08 ; kgP/kgX
Y = 0.8 ; kg X/kg S
X0 = 0 .01 ; Initial biomass inoculum, kg/m3
S0 = 10 ; Initial substrate cone., kg/m3
P0 = 0 ; Initial product conc.,kg/m3
SF = 10 ; Feed conc.,kg/m3
Pl-1. 5 ; Feed flow rate, m3/h
tfeed=22.5 ; Start time for the feed

{Initial Conditions}
init V=l
init VX=V*XO
init VS=V*SO
init VP=V*PO
206 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

{Mass balances, kg/h}


d/dt(V)=F
d/dt(VX)=RX*V
d/dt(VS)=F*SF+RS*V

d/dt(VP)=RP*V {kg/h}

{Calculation of concentrations}
X=VX/V
S=VS/V
P=VP/V

{Kinetics}
RX=U*X
U=UM*S/(KS+S)
RS=-RX/Y
RP=(K1+K2*U)*X

D=F/V {nominal dilution rate, 1/h}

{Turning the feed on at time = tfeed}


F=if time>=tfeed then Fl else 0 {batch start up}

Nomenclature

Symbols

D Dilution rate 1/h


F Flow rate m3/h
KS Saturation constant kg/m3
ki, k2 Constants in product kinetics 1/h and kg/kg
M Maintenance coefficient kg/kg h
P Product concentration kg/m3
r Reaction rate kg/m3 h
S Substrate concentration kg/m3
X Biomass concentration kg/m3
V Reactor volume m3
Y Yield coefficient kg/kg
|i Specific growth rate 1/h
T Time delay constant h
8.1 Introductory Examples 207

Indices

F Refers to feed
P Refers to product
S Refers to substrate
X Refers to biomass

Exercises

Results

Operation begins under initial batch conditions, and feeding of substrate is


started at tfeed=22.5 h. In Fig. 2, the break in the batch growth transient, as
semi-batch feeding starts is very apparent, with the transient continuing to an
apparent "quasi" steady state operating condition. Under these conditions the
biomass concentration becomes constant, while the substrate concentration (not
shown) is below the KS value and decreases very slowly. As seen in the zoom
of Fig. 3, the values of D (= F/V) also decrease since V increases due to the
incoming feed, and D eventually becomes equal to p when S falls below K$.
The total biomass is determined by the yield coefficient times the total amount
of substrate that has been consumed, which is approximately equal to the
amount in the reactor initially plus the amount added during the feeding
period. During the quasi-steady state, the total biomass will increase linearly
with time if, as in this case, the feeding flow rate is constant. This is a "linear
208 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

growth" situation in which the growth rate is limited by the feeding rate. In
Fig. 3 the values of X, S, and P are plotted versus T for a switch from batch
(F = 0) to fed batch (F = 5) at time T = 20 h. The product production rate
depends linearly on biomass concentration, and thus even when ja becomes very
low, P will continue to increase linearly in mg/m3 amounts.
TIME= 34.13 X = 12.34
10-.- .^

10 20 30 40 50 60 70 80 90 100

Figure 2. Transients during the fedbatch fermentation.

Run 1: 5000 steps in 0.1 seconds


0.4.

0.35-

0.3-
-~, I
3
0.25.

Q 0.2-
V
CO
0.15-

0.1- -J.
T
0.05-

0-
27 28 29 30 31 32 33 34 35 36 37
TIME

Figure 3. Zooming in on the quasi-steady state.


8.2 Batch Reactors 209

8.2 Batch Reactors

8.2.1 Kinetics of Enzyme Action (MMKINET)

System

The intermediate enzyme-substrate complex is the basis for the simplest form
of enzymatic catalysis (Fig. 1):

E +S ^ ES *- E +P
k2

Figure 1. Mechanistic model for enzymatic reaction.

Model

The equations for substrate, enzyme-substrate complex and product in a batch


reactor are:
- = ki E S - k 2 ES
dt
dFS
^ = ki E S - (k2 + k3) ES

dt
Using the steady state approximation for the change of active complex,

dt
the Michaelis-Menten equation is obtained.

_dS _
K
~ dt ~ M +S
210 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

where vmax = k3 E0 and KM = (k2+k3)/ki.

Program

The program with the detailed mechanism is on the CD-ROM.

Nomenclature

Symbols

E Enzyme concentration mol/m3


ES Enzyme-substrate complex concentration mol/m3
k Reaction rate constants various
KM Michaelis-Menten constant mol/m3
P Product concentration mol/m3
S Substrate concentration mol/m3
Vmax Maximum velocity mol/m3 h

Indices

0 Refers to initial values


1 Refers to reaction 1
2 Refers to reaction 2
3 Refers to reaction 3
S Refers to substrate
Mm Refers to Michaelis-Menten

Exercises
8.2 Batch Reactors 211

Results

Figs. 2 and 3 give the results of the full model and the Michaelis-Menten
simplification, respectively
Run 1:119 steps in 0.0167 seconds

0.009. Lx""""^"~ .0.9

0.008- \ f' -0.8

0.007- \ / 0.7

v
:. f ^..]

0.006 ... 8:1 0.6


tn ES:1
a
^0.005- -- P:1 0.5 *
LU riL to
0.004- t ~m 0.4
f \

0.003- \ 0.3
1
0.002 0.2
*i* fc
\. i
0.001 - i *v. " %^ 0.1
""""%-, ^*"'..._
.0
10 20 30 40 50 70 80 90 100
TIME

Figure 2. Results from the full model


212 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Run 1: 5000 steps in 0.0333 seconds

\
\

10 20 30 40 50 60 70 80 90 100
TIME

Figure 3. Results from the Michaelis-Menten simplification.

8.2.2 Lineweaver-Burk Plot (LINEWEAV)

System

This program simulates the batch uptake of substrate using Michaelis-Menten


kinetics, of the form,

r
s = K^TS-

The inverse rate is plotted versus the inverse concentration (Fig. 1).
Comparison of this plot with the concentration-time plot together with the Km
value, demonstrates the importance of data in the Km region and the difficulty
of obtaining this in a batch reactor. It is useful to make specially-scaled graphs
in the KM region.
8.2 Batch Reactors 213

Figure 1. Lineweaver-Burk plot to determine vm and

Model

The model is that of a batch reactor with Michaelis-Menten kinetics.


dS
dF = ~ r s

Program

To make the Lineweaver-Burk plot, the inverse values of S and rs are calculated
in the program on the CD-ROM.

Nomenclature

Symbols

KM Michaelis-Menten constant kg/m3


r Reaction rate kg/m3
S Substrate concentration kg/m3
214 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Si Inverse substrate concentration m3/kg


V Reaction velocity or rate kg/m3 h
Inverse reaction velocity or rate m3 h/kg

Indices

0 Refers to feed
m Refers to maximum
S Refers to substrate

Exercises

Results

The results are shown in Fig. 2 (rates and concentrations versus time) for a
range of Michaelis-Menten constants KM and in Fig. 3 the corresponding
Lineweaver-Burk plots.
8.2 Batch Reactors 215

Run 4:13710 steps in 0.133 seconds


'0.5

0.45

.0.4

0.35

.0.3

.0.25

0.2
-0.15

0.1

0.05

140 160

Figure 2. Rate and concentration plots for KM = 0.2, 0.5, 1.0 and 2.0 (bottom to top curves).

Run 4:13710 steps in 0.133 seconds

Figure 3. Lineweaver-Burk plots for KM = 0.2, 0.5, 1.0 and 2.0 (bottom to top curves).

8.2.3 Oligosaccharide Production in Enzymatic


Lactose Hydrolysis (OLIGO)

System

Some enzyme catalyzed reactions are very complex. For this reason their
rigorous modelling leads to complex kinetic equations with a large number of
constants. Such models are unwieldy and are usually not suitable for practical
216 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

purposes. One approach to simplify them is to neglect formation of enzyme-


substrate complexes altogether and to deal only with overall reactions of the
react ants to products.
An example of such a reaction is the enzymatic lactose hydrolysis, a
complex process involving a multitude of sequential reactions leading to higher
saccharide (oligosaccharides) intermediates. The mechanistic model is rather
complex even when only trisaccharides are considered (Fig. 1).

La + E ^ ^ LaE ^- Ga + GI + E

Ga E + La ^ ** E + Tr

GaE + H2O ^ E + Ga

Figure 1. Complex and simplified models for the enzymatic hydrolysis of lactose, where the
symbols are La for lactose, Ga for galactose, Gl for glucose, Tr for trisaccharide and E for
enzyme.

Neglecting the enzyme complexes, however, gives a simplified model (Fig. 2)


requiring only three constants:

K
!
1L.a
a to Ga T. V3II
f^i

K
1
La -i- Ga
fcaCl ^ Tr
i
K
2

Figure 2. Simplified model for the enzymatic hydrolysis of lactose.

The simulation of this model is easy, and the constants can be adjusted to
achieve good agreement with experimental data.

Model

This simple batch reactor model is equivalent to the Michaelis-Menten product


inhibition model.
8.2 Batch Reactors 217

dLa
-gj- - K! La - KI La Ga + K2 Tr

dGa
= Kj La - KI La Ga + K2 Tr

dTr
= KI La Ga - K2 Tr

Initial conditions: Lao =150 mmol/m3, Gao = 0, Trg = 0


Range of the kinetic constants: KI = 0.02 - 0.06 miir1, KI = 0.02 - 0.1
L/mmol min, K2 = 1 - 50 min"1.

Program

It was found that K2 must be two orders of magnitude greater than KI in order
to bring the simulation into agreement with the experimental data. The
program is on the CD-ROM.

Nomenclature

Symbols

Ga Galactose concentration mmol/L


Gl Glucose concentration mmol/L
Reaction rate constant (La > Ga + Gl) 1/min
Reaction rate constant (La + Ga -> Tri) L/(mmol min)
K2 Reaction rate constant (Tri -> La + Ga) 1/min
La Lactose concentration mmol/L
Tr Trisaccharide concentration mmol/L

Indices

0 Refers to initial concentration


218 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Exercises

Results

The outputs in Figs. 3 and 4 show the influence of KI, KI and Lao on the sugar
concentration profiles.

Run 1:10000 steps in 0.05 seconds


100. r100

90.

80.

70

60.

. 50.

40

30

20

10
0
20 100 180 200
TIME

Figure 3. Sugar concentrations with Kr = 0.04, K{ = 0.05, La0 = 100.


8.2 Batch Reactors 219

Run 1: 10000 steps in 0.15 seconds


160

80
( '

80 100 120 140 160 180 200

Figure 4. Sugar concentrations with Kj = 0.06, KI = 0.1 Lao = 160.

Reference

Prenosil, J. E., Stuker, E. and Bourne, J. R. (1987) "Formation of


Oligosaccharides during an Enzymatic Lactose Hydrolysis Process", Parts I and
II: Biotechnol. Bioeng. 30, 1019-1031.

8.2.4 Structured Model for PHB Production (PHB)

System

Heinzle and Lafferty (1980) have presented a structured model to describe the
batch culture of Alcaligenes eutrophus under chemolithoautotrophic growth
conditions, as discussed in Case C, Sec. 3.3.1. Growth and storage of PHB are
described as functions of limiting substrate S (NH4+), residual biomass R and
product P (PHB) concentrations.
220 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Figure 1. Structured kinetic model for PHB synthesis.

Model

In the model seen in Fig. 1 the whole cell dry mass (X) consists of two main
parts, namely PHB (P) and residual biomass (R), where R is calculated as the
difference between the total cell dry weight and the concentration of PHB (R =
X - P). R can be considered as the catalytically active biomass, including
proteins and nucleic acids. With constant concentrations of the dissolved gases,
two distinct phases can be recognized: growth and storage. During the growth
phase there is sufficient NH4+ to permit protein synthesis. When the limiting
substrate NH4+ (S) is exhausted, the protein synthesis ceases, and the
production rate of PHB is increased. During the storage phase only PHB is
produced. The limiting substrate NH4+ (S) is essential to produce R and limits
its synthesis at low concentrations.
For the batch process,
dR
= r
dF R = MR
where TR is the rate of synthesis of R and (j is the specific rate of synthesis of R,
where
S (S/Ks,2)n
+ S) + ^m,2 ! + (S/KS,2)n

where n is the empirical Hill coefficient (see Sec. 3.1.2), having a value of 4 in
this example.
This is based on the postulate that there are two different mechanisms for the
assimilation of NH4+ in procaryotes. This formulation is not a mechanistic one,
8.2 Batch Reactors 221

since in reality the enzyme system, using energy to assimilate NH4+, is


repressed by high concentrations of NH4+.
For the substrate
dS 1
dF = rs = -YR/S **
The rate of synthesis of P(rp) is assumed to be the sum of a growth associated
term (rpj) and a biomass associated term (rp,2) and is given by,

dP
df = rp = r P j + rP,2

where r P j = YP/R rR
The non-growth associated term of the synthesis of P(rp,2) is assumed to be a
function of the limiting substrate S, of the residual biomass R and of the
product P. When the PHB content in the cells is high, the rate of synthesis of P
is decreased, which can be formally described as an inhibition.

Program

The program is found on the CD-ROM.

Nomenclature

Symbols

KI Inhibition constant, for (NH^SC^ kg/m3


KS Saturation constant kg/m3
n Hill Coefficient
P Product concentration (PHB) kg/m3
R Residual biomass concentration kg/m3
rp Rate of synthesis of PHB kg/m3
TR Rate of synthesis of R kg/m3
rs Rate of substrate uptake kg/(m3 h)
222 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Limiting substrate concentration kg/m3


NHj as (NH4)2S04
X Biomass concentration kg/m3
YP/R Yield coefficient kg/kg
YR/S Yield coefficient, kg/kg
Specific rate of synthesis of R (rR/R) 1/h
Specific rate of synthesis of P (rp/P) 1/h

Indices

1 Refers to reaction 1
2 Refers to reaction 2
m Refers to maximum

Exercises
8.2 Batch Reactors 223

Results

Run 1: 416 steps in 0.0167 seconds


4-1 -16

3.5- 14
/ ^"~'"~*" "
3- f -*' 12

2.5- / / ISi$% -10


-.^ T M* "."T.-'V/T
'"U--b / /
2 -8 a.

y *
of - *"'"
'"" J
I -'
'*"

1.5- -6
y " f'
1-
/ / " '. .'
4

0.5-

0-
^^
*"^ V
_j__ *%
-2

_n
0 5 10 15 20 25 30 35 40
TIME

Figure 2* Profiles of residual biomass concentration R, substrate S and product P in the batch
fermentation.
Run 4: 41 6 steps in 0.01 67 seconds
35- ... 5
'"'V^
^.^.- 4.5
30-
'v -4
25- \ / .._.. 3:3(2,3) 3.5
\ / ~- P:3(2.3)
3
20- 1 / P:4 (5)
a \ -2.5 (/)
-*. \ /
15-
-2
"""-, \ f .*""r-'" -1.5
10-

5
"\ "" \"' /^i''^' -1

-0.5
0. ^^";:^ -n
0 5 10 15 20 25 30 35 40
TIME

Figure 3. PHB formation at two different initial substrate concentrations.

References

Heinzle, E., and Lafferty, R. M. (1980) Continuous Mass Spectrometric


Measurement of Dissolved H2, O2, and CC>2 during Chemolitho-autotrophic
224 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Growth of Alcaligenes eutrophus strain H16. Eur. J. Appl. Microbiol.


BiotechnoL 11, 8.1

8.3 Fed Batch Reactors

8.3.1 Variable Volume Fermentation (VARVOL and


VARVOLD)

System

Semi-continuous or fed batch cultivation of micro-organisms is common in the


fermentation industries. The fed batch fermenter mode is shown in Fig. 1 and
was also presented in the example FEDBAT. In this procedure a substrate feed
stream is added continuously to the reactor. After the tank is full or the
biomass concentration is too high, the medium can be partially emptied, and
the filling process repeated. Since the variables, volume, substrate and biomass
concentration change with time, simulation techniques are useful in analyzing
this operation. This example demonstrates the use of dimensionless equations.

Figure 1. Filling and emptying sequences in a fed batch fermenter.

Model

The balances are as follows:

Volume,
dv
dT = FO
Substrate,
224 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Growth of Alcaligenes eutrophus strain H16. Eur. J. Appl. Microbiol.


BiotechnoL 11, 8.1

8.3 Fed Batch Reactors

8.3.1 Variable Volume Fermentation (VARVOL and


VARVOLD)

System

Semi-continuous or fed batch cultivation of micro-organisms is common in the


fermentation industries. The fed batch fermenter mode is shown in Fig. 1 and
was also presented in the example FEDBAT. In this procedure a substrate feed
stream is added continuously to the reactor. After the tank is full or the
biomass concentration is too high, the medium can be partially emptied, and
the filling process repeated. Since the variables, volume, substrate and biomass
concentration change with time, simulation techniques are useful in analyzing
this operation. This example demonstrates the use of dimensionless equations.

Figure 1. Filling and emptying sequences in a fed batch fermenter.

Model

The balances are as follows:

Volume,
dv
dT = FO
Substrate,

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
8.3 Fed Batch Reactors 225

- = F0S0
Biomass,
d(VX)
dt = rx
The kinetics are
rx = MX
|LimS
** - (Ks + S)
and
rx
rs = -Y
The dilution rate is defined as

In order to simplify the equations and to present the results more generally, the
model is written in dimensionless form. Defining the dimensionless variables:

v V
=
X
X< =

s s

=
F =

- _ JL
Mm

F
tf-

t' = t
226 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Expanding the derivatives gives

d(V S) = V dS + S dV
and
d(V X) = V dX + X dV

Substituting, the dimensionless balances now become:

Volume
dV'

Biomass
dX'
dt'
Substrate
dS'
dt = (l-S)D-jiX
The Monod equation is:

KS +S

In Fig. 2 a computer solution shows the approach to and attainment of the


quasi-steady state of the dimensionless fed-batch model.
8.3 Fed Batch Reactors 227

Quasi- steady

Figure 2. Dynamic simulation results for a fed batch culture.

Programs

The program VARVOL is based on the model equations with normal


dimensions. The program VARVOLD is based on the dimensionless equations
as derived above. Both are on the CD-ROM.

Nomenclature

Symbols

D Dilution rate 1/h


F Flow rate m3/h
KS Saturation constant kg/m3
r Reaction rate kg/m3 h
S Substrate concentration kg/m3
V Reactor volume m3
228 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

X Biomass concentration kg/m3


Y Yield coefficient kg/kg
Specific growth rate 1/h

Indices

0 Refers to feed and initial values


f Refers to final
m Refers to maximum
S Refers to substrate
X Refers to biomass
Refers to dimensionless variables

Dimensionless Variables

Dimensionless flow rate


Dimensionless saturation constant
S' Dimensionless substrate concentration
V Dimensionless volume
X1 Dimensionless biomass concentration
t' Dimensionless time
Dimensionless specific growth rate

Exercises
8.3 Fed Batch Reactors 229

Results

During the quasi-steady state, \l becomes equal to D, and this requires that S
must decrease steadily in order to maintain the quasi-steady state as the volume
increases (Fig. 3). Increasing flow rates from 0.01 to 1.0 causes a delay in the
onset of linear growth and causes the final biomass levels to be higher (Fig. 4).
Run 1:105 steps in 0 seconds
4.5

Figure 3. Fed batch concentration and growth rate profiles, showing quasi-steady state.

Run 7:105 steps in 0 seconds


5

2.5 C/>

10
230 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Figure 4. Influence of flow rate on growth. Flow rate increase from 0.01 to 1.0.

References

Dunn, I.J., and Mor, J.R. (1975) Variable Volume Continuous Cultivation.
Biotechnol. Bioeng. 17, 1805.

Keller, R., and Dunn, I.J. (1978) Computer Simulation of the Biomass
Production Rate of Cyclic Fed Batch Continuous Culture. J. AppL Chem.
Biotechnol. 28, 784.

8.3.2 Penicillin Fermentation Using Elemental


Balancing (PENFERM)

System

This example is based on the publication of Heijnen et al. (1979), and


encompasses all the principles of elemental balancing, rate equation
formulation, material balancing and computer simulation. A fed batch process
for the production of penicillin as shown in Fig. 1 is considered with
continuous feeding of glucose. Ammonia, sulfuric acid and o-phosphoric acid
are the sources of nitrogen, sulfur and phosphorous respectively. O-
phosphoric acid is sufficiently present in the medium and is not fed. Oxygen
and carbon dioxide are exchanged by the organism. The product of the
hydrolysis of penicillin, penicilloic acid, is also considered, thus taking the slow
hydrolysis of penicillin-G during the process into account.
8.3 Fed Batch Reactors 231

Glucose Carbon dioxide

Oxygen

Precursor
Phenylacetic acid

Sulfuric acid

Ammonia

Figure 1. Streams in and out of the penicillin fed batch reactor.

Table 1. lists the components and their conversion rate designation.


232 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Table 1. Component properties and rate designations.


Compound Chemical formula Mol wt. Enthalpy Conversion
(Daltons) (kcal/mol) rate (mol/h)
Glucose C6H1206 180 - 303 Rl
Mycelium CHi.64Oo.52No. 16 24.52 - 28.1 R2
So.0046P<).0054
Penicillin C16H1804N2S 334 - 115 R3
Penicilloic acid C16H2005N2S 352 - 183 R4
Oxygen 02 32 0 R5
Carbon Dioxide CO2 44 -94 R6
Ammonia NH3 17 - 19 R8
Sulfuric Acid H2SO4 98 - 194 R9
Phosphoric Acid H3PO3 98 - 319 RIO
Phenylacetic Acid C8H802 136 -69 RH
Water H2O 18 -68 Rl2

Model

a) Elemental Balancing

Knowing the composition of all chemical substances and the biomass mycelium
(Table 1) allows the following steady state balances of the elements in terms of
mol/h:

For carbon

6 RI + R6 + 16 R3 + 8 RH + 16 RH + R2 = 0

For oxygen

6 RI + 2 R5 + 2 R6 + R12 + 4 R3 + 4 R9 + 4 R10 + 2 RH + 5 R4 + 0.52 R2 = 0

For nitrogen

0.16R 2 + 2 R 3 + 2R 4 + R8 = 0

For sulfur

0.00 46 R2 + R3 + R4 + R9 = 0
8.3 Fed Batch Reactors 233

For hydrogen

12 RI + 1.64 R2 + 18 R3 + 20 R4 + 3 R8 + 2 R9 + 3 RIO + 8 R n + 2 R12 = 0

For phosphorus

0.0054 R2 + RIO = 0

A steady state enthalpy balance gives the following


- 303 RI - 28.1 R2 - 115 R3 - 183 R4 - 94 R6 - 19 R8 - 194 R9 -

- 3 1 9 R i o - 6 9 R n -68Ri 2 + rH = 0

where TH is the rate of heat of production (kcal/h).


A total of 12 unknowns (Ri through R6, Rg through Ri 2 and TH) are involved
with a total of 7 equations (6 elemental balances and one heat balance). The
five additional equations are provided by five reaction kinetic relationships.
The remaining rates can be expressed in terms of these basic kinetic equations.

From the carbon balance


- R6 = 6 RI + R2 + 16 R3 + 16 R4 + 8 RH

From the nitrogen balance


- R 8 = 0.16R2 + 2R 3 + 2R 4

From the sulfur balance


- R9 = 0.0046 R2 + R3 + R4

From the phosphor balance

- R i o = 0.0054 R2

From the hydrogen and nitrogen balances


- R5 = -6 RI - 1.044 R2 - 18.5 R3 - 18.5 R4 - 9 R n

From the enthalpy balance


rH = - 669 RI - 110.1 R2 - 1961 R3 - 1961 R4 - 955 RH
234 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

To complete the model, equations for glucose uptake rate (-Ri), biomass
formation rate (R2), rate of penicillin formation (Rs), precursor consumption
rate (-Rn), and rate of penicillin hydrolysis (R4) must be known. Note that the
reaction rates are defined with respect to total broth weight, since the process is
the fed-batch type and broth weight is variable with respect to time.

b) Formulation of the Kinetic Equations

Substrate (Glucose) Uptake Rate:


A MONOD type equation for the uptake of sugar by P. Chrysogenum is used.

-QlCiM2

Biomass Formation Rate:


A linear relationship between the glucose consumption rate and growth rate of
biomass is assumed. Hence,
1
- Rl = y^ ^2 + m M2

where Y2 is the maximum growth yield and m is the maintenance rate factor
(mol glucose/mol mycelial biomass h).
Some sugar is used in the formation of the product. Hence,

- Rl = Yj R2 + m M2 + YJ (R3 + R*)

where 3 is the conversion yield for glucose to penicillin (mol penicillin/mol


glucose).
The total rate of biomass formation equals the net rate of formation,
corrected for the amount transformed to penicilloic acid. Therefore,

R2 = - Y 2 R i - Y 2 m M 2 - yf (Rs + 4)

Precursor Conversion Rate


It is assumed that the precursor is only used for penicillin synthesis. Thus

- R l l = R 3 + R4
where - RH is the precursor consumption rate.
8.3 Fed Batch Reactors 235

Rate of Penicillin Synthesis


The specific rate of penicillin synthesis is assumed not to be a function of
specific growth rate. So that

R 3 = Q3 M2 - R4

where Q3 is the maximum specific rate of penicillin synthesis (mol/mol h),

Equation for the Rate of Penicillin Hydrolysis


The hydrolysis of penicillin takes place by a first-order reaction.

R 4 = K 3 M3

c) Balance Equations

Total Mass Balance


The individual feed rates of glucose, sulfuric acid and ammonia are adjusted to
equal their molar consumption rates. Water lost by evaporation is neglected.
The change in mass due to gas uptake and production is neglected. The mass
flow rates are calculated from the molecular weights, the uptake rates and the
mass ratio compositions.

Feed rate of glucose stream (kg/h)

180 F
F F =
l = 500 2.78
where F = mol glucose /h.
Feed rate of NH3 stream (kg/h)

F8 = R 8 25Q = T4JT

Feed rate of [2804 stream (kg/h)


18 R9
F9 - R95QO ~ 2.55

The total mass in reactor G (kg/h) changes with time according to

dG F R9 Rg
+ +
"dT = Tn 235" TTTT
Component Balances
Expressed in mol/h the dynamic balances are,
236 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Glucose

~3T = R l + F
Biomass
dM2
~ar = R2
Penicillin
dM3 .

Penicilloic acid
dM4

The concentrations in mol/kg are as follows:

M2
Tr
M3

M4
c4 =
where the masses MI, M2, M3 and M4 are in mol units.

d) Metabolism Relations

The various metabolic relationships are given from

Specific growth rate for cells


R2
* = Ml
Respiration quotient
R
Re
Q = R7
Oxygen uptake rate
OUR = -R 5
CO2 production rate
CPR = R6
Fraction of N2 in mycelium
8.3 Fed Batch Reactors 237

R2
f 2 = 0.16 R|
N2 fraction in penicillin
f3 = 1-F 2

Fraction of sulfur used for mycelium


R2
f 4 = 0.046 R|

Sulfur fraction used for penicillin

f5 = 1 - F4

Fraction of glucose for cell growth


R2
=
Fraction of glucose for penicillin
R3 + R4
S3 = - Y3 R!
Fraction of glucose for maintenance
M2
g4 = -M R^-

Program

The Madonna program covers a fermentation time of 200 h starting from the
initial conditions of 5500 mol glucose, 4000 mol biomass, 0 mol penicillin and
0.001 mol penicilloic acid in an initial broth weight of IxlO 5 kg. The program
is on the CD-ROM.

Nomenclature

Symbols

a, b Flow rate variables various


C Component concentration mol/kg
CPR Carbon dioxide production rate mol/h
F Feed rate kg/h
238 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

h Fraction of nitrogen in mycelium


Nitrogen fraction in penicillin -
fl Fraction of sulfur used for mycelium -
f5 Fraction of sulfur used for penicillin -
G Mass in reactor kg
82 Fraction of glucose for cell growth
g3 Fraction of glucose for penicillin
g4 Fraction glucose for maintenance
Kl Saturation constant mol/kg
K3 Hydrolysis rate constant 1/h
M Mass of individual components mol
m Maintenance rate factor mol/(mol h)
OUR Oxygen uptake rate mol/h
Q Maximum specific rates mol/(mol h)
R Conversion mol/h
RQ Respiration quotient -
Heat production rate kcal/h
rq Respiratory quotient -
Y Yield coefficient -
Specific growth rate 1/h

Indices

0 initial
1 glucose
2 biomass
3 penicillin
4 penicilloic acid
5 oxygen
6 carbon dioxide
8 ammonia
9 sulfuric acid
10 phosphoric acid
11 phenylacetic acid
12 water

Exercises
8.3 Fed Batch Reactors 239

Results

The results of Fig. 2 show the substrate MI to pass through a maximum, while
the penicillin M2 develops linearly, for this constant feeding situation.
Increasing the feeding linearly with time (F = 500 + 5* time) gave the results in
Fig. 3, where it is seen that maintenance accounts for about 70 % of glucose
consumption at the end of the fermentation.
Run 1:215 steps in 0 seconds

0 20 40 60 80 100 120 140 160 180 200


TIME

Figure 2. Penicillin fed batch fermentation with total masses of glucose (M]) and biomass
(M2).
240 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Run 1:215 steps in 0 seconds


0.9-,

0.8-

0.7-

0.6-
8
r-
- 0.4-

0.3-

0.2-

0.1-

0-
20 60 80 100 120
TIME

Figure 3. Linear increase of feeding with time F = 500 + 5*T.

Reference

Heijnen, J., Roels, J. A., and Stouthamer, A.H. (1979). Application of Balancing
Methods in Modeling the Penicillin Fermentation. Biotechnol. and Bioeng., 21,
2175-2201.

8.3.3 Ethanol Fed Batch Diauxic Fermentation


(ETHFERM)

System

Yeast exhibits diauxic behavior with respect to the glucose and ethanol in the
medium as alternative substrates. In addition, the glucose effect, when glucose
levels are high, will cause fermentation, instead of respirative oxidation, to take
place, such that the biomass yields are much reduced (Fig. 1). In this example
the constant a designates the fraction of respiring biomass and (1 - a) the
fraction of biomass that ferments. The rates of the process are controlled by
three enzymes.
8.3 Fed Batch Reactors 24 1

^^ C02 + X

Glucose
^^^*- Ethanol + X

Figure 1. Pathways of aerobic ethanol fermentation.

Model

The rates of the processes are as follows:

Respirative oxidation on glucose,

R, =
Glu+K sl
Fermentation to ethanol,

R2 = K2 (1 - a) X
Glu + KS2

Conversion of ethanol to biomass,

Enzyme activation for the transformation of ethanol to biomass is assumed to


involve an initial concentration of starting enzyme EQ, which is converted to en-
zyme 2 and which catalyzes growth on ethanol through an intermediate
enzyme EI.
Thus, the production rate of enzyme EI is inhibited strongly by glucose,

R4 = - -rXEo
K S4 +Glu 3

and the production rate of enzyme 2 controlling the conversion of biomass to


ethanol depends on EI,

R5 = K 5 X E i

The mass balances for the biomass, substrates and enzymes are those for a fed
batch with variable volume.
242 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

For the total mass balance with constant density,

dt
The component balances are written by separating the accumulation term,
noting that

d(VC) _ VdC CdV _ VdC


- - + - -
dt dt dt dt
Thus,

dt V

^o = _ E0Q
dt V

f -*-.-*

Program

Note that the program on the CD-ROM is formulated in terms of C-mol for the
biomass. This is defined as the formula weight written in terms of one C atom,
thus for yeast CHL667Oo.5No.i67-
8.3 Fed Batch Reactors 243

Nomenclature

Symbols

C Component concentration mol/m3


E Enzyme concentration mol/m3
EtOU Ethanol concentration mol/m3
Glu Substrate feed concentration mol/m3
K Rate constants various
Q Feed flow rate m3/h
R Reaction rate mol/m3 h
V Reactor volume m3
X Biomass concentration C-mol/m3
Y Yield coefficient mol/mol
a Fraction of respiring biomass

Indices

0 Refers to feed
1 Refers to reaction 1
2 Refers to reaction 2
3 Refers to reaction 3
4 Refers to reaction 4
5 Refers to reaction 5

Exercises
244 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

Seen in Fig. 3 are the simulation results giving the concentrations (glucose,
ethanol and biomass) during the fed batch process. In Fig. 4 the maximum in
ethanol concentration as a function of feedrate is given from a Parameter Plot.
Run 1: 605 steps in 0.0167 seconds
30

25

60

Figure 3. Batch yeast fermentation.


8.3 Fed Batch Reactors 245

Run 2:12100 steps in 0.333 seconds


30

0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

Figure 4. Influence of flowrate on the maximum ethanol concentration.

Reference

This example was contributed by C. Niklasson, Dept. of Chemical Reaction


Engineering, Chalmers University of Technology, S - 41296 Goteborg,
Sweden.

8.3.4 Repeated Fed Batch Culture (REPFED)

System

A single cycle of a repeated fed batch fermentation is shown in Fig. 1. In this


operation a substrate is added continuously to the reactor. After the tank is full,
the culture is partially emptied, and the filling process is repeated to start the
next fed batch. The operating variables are initial volume, final volume,
substrate feed concentration and flow rates of filling and emptying.
246 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Figure 1. One cycle of a repeated fed batch.

Model

The equations are the same as given in the example FEDBAT (Section 8.1.3),
where the balances for substrate and biomass are written in terms of masses,
instead of concentrations. The only difference is that an outlet stream is
considered here to empty the fermenter at the end of the production period.

Program

Since in a Madonna program, the initial conditions cannot be reset, an outlet


stream is added. The inlet and outlet streams are controlled by conditional
statements as shown below. The full program is on the CD-ROM.

{Statements to switch the feed and emptying streams)


Fin=if time> = 10 then Flin else 0 {batch start up}
Flin= if time> = 33 then 0.5 else if time> = 32 then 0
else if time> = 21 then 0.5 else if time> = 20 then 0
else 0.5
Fout= if time>=33 then 0 else if time>=32 then 5.39
else if time> = 21 then 0 else if time> = 20 then 5.39
else 0
8.3 Fed Batch Reactors 247

Nomenclature

Symbols

D Dilution rate 1/h


F Flow rate m3/h
Kl and K2 Product kinetic constants various
KS Saturation constant kg/m3
P Product concentration g/m3
S Substrate concentration kg/m3
X Biomass concentration kg/m3
V Reactor volume m3
V
0 Initial volume of liquid m3
VX Biomass in reactor kg
VS Substrate in reactor kg
Y Yield coefficient kg/kg
|i Specific growth rate 1/h

Indices

S Refers to substrate
X Refers to biomass
0 (zero) Refers to initial and inlet values
initial Refers to initial values
in Refers to inlet
out Refers to exit

Exercises
248 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

Shown below are results of a simulation with three filling cycles.

Run 1: 5004 steps in 0.15 seconds


60 -| .-80

50- A pr^Ti
| "vsi|
/- 70
, go
/ I
40-
/ I /"( / .50

30-
/ I ' J -40
f I / \
20- / I / i /-V
/ I/ I / ^\
10-
\l XV ,* \ -10
0- --*'"-' L -Q
0 5 10 15 20 25 30 35 40 45 50
TIME

Figure 2. Masses of substrate and biomass during filling and emptying cycles.
8.3 Fed Batch Reactors 249

Run 1: 5004 steps in 0.35 seconds


10

Figure 3. Concentrations of product, substrate and biomass during filling and emptying
cycles. The volume is also shown.

References

Dunn, I.J., Mor, J.R., (1975) Variable Volume Continuous Cultivation


Biotechnol. Bioeng. 17, 1805.

Keller, R., Dunn, I.J. (1978) Computer Simulation of the Biomass Production
Rate of Cyclic Fed Batch Continuous Culture J. AppL Chem. Biotechnol. 28,
784.

8.3.5 Repeated Medium Replacement Culture


(REPLCUL)

System

Slow-growing animal and plant cell cultures require certain growth factors and
hormones which begin to limit growth after a period of time. To avoid this,
part of the entire culture is replaced with fresh medium. A single cycle of
repeated replacement culture is shown in Fig. 1. In this procedure part of the
medium volume (with cells) is removed after a certain replacement time and
replaced with fresh medium. Each cycle operates as a constant volume batch in
250 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

which the concentration of substrate decreases, while that of biomass increases.


The operating variables are replacement volume, replacement time, and
substrate concentration in the replacement medium. The initial conditions for
each cycle are determined by the final values in the previous cycle and the
replacement volume and concentration.

Replacement

Final Conditions VX VS
Initial Conditions

Figure 1. One cycle for medium replacement culture.

Model

The equations are those of batch culture, where for convenience the total
masses are used.
dVS = r
"dT sV
dvx

Monod kinetics is used.


The effective starting conditions for each batch can be calculated using the
final conditions of the previous cycle from the volume replaced, VR? and the
total volume, V, by the equations,

* VR
f=

VX = (1 - ) VXF
8.3 Fed Batch Reactors 251

VS= ( l - f ) V S F + V R S 0

where f is the volume fraction replaced.

Program

The program as shown on the CD-ROM makes use of the PULSE function to
vary the biomass and substrate concentrations corresponding to the
replacement of a fraction F of the culture medium. The time for each batch is
the value of INTERVAL.

Nomenclature

Symbols

D Dilution rate 1/h


f Fraction of volume replaced
KS Saturation constant g/m3
s Substrate concentration g/m3
X Biomass concentration g/m3
V Reactor volume m3
v Initial volume of liquid m3
o
vx Biomass in reactor kg
vs Substrate in reactor kg
VR Volume replaced m3
Y Yield coefficient
Specific growth rate 1/h

Indices

F Refers to final values at end of the cycle


S Refers to substrate
X Refers to biomass
0 Refers to initial and inlet values
252 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Exercises

Results

Fig. 2 shows how the biomass increases, until after six cycles the time profiles
become almost identical.
TIME= 19.29 X = 1.26

10 20 30 40 50 60 70 90 100
TIME

Figure 2. Oscillations of biomass and substrate concentrations with replacement cycles for
Interval 10 and F=0.8
8.3 Fed Batch Reactors 253

8.3.6 Penicillin Production in a Fed Batch


Fermenter (PENOXY)

A fed batch process is considered for the production of penicillin, as described


by Muttzall (1), The original model was altered to include oxygen transfer and
the influence of oxygen on the growth kinetics.

Figure I. Fed batch reactor showing nomenclature.

Model

As explained in the example FEDBAT the balances are:

Total mass

dt
Biomass:
d(MassX)
= Vr-X
dt
Substrate:
d(MassS)
= FSf+Vrs
dt
Product:
254 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

d(MassP) _.
= V fp
p
dt
Dissolved oxygen, neglecting the content of the inlet stream is calculated from

d(MassO)
= K L a*(O sat -0) + Vr 0
dt
The influence of biomass concentration on the oxygen transfer is
approximated here by

KX+X

The concentrations are calculated from

_MassX MassS MassP MassO


1\. """""""""""""^ , J , L , U
V V V V
The growth kinetics take into account the oxygen influence

The substrate uptake kinetics includes that amount used for growth, for product
and for maintenance

J*o ~ Jft o _/V


^ V
Y V"
Y ^
XS PS

Product production involves two terms whose constants are turned on and off
according to the value of |ii, as seen in the program.

Oxygen uptake includes growth and maintenance

=-TT
Y
xo
8.3 Fed Batch Reactors 255

Program

The program is on the CD-ROM.

Nomenclature

Symbols

F Feed flowrate m3/h


KLa Oxygen transfer coeff. 1/h
Ko Monod constant for oxygen kg/m 3
KS Monod constant for glucose kg/m 3
KX Constant for biomass effect on kg/m3
Mass Component mass kg
mo Maintenance coeff. for oxygen kg O/kg X h
ms Maintenance coeff. for glucose kg S/kg X h
Osat Saturation for oxygen kg/m3
Sf Feed cone, of glucose kg/m3
V Volume m3
'max Maximum volume m3
YPS Yield product to substrate kg/kg
YXO Yield biomass to oxygen kg/kg
YXS Yield biomass to substrate kg/kg
H-max max.specific growth rate 1/h

Exercises
256 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

II!

References

K. Mutzall, "Modellierung von Bioprozesses", Behr's Verlag, 1994.

Program and model developed by Reto Mueller, ETH Zurich.

Results

Run 1: 2023 steps in 0.117 seconds


0.008

20 40 60 80 100 120 140 160 180 200

Figure 2. Dynamics of the fed batch reactor.


8.4 Continuous Reactors 257

Run 3: 2021 steps in 0.15 seconds


120 -0.008

0.007
100
! I 0.006
80 i I 0.005

- 60
: I 0.004 O
Li I -0.003
40 I
-0.002
20
-0.001

0
0 20 40 60 80 100 120 140 160 180 200
TIME

Figure 3. Influence of initial KLa value from 100 to 160 h"^ on the S and O profiles.

8.4 Continuous Reactors

8.4.1 Steady-State Chemostat (CHEMOSTA)

System

The steady state operation of a continuous fermentation having constant


volume, constant flow rate and sterile feed is considered here (Fig. 1).
8.4 Continuous Reactors 257

Run 3: 2021 steps in 0.15 seconds


120 -0.008

-0.007
100 ^
1I 1 -0.006
80 I
I 1
-0.005

\ . !i i1
. ll -0.004 O
v i
- 60
'-, 1
40 \\1 1 |\ i
I i '\ *
-0.003

-0.002
20
\ \! \l -0.001
!
\\ \
-0
20 40 60 80 100 120 140 160 180 200
TIME

Figure 3. Influence of initial KLa value from 100 to 160 h"^ on the S and O profiles.

8.4 Continuous Reactors

8.4.1 Steady-State Chemostat (CHEMOSTA)

System

The steady state operation of a continuous fermentation having constant


volume, constant flow rate and sterile feed is considered here (Fig. 1).

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
258 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

D,S

Figure 1. Chemostat fermenter with model variables.

Model

The dynamic balance equations may be modified to apply only to the steady
state by setting the time derivatives equal to zero. The corresponding equations
are then:

For biomass,
0 = - D X + rx
For substrate,
0 = D (S0 - S) + rs
Growth kinetics,
rx = ^ X

Substituting into the biomass balance gives

\i = D
where S is determined by the kinetics

\i = f(S)
The Monod relation results in,

S =
The substrate balance gives,
X = Y(S 0 -S)
8.4 Continuous Reactors 259

The productivity of the reactor for biomass is X D.

The above equations represent the steady state model for a chemostat with
Monod kinetics. Using them it is possible to calculate the values of S and X,
which result from a particular value of D, and to investigate the influence of the
kinetic parameters.

Program

In Madonna programs, time can be used as a variable which will increase from
the starting time. Here it is renamed D. Thus equations will be solved for
increasing values of the dilution rate. Fortunately X and S can be explicitly
solved for in this problem. If not, the ROOT FINDER facility of Madonna can
be used. The program is found on the CD-ROM.

Nomenclature

The nomenclature is the same as the example CHEMO, Sec. 8.1.2.

Exercises
260 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

The steady state curves of X, S, and XD versus D are given Fig. 2. The results
in Fig. 3 were obtained by varying K$ in each run. An interesting effect can be
observed on the position of the washout point.
Run 1:113 steps in 0 seconds
10 - -4

9-
3.5
8-
*S**~\ !I -3
7-
m ^r Y.-I
\1 i
6-
f*r
i ~s;i
mm .
\! -2.5

v- 5
" S" 1 2 S
4-
1.5
3-
1
2-
// /{
1-
0.5

0- 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Figure 2. Steady state curves of X, S and XD versus D.


8.4 Continuous Reactors 261

Run 5:113 steps in 0 seconds

Figure 3. Runs obtained by varying KS from 0.2 to 1.0.

8.4.2 Continuous Culture with Inhibitory Substrate


(CONINHIB)

System

Inhibitory substrates at high concentrations reduce the specific growth rate


below that predicted by the Monod equation. The inhibition function may be
expressed empirically as

where KI is the inhibition constant (kg/m3).

If substrate concentrations are low, the term S2/Kj is lower in magnitude than
KS and S, and the inhibition function reduces to the Monod equation. In batch
cultures the term S2/Kj may be significant during the early stages of growth,
even for higher values of K[. The inhibition function passes through a
maximum at Smax = (Kg Ki)-5. A continuous inhibition culture will often lead
262 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

to two possible steady states, as defined by the steady state condition JLI = D and
as in shown Fig. 1.

D=

Figure 1. Possible steady states for a chemostat with inhibition kinetics.

One of these steady states (A) can be shown to be stable and the other (B) to be
unstable. Thus, only state A and the washout state (S = SQ) are possible.

Model

A model of a chemostat with its variables is represented schematically in Fig. 2.


F,S 0

+- F,S,X

Figure 2. Model variables.


8.4 Continuous Reactors 263

Cell material balance,

VdX
jj- = ^ i V X - F X
or,

where D is the dilution rate = F/V.

Substrate material balance,

VdS ^VX
= F (S 0 -S) -~^f
or,
dS MX
dT = D (S0 - S) -

where Y is the yield factor.

Program

When the system equations are solved dynamically, one of two distinct steady
state solutions are obtained, the stable condition A and the washout condition.
The initial substrate and organism concentrations in the reactor will determine
the result. This is best represented as a phase-plane plot X versus S. All results
indicate washout of the culture when the initial cell concentration is too low;
higher initial substrate concentrations increases the likelihood of washout.

Nomenclature

Symbols

D Dilution rate 1/h


KI Inhibition constant kg/m3
KS Saturation constant kg/m3
264 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

S Substrate concentration kg/m3


Smax Maximum in S for inhibition function kg/m3
X Biomass concentration kg/m3
Y Yield coefficient kg/kg
Specific growth rate 1/h

Indices

0 Refers to inlet
I Refers to initial value
m Refers to maximum

Exercises
8.4 Continuous Reactors 265

Results

Run 1:2000 steps in 0 seconds


1-1-2

10 15 20 25 30 35 40
TIME

Figure 3. Time course of X, S and U.

Run 10: 2000 steps in 0.0167 seconds


5 i

4.5 -

3.5

e/> 2.5

1.5

0.5

0
0.5 2.5

Figure 4. Phase-plane plot of X versus with varying ST from 0 to 5 kg/m3 using Batch Runs
with overlay.
266 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Run 9: 2000 steps in 0 seconds


5-
X,
4.5-

4 - ^^^>
3:2 (0.5)
*" t**
1 3:3 (0.5857)
3.5- ^ *t %% _ - - 3:4(0.6714)
\ * * C^1 v 3:5(0.7571)
3 "| "% % V 3:6(0.8429)
\ ^ x 3:7(0.9286)
3:8(1.014)
</) 2 . 5 - i ! l 3:9(1.1)

f/ // // /;^',--^,
2

1.5-

1 / / J
. *x
j* S ^ "^ -s^^-
K
f f / t / ,- rf'^T "^"^>*i'^' ** -^
0.5- f
:' / / * / Jf*'^ ^^^*^^$^S? **
!
0 * \ f 1 C & ( ^^*^^ **
0.5

Figure 5. Phase plane plot of influence of the initial biomass Xi from 0.5 to 1.1 for = 0.0.
Steady states upper left and lower right.

Run 20:2000 steps in 0.0167 seconds

Figure 6. Influence on the inhibition function made by varying KI between 1 and 3.

Reference

Edwards, V.H, Ko, R.C. and Balogh, S.A. (1972). Dynamics and Control of
Continuous Microbial Propagators Subject to Substrate Inhibition Biotechnol.
and Bioeng. 14, 939-974.
8.4 Continuous Reactors 267

8.4.3 Nitrification in Activated Sludge Process


(ACTNITR)

System

Nitrification is the process of ammonia oxidation by specialized organisms,


called nitrifiers. Their growth rate is much slower than that of the heterotrophic
organisms which oxidize organic carbon, and they can be washed out of the
reactors by the sludge wastage stream (Fs). In an activated sludge system
(Fig. 1) when the organic load (F So/V) is high, then the high biomass growth
rates require high waste rates. Nitrification will not be possible under these
conditions because the concentration of nitrifiers (Ni) will become very low.

O,F O 2, F4

Reieto*

2, F3

Figure 1. Configuration and streams for the activated sludge system.

Model

The dynamic balance equations can be written for all components around the
reactor and around the settler. The settler is simplified as a well-mixed system
with the effluent streams reflecting the cell separation.
Organic substrate balance for the reactor:

= F 0 So + F 2 S 2 -
268 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Ammonia substrate balance in the reactor:

R2Vt
= FQ AQ + F2 A2 - FI A

Reactor balance for the heterotrophic organisms:

pj7 = p2 O2 FI QI + RI Vj_

Reactor balance for the nitrifying organisms:

l
di = F 2 N 2 - F i N i + R2Vi

Organic substrate balance in the settler:

V 2 dS 2
j^ = F i S i - F3S2 - F4S2

Ammonia substrate balance for the settler:

V 2 dA 2
= F A - F4A2
3t I I -
Balance for heterotrophic organisms in the settler:

V 2 d0 2
dt = Fl l - F3 02

Balance for nitrifying organisms in the settler:

V2 dN2
34 = F i N i - F 3 N 2

The equations for the flow rates are given below.

Recycle flowrate:
F2 = F 0 R
where R is the recycle factor.

Reactor outlet flow:


FI = F2 + F0 = F O R + FO
Flow of settled sludge:
8.4 Continuous Reactors 269

where C is the concentration factor for the settler.

Flow of exit substrate:


F4 = FI - F3>
Flow of exit sludge wastage:
F5 = F3 - ?2.

Note that C and R must be chosen so that F5 is positive.

Monod-type equations are used for the growth rates of the two organisms.

R =
l

R
l^2max
2 = ^Ni =

Program

The program is given on the CD-ROM.

Nomenclature

Symbols

A Ammonia substrate concentration kg/m3


C Concentrating factor for settler -
F Flow rate m3/h
Fo-5 Flow rates, referring to the figure m3/h
KI Saturation constant of heterotrophs kg/m3
K2 Saturation constant of nitrifying organisms kg/m3
N Concentration of nitrifiers kg/m3
O Concentration of heterotrophs kg/m3
R Recycle factor -
270 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Rl Growth rate of heterotrophics kg/m3h


R2 Growth rate of nitrifying organisms kg/m3h
s Organic substrate concentration kg/m3
V Volumes m3
Y Yield coefficients kg/kg
Hi Specific growth rate of heterotrophs 1/h
Specific growth rate of nitrifying organisms 1/h

Indices

Flow and concentration indices referring to Fig. 1 are as follows:


0 Refers to feed and initial values
1 Refers to reactor and organic oxidation
2 Refers to settler and ammonia oxidation
3 Refers to recycle
4 Refers to settler effluent
5 Refers to sludge wastage
m Refers to maximum

Exercises
8.4 Continuous Reactors 271

Results

The results in Fig. 2 demonstrate the influence of flow rate on the effluent
organics 82- The ammonia in the effluent A2 is seen, in Fig. 3, to respond
similarly to FQ, but for a very high value of FQ = 1000 m3/h the nitrification
ceases, and A2 becomes the same as the inlet value AQ. This corresponds to
washout of the nitrifiers, which would be seen by plotting NI versus time.

Run 4: 405 steps in 0.0167 seconds


0.9 ,

0.8 -I ,/"

M if
If J
II /
It
0.3- rr
J 82:1(20)
02. II 82:2(180)
" 82:3(340)
0-1 JM | 82:4(500)

" .
6 2 4 6 8 10 12 14 16 18 20
TIME

Figure 2. Transient of S2 at various flow rates F0 (20 to 500m3/h, bottom to top).

Run 4:405 steps in 0.0167 seconds


0.1 -I

0.09-

0.08-

3 0.06 -|-_- A2:1(20)


A2:2(180)
005 J*
1
-_A2:3(340)
S. I A2:4(5QO)

0.04-

0.03-

0.02J
0 2 4 6 8 10 12 14 16 18 20
TIME

Figure 3. Ammonia in the effluent (A2) at various flow rates F0 (5 to lOOOm^/h, bottom to
top).
272 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

8.4.4 Tubular Enzyme Reactor (ENZTUBE)

System

A tubular, packed-bed, immobilized-enzyme reactor is to be investigated by


simulation. The flow is assumed to be ideal plug flow. The distribution of the
enzyme is not uniform and varies linearly from the inlet to higher values at the
outlet, as shown in Fig. 1.

Enzyme
concentration
Enzyme distribution

Distance along reactor, Z

Figure 1. Distribution of enzyme along the tubular reactor.

Model

The equations for steady state operation are given below.

Substrate balance,
dS 1
=
dZ ~v
Kinetics,

The linear flow velocity is increased by the presence of the solid enzyme carrier
particles according to
8.4 Continuous Reactors 273

F
V7
L = ~
Ae

The reaction velocity depends on the enzyme concentration,

vm = KE

and the linear distribution of enzyme distribution given by,

E = E0 + mZ

Program

The model is solved by renaming the independent variable, TIME, to be the


reactor length coordinate Z. The program is given on the CD-ROM.

Nomenclature

Symbols

A Reactor tube cross section m2


F Flow rate m3/h
K Rate constant 1/h
KM Michaelis-Menten constant kg/m3
m Enzyme distribution constant kg/m3 m
r Reaction rate kg/m3 h
S Substrate concentration kg/m3
vm Maximum reaction velocity kg/m3h
vz Linear flow velocity m/h
Z Reactor length m
e Void volume fraction of packing
E Enzyme concentration kg/m3
274 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Indices

0 Refers to inlet
S Refers to substrate

Exercises

Results

Flow rate is the primary operating variable, along with enzyme loading and
inlet concentration. In Fig. 2 the influence of F is seen in the steady-state, axial,
substrate profile.
8.4 Continuous Reactors 275

Run 6:1000 steps in 0.05 seconds


7

10 12 14 16 18 20

Figure 2. Substrate profile under the influence of F (1 to 10 m^/h, bottom to top).

8.4.5 Dual Substrate Limitation (DUAL)

System

In defined-nutrient growth media, one substrate can usually be made to be


limiting by adjusting its concentration relative to those of the other medium
components. In general, however, more than one substrate may limit the cell
growth rate. In this case the yield coefficients for the various components,
Yxsi> may vary depending upon the growth regime. This situation was
discussed by Egli et al. (1989), who examined results at steady state with dual
nutrient limitation. The present mathematical model simulates the transient
behaviour of such a dual (Si -carbon, 82 -nitrogen) nutrient-limited system
when carried out in a chemostat. The model assumes that the yield coefficients
are each a function of the ratio 81/82, i.e. the ratio of the carbon-nitrogen
substrate concentrations in the vessel. The original paper took the carbon-
nitrogen ratio in the feed stream as the controlling parameter. Here the
concentrations in the reactor are assumed to be controlling.

Model

Assuming a perfectly mixed, constant volume continuous-flow stirred-tank


reactor, the mass balance equations for the cells and for the two limiting
substrates are as follows:
276 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

= D
1
(SlFeed - Si) -

/O O \ I ^__i^^_ I i i "V
D (o2pee(j o2) lv YQO J ]LlA

where D = F/V.

The specific growth rate is modelled as


S
Si V 2

The yield coefficients are assumed to vary with the carbon-nitrogen ratio in the
reactor.
Si
RATIO = ^

The yield coefficients are varied according to RATIO using the following logic:

and YXS2 = Y2min if RATIO < B i


Y X Sl=Yi m i n and YXS2 = Y2max if RATIO > B2
where,
_ Y2min Y 2max
1 - and Bo = 1v~,T"
Bi = \r
Imax - 1mm

The boundaries of the three growth regimes in Fig. 1 are defined by the
quantities BI and B2.
8.4 Continuous Reactors 277

C limitation Double limitation N limitation


10

XSi
0.8
r
XS1

B2
S2

Figure 1. Limitation regions for carbon and nitrogen showing influence on yield.

The yield coefficients for biomass on nitrogen and carbon take maximum or
minimum values when only one substrate is limiting and vary linearly with
opposing tendencies in the double-limitation region.

Program

Note that the programing of this example is rather more complicated than usual
owing to the need to allow for the logical conditions of carbon limitation,
nitrogen limitation or both substrates together causing limitation. A partial
listing is seen below and the full program is on the CD-ROM.

(CALCULATION OF YIELD VALUES)


YXSl=if (RATIO < Bl) then YlMAX else ( if (RATIO >
B2) then Y1MIN else (Y1MAX+(RATIO-B1)/(B2-
B1)*(Y1MIN-Y1MAX)) )
YXS2 = if (RATIO < Bl) then Y2MIN else ( if (RATIO >
B2) then Y2MAX else (Y2MIN+(RATIO-B1)/(B2-
Bl)*(Y2MAX-Y2MIN)) )
278 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Nomenclature

Symbols

Bi Ratio of Y 2min /Yi max _


B2 Ratio of Y 2max /Yi min -

Cc Carbon source concentration kg/m3


Cn Nitrogen source concentration kg/m3
D Dilution rate 1/h
F Volumetric feed rate m3/h
KS Affinity constant kg/m3
R Reaction rates kg/m3 h
Si Carbon source concentration kg/m3
S2 Nitrogen source concentration kg/m3
X Biomass concentration kg/m3
Y Yield coefficient kg/kg
H Specific growth rate 1/h

Indices

1 Refers to carbon source


2 Refers to nitrogen source

Exercises
8.4 Continuous Reactors 279

Results

The startup of a continuous culture is shown in Fig. 2. Note that the nitrogen
level 82 in the reactor drops to a low level after 15 h and causes a change in the
yield coefficients. The influence of dilution rate on the system was investigated
by varying D from 0 to 1.5 as shown in Fig. 3.

Run 1: 305 steps in 0.0333 seconds


3-c 1

Figure 2. Startup of a continuous culture.

Run 4: 305 steps in 0 seconds

X 1.5

Figure 3. Variation of D from 0.1 to 1.5 (top to bottom).


280 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Reference

Egli, Th., Schmidt, Ch. R. (1989). "On Dual-Nutrient-Limited Growth of


Microbes, with Special Reference to Carbon and Nitrogen Substrates", in
Proceed. Microb. Phys. Working Party of Eur. Fed Biotech. Eds. Th. Egli, G.
Hamer and M. Snozzi, Hartung-Goree, Konstanz, 45-53.

This example was developed by S. Mason, ETH-Zurich.

8.4.6 Dichloromethane in a Biofilm Fluidized Sand


Bed (DCMDEG)

System

The process involves the removal of dichloromethane (DCM) from a gas stream
and the subsequent degradation by microbial action. The reactor consists of
biofilm sand bed column with circulation to an aeration tank, into which the
substrate and oxygen enters in the gas phase, or the substrate can be fed in a
liquid stream, as shown in Fig. 1. The column is approximated by a series of six
stirred tanks. The reaction is treated with homogeneous, double saturation
kinetics with dichloromethane (DCM) inhibition. The oxidation of one mole of
DCM produces 2 moles of HC1, making a hydrogen ion balance for pH im-
portant. The yield with respect to oxygen is 4.3 mg DCM/mg 62. In practice,
care must be taken to prevent stripping of DCM to the air stream.
8.4 Continuous Reactors 281

C
SR6>

SRin C jn , pH jn

Figure 1. Schematic of fluidized bed column with external aeration vessel.

Model

The model does not include a gas balance on the aeration tank, since it is
assumed that the gas phase dynamics are comparatively fast and hence an
equilibrium with the inlet concentration of oxygen and DCM may be assumed.
The biomass is assumed to grow slowly, and growth rates are therefore also not
modelled. The model for pH changes does not include buffering effects.

For the inlet section 1 at the bottom of the column the balances are as follows:

O2 balance,
dCQ1 ^
dt
282 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

DCM balance,
C
dC
Srl _ Srin~ C Srl
dt t
+
H ion balance,
dCHi i -CHI 2r S i
84900

Here T is the residence time of the liquid in one section of the column. The
constant 84,900 converts grams to moles and includes the stoichiometry.

pHi = -0.434 log |Cm|

Evaluation of rates for the inlet section 1:

V
maxCSrl -01

KI )

For the aeration tank the 62 and DCM balances are:

K L a 0 2(Co2eq-Coin)

dC R
-~ = (CSr6 - CSrin ) DCM (Cs2eq - Tr- (CSFO ~ CSrin
at V

Program

The program constants describe DCM entering the reactor in the gas stream.
The DCM concentration in the liquid feed is set to zero. The program is on the
CD-ROM.
8.4 Continuous Reactors 283

Nomenclature

H+ ion concentration in section n kg mol/m3


oin Inlet dissolved oxygen concentration g/m3
Oxygen saturation constant g/m3
DCM saturation constant g/m3
srin DCM inlet concentration g/m3
DCM concentration in section n g/m3
Oxygen concentration in section n liquid g/m3
CSFO DCM concentration in feed g/m3
CSG DCM gas concentration g/m3
F Feed rate m3/h
KI Inhibition constant g/m3
KLa Transfer coefficients for DCM and 2 1/h
KS Saturation constants g/m3
pHn pH in n section n pH units
R Recirculation rate m3/h
Oxygen uptake rate in section n g/m3 h
Substrate uptake rate in section n g/m3 h
VR Reactor volume m3
VT Volume of aeration tank m3
Vmax Maximum degradation rate g/m3 h
YSO Yield coefficient for DCM/oxygen
Liquid residence time in one section h

Exercises

11
284 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

The concentrations in the stream leaving the top of the column (CSr6) during
startup of the fluidized bed are shown in Fig. 2 for four values of F (0.5 to 10)
The change of the pH for one flow rate (F = 0.5) is shown in Fig. 3.
8.4 Continuous Reactors 285

Run 4: 55 steps in 0.0167 seconds

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
TIME

Figure 2. Fluidized bed startup for four values of F (0.5 to 10, bottom to top).

Run 1:55 steps in 0 seconds


3.5

2.5

CSR6:1
: 2 ... PH6:1

1.5

0.5

0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
TIME

Figure 3. Change of carbon substrate and pH in the top section 6 during startup.

Reference

D. Niemann Ph.D. Dissertation 10025, ETH, 1993.


286 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

8.4.7 Two-Stage Chemostat with Additional Stream


(TWOSTAGE)

System

Two chemostats are arranged in series (Fig. 1) with the intention that the first
operates at a relatively high rate of cell growth, while the second operates at low
growth rate, but high cell density, for secondary metabolite production.
Additional substrate may be fed to the second stage.

, 810

Hi
X1.S! US

Figure 1. Two-stage chemostat with two feed streams.

Model

The balance equations are written for each component in each reactor.

Stage 1 with sterile feed,

= F[S O -S!] -
8.4 Continuous Reactors 287

Stage 2 with additional substrate feed and an input of cells and substrate from
Stage 1,

V2 - [F + Fi]X2

V2 ^2.= F [Si - S2] + F! [Sio - S2] -


dt Yv

KS + S2

Productivity for biomass:

First stage,
Prodi =
V,
Both stages,
Prod2 =

Program

The program is on the CD-ROM.

Nomenclature

Symbols

F Volumetric feed rate m3/h


KS Saturation constant kg/m3
Prod Productivity for biomass kg/m3 h
S Substrate concentration kg/m3
3
V Reactor volume
X Biomass concentration kg/m3
Y Yield coefficient kg/kg
Specific growth rate 1/h
288 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Indices

0 Refers to tank 1 inlet


1 Refers to tank 1 and inlet of tank 2
2 Refers to tank 2 and outlet of system
10 Refers to separate feed for tank 2
m Refers to maximum

Exercises

Results

The results in Fig. 2 give biomass concentrations and productivities for both
tanks during a startup with a constant feed stream to the first tank (F = 0.5). In
Fig. 3 the influence on X2 of feed to the second tank Fl (0 to 1.0) with
constant F is shown.
8.4 Continuous Reactors 289

Run 1: 805 steps in 0.0333 seconds


T5

35 40

Figure 2. Biomass (Xj X2) and productivities for both tanks (F = 0.5).

Run 4: 805 steps in 0.0333 seconds


5

4.5

3.5

32.5
2

1.5

0.5

0
10 15 20 25 30 35 40
TIME

Figure 3. Influence on X2 of feed to the second tank (Ft = 0 to 1.0, curves right to left).
290 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

8.4.8 Two Stage Culture with Product Inhibition


(STAGED)

System

Products may inhibit growth rates. Under such conditions a multi-staged


continuous reactor as shown in Fig. 1 will have kinetic advantages over a single
stage. This is because product concentrations will be lower and consequently
the rates in the first tank will be higher as compared with a single tank. This
effect may be conveniently investigated by simulation. Batch cultures can be
expected to have similar kinetic advantages for product inhibition situations.

Figure 1. Two-stage chemostat with product inhibition.

Model

The inhibition function is expressed empirically as

When product concentrations are low, the equation reduces to the Monod
equation.
The product kinetics are according to Luedeking and Piret, with dependence
on both growing and non-growing biomass,

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
8.4 Continuous Reactors 291

rpn = (On + (3n \ln) Xn

In addition, the non-growth term, an, is assumed to be inhibited according to,

an - a"Q
~ 1i-r+rPn

When product concentrations are low, a = ano.

Kinetics for growth:

Kinetics for substrate consumption (neglecting consumption for product):

_ _rxn

where Y is the yield factor.

Mass balances:

Stage 1,

j- = F[So-Si] +r S iVi

jp = F[P 0 -Pi] + rp^j

Stage 2 with additional substrate feed FI,

dX2
V 2 -gjT- = F Xj - [F + F!]X2 + rX2V2

dS2
V2 -gj- = F [Si - S2] + FI [Sio - S2] + rS2V2

dP2
- =FPl- [F + Fi]P2 + rp2V2

Productivity for product:


First stage,
292 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Prodi =

Both stages,

Program

The program is on the CD-ROM.

Nomenclature

Symbols

F Volumetric feed rate m3/h


KI Inhibition constant
KS Saturation constant kg/m3
P Product concentration kg/m3
Prod Productivity for product kg/m3 h
r Reaction rate kg/m3 h
S Substrate concentration kg/m3
V Reactor volume m3
X Biomass concentration kg/m3
Y Yield coefficient kg/kg
a Non-growth product rate term kg P/kg X h
OC0 Non-growth term with no inhibition kg P/kg X h
P Growth dependent product yield kg/kg
Specific growth rate 1/h
Maximal specific growth rate 1/h

Indices

n Refers to tank n
0 Refers to tank 1 inlet
1 Refers to tank 1 and inlet of tank 2
8.4 Continuous Reactors 293

2 Refers to tank 2 and system of outflow


10 Refers to inlet concentration of tank 2

Exercises

Results

The startup and approach to steady state for the two stages is shown in Fig. 2.
The influence of the inhibition can be tested by varying KI from 0.1 to 10.0, as
shown in Fig. 3. The higher the KI the lower is the degree of inhibition and the
greater is the product concentration P2-
294 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Run 1: 255 steps in 0 seconds


4 r 10

3.5 ,
3

2.5

,
1.5

0.5

0
10 15
TIME

Figure 2. Startup and approach to steady state for the two stages.

Run 4: 255 steps in 0.0167 seconds


1.4.,

1.3

1.2.

1.1
I
1

0.9.

0.8

0.7 J
10 12 14 16 18 20 22 24 26
TIME

Figure 3. Product concentration P at various values of KI (1 to 5), curves bottom to top.

Reference

Herbert, D. (1961). A Theoretical Analysis of Continuous Culture Systems.


Soc. Chem. Ind. Monograph No. 12, London, 2L
8.4 Continuous Reactors 295

8.4.9 Fluidized Bed Recycle Reactor (FBR)

System

A fluidized bed column reactor can be described as 3 tanks-in-series (Fig. 1).


Substrate, at concentration SQ, enters the circulation loop at flow rate F. The
flow rate through the reactor due to circulation is FR. Oxygen is absorbed in a
well-mixed tank of volume VT. The reaction rate for substrate (r$) depends on
both S and dissolved oxygen (CL)- The rate of oxygen uptake (ro) is related to
S by a yield coefficient (Yos)- The gas phase is not included in the model,
except via the saturation concentration (CLS)- The oxygen uptake rate of
reactor can be determined by the difference in CL inlet and outlet values.

? So , Fn

Fluidized
Bed

F,S

Figure 1. Biofilm fluidized bed with external aeration.


296 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Model

The model balance equations are developed by considering the individual tank
stages and the absorber separately. The gas phase in the absorber is assumed to
be air.

Substrate balances:
For the absorption tank

dS FR
dF =
For each stage n

dSn FR
-3T = -^(Sn-!-Sn)- rsn
Oxygen balances:
For the absorption tank

r = ^(C L 3-C L )+K L a(C L s-C L )


VT
For each stage
dCLn FR
= (CLn
~dT" V -! ~ C Ln) ~ rOn
Kinetics for stage n:
V Tm
K n +Sn K 0 +C Ln

Program

The program is on the CD-ROM.


8.4 Continuous Reactors 297

Nomenclature

Symbols

CL Dissolved oxygen concentration g/m3


CLS Saturation oxygen concentration g/m3
F Feed flow rate m3/h
FR Recycle flow rate m3/h
KLa Transfer coefficient 1/h
Ks Saturation constant kg/m3
Ko Saturation constant for oxygen g/m3
r Reaction rate kg/m3 h
S Substrate concentration kg/m3
V Reactor volume of one stage m3
VT Volume of absorber tank m3
^m Maximum velocity kg/m3 h
x Biomass concentration kg/m3
Y Yield coefficient kg/kg and g/kg
T Inverse liquid residence time 1/h

Indices

0 Refers to feed
l,2,3,n Refer to the stage numbers
m Refers to maximum
O Refers to oxygen
S Refers to substrate
T Refers to aeration tank
X Refers to biomass

Exercises
298 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

Note from the results below that the steady state for oxygen is reached rather
quickly, compared to that of substrate.

Run 1:1003 steps in 0.0333 seconds

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Figure 2. Oxygen concentrations in fluidized bed reactor. Top of column is the lower curve.
8.4 Continuous Reactors 299

Run 1:10003 steps in 0.4 seconds


35

tf

Figure 3. Substrate concentrations from the run as in Fig. 2.

8.4.10 Nitrification in a Fluidized Bed Reactor


(NITBED)

System

Nitrification is an important process for wastewater treatment. It involves


the sequential oxidation of NFLt"1" to NO2~ and NC>3~ that proceeds
according to the following reaction sequence:

NH4+ + 1 02 -> N02- + H20 +2H+

NO2~ + O2 - NO3~
The overall reaction is thus

NH4+ + 2O2 NO3- + H2O + 2H+

Both steps are influenced by dissolved oxygen and the corresponding nitrogen
substrate concentration. Owing to the relatively slow growth rates of nitrifiers,
treatment processes benefit greatly from biomass retention.
300 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

In this example, a fluidised biofilm sand bed reactor for nitrification, as


investigated by Tanaka et al. (1981), is modelled as three tanks-in-series with a
recycle loop (Fig. 1). With continuous operation, ammonium ion is fed to the
reactor, and the products nitrite and nitrate exit in the effluent. The bed
expands in volume because of the constant circulation flow of liquid upwards
through the bed. Oxygen is supplied external to the bed in a well-mixed gas-
liquid absorber.

Model

The model balance equations are developed by considering, separately, the


individual tank stages and the absorber. Component balances are required for all
components in each section of the reactor column and in the absorber, where the
feed and effluent streams are located. Although the reaction actually proceeds
in the biofilm phase, a homogeneous model with apparent kinetics is employed
rather than a biofilm model, as in the example NITBEDFILM.

03.

Fluidized
bed

Figure 1. Biofilm fluidised-bed recycle loop reactor for nitrification.

In the absorber, oxygen is transferred from the air to the liquid phase. The
nitrogen compounds are referred to as Si, 82, and 83, respectively. Dissolved
8.4 Continuous Reactors 301

oxygen is referred to as O. Additional subscripts, as seen in Fig. 1, identify the


feed (F), recycle (R) and the flows to and from the tanks 1, 2 and 3, each with
volume V, and the absorption tank with volume VA-
The fluidised bed reactor is modelled by considering the component
balances for the three nitrogen components (i) and also for dissolved oxygen.
For each stage n, the component balance equations have the form

Similarly for the absorption tank, the balance for the nitrogen-containing
components include the input and output of the additional feed and effluent
streams, giving

The oxygen balance in the absorption tank must account for mass transfer from
the air, but neglects the low rates of oxygen supply and removal of the feed and
effluent streams. This gives

For the first and second biological nitrification rate steps, the reaction kinetics
for any stage n were found to be described by
v
r = ml Sin Qn
K + S K + O

V
r2n = m2 S2n n
K S K
2+ 2n O2+n

The oxygen uptake rate is related to the above reaction rates by means of the
constant yield coefficients, YI and 2, according to

ron = - H n Y i -r 2 n Y 2

The reaction stoichiometry provides the yield coefficient for the first step
302 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

YI = 3.5 mg O2/(mg NNH4)


and for the second step
Y2 = LI mg O2/(mg NNO2)

Program

The program is found on the CD-ROM.

Nomenclature

Symbols

F Feed and effluent flow rate L/h


FR Recycle flow rate L/h
Kj^a Transfer coefficient h
K Saturation constants mg/L
KI Saturation constant for ammonia mg/L
K2 Saturation constant for ammonia mg/L
O Dissolved oxygen concentration mg/L
Os and O* Oxygen solubility, saturation cone. mg/L
OUR Oxygen uptake rate mg/L
r Reaction rate mg/L h
S Substrate concentration mg/L
V Volume of one reactor stage L
VA Volume of absorber tank L
vm Maximum velocity mg/L h
Y Yield coefficient mg/mg

Indices

1,2,3 Refer to ammonia, nitrite and nitrate, resp.


1,2,3 Refer to stage numbers
A Refers to absorption tank
F Refers to feed
ij Refers to substrate i in stage j
m Refers to maximum
8.4 Continuous Reactors 303

Ol and O2 Refer to oxygen in first and second reactions


S1,S2 Refer to substrates ammonia and nitrite
S and * Refer to saturation value for oxygen

Exercises
304 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

Run 1: 519 steps in 0.2 seconds


6

10 15 20 25 30 35 40 45 50

Figure 2. Dynamic startup of continuous operation showing oxygen concentrations and


nitrogen compounds at the top of the column.

Run 2:10386 steps in 4.83 seconds


280 P 2.5

270

260

250

1^240.
5
,230-
c
<
220
M I
210

200 0.5

190

180
15 20 25 30 35 40
KLA

Figure 3. Parametric run of continuous operation showing oxygen and ammonia in the effluent
versus
8.4 Continuous Reactors 305

8.4.11 Continuous Enzymatic Reactor (ENZCON)

System

This example, schematically shown in Fig. 1 involves a continuous, constant


volume, enzymatic reactor with product inhibition in which soluble enzyme is
fed to the reactor.

EO.FE I S 1f P 1 F1

Figure 1. Continuous enzymatic reactor with enzyme feed.

Model

The mass balance equations are formulated by noting the two separate feed
streams and the fact that the enzyme does not react but is conserved.

Total flow:
FS + FE =

Mass balances:
dSi
= FsSo-FiS1+rsV
306 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

r = -FiP1+rpV

Kinetics with product inhibition:

r
S - - v mK M + S + (P/Ki)

vm = EI K2

rP = - 2 r s

Program

The program is found on the CD-ROM.

Nomenclature

Symbols

E Enzyme concentration kg/m3


F Flow rate m3/h
KI Inhibition constant
KM Saturation constant kg/m3
K2 Rate constant 1/h
P Product concentration kg/m3
r Reaction rate kg/(m3 h)
S Substrate concentration kg/m3
V Reactor volume m3
vm Maximum rate kg /(m3 h)

Indices

0 Refers to inlet values


8.4 Continuous Reactors 307

1 Refers to reactor and outlet values


E Refers to enzyme
P Refers to product
S Refers to substrate

Exercises

Results

Variations in the flows FE (Fig. 2) or Fs (Fig. 3) cause the product levels to


change.
308 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Run 3: 8004 steps in 0.1 83


SeC ndS
1.8

1.6

--- P1:2(0.2)
P1:3(0.3)

0.8
s
.0.6 / / ,-'"

0.4 tS
0.2

0 10 20 30 40 50 60 70 80
TIME
Figure 2. Performance for three values of FE.

Run 3: 8004 steps in 0.233


seconds
r4.5

3.5

3 x--'
2.5

.2

-1.5 r *>"* -^ - **" ~ ""

<*/r
jft
I . P1:2(1.5)
.. P1:3(2)

0 10 20 30 40 50 60 70 80
TIME

Figure 3. Performance for three values of Fs.

8.4.12 Reactor Cascade with Deactivating Enzyme


(DEACTENZ)

System

Biocatalysts usually deactivate during their use, and this has to be considered in
the bioreactor design. One of the methods to keep productivity fluctuations
low, and hence to efficiently utilize the biocatalyst, is to use a series of reactors
with biocatalyst batches having different times-on-stream in each reactor. In
8.4 Continuous Reactors 309

this example a series of three stirred tanks of a constant equal volume with
biocatalyst deactivating by first order reaction kinetics is investigated (Fig. 1).
After a time period ILAG? ^e biocatalyst from the tank with longest time-on-
stream (first tank in the cascade) is discarded and replaced by a fresh batch.
The streams are switched over so that tank 1 becomes tank 3, the last reactor in
the series. Other tanks are switched over correspondingly. This is equivalent to
replacing the used enzyme with fresh enzyme in tank 3 and moving the used
enzyme upstream from tank 3 to tank 2 to tank 1, which is easier to simulate.
(3-galactosidase was taken as an example of the biocatalyst. This obeys
Michaelis-Menten kinetics with competitive product inhibition, and the kinetic
constants were determined with considerable accuracy. The same constants are
used also in this substrate inhibition model.

F,S 0

F,Si F,S 2 F,S 3

Figure 1. Tanks in series reactor with immobilized enzyme.

Model

Using the stoichiometry, S > P, the mass balances for the ith tank (i = 1, 2, 3)
with the volume V can be written

Substrate

Product
= F(P M -Pi)
dt
Enzyme (active)
310 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

=r Ei v
where rate of substrate consumption is given by product inhibition
(competitive)
s
r i
Si = ' v max b i -7-Z~^\
K
V inh

According to the molar stoichiometry

R P i = -R Si

The rate of enzyme deactivation is assumed to be:

rEi = - kD EI

For each batch of enzyme in tank i


dE
i -i c

This equation can be applied by changing the initial conditions for each tank
when the enzyme is moved from tank to tank. Thus the final value in tank n
becomes the initial condition in tank i-1. The initial conditions can also be
calculated by analytical integration of the enzyme deactivation equation at
times corresponding to the respective ages of the biocatalysts in the respective
reactors (multiples of TLAG)- Fresh enzyme with the activity EQ is in the third
tank. The other tanks start with the following enzyme activities:

EI = E0 e C- (3 - i) ko TLAG]

Program

In the program on the CD-ROM note that the cost calculation at the end of the
program is included only as a comment but could be incorporated into the
program with the corresponding values for the constants.
8.4 Continuous Reactors 311

Nomenclature

Symbols

COST Specific product costs $/kg


E Enzyme concentration kg/m3
ECOST Enzyme cost $/kg
F Flow rate m3/h
ICOST Investment cost $/kg
kD Deactivation constant 1/h
Kinh Inhibition constant mol/m3
Km Michaelis - Menten constant mol/m3
OCOST Operating cost $/kg
P Total amount of product mol
RC Reactor refill cost $/kg
m Reaction rate of deactivation kg/(m3 h)
rs Reaction rate of substrate mol/(m3 h)
S Substrate concentration mol/m3
T Residence time h
t Time h
TDOWN Down time h
TLAG Time-on-stream difference h
v
max Maximum specific reaction rate mol/kg h

Indices

0 Refers to initial, feed


i Refers to reactor number

Exercises
312 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

The results from DEACTENZ show an exponential decrease of the biocatalyst


activity (Fig. 2), which causes dynamic changes in the substrate and product
concentrations (Fig. 3) in all three reactors.
8.4 Continuous Reactors 313

Run 1: 50000 steps in 0.917 seconds


0.5-,\ -4000

0.45-
\
s *** -3500
0.4- \ j**
% r* -3000
0.35- *. i**
"% ^

a ' 03 x
x "'' -2500 3
I
-2000 Q_
.0.25-
s . E2:1
"* 0.2- -. -' _- Totalproduct:1 -1500 p
*"">cr H
^r "'"'"-'"*. |
0.15-
-1000
0.1- _/ j+ " . -i.. .500
0.05- i...
0- ^ -" -0
0 100 200 300 400 500 600 700 800 900 1000
TIME

Figure 2. Exponential biocatalyst deactivation and total product during one run.

Run 1: 50000 steps in 0.933 seconds


140-i 120

120- 100

100-
80

8
l
C/l
-
a-
^ 60
cn
40
40-

20
20

0
0 100 200 300 400 500 600 700 800 900 1000
TIME

Figure 3. Dynamic changes in the substrate and product concentrations.


314 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

References

Prenosil, I.E., Peter, J., Bourne, J.R. (1980). Hydrolytische Spaltung des
Milchzuckers der Molke durch immobilisierte Enzyme im Festbett-Reaktor.
Verfahrenstechnik 14, 392.

Prenosil, J.E. (1981). Optimaler Betrieb fur einen Festbett- und einen Fliessbett-
Reaktor mit desaktivierendem Katalysator. Chimia 35, 226 .

Prenosil, J.E., Hediger, T. (1986). An Improved Model for Capillary-


Membrane, Fixed-Enzyme Reactors. In Membranes and Membrane Processes,
Plenum, N. Y., 515.

8.4.13 Continuous Production of PHB in a Two-Tank


Reactor Process (PHBTWO)

System

This example considers a two-stage process for the production of PHB, a


biopolymer. The kinetics of this fermentation is presented in the example PHB.
The structured kinetic model involves a Luedeking-Piret-type expression and
also an inhibition by the product. From this it might be expected that two stages
would be better than one, and it is the goal of this example to optimize the
process. The volume ratio and the feed rate are the obvious design and
operating parameters.

Sfeed,

> 82, F0

Figure 1. Configuration of the two-tank system.


8.4 Continuous Reactors 315

Model

The details of the structured model will not be repeated here (See PHB). The
biomass consists of a synthesis part R and the intracellular product P. The
biomass growth rate of R is proportional to the specific growth rate, which is
given by a two-part expression
S (S/Ks,2)n
(KS,i + S) -*

The synthesis rate of PHB is given by a two-part expression

The term -kiP represents a product inhibition.

The model requires component balances for P, R and S for both tanks, as seen
in the program. The relative reactor volumes are determined by the parameter
Vrat. The volumetric productivities are calculated to compare the results.

Program

The program is found on the CD-ROM

Nomenclature

Symbols

FO Feed flow rate m3/h


KI Inhibition constant, for (NH4>2SO4 kg/m3
KS Saturation constant kg/m3
n Hill Coefficient
P Product concentration (PHB) kg/m3
PROD Productivity kg/(m3h)
R Residual biomass concentration kg/m3
rp Rate of synthesis of PHB kg/m3
TR Rate of synthesis of R kg/m3
316 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Rate of substrate uptake kg/(m3 h)


Limiting substrate cone. NH4+ as
(NH4)2S04 kg/m3
Sfeed Feed concentration kg/m3
Vi and V2 Reactor volumes m3
X Biomass concentration kg/m3
YP/R Yield coefficient kg/kg
YR/S Yield coefficient, kg/kg
Specific rate of synthesis of R (rR/R) 1/h
MP Specific rate of synthesis of P (rp/P) 1/h

Indices

1 Refers to reaction 1 and tank 1


2 Refers to reaction 2 and tank 2
m Refers to maximum

Exercises
8.4 Continuous Reactors 317

Results

Run 1:119 steps in 0.0167 seconds


4

90 100

Figure 2. A run showing the dynamic approach to steady state for X, S, P in both tanks.

Run 20: 20380 steps in 5.78 seconds

0.2

Figure 3. Here with FO set at the optimum value of 1.24, the influence of VRAT is investigated
giving a value for the maximum in PROD corresponding to the OPTIMIZE results. VRAT is seen
not to be very important. Thus equal-sized tanks are adequate.
318 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

8.5 Oxygen Uptake Systems

8.5.1 Aeration of a Tank Reactor for Enzymatic


Oxidation (OXENZ)

System

The influence of gassing rate and stirrer speed on an enzymatic, aerated reactor,
as shown in Fig. 1, is to be investigated. The outlet gas is assumed to be
essentially air, which eliminates the need for a gas balance for the well-mixed
gas phase.

gas

Hi
Ill + 02

ii
air

Figure 1. Schematic of the enzymatic oxidation batch reactor.

Model

The reaction kinetics are described by a double Monod relation:

_S CL

The batch mass balances lead to:

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
8.5 Oxygen Uptake Systems 319

dS
= r
dF S

dCL *
= K L a(C L * -CL) - r s Y o /s

dP
=
dF ~ r s Y P/s

KLa varies with stirring speed (N) and aeration rate (G) according to:

KLa = kN 3 G- 5

where k = 4.78 x 10-13 with N in 1/h, G in m3/h and KLa in 1/h.

Program

The program is found on the CD-ROM.

Nomenclature

Symbols

CL Dissolved oxygen concentration g/m3


CLS>CL* Saturation oxygen concentration g/m3
G Aeration rate m3/h
KCL Saturation constant for oxygen kg/m3
^a Transfer coefficient 1/h
KS Saturation constant kg/m3
k Constant in K^a correlation complex
N Stirring rate 1/h
P Product concentration kg/m3
r Growth rate kg/(m3 h)
S Substrate concentration kg/m3
vm Maximum degradation rate g/(m3 h)
Y Yield coefficient kg/kg
320 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Indices

0 Refers to feed
o Refers to oxygen
p Refers to product
s Refers to substrate

Exercises

Results

The results in Fig. 2 show the influence of stirrer speed N on the dissolved
oxygen level. Variations from 30,000 to 5,000 1/h were made with the Batch
Run facility. Runs to obtain the results in Fig. 3 were made by varying the gas
flow rate G from 25 to 5 m3/h (curves top to bottom).
8.5 Oxygen Uptake Systems 321

Run 4: 1004 steps in 0.0167 seconds

7s"
8

7 -
_..-....._..-..._ --'/r
f
*
.*** *
6 -

5
i

CL1 (3e+4) *
CL:2 (2.1676+4) 1
3 -- CL:3 (1.3336+4) /
-^_CL:4 (5000) 4
2 -

{ -"-''

0
C 1 2 3 4 5 6 7 8 9 1 0
TIME

Figure 2. Influence of stirrer speed on dissolved oxygen levels.


Run 5:1004 steps in 0.0167 seconds

<j 6.5

1 2 3 4 5 6 7
TIME

Figure 3. Influence of the gas flow rate on dissolved oxygen levels.

8.5.2 Gas and Liquid Oxygen Dynamics in a


Continuous Fermenter (INHIB)

System

Cell growth is limited by the oxygen mass transfer rate, and hence by the
dissolved oxygen concentration. It is also inhibited by an inhibitory substrate S.
Liquid phase balances for X, S and 62 in the liquid phase are therefore used,
together with a gas phase oxygen balance to determine the rate of O2 supply.
322 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

To avoid washout of cells, it is important that the reactor should never enter the
range of inhibitory behavior. Schematic representation of a continuous aerated
fermenter is given in Fig. 1.

feed F, S 1 gas G, CG2

liquid F,S 2 , X

air, G, CG1

Figure 1. Schematic of the continuous fermentation with oxygen transfer.

Model

The liquid phase mass balances are as follows:

For biomass,
dX
rxVL
For substrate,
dS2

For oxygen,

dCL2

The kinetics are as follows:

L2
KS+S2+(S22/KI)K0+CL2

rx = |a X
8.5 Oxygen Uptake Systems 323

rs =
~

The balance for oxygen in the gas phase is:

The oxygen equilibrium relates the concentration in the gas phase to the liquid
phase saturation concentration,

CL2* = MC G2
The gas holdup fraction is,
VG = eV L
Proportional control of the feed rate, based on exit substrate concentration, can
be added with,
F = Fo + KpE

with E = S2set ~ $2- Here the sign must be adjusted depending on the substrate
region above or below the maximum kinetic rate.

Program

The program is on the CD-ROM.

Nomenclature

Symbols

CL Dissolved oxygen concentration mg/L


CLS Saturation oxygen concentration nig/L
E Control error g/L
F Flow rate L/h
G Gassing rate L/h
KI Inhibition constant g/L
324 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

KLa Transfer coefficient 1/h


Ko Oxygen saturation constant mg/L
KP Proportional control constant L/h/g/L
KS Saturation constant g/L
M Equilibrium coefficient -
OUR Oxygen uptake rate mg/h
r Reaction rate g/Lh
S Substrate concentration g/L
V Reactor volume L
X Biomass concentration g/L
Yx/s Yield coefficient g/g
Yo Mole fraction of oxygen in outlet gas -
8 gas/liquid volume ratio -
H Specific growth rate 1/h

Indices

0 Refers to feed
1 Refers to inlet
2 Refers to outlet
G Refers to gas
I Refers to inhibitor
L Refers to liquid
m Refers to maximum
O Refers to oxygen
P Refers to product
S Refers to substrate
X Refers to biomass
Refers to equilibrium

Exercises
8.5 Oxygen Uptake Systems 325
326 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

Run 4: 10003 steps in 0.367 seconds


8- 0.1

^ 0.09
7. . L

S2:1 (1) -0.08


\
.. CL2:1 (1)
6- S2:2 (4)
s -0.07
S2:3 (7) -0.06
5- CL2:3(7)

a'4-
|N ^^
% *%- N**
82:4(10)
-CL2:4(10)
-0.05

0.04

3- -0.03

" -^ .^^ -0.02


0.
"-""' .**-*uT-ZZ_H.r. r 0.01

1. 0
i) 1 2 3 4 5 6 7 8 9 K)
TIME

Figure 2. Dissolved oxygen versus time at various feed rates.

Run 1:10003 steps in 0.417 seconds


.70

DJ

2 3 4 5
TIME

Figure 3. Influence of the control on the reactor. The setpoint 82 is 5.0 kg/m^.
8.5 Oxygen Uptake Systems 327

8.5.3 Batch Nitrification with Oxygen Transfer


(NITRIF)

System

Nitrification in a biofilm fluidized bed is to be modelled. The sequential


oxidation of NtLj.* to NC>2~ and NC>3" proceeds according to:

NH4+ + 02 -> N02-


O2 -> NO3-

The two steps are shown schematically in Fig. 1.

Ammonium ion -^ Nitrite ion -> Nitrate ion

Figure 1. Reaction sequence for nitrification.

The stoichiometry is for the first step YI = 3.5 g O2/ (g NPfy-N) and for the
second step Y2 = 1.1 g O2/(g NO2-N).

Model

Neglecting the details of the biofilm diffusion, the apparent kinetics of this
biofilm process can be approximately described with homogeneous kinetics
that follow a double Monod limitation:

Si CL
*
I + S| KOI +C

S2 CL
= vm2
K 2 +S 2

The batch balances are as follows:

ForNH 4 + (Si):
328 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

~dT = - f i
For NO2~ (S2):
dS2
dt = ri - r2
For NO3- (S3):
dS3
~3T = r2
For oxygen (CL):
dCL
= - Y i r i - Y 2 r 2 + K L a(C L *-C L )

Program
The program is found on the CD-ROM.

Nomenclature

Symbols

CL Dissolved oxygen concentration g/m3


CLS Saturation oxygen concentration g/m3
K Saturation constants g/m3
KLa Oxygen transfer coefficient 1/h
r Reaction rate kg/m3
Si Concentration of NH4+ - N g/m3
S2 Concentration of NC>2~ - N g/m3
S3 Concentration of NO3~ - N g/m3
Vm Maximum degradation rates g/m 3 h
Yield coefficients g/g

Indices

0 Refers to feed
1,2,3 Refer to reaction steps
O Refers to oxygen
s Refers to substrate
8.5 Oxygen Uptake Systems 329

Exercises
330 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

Run 1:313 steps in 0 seconds


100
+* -
90 X *'
80

70
...X'-''
CO 60

$ 50
5) 40
30
20

10
0
1.5
TIME

Figure 2. NH4+, NO2~ and NC>3" and dissolved oxygen in a batch nitrification with KLa = 40 h"1/

Run 3: 313 steps in 0.0167 seconds


8

2.5

Figure 3. NH 4 + and dissolved oxygen in batch nitrification using three values of KLa from 20
8.5 Oxygen Uptake Systems 331

8.5.4 Oxygen Uptake and Aeration Dynamics


(OXDYN)

System

The aeration of a batch culture (with essentially constant biomass X) is stopped


and the dissolved oxygen (CL) is allowed to fall zero before re-aerating. The
slope of the CL curve is the oxygen uptake rate and it is approximated by the
slope of the electrode response curve CE curve. The dynamics of the electrode
are known, and it is desired to investigate the lag effects as shown in Fig. 1.

Model

The following equations represent the model:

Oxygen uptake rate,


OUR = q 0 2X

Specific OUR,
qo2m CL
102 - KQ + CL
Oxygen balance,
dCL *
T = K L a(C L *-C L ) -OUR
332 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

C (mg/L)
6

0 20 40 60 80 time(s)

Figure 1. Typical response of the batch oxygen uptake and reaeration experiment.

Measurement dynamics for the liquid film may be important with a viscous
culture,
dCp CL - Cp
TF
and for the electrode lag,

dt =

Program

Experimental data, in the file OXDYNDATA, and the program are found on
the CD-ROM.

Nomenclature

Symbols

C Oxygen concentrations g/m3


KLa Transfer coefficient 1/h
8.5 Oxygen Uptake Systems 333

KO Saturation constant for oxygen g/m3


OUR Oxygen uptake rate g/m3s
Q Specific oxygen uptake rate g/kgs
X Biomass concentration kg/m3
T Time constants

Indices

E Refers to electrode
F Refers to liquid film
L Refers to liquid
m Refers to maximum
02,0 Refer to oxygen
S Refers to saturation
* Refers to saturation

Exercises
334 g Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

The results shown in Fig. 2 demonstrate the effect of changing K^a. Runs
varying the electrode time constant TE gave the results of Fig. 3.

Run 2:10004 steps in 0.183 second:


8

70 80 90 100

Figure 2. Aeration turned on at low CL for two KLa values.


8.5 Oxygen Uptake Systems 335

Run 3: 10004 steps in 0.183 seconds


8

10 20 30 40 50
TIME

Figure 3. Variation of the electrode time constant, TE from 1 to 25.

8.5.5 Dynamic Oxygen Electrode Method for KLa


(KLADYN, KLAFIT and ELECTFIT)

System

A simple and effective means of measuring the oxygen transfer coefficient


(^a) in an air- water tank contacting system involves first degassing the batch
water phase with nitrogen (Ruchti et al., 1981). Then the air flow is started and
the increasing dissolved oxygen concentration is measured by means of an
oxygen electrode.
336 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

VQ.CQ

Hii

V
G> CGO

Figure 1. Aerated tank with oxygen electrode.

As shown below, the influence of three quite distinct dynamic processes play a
role in the overall measured oxygen concentration response curve. These are
the processes of the dilution of nitrogen gas with air, the gas-liquid transfer and
the electrode response characteristic, respectively. Whether all of these processes
need to be taken into account when calculating K^a can be determined by
examining the mathematical model and carrying out simulations.

Measurement
CF

Gas phase Liquid phase Electrode Electrode


diffusion film

Figure 2. Representation of the process dynamics.

Model

The model relationships include the mass balance equations for the gas and
liquid phases and equations representing the measurement dynamics.

Oxygen Balances
The oxygen balance for the well-mixed flowing gas phase is described by

dCG_ _
VG = G (CGO - CG) - KLa (CL* - CL) VL
dt
8.5 Oxygen Uptake Systems 337

where VG/V = TG , and K^a is based on the liquid volume.


The oxygen balance for the well-mixed batch liquid phase, is

= K L a(C L *-C L )V L
dt
The equilibrium oxygen concentration CL* is given by the combination of
Henry's law and the Ideal Gas Law equation where

RT
r
CL * = rCG

and CL* is the oxygen concentration in equilibrium with the gas concentration,
CG- The above equations can be solved in this form as in simulation example
KLAFIT. It is also useful to solve the equations in dimensionless form.

Oxygen Electrode Dynamic Model


The response of the usual membrane-covered electrodes can be described by
an empirical second-order lag equation. This consists of two first-order lag
equations to represent the diffusion of oxygen through the liquid film on the
surface of the electrode membrane and secondly the response of the membrane
and electrolyte:

dC F _ C L -C F
dt TF
and
dCg Cp ~Cg
dt " TE

Tp and TG are the time constants for the film and electrode lags, respectively. In
non- viscous water phases Tp can be expected to be very small, and the first lag
equation can, in fact, be ignored.

Dimensionless model equations


Defining dimensionless variables as

CG
C = CL =
CGO C GO (RT/H) TG
the component balance equations then become
338 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

u
dt' V ' VG H
and

Initial conditions corresponding to the experimental method are,

t1 = 0 ; C' L =C' G =0

In dimensionless form the electrode dynamic equations are

C
= L-CF
dt
and
dC C
E = F-CB
dt' TE/^G
where Cp is the dimensionless diffusion film concentration.

C GO (RT/H)

and CE is the dimensionless electrode output

CE
C GO (RT/H)

As shown by Dang et al. (1977), solving the model equations by Laplace


transformation gives
1 ,RTVL
= =
KLa + ( H
V-. + 1) TO + TE + TF
G

where oc is the area above the CE versus t response curve, as shown in Fig. 3.
8.5 Oxygen Uptake Systems 339

1.0

CE'

Time (s)

Figure 3. Determination of the area a above the CE' versus time response curve.

Program

The program KLADYN can be used to investigate the influence of the various
experimental parameters on the method, and is formulated in dimensionless
form. The same model, but with dimensions, is used in program KLAFIT and
is particularly useful for determining K^a in fitting experimental data of CE
versus time. A set of experimental data in the text file KLADATA can be used
to experiment with the data fitting features of Madonna. All are on the CD-
ROM.
The program ELECTFIT is used to determine the electrode time constant in
the first-order lag model. The experiment involves bringing CE to zero by first
purging oxygen from the water with nitrogen and then subjecting the electrode
to a step change by plunging it into fully aerated water. The value of the
electrode time constant, TE can be obtained by fitting the model to the set of
experimental data in the file, ELECTDATA. The value found in this
experiment can then be used as a constant in KLAFIT.

Nomenclature

Symbols

Oxygen concentration g/m3


340 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

G Gas flow rate m3/s


H Henry coefficient Pa m3/mol
KLa Oxygen transfer coefficient 1/s
RT/H (Gas constant)(Abs. temp.)/Henry coeff. -
t Time s
V Reactor volume m3
a Area above Cn'-time (s) curve s
Time constant s

Indices

E Refers to electrode
F Refers to film
G Refers to gas phase
L Refers to liquid
Prime denotes dimensionless variables

Exercises
8.5 Oxygen Uptake Systems 341

Results

Run 1: 206 steps in 0 seconds

0 20 40 60 80 100 120 140 160 180 200


TIME

Figure 4. Response of CG, CL and CE versus Ttime from KLAFIT.


342 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Run 1: 206 steps in 0 seconds

TIME

Figure 5. A fit of experimental data (open circles) as dimensionless CE versus time (s) to
determine KLA using KLAFIT, which gives 0.136 1/s.

References

Dang, N.D.P., Karrer, D.A. and Dunn, IJ. (1977).Oxygen Transfer Coefficients
by Dynamic Model Moment Analysis, Biotechnol. Bioeng. 19, 853.

Ruchti, G. Dunn, IJ. and Bourne J.R. (1981). Comparison of Dynamic Oxygen
Electrode Methods for the Measurement of KLa, Biotechnol. Bioeng., 13, 277.

8.5.6 Biofiltration Column for Removing Two


Inhibitory Substrates (BIOFILTDYN)

System

Biofiltration is a process for treating contaminated air streams. Moist air is


passed thrpugh a packed column, in which the pollutants in the contaminated
air are adsorbed onto the wetted packing. There in the biofilm solid phase the
resident population of organisms oxidizes the pollutants.
8.5 Oxygen Uptake Systems 343

L, S

r
G, C1T6

Tank6 Transfer
wifsp#K
iS:Tl;niiiS?'
Gas
TriT6 Illllpll

G.Q1T5

'"';:"H? H'MK'SK
*
Transfer
Tanks illlii
Gas
llllll
I
Tank4
Transfer
tlwilffi
Gas I l
lll
lll

I
fankS
Transfer
illlll
i
Gas Illlllll
I
Tank 2
Transfer
^
: : ' : . ; : - : ^ v \ r : : ^ -^ ;.;ji: ' ;' ';':'

^^iiiriS
Gas i;|tii^|;i;||

I
Tank1
Transfer
^^iinS^p:
I
^
Gas TriT1 OliiulllI

L, S1T1

Figure 1. Biofiltration countercurrent column.

Such columns can be run with a liquid phase flow (bio-trickling filter) or only
with moist packing (biofilter). The work of Deshusses et al. (1995) investigated
the removal of two ketones, methyl isobutyl ketone (MIBK) and methyl ethyl
ketone (MEK), in such a biofilter. The kinetics of this multi-substrate system is
especially interesting since both substances exhibit mutual inhibitory effects on
their rates of degradation.
344 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Model

The model requires stagewise mass balances in the gas and liquid phases for
both components. Transfer takes place from the gas to liquid phases with
reaction in the liquid phase. The symbols used for the concentrations of
substrates 1 and 2 in the n th tank are for the liquid phase Sixn and S2Tn and
for the gas phase CiTn and C2Tn- The reaction rates are RiTn and R2Tn and the
transfer rates are designated TriTn and Tr2Tn-
G, C1Tn, C2Tn L, S1Tn+1, S2Tn+1

G,C1Tn-1,C2Tn-1 L, S1Tn, S2Tn

m
nth
Figure 2. Single n stage for the biofiltration countercurrent column.

Referring to above figure, the mass balances for a single tank can be written as:

Gas phase

^OL~(G(C2Tn.1-C2Tn)-Tr2Tn
Liquid phase

-^ = ^- (L(SlTn+l - SlTn ) +TrlTn -rlTn VS )


8.5 Oxygen Uptake Systems 345

1Q 1

n =
d ^~ (L(S2Tn+l - S2Tn ) + Tr2Tn -r2Tn VS )

Here the reaction is assumed to occur in a solid phase of volume Vs.

Vs = (1 - EG - e L ) ^E.
For the transfer terms

Vc
Tr
2Tn = K L a ( S 2EQn - S2Tn)~

For the gas-liquid equilibria:

For the reaction rate terms the following equations are used to describe the
mutual inhibition. Note that oxygen is assumed to be in excess.

For substrate 1 (MEK) in tank n:


v
r
mi SlTn
lTn = "

For substrate 2 (MIBK) in tank n:


V
r
m2 S2Tn
2Tn -
V i i SlTn |
|1+

Program

The program developed by M. Waldner, ETH, is given on the CD-ROM.


346 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Nomenclature

Symbols

C Concentration in gas phase kg/m3


G Gas flow rate m3/s
Ki Inhibition constant kg/m3
KLa Mass transfer coefficient 1/s
Km Monod coefficient kg/m3
L Liquid flowrate m3/s
M Partition coefficient
N Number of tanks
r Reaction rate kg/m3s
S Concentration in liquid phase kg/m3
Tr Transfer rate kg/s
VC Volume of column m3
VG Volume of gas phase m3
VL Volume of liquid phase m3
vm Maximum reaction velocity kg/m3s
VS Volume of solid phase m3
z Length or height m
e Volume fraction

Indices

Eq Refers to equilibrium value


G Refers to gas
in Refers to inlet
L Refers to liquid
M Refers to maximum
n Refers to nth stage
Tn Refers to nth tank
1 Refers to methyl ethyl ketone (MEK)
2 Refers to methyl isobutyl ketone (MIBK)
8.5 Oxygen Uptake Systems 347

Exercises

Reference

Deshusses, M. A, Hamer, G. and Dunn, I. J. (1995) Part I, Behavior of Biofilters


for Waste Air Biotreatment: Part I, Dynamic Model Development and Part II,
Experimental Evaluation of a Dynamic Model, Environ. Sci. Technol. 29,
1048-1068.
348 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

Run 1:533 steps in 0.667 seconds


1.2

0.0025-

^ 0.002- 0.8 ^

. 0.0015- 0.6
&-

0.001 - ,0.4 W

H
'0.2 (0

0 5e+5 1e+6 1.5e+6 2e+6 2.5e+6 3e+6 3.5e+6 4e+6 4.5e+6 5e+6
TIME

Figure 3. Dynamic startup of the column.

Run 2: 20660 steps in 25.2 seconds


0.1

5e-5 1e-4 1.5e-4 2e-4 2.5e-4 3e-4 3.5e-4 4e-4 4.5e-4 5e-4

Figure 4. Influence of gas flowrate on the steady state fraction removed.


8.5 Oxygen Uptake Systems 349

8.5.7 Optical Sensing of Dissolved Oxygen in


Microtiter Plates (TITERDYN and
TITERBIO)

System

Measurement of dissolved oxygen in microtiter plates is of potential interest for


the screening of oxygen-consuming enzymes (e.g., oxidases), aerobic cell
activities, and biological degradation of pollutants, and for toxicity tests. John et
al. developed microtiter plates with the integrated optical sensing of dissolved
oxygen by immobilization of two fluorophores at the bottom of 96-well
polystyrene microtiter plates. The oxygen-sensitive fluorophore responded to
dissolved oxygen concentration, whereas the oxygen-insensitive one served as
an internal reference. As modelled in TITERDYN, oxygen transfer coefficients
were determined by a dynamic method in a commercial microtiter plate reader
with an integrated shaker. For this purpose, the dissolved oxygen was initially
depleted by the addition of sodium dithionite and, by oxygen transfer from air,
it increased again after complete oxidation of the dithionite. Available
commercial readers have an intermittent operation. After a certain period of
shaking, the plate is moved around to measure dissolved oxygen concentration.
During this period the plate moves more slowly and oxygen transfer rate is
reduced. This may lead to oxygen depletion during the measurement process.
It is essential to know the size of the errors that are introduced by this
intermittent procedure. This is evaluated by the simulation example TITERBIO.

Microtiter plate

Filter

Light

Figure 1. Microtiter well showing light path and sensor layer.


350 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Model

Experiments involved measuring the oxygen uptake rate by removing the


oxygen from the liquid using a chemical reaction (oxidation of sodium
dithionite). Oxygen is depleted immediately after the addition of dithionite.
After the consumption of the dithionite the oxygen transfer increased the
oxygen in the liquid. The following model was used to evaluate the KLa value.

dCT
dt

where CL is the dissolved oxygen concentration, CL* is the saturation value and
OUR is the oxygen uptake rate in mM/min.
The experiment starts with high values of dissolved oxygen concentration,
CL After addition of dithionite OUR increases as calculated by

OUR = ko CL CD

As oxidation proceeds the dithionite concentration changes according to

dC D ^ 2 f
dt "" 3
In order to account for some time delay of the sensor a first order equation was
used

dC E ^C L -C E
dt TE
The time constant TE was estimated to be about 1 s.

In further experiments this method was also applied to simulate a microbial


cultivation in the wells of a microtiter plate. In this case the OUR value was
taken to be a constant value as measured in a larger fermentation vessel. KLa
varied periodically simulating the high value during shaking and the lower
value during the measurement period. The questions of interest are how much
the measured OUR or KLa would differ from the actual one provided KLa or
OUR were already known.
8.5 Oxygen Uptake Systems 351

Program

Two separate programs are given on the CDROM: TITERDYN for the chemical
oxidation with re-aeration and TITERBIO for the biological oxidation and re-
aeration dynamics during a cultivation in a microplate reader. For the program
TITERDYN there is experimental data on the file TITERDYNDATA available
to allow fitting the value of KLa. In TITERBIO KLa during measurement is a
fraction of KLa during shaking and is determined by the parameter kmax. KLa
during measurement is defined as,
KLameasure=KLashaking*(kmax-l)/kmax.

In the original model settings, kmax has a value of 2. The larger kmax, the
larger the error of KLa or OUR estimation.

Nomenclature

The program TITERDYN uses minutes and TITERBIO uses seconds.


Additional symbols are defined in the programs.

Symbols

CD Dithionite concentration mM
CL Oxygen concentration mM
ko Rate constant for dithionite reaction 1/min mM
^a Transfer coefficient 1/s
KQ Saturation constant for oxygen mM
OUR Oxygen uptake rate mM/s
Q Specific oxygen uptake rate mM/ s
TE Time constant for measurement s

Indices

E Refers to electrode
D Refers to dithionite
L Refers to liquid
S and * Refer to saturation
352 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Exercises
8.5 Oxygen Uptake Systems 353

Results

1:1019 steps in 0.0167 seconds


'0.1

0 100 200 300 400 500 600 700 800 900 1000
TIME

Figure 2e Dynamics of biological uptake and reaeration. Program TITERBIO.

Run 1: 834 steps in 0.0333 seconds

it 'VtV'tj " *

Figure 3. Data fitting using TTTERDYN, yielding KLa=0.201, Calcfact=102 and


Duration=0.48.

Reference

John, G.T., Klimant, I., Wittmann, C., Heinzle, E. (2003). Integrated Optical
Sensing of Dissolved Oxygen in Microtiter Plates - A Novel Tool for Microbial
Cultivation, Biotechnol. Bioeng., 81, 829-836.
354 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

8.6 Controlled Reactors

8.6.1 Feedback Control of a Water Heater


(TEMPCONT)

System

A simple feedback control system involving a stirred tank, temperature


measurement, controller and manipulated heater is shown in Fig. 1.

T 0 ,F
IT*
F,T R

ip

Figure 1. Feedback control of a simple continuous water heater.

Model

The energy balance for the tank is


dTR F Q

where Q is the delayed heat input from the heater represented by a first order
lag

dt TQ

The measurement of temperature is also delayed by a sensor lag given by

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
8.6 Controlled Reactors 355

u*sens .... *R ~ *sens


dt Tsens

A proportional-integral feedback controller can be modelled by

where the control error is given by

Program

Random disturbances in flowrate or feed temperature can be generated using


the RANDOM function in Madonna, as explained in the HELP on the CD-
ROM.

Nomenclature

Symbols

cp Specific heat kJ/(kg C)


f Frequency of oscillations 1/h
F Flow rate m3/h
Kp Proportional control constant kJ/(h C)
Q Heat input kJ/h
T Temperature C
V Reactor volume m3
8 Error C
p Density kg/m 3
TD Differential control constant h
TI Integral control constant h
TQ Time constant for heater h
Time constant for measurement h
356 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Indices

C Refers to controller
R Refers to reactor
sens Refers to sensor
set Refers to setpoint
0 Refers to inlet or initial

Exercises
8.6 Controlled Reactors 357

Results

Run 1: 14286 steps in 0.0667 seconds


7000

6000

3000

1000
70 80 90 100

Figure 2. Approach to steady state for a setpoint of 80C.


358 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Run 1: 28571 steps in 0.2 seconds


-1.2e+4

1e+4

8000

100.
6000

80 4000

2000
60
0

-2000

-4000
20 40 80 100 120 140 160 180 200
TIME

Figure 3. Response to a step change in the inlet temperature TO at 120 h. The controller
constant Kp was set higher than in the run of Fig. 2.

8.6.2 Temperature Control of Fermentation


(FERMTEMP)

System

Heat effects in fermentation can be important, especially on a large scale.


Shown in Fig. 1 is a batch fermentation process, during which the cooling water
flowrate is controlled by a feedback controller. The rate of heat generation is
related to rate of substrate uptake by a constant yield factor YQS. The cooling
coil is modelled as a well-mixed system.
8.6 Controlled Reactors 359

Water

Figure 1. Feedback control of the temperature in during a fermentation.

Model

The batch fermentation model is given by,


dX
df =

dS -H
dt = Y

^ = Ks+S

The energy balance equation for the reactor is,


dT
dtR _ rQ UA
~ (TR-Tc)
VpCp
360 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

where,
TQ = ^XY Q S /Y

For the well-mixed cooling coil, the energy balance equation is:

dTr F (Tcin TC) + UA


- (TR ~ Tc)
The controller is a proportional-integral type

F = F0 + K P 8 +

Program

As seen on the CD-ROM and below, the control equations can be written in
terms of the error and its integral.

{CONTROL EQUATIONS FOR FLOWRATE}


d/dt(EInt)=E
F=FO+KP*E+(KP/TI)*EInt
limit F> = 0
E=TR-TSET

Nomenclature

Symbols

Cp Heat capacity kcal/(kg C)


F Flow rate m3/h
Kp Controller constant m3/(h C)
KS Saturation coefficient kg/m3
UA Reactor transfer-area constant kcal/kg
r Rate of heat production and transfer kcal/(m3 h)
V Reactor volume m3
8.6 Controlled Reactors 361

X Biomass concentration kg/m3


Y Yield coefficient kg/kg
YQS Heat yield for substrate kcal/kg
P Density kg/m3
T Time constant of controller h
s Substrate concentration kg/m3
e Temperature error C
H Specific growth rate 1/h

Indices

C Refers to coolant and cooling


I Refers to integral control
m Refers to maximum
Q Refers to heat
R Refers to reactor
S Refers to substrate
O Refers to normal value and inlet value
P Refers to proportional
set Refers to setpoint (desired value)

Exercises
362 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

TIME= 3.452 S = 18.25 TR= 24.6

Figure 2. Cooling flow starts when TR > Tset (25 C); after batch growth finishes at time=4.6 h
the reactor cools. Here Kp=0.6 and TI = 0.6.
8.6 Controlled Reactors 363

Run 6: 9480 steps in 0.217 seconds


'0.2

100 150 200 250 300 350 400 450 500

Figure 3. With Parameter Plot, the integral of the error and minimum water temperature versus
Kp for a fixed value of Tj=9.

8.6.3 Turbidostat Response (TURBCON)

System

Although not so widely used as the chemostatic type of operation of


continuous culture, the turbidostat may offer advantages for the investigation of
particular problems. As shown in Fig. 1, the flow rate of the incoming substrate
is controlled by the biomass concentration (more correctly, the turbidity) in the
vessel. In practice, this control is usually on-off or proportional, but more
sophisticated control would be simple to implement.
364 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

F,S0

Feed pump
X,S

Turbidometer

Figure 1. Feedback control of the biomass concentration using a turbidostat.

Model

For the well-mixed tank with Monod growth:

dS M_X
dt" = F(So-S) - -y~
dX FX
dT = -IT
Considering product production with Luedeking-Piret kinetics:
dP FP
dT = -"V" +

The turbidometer control is modelled by:


KPP ,.
F = F0 + K P e + fedt
s
8.6 Controlled Reactors 365

8 = (X-X S et)

Program

The program is on the CD-ROM.

Nomenclature

Symbols

A Growth-associated constant
B Nongrowth-associated constant 1/h
F Flow rate m3/h
Fo Normal feed flow rate m3/h
Kp Proportional controller constant m6/h kg
KS Saturation constant kg/m3
P Product concentration kg/m3
S Substrate concentration kg/m3
V Reactor volume m3
X Biomass concentration kg/m3
Y Yield coefficient kg/kg
e Error kg/m3
Specific growth rate 1/h
Integral control time constant

Indices

m Refers to maximum
P Refers to proportional control
S and set Refers to setpoint
0 Refers to inlet stream
366 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Exercises

Results

Run 1:105 steps in 0 seconds


1-5

,3.5 c/>

1
5
TIME

Figure 2. Startup and response of the controlled reactor.


8.6 Controlled Reactors 367

Run 1: 205 steps in 0.0167 seconds


3.5- -5

45
3.

4
2.5-

-3.5
2" _X:1
-3 C^
X" F:
1.5. " '
-2.5
1 ,m -2
n e
\
VJ.O
r i T... -1.5

0- j u ' " ...^. ._..._. ._..._,,....


-1
4 6 8 10 12 14 16 18 20
TIME

Figure 3. Response of the controlled reactor to a step change in X se t o

8.6.4 Control of a Continuous Bioreactor with


Inhibitory Substrate (CONTCON)

System

The continuous fermenter is equipped with feedback control based on substrate


measurement, as shown in Fig. 1. This type of controlled fermenter has been
referred to as an auxostat.
368 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

F,S,
Feed
pump X,S
' Substrate
measurement

Controller

Figure 1. Flow diagram of a feedback loop to control substrate concentration.

Model

The biomass and substrate mass balances are the same as in the previous
TURBCON model.

Kinetics:

Biomass balance,
V = VX-FX
dt
or,
f

where D is the dilution rate (= F/V). Thus steady-state behaviour, where dX/dt
= 0, is represented by the conditions that |u = D.
Substrate mass balance,

dt
or,

f "><*>--f
8.6 Controlled Reactors 369

where Y is the yield factor for biomass from substrate. Also from this equation
at steady state, since (j = D and dS/dt = 0, the steady-state cell concentration is
given by
X = Y(S 0 -S)

A continuous inhibition culture will often lead to two possible steady states, as
defined by the steady-state condition (a = D, as shown in Fig. 2.

Control equations:
=
Sset- S

Kp r
F = F0 + KP e + I edt
>

Program

When the system equations are solved dynamically, one of two distinct steady-
state solutions is obtained, i.e., the reactor passes through an initial transient but
then ends up under steady-state conditions either at the stable operating
condition, or at the washout condition, for which X=0. The initial
concentrations for the reactor will influence the final steady state obtained. A PI
controller has been added to the program, and it can be used to control a
substrate setpoint below Smax. The controller can be turned on setting by
Kp>0. The control constants Kp, and the time delay tp can be adjusted by the
use of sliders to obtain the best results. Appropriate values of control constants
might be found in the range 0.1 to 10 for Kp and 0.1 to 10 for TJ. Note that
the control does not pass Smax even though the setpoint may be above Smax.
Another feature of the controller is a time delay function to remove chatter.
The program comments on the CD-ROM should be consulted for full details.

Nomenclature

Symbols

D Dilution rate 1/h


370 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

F Flow rate m3/h


KI Inhibition constant kg/m3
KP Controller constant kg/m3
KS Saturation constant m 6 /kgh
S Substrate concentration kg/m3
V Volume m3
X Biomass concentration kg/m3
Y Yield coefficient kg/kg
H(U) Specific growth rate coefficient 1/h
T
I Controller time constant h
TF Time constant controller delay h

Indices

0 Refers to inlet
I Refers to initial value
m, max Refers to maximum

Exercises
8.7 Diffusion Systems 371

Results

Run 1: 25009 steps in 3.13 seconds


r 0.3

0.25

0.2
1.5

0.15

TIME

Figure 2. A control simulation of the process with the setpoint below Sn

References

Edwards, V. H, Ko, R. C. and Balogh, S. A. (1972) Dynamics and Control of


Continuous Microbial Propagators Subject to Substrate Inhibition, Biotechnol.
Bioeng. 14, 939-974.

Fraleigh, S. P., Bungay, H. R. and Clesceri, L. S. (1989) Continuous Culture,


Feedback Control and Auxostats. Trends in Biotechnology, 7, 159-164.

8.7 Diffusion Systems

8.7.1 Double Substrate Biofilm Reaction


(BIOFILM)

System

A biocatalyst is immobilized inside a solid matrix (gel or porous solid) through


which substrates diffuse and react. As shown in Fig. 1, for simulation purposes
8.7 Diffusion Systems 371

Results

Run 1: 25009 steps in 3.13 seconds


r0.3

0.25

0.2
1.5

0.15

TIME

Figure 2. A control simulation of the process with the setpoint below Sn

References

Edwards, V. H, Ko, R. C. and Balogh, S. A. (1972) Dynamics and Control of


Continuous Microbial Propagators Subject to Substrate Inhibition, Biotechnol.
Bioeng. 14, 939-974.

Fraleigh, S. P., Bungay, H. R. and Clesceri, L. S. (1989) Continuous Culture,


Feedback Control and Auxostats. Trends in Biotechnology, 7, 159-164.

8.7 Diffusion Systems

8.7.1 Double Substrate Biofilm Reaction


(BIOFILM)

System

A biocatalyst is immobilized inside a solid matrix (gel or porous solid) through


which substrates diffuse and react. As shown in Fig. 1, for simulation purposes

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
372 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

the matrix is divided into segments, and the diffusion flux, j, from segment to
segment, is expressed in terms of the concentration difference driving force.

Solid Biocatalyst Matrix with N Segments

Liquid

j n-1 jn

n-1 ^_ n ^ n+1
^

Figure 1. Finite-differencing of the concentration profiles within the immobilized biocatalyst


into segments 1 to N.

Model

A multicomponent reaction whose reactants and products diffuse to and from


the reaction site, for example into an immobilized enzyme or biofilm, can be
described by diffusion-reaction equations. The original problem in terms of
non-linear partial differential equations, is described by a large number of
time-dependent differential-difference equations by discretizing the length
variable.
A component mass balance is written for each segment and for each
component:

[Accumulation | (Diffusion^ _ (Diffusion^ ^Production rate^


^ rate J ^ rate in J ^ rate out J v by reaction }

dSn
' F = J n - l A - j n A +r S n A A Z
Using Pick's law,
S n -l ~ Sn
n-1 =
AZ
and dividing by A AZ,
8.7 Diffusion Systems 373

dSn (Sn.i -2S n + S n +i)


"dT = s AZ2 + rs

Thus N dynamic equations are obtained for each component at each position,
one for each element. The boundary conditions are for the above case dS/dZ =
0 at Z = L and S = So at Z = 0. The equations for the first and last elements
must be written accordingly.
The kinetics used here consider carbon-substrate inhibition and oxygen
limitation. Thus,
S O

At steady state, the overall reaction rate or consumption rate of substrate can be
calculated from the gradient at the outer surface. To find the resulting change
of bulk concentration, the liquid phase can be coupled with suitable mass
balances. For a well-mixed, continuous-flow, liquid the resulting balance
equation would be

dS0 F SQ-SI
So) - a DS ^Z

For oxygen transferred from the gas phase:

dO0
= K L a(Os-0 Q ) - a Do AZ

Program

As seen below, the program on the CD-ROM uses the array-vector form which
permits plotting the values at time=Stoptime versus the distance index. Also the
number of finite-difference elements N can be varied.

a/at (s[i. .N-i] >=DS* (s[i-u -


)/(Z*Z)+RS[i]

a/at (o[i. .N-I] )=DO* (o[i-u


2*0[i]+0[i+l])/(Z*Z)
374 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Nomenclature

Symbols

a Specific area perpendicular to the flux 1/m


A Area perpendicular to diffusion flux m2
C Concentration g/m3
D Diffusion coefficients m2/h
F Volumetric flow rate m3/h
j Diffusion flux g/ (m2 h)
K Saturation constants g/m3
KLa Oxygen transfer coefficient 1/h
O Dissolved oxygen concentration g/m3
0S Saturation concentration for oxygen g/m3
R,r Reaction rate g/ (m3 h)
S Substrate concentration of carbon source g/m3
V Volume of tank m3
vm Maximum reaction rate g/ (m3 h)
Yos Yield for oxygen uptake -
z Length of element m

Indices

0 Refers to bulk liquid


1 - 10 Refer to sections 1-10
I Refers to inhibition
O Refers to oxygen
S Refers to carbon source
n Refers to section n
Feed Refers to feed

Exercises
8.7 Diffusion Systems 375
376 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

Run 1: 5009 steps in 5.53 seconds


.1

Figure 2. Oxygen and substrate time profiles for a step change in KLA.

Run 1: 5009 steps in 5.47 seconds


1.6-1 1

1.4-

0 1 2 3 4 5 6 7 8 10

Figure 3. Oxygen and substrate distance profiles at the end of the run in Fig. 2.
8.7 Diffusion Systems 377

Run 1: 5009 steps in 5.85 seconds


5

Figure 4. Dynamic response of oxygen and substrate mid-points caused by a step change in
KLA (as Fig. 2) followed at 3 h by a step reduction in Sfeed.

8,7.2 Steady-State Split Boundary Solution


(ENZSPLIT)

System

A rectangular slab of porous solid supports an enzyme. For reaction, substrate


S must diffuse through the porous lattice to the reaction site, and, as shown in
Fig. 1, it does so from both sides of the slab. Owing to the decreasing
concentration gradient within the solid, the overall rate is generally lower than
that at the exterior surface. The magnitude of this gradient determines the
effectiveness of the catalyst.
378 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Biocatalyst Matrix

Substrate diffusion

diffusion

X = L -* X = 0 ^ X=L

Figure 1. Symmetrical concentration gradients for substrate and product.

Model

Under steady state conditions:

f Rate of diffusion of Rate at which reactant \


=
Uubstrate into the slab ^is consumed by reaction^

dX
A quasi-homogeneous form for the reaction term is assumed.
The boundary conditions are given by:

At X = L:
S = S0 , P = P0
At X = 0:

dX dX
8.7 Diffusion Systems 379

The external concentration is known, and the concentration profile throughout


the slab is symmetrical.
The reaction rate is expressed by the Michaelis-Menten equation with
product inhibition
= kES
KM(I+P/KI)+S
where k, KM and K\ are kinetic constants and E and P are the enzyme and
product concentrations. At steady state, the rate of diffusion of substrate into
the slab is balanced by the rate of diffusion of product out of the slab.
Assuming the simple stoichiometry S > P

dS dP
=
SdX -PdX
which on integration gives
DS
P = (So-S)

where P is assumed zero at the exterior surface.


Defining dimensionless variables

S< P
=^ ' ' =^ and X' =
gives
d^' L2R'
dX'2 DSS0 ~
where,
kES'
R' =
(K M /S 0 )(1 +(S 0 F/K I ))
and,
P = (1 - S')

with boundary conditions at


X' = 1 S' = 1

X' = 0 dS'/dX' = 0

The catalyst effectiveness may be determined from

DS Sp (dSVdX')x=l
=
^ L2R0
380 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

where RQ is the reaction rate determined at surface conditions,

kES0
~K M (i +P O /K I ) + SO

Program

The dimensionless model equations are used in the program on the CD-ROM.
Since only two boundary conditions are known, i.e., S at X = L and dSVdX' at
X' = 0, the problem is of a split-boundary type and therefore requires a trial
and error method of solution. Since the gradients are symmetrical, as shown in
Fig. 1, only one-half of the slab must be considered. Thus starting at the mid-
point of the slab at X1 = 0, where dSVdX' = 0, an initial value for S1 is assumed
(SGUESS). After integrating twice, the computed value of S is compared with
the known value of SQ at X' = 1. A revised guess for S' at Xf = 1 is then made.
This is repeated until convergence is achieved.

Nomenclature

Symbols

D Diffusion coefficient m2/h


E Enzyme concentration mol / m3
K Kinetic constant kmol / m3
k Reaction rate constant 1/h
L Distance from slab center to surface m
P Product concentration kmol / m3
R Reaction rate kmol /(m3 h)
S Substrate concentration kmol / m3
X Length variable m
Effectiveness factor

Indices

I Refers to inhibition
M Refers to Michaelis-Menten
8.7 Diffusion Systems 381

P Refers to product
S Refers to substrate
!
Refers to dimensionless variables
0 Refers to bulk concentration
GUESS Refers to assumed value

Exercises
382 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

Run 5:1000 steps in 0


seconds

0.2 0.3 0.5 0.6 0.7


X

Figure 2. Substrate profiles generated by manual slider iterations.

Run 5:1000 steps in 0


seconds

1.86

1.84

Q. 1.82

1.8

1.78

1.76

1.74

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9


X

Figure 3. Product profiles for the runs in Fig. 2.


8.7 Diffusion Systems 383

8.7.3 Dynamic Porous Diffusion and Reaction


(ENZDYN)

System

This example involves the same diffusion-reaction situation as the previous


example ENZSPLIT, except that here a dynamic solution is obtained by finite
differencing. The substrate concentration profile in the porous biocatalyst is
shown in Fig. 1.

Model

With complex kinetics a steady state split boundary problem of the type of
Example ENZSPLIT may not converge satisfactorily, and the problem may be
reformulated in the more natural dynamical form. Expressed in dynamic
terms, the model relations become,

3S
dt =

ap a2p +R

where at the center

dX~dX
384 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Outside Center

S2
S3
S4

Figure 1. Finite-differencing for ENZDYN.

Using finite differencing techniques (refer to Sec. 6.2.1), these relations may be
expressed in semi-dimensionless form for any given element n by

dS'n n+l - 2S'n + S'n.f


W
_= 2iL(*-
2
R'n
L I AX'2 ) Sl

f 1
'P
dP'n r
n+l OP
^r n _i_ P'
"* r n+ A R'n
ar AX' 2 ' + S
I

where
D S S 0 1-S
L 2 R n AX'

and
S'n = Sn/SI; Fn = Pn/Si andAX' = AX/L

Sj is the external substrate concentration and AX is the length of the finite


difference element. Boundary conditions are given by the external
concentrations Sj and PI and at the slab center by setting SN+I = SN and PN+I =
PN.
8.7 Diffusion Systems 385

Catalyst effectiveness may be determined according to two different


methods:

(a) the effectiveness factor based on the ratio of actual rate to maximum rate
(here for eight segments).
R R
Ri + R29 + Ra3 + RA4 + RS5 + R*
-J 6 + 77 + a8.
R
o
(b) an estimate of the slope of the substrate concentration at the solid surface

o I c)
_ D OQ 1~ O1

^"L^RO AX'
Where the rate at the bulk conditions is

kES 0
+ P 0 /K I ) + S0

The same constant values are used as in Example ENZSPLIT.

Program

The numerical results of example ENZSPLIT and should be essentially the


same as the steady state of ENZDYN. Both programs are on the CD-ROM.

Nomenclature

The nomenclature is the same as ENZSPLIT with additional symbols and


indices:

Symbols

AX Increment of length m
r|2 Effectiveness factor based on rates -
386 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

111 Effectiveness factor based on flux


K2 Same as KM

Indices

refers to segment n

Exercises

References

Blanch, H.W., Dunn, I.J. (1973) Modelling and Simulation in Biochemical


Engineering in Advances in Biochemical Engineering, Eds. T.K. Ghose, A.
Fiechter, N. Blakebrough, 3, Springer.

Goldman, R., Goldstein, L. and Katchalski, Ch.L (1971) in Biochemical Aspects


of Reactions on Solid Supports, Ed. G.P. Stark, Academic Press.
8.7 Diffusion Systems 387

Results

Run 1:1005 steps in 0.0833 seconds


1

0.9

* 0.8

B?' 0.7

0.6
81:1
,.. 82:1
.. 83:1

tf"
tfO.3
. 84:1
- 85:1
_ _S6:1
, 87:1
_S8:1
W0.2
</)
0.1

0
30 50 80
TIME

Figure 2. Substrate concentrations in porous enzyme catalyst during dynamic solution.

Run 1:1005 steps in 0.0833


seconds

P5:1
-- P6:1
-ST-J7U.
_ -P8:1
, P3:1
-P4:1

10 20 30 50 60 70 80 90 100
TIME

Figure 3. Product concentrations in porous enzyme catalyst.


388 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

8.7.4 Oxygen Diffusion in Animal Cells


(CELLDIFF)

System

This example treats a diffusion-reaction process in a spherical biocatalyst bead.


The original problem stems from a model of oxygen diffusion and reaction in
clumps of animal cells by Keller (1991), but the modelling method also applies
to bioflocs and biofilms, which are subject to potential oxygen limitation.
Sphere
Oxygen

Product N

I/I
Substrate AV

Rp
Rp

Figure 1. The finite differencing of the spherical bead geometry.

Diffusion and reaction takes place within a spherical bead of volume = 4/37cRp3
and area =47iRp2. It is of interest to find the penetration distance of oxygen for
given specific activities and bead diameters. As shown, the system is modelled
by dividing the bead into shell-like segments of equal thickness. The problem
is equivalent to dividing a rectangular solid into segments, except that here the
volumes and areas are a function of the radial position. Thus each shell has a
volume of 4/3 n (rn3 - rn_i3). The outside area of the nth shell segment is 4n rn2
and its inside area is 4n r n _i 2 .
8.7 Diffusion Systems 389

in

Figure 2. The diffusion fluxes entering and leaving the spherical shell with outside radius rn
and inside radius r n _j.

Model

Here the single limiting substrate S is taken to be oxygen.


The oxygen balance for any element of volume AV is given by

The diffusion fluxes are

Ar
S
n~ S n-l
jn-l =
Ar
Substitution gives

dSn 3D
dt "'

The balance for the central increment 1 (solid sphere not a shell) is

4 3 dS = J.. ,, 22 4
4 3
= -
-

Since ri = Ar, this becomes


390 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

dS 3D
l _ S /q o N , p
I f - ^ S 2- S l) +R sl
r

The reaction rate is expressed by a Monod-type equation

S
RSn = - (

where X is the biomass concentration (cell number/m3) in the bead, OUR is the
specific oxygen uptake rate (mol/cell s) and Sn is the oxygen concentration
(mol/m3) in shell n.

Program

As shown below, segments are programmed using the array-vector facility of


Madonna, numbered from the outside to the center. The effectiveness factor,
expressing the ratio of the reaction rate to its maximum, is calculated in the
program, part of which is shown below. The number of elements N is called
Array in the program, which is on the CD-ROM.

{Shells 2 to Array-1}
a/at (S [2. . (Array-1) ] )=3*D*( ((r[i]**2)*(S[i-l]-

/(deltar*( (r[i]**3)-(r[i+l]**3))

Nomenclature

Symbols

D Diffusion coefficient m2/s


KS Saturation constant in Monod equation mol/m3
OURmax Maximum specific oxygen uptake rate mol/cell s
r Radius at any position m
Rp Outside radius of bead m
8.7 Diffusion Systems 391

RS Reaction rate in the Monod equation mol/s m3


S Oxygen substrate concentration mol/m3
X Biomass concentration cells/m3
Ar (Deltar) Increment length, r/N m

Indices

1 Refers to segment 1
2 Refers to segment 2
n Refers to segment n
P Refers to particle
S Refers to substrate

Exercises
392 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

0 10 20 30 40 50 60 70 80 90 100

Figure 3. Profiles of oxygen concentrations versus time for each shell.

Figure 4. Doubling the bead radius causes oxygen deficiency inside the bead (lower curve) as
these radial profiles show.

Reference

Keller, J. (1991) PhD Dissertation No. 9373, ETH-Zurich.


8.7 Diffusion Systems 393

8.7.5 Immobilized Biofilm in a Nitrification


Column System (NITBEDFILM)

Nitrification is the sequential oxidation of NH4+ to NO2~ and NO3" which


proceeds according to the following reaction sequence:

NH4+ +1 O2 -> NO2- + H2O +2H+

NO2- + \ O2 -> NO3-


The overall reaction is thus

NH4+ + 202 -> N0 3 -+H 2 0

Both steps are influenced by dissolved oxygen and the corresponding substrate
concentration and are catalyzed by two different organism species. Since their
growth rates are very low, nitrification as a wastewater treatment process benefits
greatly from biomass retention.
In this example, a biofilm column reactor for nitrification is modelled as three
tanks-in-series with a recycle loop (Fig. 1). Oxygen is supplied only in an
external contactor and circulates to the reaction column in dissolved form.
This is similar to the example NITBED. However, in this case the reaction takes
place within an immobilized biofilm, similar to the single tank example
BIOFILM.
394 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

O3, Si3

Fluidized
bed

SiA

Figure 1. Biofilm column reactor with recycle loop for nitrification.

Model

The column reactor is assumed to be described by three tanks. The model


balance equations for the liquid phase are developed by considering both the
individual tank stages and the absorber. Component balances are required for
all components in each section of the reactor column and in the absorber,
where the feed and effluent streams are located. For the solid biofilm phase,
where the reaction takes place, the concentrations change both with distance
and time. Therefore, a descretization of the length variable is required as
developed for the example BIOFILM.
8.7 Diffusion Systems 395

To tank n+1

t
S2ntO]
S3ntO]
On[0]

From tank n-1


Figure 2. Schematic of a single tank in the column.

Because of the complexity involving four components in two phases and four
regions care must be taken with the nomenclature. The nitrogen compounds are
referred to as Si, 82, and 83, respectively. Dissolved oxygen is referred to as O.
Referring to the above figure, a single tank n is shown with the four
components. [0] refers to the liquid phase in contact with the solid. Transfer to
the biofilm is by diffusion to the first section, denoted [1].

Figure 3. Schematic of a single section i of biofilm in tank n.

Further diffusion brings substrate to all the biofilm sections, as shown above, for
a single substrate in section i. The reactions occur in these sections.
In the absorber, oxygen is transferred from the air to the liquid phase.
Additional subscripts, as seen in Fig. 1, identify the feed (F), recycle (R) and the
flows to and from the tanks 1, 2 and 3, each with volume V, and the absorption
tank with volume VA-
The fluidised bed reactor is modelled by considering the component balances
for the three nitrogen components (i) and also for dissolved oxygen. For each
stage n, the liquid phase component balance equations have the form

dSin[0] =
dt
396 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

-J0[0]

For the absorption tank, the balance for the nitrogen containing components
include the input and output of the additional feed and effluent streams, giving
ds
iA _(S/QiF -SQiA )\
-

The oxygen balance in the absorption tank must account for mass transfer from
the air, but neglects the low rates of oxygen supply and removal by the
convective streams. This gives

For the first and second biological nitrification rate steps, the reaction kinetics
for any stage n are given by
v
= ml Slni ni
Kl+ S lni K O i + O ni

V
rs2n = m2 S2ni
K
2+ S 2ni K
02+ni

The oxygen uptake rate is related to the above reaction rates by means of the
constant yield coefficients, YI and Y2, according to

i -r S 2niY2

The reaction stoichiometry provides the yield coefficient for the first step

YI = 3.5 mg 02/(mg N-NNH4)


and for the second step
Y2 = 1.1 mg O2/(mg N-NO2)
8.7 Diffusion Systems 397

Nomenclature

Symbols

A Specific area of film 1/m


F Feed and effluent flow rate m3/h
FR Recycle flow rate m3/h
KLa Transfer coefficient h
K Saturation constants g/m 3
KI Saturation constant for ammonia g/m 3
K2 Saturation constant for ammonia g/m 3
L Biofilm thickness m
N Number of biofilm segments -
0 Dissolved oxygen concentration g/m 3
Os and O* Oxygen solubility, saturation cone. g/m 3
OUR Oxygen uptake rate g/ m3 h
r Reaction rate per volume of biofilm g/ m3 h
S Substrate concentration g/m 3
V Volume of one reactor stage m3
VA Volume of absorber tank m3
vm Maximum velocity mg/L h
Y Yield coefficient mg/mg

Indices

1,2,3 Refer to ammonia, nitrite and nitrate, resp.


1,2,3 Refer to stage numbers
A Refers to absorption tank
F Refers to feed
jn[I] Refers to substrate j in stage n in segment i
m Refers to maximum
Ol and O2 Refer to oxygen in first and second reactions
S1,S2 Refer to substrates ammonia and nitrite
S and * Refer to saturation value for oxygen
398 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Exercises

Program

The program is on the CD-ROM.


8.7 Diffusion Systems 399

Results

Run 1: 133 steps in 13.7 seconds


5

'4.5
*C
4 3
A
I 60 ' .3.5 g
j,g 5 0 - 13
a?
3
40- 2.5

.2 S
1.5
520.
1
55 io- 5
o-
50 60 70 80 90 100
TIME

Figure 4. Time profiles of the nitrogen component bulk concentrations in the first tank and
the oxygen bulk concentrations in the three tanks.

Run 1: 133 steps in 13.7 seconds

4.5
4

3.5

0.5

0
0 10 20 30 40 50 60 70 80 90 100
TIME

Figure 5. Time profiles of the oxygen concentrations within the 10 segments of biofilm in
the first tank.
400 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

8.8 Multi-Organism Systems

8.8.1 Two Bacteria with Opposite Substrate


Preferences (COMMENSA)

System

Considered here (Fig. 1) is the batch growth of a two-organism culture on two


substrates, in which both species can utilize both substrates (Kim et al., 1988),
but where the organisms have opposing substrate preferences. The two
bacterial species involved are: Klebsiella oxytoca (XA) and Pseudomonas
aeruginosa (XB). The two substrates are glucose (Y), which is preferred by K.
oxytoca, and citrate (Z), which is preferred by P. aeruginosa.

XA XB

Figure 1. Organism XA prefers substrate Y, and organism Xg prefers substrate Z.

The assumptions are as follows:

- The overall individual growth rate of each species at any time is the sum of
the rate of growth on glucose plus the rate on citrate.
- The specific growth rate on each substrate depends on the concentration
level of some key enzyme responsible for the rate-controlling step E.
- The key enzyme for the preferred substrate is assumed to be constitutive.
- The production of the key enzyme for the secondary substrate is subject to
induction and repression by the preferred substrate.
- An inhibitor I is produced from the growth of K. oxytoca on glucose and
inhibits the growth of P. aeruginosa on citrate. The inhibitor is thus a
growth-associated product.

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
8.8 Multi-Organism Systems 401

- The total rate of substrate consumption is the sum of the rates of


consumption by each organism plus the rate of consumption of substrate for
the production of inhibitor.
- The oxygen uptake rate (OUR) and carbon dioxide evolution rate (CER)
involve the sum of the individual contributions from each organism.
- The dissolved oxygen tension in percentage air saturation (DOT) is obtained
using a steady state oxygen balance.

Model

The growth rates, jny, for each organism are the sums of the growth rates on
glucose and citrate. The subscripts i and j have the following meaning: i refers
to the organisms (K. oxytoca - A and P. aeruginosa = B) and j refers to the
substrate (glucose = Y and citrate = Z). The levels of the key enzymes are
denoted by E.

The biomass balances for the batch system are

dXA =
(MAY + MAZ) XA

dXB
I
The specific growth rate equations for the two organisms on each substrate are
given by:
A*maxAYSYEAY
K
SAY + S Y

MmaxAZSZEAZ
K
SAZ +S Z

A*maxBZSZEBZ I K
I
M-BZ - K--
S K"FT
SBZ+ Z V I +

The substrate balances are given by:


402 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

dS a
Y = !
Y Y Y
SAY I ) SBY

dSz
XB XA
dt - - YSBZ " YSAZ
Inhibitor (I) production is growth associated to organism A, and its decay is
proportional to the cell concentration. The balance for the inhibitor is

dl
gj- = oc|u AY X A -pX A

The balances for the key enzymes, which control the growth on secondary
substrates are given by:

Sz KRAZ
-
T-T - --
" *^--rt\iu
kpAZf\z^
EAZ

and
dE
BY SY KRBY
- " --Y KbY

where the consecutive terms in the above equations represent induction,


repression, and dilution due to cell division, respectively. Here the enzyme
levels E are normalized with respect to the maximum levels (See reference).
Because growth on the preferred substrates is constitutive, EAy and EBZ are
equal to 1 .
The oxygen uptake rate (OUR), carbon dioxide evolution rate (CER) and
dissolved oxygen tension (DOT) are given by:

OUR =
OAY OAZ

OBY OBZ

CER =
Y Y
CAY CAZ

CBY CBZ
8.8 Multi-Organism Systems 403

OUR
DOT = 100 1-
K L aC 0

The cell mass fractions are given by:

FA=-
X A + XB

FB = 1 - F A

Program

The program is given on the CD-ROM.

Nomenclature

Symbols

CER Carbon dioxide evolution rate kg/m3 h


DOT Dissolved oxygen tension
F Cell mass fractions
I Inhibitor concentration kg/m3
K Saturation and inhibitions constants kg/m3
KLaC0* Oxygen transfer rate kg/m3 h
OUR Oxygen uptake rate (normalized) kg/m3 h
S Substrate concentration kg/m3
X Biomass concentrations kg/m3
Y Yield coefficients kg/kg
E Level of key enzyme
m Specific maintenance rates kg/kg h
a Yield constant for inhibitor kg/kg
P Inhibitor consumption rate constant kg/kg h
Specific growth rate 1/h
404 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Indices

A Refers to K. oxytoca
B Refers to P. aeruginosa
C Refers to carbon dioxide
I Refers to inhibitor
M Refers to maximum
O Refers to oxygen
P Refers to dilution due to cell division
R Refers to repression
S Refers to substrate
Y Refers to glucose
Z Refers to citrate

Exercises
8.8 Multi-Organism Systems 405

Results

The graphical results in Fig. 2 show the dynamic changes in biomass fractions
FA and FB for two values of a: 0.007 kg/kg and 0.0007 kg/kg .

Run 2: 8200 steps in 0.167 seconds


1

0.9.

0.8

0.7

0.6
e
s-0.45
0.3

0.2

0.1

0
4 5
TIME

Figure 2. Dynamic changes in biomass fractions FA and FB for a = 0.007 and 0.0007.

Reference

Kim, S. U., Kim, D. C, Dhurjati, P. (1988). Mathematical Modeling for Mixed


Culture Growth of Two Bacterial Populations with Opposite Substrate
Preferences. Biotechnol. Bioeng., 31, 144-159.

This example was developed from the original paper by J. Lang, ETH-Zurich.
406 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

8.8.2 Competitive Assimilation and Commensalism


(COMPASM)

System

The interactions between two microbial species (Ma and Mb) in a mixed
continuous culture are considered (Miura et al., 1980). The population
dynamics of the two microbes, is described by competitive assimilation of
substrate Si and commensalism, with the participation of growth factor Ga that
is excreted by microbe Ma and required by microbe Mb for growth. Mb also
consumes a second substrate 82 from the medium. These interactions are
represented in the Fig. 1.

,G

Figure 1. Interaction of two organisms and two substrates in continuous culture.

Model

For the chemostat shown above the unsteady-state material balances are as
follows:

Dilution rate:
8.8 Multi-Organism Systems 407

Organism Ma:
dXa
j- = ftia-D)Xa

Organism Mb:
dXb
-gj- = (jib - D) Xb

Substrates S\ and 82:

ar = - "a
dS2 M-b Xb + D (S2
-ar = - ~YT" "S2)
The yields for organism Mb on the two substrates are assumed here, for
simplicity, to have the same values, Yb.

The growth factor balance is

dGaa "P 11 Y
Jib Xb T-X f*

The mass balance for the growth factor, Ga, is formulated by assuming a
formation rate, Pa |ia Xa, and consumption rate, (|Lib Xb)/Ybg. Here Xa and Xb
are the concentrations of microorganisms Ma and Mb, respectively. The
constants Pa and Ybg are the biological yield constants for the formation and
consumption of Ga, respectively. The specific growth rates of microbes Ma and
M b are expressed by:

Organism Ma:
Si
KSa+Si
Organism Mb:
Si Ga

82
K Sb 2 + S2 Kg + Ga
408 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

where K$a and K$bi are the saturation constants of Ma and Mb for substrate Si,
Ksb2 is the saturation constant of Mb for substrate 82, and Kg is the saturation
constant of Mb for growth factor Ga. Setting \imb2 = 0 corresponds to the
consumption of only one substrate Si.
A rigorous stability analysis of the system has been carried out by Miura et.
al. (1980). This involves linearizing the mass balances by Taylor's method in
the vicinity of the steady state solution and determining the characteristic
eigenvalues of the resultant matrix. The following relationship for co-existence
of the two microbes can be derived for the case of a single substrate.

^sm ma D(K Sbl -K Sa )


Sio > KSa D Oima - D) + Y F
a aCmmbKSa - m ma K Sbl)

Also, a critical dilution rate, where the maximum dilution rates of the two
organisms cross-over can be written:

.
Cnt _
"
Sa Sbl

Four particular cases depending on the values of the maximum specific growth
rate and saturation constants of both microbes can be simulated for the single
substrate case (|imb2 = 0).

1- M-ma > l^mbli Ksa > K$bi: Coexistence below a certain value of D
2. |ima > Mmbi; Ksa < KSbi: No coexistence range
3. |ima < |imbi; Ksa > Ksbi: Coexistence with wider range of stable focus
K$a > Coexistence at higher D and wider range of

Program

The program is given on the CD-ROM.


8.8 Multi-Organism Systems 409

Nomenclature

Symbols

D Dilution rate 1/h


F Feed rate m3/h
G Growth factor concentration g/m3
K Saturation constants g/m3
P Yield constant
S Substrate concentration g/m3
V Reactor volume m3
X Biomass concentration g/m3
Y Yield coefficient g/g
Specific growth rate 1/h

Indices

0 Refers to feed
1 Refers to substrate 1
2 Refers to substrate 2
a Refers to organism a
b Refers to organism b
bl Refers to organism b growing on substrate 1
b2 Refers to organism b growing on substrate 2
g Refers to growth factor
m Refers to maximum

Exercises
410 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

A simulation is given in Fig. 2 for the parameters as given in the program with
feed flow rate F = 0.24 m3/h, and the feed concentrations SIQ = 500 mg/L and
820 = 1500 mg/L. The oscillating concentrations are given for Xa and S\
versus time. It is seen that this solution is stable and homes into a steady state,
corresponding to case 1.

Run 1: 3000 steps in 0.0167 seconds


300

250

200

150 CD

100

300

Figure 2. Competition and commensalism of two organisms (F = 0.24 m3/h, S10 = 500 mg/L
and S2o = 1500 mg/L), showing the biomass concentrations.
8.8 Multi-Organism Systems 411

Reference

Miura, Y., Tanaka, H., Okazaki, M. (1980). Stability Analysis of Commensal


and Mutual Relations with Competitive Assimilation in Continuous Mixed
Culture. Biotechnol. Bioeng., 22, 929.

Example developed from the original paper by S. Ramaswami, ETH-Zurich.

8.8.3 Stability of Recombinant Microorganisms


(PLASMID)

System

In genetic engineering, microorganisms are used as host cells to produce


important biochemicals by inserting a small portion of extra-chromosomal
DNA (on plasmids) into the cell. These plasmids carry the genetic instructions
to produce the desired product and tend to lose their engineered properties
during cell division because of non-uniform plasmid distribution. The
engineered or recombinant strain usually grows more slowly than the wild-type,
nonplasmid-bearing strain, so that engineered strain may be lost through
extinction. By exploiting the difference in the adaptation times of wild and
engineered strains, a possibility exists of maintaining a plasmid-bearing
population in continuous culture by cycling the substrate feed concentration or
the dilution rate. This dynamic problem is adapted from Stephens and
Lyberatos (1987 and 1988), based on the concept of plasmid stability from
Aiba and Imanaka (1981).
The Monod model assumes a balanced growth in which all cellular
components change in the same proportion at all times but does not account
for dynamic effects. Dynamic first order lag relations are added to account for
the response of the organisms to rapid changes in the medium. It is assumed
that the time constants for the two strains are different and that the responses to
changing concentrations are therefore different. As a consequence, the strain
with the smallest time constant has the advantage when the concentration of the
limiting substrate is oscillating. The simulation model based on Fig. 1 is used
to predict the stability in the competition between wild (Xi) and engineered
(X2) strains in continuous culture.
412 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

S.XLX2

Figure 1. Competitive cultures x j and X2 in a continuous system.

Model

The following dimensionless parameters are defined: s = S/K, P = 0,1/0.2, t = treai


2, Rmd = torn/Ok, Dd = D/a2, xi = Xi/Y K, x2 = X2 A" K.
The mass balances in dimensionless form are:

= (ill (zi) xi - D xi + p [12 (Z2) X2

dx2
2 - D X2

ds
g^ = D (SQ - s) - (Lli (s) Xi - |I2 (s) X2

where the time delayed specific growth rates are,

"SIT

Z2)
8.8 Multi-Organism Systems 413

and the growth rates are

In the above, z\ and Z2 represents the time-delayed substrate concentrations,


and the specific growth rates of xi and X2 are taken as functions of z.

Thus:
dz
= P(S-ZI)

where p = ai/a2, and cxi and a2 are the adaptability factors or inverse time
constants. The effect of (3 is to delay the substrate for growth in the wild and
engineered organisms according to their first order time constants. For p > 1 the
wild type is delayed with a shorter time constant. At high values of oci and OC2,
the model describes an undelayed, instantaneous Monod growth model. It is
assumed, that (li^ > H2m .
The probability factor p represents the probability (or fraction) of
conversion to the wild strain during growth of the engineered strain. Thus the
growth rate of the engineered strain is multiplied by [1 - p].

Program

In the program on the CD-ROM, the square-wave input for SQ is generated by


the Madonna Conditional Operator, using the parameter MARK (ratio of the
time during which the function has the value 1 to the time of the complete
period) and PER (period).

{Square wave feeding generated by conditional


operator}
SO=IF(Time/PER-INT(Time/PER))<=MARK THEN SI ELSE 0
414 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Nomenclature

Symbols

D Dilution rate 1/h


K Saturation constant kg/m3
MARK Ratio controlling step function
PER Time period h
R Biomass ratio
S Substrate concentration kg/m3
s Substrate concentration, dimensionless
X Biomass concentration kg/m3
X Biomass concentration, dimensionless
Y Yield coefficient kg/kg
P Probability factor
r Growth rate kg/m3 h
t Dimensionless time
X Biomass concentration, dimensionless
z Delayed substrate concentration, dimensionless -
a Adaptability factors 1/h
P Ratio of adaptability factors -
H Specific growth rate 1/h

Indices

0 Refers to inlet stream


1 Refers to wild strain
2 Refers to engineered strain
d Refers to dimensionless
i Refers to 1 or 2
m Refers to maximum
real Refers to real time
8.8 Multi-Organism Systems 415

Exercises

The stability problem can be studied by the variation of several parameters.

Results

The output in Fig. 2 gives the substrate oscillations created by the square wave
feed concentrations, showing the engineered organism X2 being washed out. A
similar situation for sine wave feeding is given in Fig. 3.
416 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Run 1:2500 steps in 0.05 seconds

100 150 200 250


TIME

Figure 2. Square wave substrate feeding, causing X2 to wash out.

Run 1:40000 steps in 0.583 seconds

100 150 200


TIME

Figure 3. Sine wave feeding. Similar to Fig. 2 but with longer period and a higher b value.

References

Aiba, S., Imanaka, T. (1981) in Annals of the New York Acad. of Sciences, 369,
1-15.
8.8 Multi-Organism Systems 417

Stephens, M.L., Lyberatos, G. (1988) Biotechnol. and Bioeng., 31, 464-469.

Stephens, M.L., Lyberatos, G. (1987) Biotechnol. and Bioeng., 29, 672-678.

Example developed by N. Mol, ETH-Zurich.

8.8.4 Predator-Prey Population Dynamics


(MIXPOP)

System

The growth of a predator-prey mixed culture in a chemostat can be described


with a reaction kinetics formulation. In this growth process, the dissolved
substrate S is consumed by organism Xi (the mouse), while species X2 (the
monster) preys on organism Xi, as shown in Fig. 1.

Figure 1. Monster attacks mouse while it unsuspectingly feeds on S.

Model

The model involves the chemostat balances for each species with the
corresponding kinetics. The variables are given in Fig. 2.
418 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

S0 D Sl,X-|,X2
^*
1
II

Figure 2. Chemostat predator-prey reactor.

Substrate balance,
dS
= D(So-Si) -
Species 1 (prey) balance,

^2X2

Species 2 (predator) balance,

dX2
- DX 2

where D is the dilution rate,

~"v
The kinetics are given by Monod relations,

Program

The program is found on the CD-ROM.


8.8 Multi-Organism Systems 419

Nomenclature

Symbols

D Dilution rate 1/h


F Flow rate m3/h
K Saturation rate constant kg/m3
S Substrate concentration kg/m3
V Reactor volume m3
X Biomass concentration kg/m3
Y Yield constants kg/kg
Specific growth rate 1/h

Indices

0 Refers to feed stream


1 Refers to prey
2 Refers to predator
m Refers to maximum

Exercises
420 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

Stable steady states for the system are shown in Fig. 3 for |LLmi = 0.5 and
|Lim2 = 0.11. Oscillations in the biomass populations are achieved by setting the
specific growth rates nearly equal (\im\ = 0.5 and |Lim2 = 0.49) as shown in
Fig. 4 and also by the phase plane plot of Fig. 5.

Run 1: 5005 steps in 0.1 seconds


4

3.5

2.5

1.5

0.5
50 100 150 200 250 300 350 400 450 500
TIME

Figure 3. Stable steady state (|iml = 0.5 and [im2 = 0.11).


8.8 Multi-Organism Systems 421

Run 1:5005 steps in 0.1 seconds


6

450

Figure 4. Oscillatory state (|iml = 0.5 and p,m2 = 0.49).

Run 1:5005 steps in 0.117 seconds

Figure 5. Phase plane plot of oscillations.


422 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

8.8.5 Competition Between Organisms (TWOONE)

System

Consider organism A and organism B with their respective specific growth rates,
(HA and ILLB, which both grow independently on substrate S. Assume:

S/(KSA + S)

S/(KSB + S)

Depending on the values of |LIM and K$, these two functions may occur in two
different forms, as shown in Fig. 1.

B M
inter

inter

Figure 1. Comparison of growth rate curves for the competitive chemostat growth.

It is clear that the curves B and A will cross each other if (IMB < MMA and KSB
< KSA- In Fig. 1, the situation on the left indicates that B will grow fastest at
any value of S. For this case, in chemostat cultures with dilution rate DI, after
an initial start up period, a substrate concentration S i will be reached at which
(LIB = DI and for which |LIA < DI. Organism A will then be washed out, and
only organism B will remain in the reactor.
The situation in the right of Fig. 1 shows (IB crossing (IA- The point of
intersection can be found easily by simple algebra where:
8.8 Multi-Organism Systems 423

Wnter =

Solving for S at the intersection,

=
Sinter

For this case a chemostat can theoretically operate stably at D = Jiinter such
that both A and B will coexist in the reactor. This however is an unrealistic
metastable condition, and with D < Hunter* A will wash out. With D > Hunter A
will grow faster, causing B to be washed out.

Model

The equations for the operation of chemostat with this competitive situation are,

d*A ,

dXB
jp = 0 - D XB + JIB XB

dT = D(S0 - S) -

In addition, the Monod relations, |IA = f(S) and |LLB = f(S), are required.
Solution of these equations will simulate the approach to steady state of A and
B competing for a single substrate.

Program

The program is given on the CD-ROM.


424 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Nomenclature

Symbols

D Dilution rate 1/h


F Flow rate m3/h
KS Saturation constants kg/m3
S Substrate concentrations kg/m3
V Reactor volume m3
X Biomass concentrations kg/m3
Y Yield coefficient kg/kg
Specific growth rates 1/h

Indices

A Refers to organism A
B Refers to organism B
M Refers to maximum
0 Refers to inlet stream
inter Refers to the intersection of the ju versus S curves

Exercises
8.8 Multi-Organism Systems 425

Results

Run 1:1004 steps in 0.0333 seconds


5

4.5

3.5

3
!
2.5 '

1.5

0.5
9 10

Figure 2. Organism A and B competing for substrate.

8.8.6 Competition between Two Microorganisms for


an Inhibitory Substrate in a Biofilm
(FILMPOP)

System

Wastewater with toxic chemicals is often treated directly at the source with
specialized microbial cultures in small-scale biofilm reactors. A model may
help to understand, optimize, and control such reactors. In a paper by Soda et
al. a simple biofilm model was developed to simulate the competition between
two microorganisms for a common inhibitory substrate. In the model the
following assumptions were made: (i) the biofilm has a uniform thickness and
is composed of 5 segments, (ii) each microorganism A and B utilizes a
common substrate, and the growth rates are expressed by Haldane kinetics with
a spatial limitation term but is otherwise independent of the other
microorganism and (iii) the diffusion of the substrate, movement of the
microorganisms, and continuous loss of the biomass by shearing are expressed
by Pick's law-type equations.
8.8 Multi-Organism Systems 425

Results

Run 1:1004 steps in 0.0333 seconds


5

4.5

3.5

3
!
2.5 '

1.5

0.5
9 10

Figure 2. Organism A and B competing for substrate.

8.8.6 Competition between Two Microorganisms for


an Inhibitory Substrate in a Biofilm
(FILMPOP)

System

Wastewater with toxic chemicals is often treated directly at the source with
specialized microbial cultures in small-scale biofilm reactors. A model may
help to understand, optimize, and control such reactors. In a paper by Soda et
al. a simple biofilm model was developed to simulate the competition between
two microorganisms for a common inhibitory substrate. In the model the
following assumptions were made: (i) the biofilm has a uniform thickness and
is composed of 5 segments, (ii) each microorganism A and B utilizes a
common substrate, and the growth rates are expressed by Haldane kinetics with
a spatial limitation term but is otherwise independent of the other
microorganism and (iii) the diffusion of the substrate, movement of the
microorganisms, and continuous loss of the biomass by shearing are expressed
by Pick's law-type equations.

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
426 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Sf S0

Wall

Figure 1. Schematic of the continuous reactor with biofilm, showing the descretization into
five layers.

Model

Fig. 1 illustrates an idealized flat biofilm with a uniform thickness Lf (m). The
biofilm is divided into N segments for simulation purposes and each has a
thickness of AZ = Lf/N (m). Wastewater containing the substrate is fed to the
reactor at a constant feed rate and a concentration Sf (mg/L). The bulk liquid in
the reactor is mixed throughout the tank and the substrate diffuses into the
biofilm. The substrate is transported from the bulk liquid having a
concentration S[0] (mg/L) to the surface of the biofilm having a concentration
S[l] (mg/L). A diffusion layer of a thickness Lj (m) is used to represent the
external mass transport resistance.
Using the same approach as in the example BIOFILM, the mass balances in
the bulk liquid for the substrate and microorganisms A and B with a continuous
flow are simply described as following:
8.8 Multi-Organism Systems 427

dS[0] Wll _, S[0]-S[1] m [01X fC] mB[0]XB[0]


v f
dt ' " DZ YA YB

, XA[0]-XA[1] , }
- -DXA|0] - aDXA (mALO] - bA JXA[0]

/\/1

where S is substrate concentration (mg/L). XA and XB are biomass of


microorganisms A and B (mg/L), respectively. Each number in the brackets
refers to the bulk liquid or a segment illustrated in Fig. 1 . DX, b, Y, and |Li are
diffusion coefficient of microorganisms (m2/day), biomass decay coefficient
(day1), yield coefficient (-), and net specific growth rate (day1). Subscripts A
and B refer to microorganisms A and B. D, Ds, a, and t are dilution rate (day"1),
diffusion coefficient of substrate (m2/day), specific area perpendicular to the
flux (nr1), and time (day), respectively.

Reactions within the biofilm are described by diffusion reaction equations. The
mass balances of the surface segment are described as following:

= s[0]-sm
S S
dt LjAZ AZ2

dX A [l] XA[0]-XA[1] XA[1]-XA[2]


at ~ - - -- DXA -
LZXZ,

dX B [l] XB[0]-XB[1] n XB[1]-XB[2]


=: \j v"-p LJ YT3 ^
XB XB
dt L,AZ AZ2

Component mass balances are written for each segment (i = 2, .., N-l), where:
428 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

dS[i] 5>[i-l]-2S[i]+S[i+]L] MA[i]XA[i] /iB[i]XB[i]


DS
dt AZ2 YA YB

dXA[i] X A [i-l]-2X A [i] + XA[i


t~DXA 2

dXB[i] XB[i-l]-2XB[i] + XB[i


^t-DXB 2

The mass balances of the boundary segment on the support wall are described
by the following equations:

dS[N] = p S[N-1]-S[N]
S
dt AZ2

The "diffusion" coefficients of microorganisms, DXA and DXB represent


displacement by cell division and by shearing off at the film boundary
contacting the bulk liquid.

Growth Kinetics Of Microorganisms

The inhibitory influence of high substrate concentration was described by the


Haldane kinetics. The two types of microorganisms compete for substrate, but
in the biofilm they also have to compete for the limited space available.
Therefore, growth of the microorganisms was described by the Haldane kinetics
with a spatial limitation term which was originally proposed as cell inhibition
kinetics by Han and Levenspiel (1988).
8.8 Multi-Organism Systems 429

.. m_

K IA

[i]=
K IB

where KI? Ks, and (im are inhibition constant (mg/L), half saturation constant
(mg/L), and maximum specific growth rate (day" ). Xm (mg/L) is the maximum
capacity of total biomass of microorganisms A and B in a segment.
The formulation of the spatial limitation term used here is the most simple
one possible with non-restricted growth at zero biomass concentration and zero
growth at maximal biomass concentration Xm.
Applying the above model it was found (Soda et al., 1999) that the
qualitative behavior of the biofilm reactor is characterized by 5 regions,
depending on the operating conditions, the substrate concentration in the feed
and the dilution rate. In region I, both microorganisms are washed out of the
biofilm reactor. In region II, microorganism B is washed out, and in region III,
microorganism A is washed out of the biofilm. In region IV, both
microorganisms coexist with one another. In region V, both microorganisms
coexist with a sustained oscillatory behavior. Convergence to regions I-V
depends on the initial conditions. In regions II-V, washout of either or both
microorganisms is also observed when the initial conditions are too far away.

Nomenclature

Symbols

a specific area perpendicular to the flux,


related to bulk liquid volume m"
b biomass decay rate day"
D dilution rate day"
DS diffusion coefficient of substrate m /day
Dx diffusion coefficient of microorganisms m /day
K! inhibition constant mg/L
KS saturation constant mg/L
430 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

LI thickness of diffusion layer m


Lf thickness of biofilm m
N number of segments in biofilm
S[i substrate concentration in element i mg/L
Sf substrate concentration in feed mg/L
biomass mg/L
spatial capacity of total biomass
of microorganisms A and B in a segment mg/L
Y yield coefficient
AZ thickness of each segment m
"
maximum specific growth rate day

Indices

A refers to microorganism A
B refers to microorganism B

Numbers in brackets

0 refers to bulk liquid


1-5 refer to segments 1-5

Exercises
8.8 Multi-Organism Systems 431

References

Soda, S, Heinzle, E,, Fujita, M. (1999) "Modeling and simulation of


competition of two microorganisms for a single inhibitory substrate in a
biofilm reactor." Biotechnol. Bioeng., 66, 258-264.

Han, K. and Levenspiel, O. 1988. "Extended Monod kinetics for substrate,


product, and cell inhibition." Biotechnol. Bioeng. 32: 430-437.

Program

Shown below is a portion of the program. The full program is on the CD-ROM.

{BALANCES FOR BIOFILM IN 10 SEGMENTS}


d/dt (S[2. .nslabs-1] )=DS* (S[i-l] -
2*S[i]+S[l+l])/ (Z*Z)-UA[i] *XA[i] /YA-UB[i] *XB[i] /YB
d/dt (XA[2. .nslabs-1] ) =DSA* (XA[i-l]-
2*XA[i] +XA[i+l] )/(Z*Z)+(UA[i] -kdA) *XA[i]
d/dt ( X B [ 2 . . n s l a b s - 1 ] ) =DSB* ( X B [ i - l ] -
2*XB[i] ) / ( Z * Z ) + ( U B [ i ] -kdB) * X B [ i ]
432 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

Run 1:1017 steps in 0.967 seconds


sr^^"
^^' '
/ \ /- 0.035

-0.03
\,
2 40-
-0.025
m /
X
0.02
\ XAmid:1
(A
. . XBmid:1
I Smid:1 -0.015
x 20 i
\ -0.01
g.
-0.005
---. '--.
10 20 30 40 50 60 70 80 90 100
TIME

Figure 2. Results corresponding to Case 2: UmB=0.4, KSB=0.1, KIB=10.

Run 1:1058 steps in 1 seconds


0.05

0.045

0.04

0.035

0.03

0.025
(A
0.02

0.015

0.01

0.005

10 20 30 40 50 60 70 80 90 100

Figure 3. Results corresponding to Case 4: UmB=1.8, KSB=0.01, KIB=0.01.


8.8 Multi-Organism Systems 433

8.8.7 Model for Anaerobic Reactor Activity


Measurement (ANAEMEAS)

System

As already discussed in Chapter 3, anaerobic processes can be described by


multi-substrate, multi-organism kinetics. As shown in Table 1, organic acids
are formed from monomeric and polymeric substrates contained in wastewater.
These are then converted into hydrogen, CO2 and acetic acid. In a last step,
acetic acid and H2 with CO2 form methane.

Table 1. Stoichiometry of Anaerobic Reactions.

Step 1: Hydrolysis (example: carbohydrate-hexoses)


(C6Hi2O6)n + n H2O -> nC 6 Hi 2 O 6

Step 2: Acid production (example glucose)


C6Hi2O6 -> CH3(CH2)2COOH + 2 H2 + 2 CO2
C6Hi2O6 + 2 H2 -> 2 CH3CH2 COOH + 2 H2O
2H 2 O H 2CH3COOH + 4 H2 + 2 CO2

Step 3: Acetic acid production


CH3(CH2)2COOH + 2 H2O -> 2 CH3COOH + 2 H2
CH3CH2COOH + 2 H20 -> CH3COOH + 3 H2 + CO2

Step 4: Methane production


CH3COOH -> CH4 + CO2
4 H2 + C02 -> CH4 + 2 H20

In order to evaluate the activity of an anaerobic reactor and to evaluate the


correctness of the reactions in Table 1, an off-line measurement system has
been designed. This involves a small batch reactor coupled to a mass
spectrometer. A sample of biomass with medium is taken from the larger
continuous anaerobic reactor and put into the small batch reactor. Dissolved
gases are stripped by helium, all gas bubbles are removed and substrate is
added to start the reaction. The accumulation of organic acids, CO2, H2 and
CtLj. is measured. pH is adjusted according to total acid concentration and
buffer capacity. Biomass concentration is constant throughout the experiment.
434 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

The model was developed to aid the design of the measurement system and in
the interpretation of the data.

Model

A five organism model with lumped hydrolysis and acid-generating bacteria


was established. Substrate, intermediate and product balances of the batch
reactor are
dSi
"dT = IrSi

where rsi are the rates of consumption and synthesis of S[.


The respective reaction rates rj for the consumption of substrate Si and for
the formation of product Pj in each step are those from the reactions in Table 1
as follows:

and the specific growth rates take the Monod form,

M-imax Si
Mi = KSi + Si

or modified in the case of substrate inhibition for acetate,

M^maxi Si

The individual equations for each substrate Si are given in the program,
Thermodynamic equilibrium constraints on the Step 3 reactions (Table 1)
are also included in the model.
8.8 Multi-Organism Systems 435

Reaction Equilibrium

In the acetogenic step (Step 3 reaction in Table 1), acetic acid, hydrogen and
carbon dioxide are produced from propionic and butyric acid. The
thermodynamic equilibria for these reactions are incorporated by estimating
the chemical equilibrium limits for butyric acid:

CH3(CH2)2COO- + 2 H2O 2 CH3COO- + 2 H2 + H+ AGo = 48.3 kJ

From this the equilibrium constant is KBU = 2.02 x 10~3 (mol4 nr12) given by

KBU
"
For propionic acid similarly,

CH3CH2COO- + 2 H2O 2 CH3COO- + 3 H2 + CO2 AG^ = 76.1 kJ

and a equilibrium constant of Kpro = 1.35 x 10~12 (mol4 nr12).

4 CEAc2 CEH23 CEC02


K
Pro - 3 CEPro

The factor 4/3 is necessary because concentrations here are given in C-mol.
An empirical approach was chosen to slow the reactions down on
approaching the equilibrium, and they were not allowed to proceed to the right
side when the equilibrium condition was reached. Using the actual
concentrations, the parameters KBU* and Kpro* were estimated.

* C A c 2 C H2 2 CH +
KBU
~~ CBu

_ 4 CAc2 CH23 CCQ2


- 3C Pro

The ratio of these values to the true equilibrium constant, K*BU/KBU, and
K*pro/Kpro will be greater than unity if the equilibrium has not yet been
reached. Using these ratios with the empirical S-shaped curve of Fig. 1, the
factor FEQ was determined and was used to modify the growth rates. This
somewhat arbitrary function starts from FEQ = 0 at K*/K < 1 and rises to FEQ
= 1 at K*/K > 2. The factor FEQ causes the reaction to the right to stop when
436 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

the equilibrium is reached, but there is no reverse reaction when concentrations


of acetate, hydrogen, CO2 and H+ exceed the equilibrium values caused by
other reactions. The reaction proceeds irreversibly further away from the
equilibrium (K*/K > 2).
1.2
1.0-
0.8.

a 0.6-
uj
u_
0.4-

0.2.

0.

-0.2-

K*/K

Figure 1. Equilibrium factors (FEQ) to slow the growth rates near equilibrium.

The kinetics of biomass growth butyric acid, and propionic acid were modified
by these empirical equilibrium factors, FEQ, according to

i =FEQi

Ion Charge Balance to Estimate pH

As discussed in Sec. 1.3.7, in calculating the pH an ion charge balance can be


written to account for the acid-base dissociation buffer effects. The ion balance
represents an implicit non-linear equation in the dynamic model and must be
solved by iteration for each time interval, such that 8 = 0 in the equation

I + CH+

Thus CH+ is varied iteratively until 8 becomes essentially zero. This numerical
solution is not always trivial using conventional methods for non-linear
8.8 Multi-Organism Systems 437

algebraic equations (e.g., Newton-Raphson, and Regula falsi). Fortunately this


type of equation can be handled conveniently by the root finding feature of
BerkeleyMadonna, as shown in the program on the CD-ROM and below.
If base is added to control pH, an additional balance for cations of strong
bases (K+, Na+, ...) and anions of strong acids (Cl% SC>42~, ....) becomes
necessary as follows:
dCz
= F
titr Qtitr

Program

The program nomenclature is rather extensive and is defined within the


program. The Berkeley Madonna ROOTS feature is used to calculate the pH,
as shown below. The full program is on the CD.

(PH>
GUESS CHPLUS = le-4
ROOTS CHPLUS =
KW/CHPLUS+KdBu/(KdBu+CHPLUS)*BU/4+KdPr/(KdPr+CHPLUS)
*Pr/3 +KdAc/(KdAc+CHPLUS)*Ac/2 +KdC/(KdC+CHPLUS)*Cg
+KdBuf/(KdBuf+CHPLUS)*BUFFER-lonen-CHPLUS

LIMIT CHPLUS >= 0


LIMIT CHPLUS <= 1000
pH=-loglO(chplus)+3

Nomenclature

The nomenclature of the program is partially defined within the program.

Symbols

C Concentration C-mol/m3
Cons Consumption rate C-mol /m3h
F Stoichiometric coefficients (-)
FEQ Equilibrium factor (-)
Ftitr Titration flow rate m3/h
438 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

KD Dissociation constant mol /m3


Inhibition constant C-mol /m3
Ks Saturation constant C-mol/m3
S Substrate concentration C-mol /m3
P Product Concentration C-mol /m3
Pro Production rate C-mol /m3 h
X Biomass concentration C-mol /m3
YP/S Yield coefficient, product from substrate C-mol /C-mol
YX/S Yield coefficient, biomass from substrate C-mol /C-mol
1^ Specific rate of biomass synthesis 1/h

Indices

Ac Refers to acetic acid


Bu Refers to butanediol
Buf Refers to buffer
d Refers to death rate
Hy Refers to hydrogen gas
i Refers to reaction i
in Refers to initial
Mo Refers to whey substrate
Pr Refers to propionic acid
titr Refers to titration
Tot Refers to total
Z Refers to difference between cations and ions

Results

The first of the three graphs in Fig. 2 shows dynamic profiles of substrate whey
(Mo), CH4 (CH), dissolved CO2 (CO) and dissolved hydrogen (Hy). The whey
is almost instantaneously consumed. Hy reaches a maximum very soon and is
then quickly reduced to almost zero. CH4 is produced with varying rates. CC>2
reaches a maximum, which is partly caused by pH changes and by
consumption by hydrogen-consuming organisms. The peaks in the CC>2 curve
originate from numerical inaccuracies in the stiff system. In the second graph,
Fig. 3, the total concentration of volatile acids acetate (Ac), propionate (Pr) and
butyrate (Bu) are given. The thermodynamic inhibition of acetogenesis is
clearly seen in the early phase of the experiment. Ac reaches a maximum much
later than Pr and Bu, since it is produced from these two acids. In the third
graph, Fig. 4, the pH versus time profile is given, exhibiting an early decrease,
followed by almost constant pH during the rest of the simulation.
8.8 Multi-Organism Systems 439

Run 1:4389 steps in 1.43 seconds


0.06 -0.3

0.05 -0.25

0.04 -0.2

s' 0.03 0.15

0.02 -0.1

0.01 0.05

0.005 0.025 0.03

Figure 2. Dynamic profiles of substrate whey (Mo), CH4 (CH), dissolved CO2 (CO) and
dissolved hydrogen (Hy). Zoomed to show the early period.
Run 1:4389 steps in 1.43 seconds
0.3

0.25

r 0.2

0.15 -

-0.05

0.35

Figure 3. Total concentration of volatile acids acetate (Ac), propionate (Pr, lower curve) and
butyrate (Bu). The whey (Mo) peak is hardly visible at T = 0.
440 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Run 1:4389 steps in 1.47 seconds


6.18 -0.2

6.16 0.18

6.14 0.16

6.12 .0.14

6.1 .0.12

: 6.08 0.1
6.06 0.08

6.04 0.06
..'
6.02 0.04

6 0.02

5.98 0
0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

Figure 4. Variation of pH and acetate with time.

Exercise

References

Heinzle, E., Dunn, I.J. and Ryhiner, G. (1993) "Modelling and Control for
Anaerobic Wastewater Treatment." Adv. Biochem. Eng. 48, 79-114.

Yamada, N., Heinzle, E. and Dunn, I.J. (1991) "Kinetic Studies on


Methanogenic Cultures Using Mass Spectrometry." in: Biochemical
Engineering - Stuttgart (Eds. Reuss, M., Chmiel, H., Gilles).
8.8 Multi-Organism Systems 441

Ryhiner, G. Heinzle, E., Dunn, I.J. (1992) "Modelling and Simulation of


Anaerobic Waste water Treatment and Its Application to Control Design: Case
Whey," Biotechnol. Progr. 9, 332-343.

8.8.8 Oscillations in Continuous Yeast Culture


(YEASTOSC)

System

Oscillations in continuous cultures of baker's yeast have often been observed.


An example of measurements is shown in Chapter 3, whose oscillations were
modelled by the reaction scheme in Fig. 1.

Figure 1. Pathways of proposed model for yeast culture oscillations.

Model

The balance equations for continuous culture with dilution rate D are as
follows:
dR

dE
= - D E + [ rGE(R,G,E) + rSE(S) - rEX(E) ] R

- = D (SF - S) - [ rSE(S) + rSG(S,E) ] R


442 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

dG
= - D G + [ rSG(S,E) - rGE(R,G,E) ] R

dT
= - D T + [ rTi(R,G,E) - rT2(R,G,E) ] R

The species in parenthesis indicate the dependencies of the rates. The kinetic
expressions used in the balance equations are as follows:

r
GEm E
/KGX\n
l /

S
Ks + S

ME)
TEX = -

1 + ( KG/G + KET/E )n
rT2 =

Many of the parameters were determined independently from experiments,


some were taken from the literature, and some, especially those describing the
enzyme activity (T), had to be chosen during simulations. This model leads to
oscillations whose existence and dependency on operating conditions
qualitatively agree with experimental results. Also the directions in the phase
plane plot agree with the experiments.
8.8 Multi-Organism Systems 443

Nomenclature

Symbols

D Dilution rate 1/h


E Ethanol concentration kg/m3
G Storage material kg/m3
K Growth rate constants kg/m3
n Empirical exponent in rate model -
R Residual biomass without G g/m3
r Growth rates kg/m3 h
S Glucose concentration kg/m3
sig Rate constants. Example: sigGEm = TGEM various
T Enzyme concentration g/m3
X Biomass concentration kg/m3
Specific growth rate 1/h

Indices

E Refers to ethanol
G Refers to storage material
m Refers to maximum
S Refers to glucose
T Refers to enzyme
X Refers to biomass

Exercise
444 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

The influence of dilution rate is given below in plots from simulations: Fig. 2
with D = 0.05 and Fig. 3 with D = 0.1. In Fig. 4 the phase plane from the run
of Fig. 2 is shown.

Run 1:40012 steps in 3.3 seconds


35- -0.45

0.4
30-
0.35
25-
0.3

20- 0.25
(
0.2
15-

-0.15
10
0.1
5-

0-
M..JL.JL...
20 40 60
0.05

-0
100 120 140 160 180 200
TIME

Figure 2. Biomass and substrate oscillations for D = 0.05.

Run 1:40012 steps in 5.65 seconds


.2

1.8

-1.6

1.4

1.2

-1 </>

0.8

.0.6

.0.4

0.2

20 40 60 80 100 120 140 160 180 200


TIME

Figure 3. Biomass and substrate oscillations for D = 0.1.


8.8 Multi-Organism Systems 445

Run 1: 40012 steps in 4.88 seconds


0.09-|

0.08-

0.07-
0.06-

0.05-
i
0.04-

0.03-

0.02-

0.01 -

0-
12 16 18 20 22 24 26 28 30
X

Figure 4. Phase plane giving S versus X from the run of Fig. 3. Zoomed in for detail.

Reference

Heinzle, E., Dunn, I.J., Furukawa, K. and Tanner, R.D. (1983). Modelling of
sustained oscillations observed in continuous culture of Saccharomyces
cerevisiae. in Modelling and Control of Biotechnical Processes (ed. A.Halme),
Pergamon Press, London, p.57.

8.8.9 Mammalian Cell Cycle Control


(Mammcellcycle)

System

Modeling of mammalian cell cycle control is of great importance for


understanding development and tumor biology. Hatzimanikatis et al. (1999)
presented a model in the literature using simplified molecular mechanisms as
depicted in Fig. 1.
446 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

cycE+cdk2 CycE:cdk2-P+1 f> cycE:cdk2-P:1

Rb+E2F Rb-P+E2F

Figure 1. Schematic representation of the molecular mechanism of components and


interactions believed to be most important in controlling the Gl-S transition. cycE- cyclin E;
cdk2 - cyclin dependent kinase 2; Rb - pRb, a pocket protein; E2F - a transcription factor; P -
phosphate.

Model

For this reaction scheme the dynamic mass balances become

= V2 -Vl
dt ~

dK

dKr
dt
8.8 Multi-Organism Systems 447

dRE
dt

^=v 6 > r -v 6 , f

The symbols are defined as follows:


V are reaction rates.
C is the cyclin E concentration.
K is the cdk2 concentration.
KP is the phosphorylated cyclin E-cdk2 complex concentration.
K P I is the concentration of cyclin E-cdk2 phosphorylated complex bound to
inhibitor.
R is the concentration of the hypophosphorylated form of pRb.
RP is the concentration of the hyper-phosphorylated form of pRb.
RE is the concentration of the hypo-phosphorylated form of pRb that binds
to E2F.
E is the E2F concentration.
I is the concentration of the cyclin E-cdk2 complex inhibitor.
The subscipts "f' and "r" denote the forward and the reverse step,
respectively, of the reversible reactions.

The assumption of near equilibrium operation of reversible reactions (V5 and


V6) and of invariant total amounts of cdk2, pRb, E2F and inhibitor gave the
following dimensionless equations, consisting of 3 differential and 6 algebraic
equations.
=v
dT~ s vd y\

dk

drp
-= Va3 VA
4
di
k + k p + k p j =1

r + rp + r^ = 1
448 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

=1

i + A,kPI=l

re

0I= JPJL
kpi

Where c is the dimensionless concentration of cyclin, k is the dimensionless


concentration of cdk2 and rp is dimensionless concentration of the hyperphos-
phorylated form of pRb. All details about the kinetic equations and transposing
them into dimensionless form are given in the paper of Hatzimanikatis et al.
(1999).

Nomenclature

Dimensionless symbols as used in the program are listed here.

c Cyclin E concentration
e E2F concentration
i Concentration of cyclin E-c
k ckd2 concentration
kP Phosphorylated cyclin E-cdk2 complex
concentration
kP,I Concentration of phosphorylated cyclin
E-cdk2 complex bound to inhibitor
r Concentration of hypophosphorylated form
of pRb
rE Concentration of hypophosphorylated form
of pRb that binds to E2F
rP Concentration of hyperphosphorylated form
of pRb
g Ratio of total concentrations of cdk2 and cyclin E
s Ratio of total concentrations of pRb and E2F
1 Ratio of total concentrations of cdk2 and inhibitor
t Dimensionless time
8.8 Multi-Organism Systems 449

Exercises

Program

The program is given on the CD-ROM.

Results

Run 1: 50233 steps in 3.08 seconds


0.5

Figure 2. Profiles of concentrations k, rp and c versus time as obtained from the rate constants
in the program.
450 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Run 1: 50233 steps in 3.03 seconds


1
0.9
0.8
0.7
0.6
"o.5
0.4
0.3
0.2
0.1

0.32 0.33 0.34^ 0.35 0.36 0.37 0.38 0.39 0.4 0.41 0.42

Figure 3. Phase plane plot of k and rp versus c.

Reference

V. Hatzimanikatis, K. H. Lee, and J. E. Bailey. (1999) "A mathematical


description of regulation of the Gl-S transition of the mammalian cell cycle".
Biotechnol. Bioeng., 65, 631-637.
8.9 Membrane and Cell Retention Reactors

8.9.1 Cell Retention Membrane Reactor (MEMINH)

System

Consider a reactor whose outlet stream passes through a membrane that retains
only the biomass as seen in Fig. 1.
The growth is assumed to follow substrate inhibition kinetics with constant
yields. The oxygen transfer rate influences the growth at high cell density
according to a Monod function for oxygen.

F,S 0

i Gas

Membrane modi \"


module
F

illiiil

Air

Figure 1. Biocatalyst retention on a continuous reactor.

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
452 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Model

The reactor-membrane system is modelled as a well-mixed tank, except that


biomass is retained in the system batchwise. The balance region is chosen to
include both the reactor and the membrane separator, but the separator volume
is neglected.

Biomass balance:
dX
dT = rx
Substrate balance:
dS F
-r s

Oxygen balance (neglecting the oxygen transported by flow):

dCL
-gj- =K L a(CLS-C L )-ro
Kinetics:
S __ CL
rx
~ ^m x

r
s = rx YS/X + MS X
r = r
o x YQ/X + MO X

where the maintenance coefficients are related by

MS _ YS/X
=
M0

Program

The program is on the CD-ROM.


8.9 Membrane and Cell Retention Reactors 453

Nomenclature

Symbols

CL Dissolved oxygen concentration g/m3


CLS Saturation oxygen concentration g/m3
F Flow rate m3/h
KI Inhibition constant g/m3
KLa Transfer coefficient 1/h
Ko Saturation constant for oxygen g/m3
KS Saturation constant kg/m3
M Maintenance coefficients kg/(kgh), g/(kgh)
r Reaction rate kg/(m3 h)
S Substrate concentration kg/m3
V Reactor volume m3
X Biomass concentration kg/m3
Y Yield coefficient kg/kg
ILL Specific growth rate 1/h

Indices

0 Refers to feed
1 Refers to reaction 1
m Refers to maximum
O Refers to oxygen
S Refers to substrate
X Refers to biomass

Exercises
454 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

Oxygen transfer has a pronounced influence on performance as seen in Fig. 2


for variations of K^a. The dissolved oxygen may reach values below KQ as
shown in Fig. 3 for K^a values from 0.5 to 5.

Run 4: 2004 steps in 0.05 seconds

20

Figure 2. Influence of oxygen transfer coefficient (KLa = 0.1 to 1.0).


8.9 Membrane and Cell Retention Reactors 455

Run 4: 2004 steps in 0.05 seconds

\\.
0 2 4 6 10 12 14 16 18 20
TIME

Figure 3. Profiles of dissolved oxygen influenced by KLB (0.5 to 5, curves bottom to top).

8.9.2 Fermentation with Pervaporation (SUBTILIS)

System

The metabolic pathways for the production of acetoin and butanediol are well
known, as shown in Fig. 1. A kinetic model for a Bacillus subtilis strain has
been established from continuous culture experiments using an approach
involving overall stoichiometric relationships and energetic considerations. The
influence on the culture of product removal by pervaporation was investigated
by simulation methods.

Knowledge of these pathways allowed the following overall equations to be


written:
Respiration (reaction RI):
C6Hi2O6 > 6CO 2 + 6H 2 O

Formation of biomass (reaction R2):


1.2 NH3 > 6 CHi.8O0.5No.2 + 0.3 O2 + 2.4 H2O
456 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Formation of acetoin (reaction


C 6 Hi 2 0 6 +10 2 > C4H802 + 2 C02 + 2 H2O

Conversion of acetoin to butanediol (reaction R4):


C4H8O2 + H2O -> C4HioO2 + 0.5 O2

Complex Compounds

Biomass Carbon Dioxide


(Total Oxidation)

Substrate ^ Pyruvate ^ Acetoin + Carbon


Sugars v i Dioxide

Lactate Butanediol

Figure 1. Reaction scheme for the acetoin - butanediol formation.

The continuous reactor was coupled to a pervaporation membrane module and


was influenced by the membrane performance, owing to the removal of
products and the retention of biomass. Since only the volatile products and
water can pass through the membrane, a purge stream was needed to remove
biomass and salts. The recycle between the reactor and membrane module,
shown in Fig. 2, was high enough to provide complete mixing.

Model

Growth

The sugar (S) and dissolved oxygen effects were described in terms of a double
Monod function and the product inhibition by a simple inhibition kinetic term.
Diauxic effects were observed, but the diauxic components were unknown, and
for simplification only one diauxic switchover was assumed. The preferred
8.9 Membrane and Cell Retention Reactors 457

component is referred to as Ci, and the second component is called C2- An


empirical kinetic relation was devised such that the utilization of C2 was
inhibited by the presence of Ci in amounts greater than a repression constant
kRpi. The formulation then involves considering the growth as the sum of two
terms,

where,

and
C2
*pl
Recycle Membrane
Feed
Pervaporation Module

Condenser
Bioreactor
Permeate
Fpe
ACpe , Bu

Pump

Figure 2. Bioreactor and membrane pervaporator, showing process variables.

Each product was assumed to inhibit separately,

1 1
and
Bu_cc Bu

The inhibition constants, kinh,Ac and kinh,BU> an<l the exponents, and
were estimated from shake flask experiments.
The complete expression for specific growth rate was thus,
DO i i
1+ [_^_]Ac 1+ | Bu_iaBu
458 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Formation of Acetoin and Butanediol

The products were assumed to be formed when the respiration was not
sufficient to cover the energy requirement for growth. While the distribution of
the two products was assumed to be dependent on oxygen, through a rapid
equilibrium, according to an S-shaped empirical function described by

Ac fAcBu,max (DO)2
Bu -

To avoid an algebraic loop, the kinetics for the rate of butanediol production
was assumed to be dependent on the deviation from equilibrium,

TBU = qX4 X = kAcBu (Ac - Bu fAcBu)

The constant kAcBu was set high enough to ensure equilibrium conditions.
The specific reaction rate q^3 for product formation was obtained as,

~ YR2/R1 qX,l - YR2/R4

Reaction Rates

The specific reaction rates for each component were obtained from the specific
reaction rates q^ and the stoichiometric coefficients, Vy (component j in
reaction i) as follows,

i
The volumetric rates rj [mol L"1 Ir1] for components X, S, Ac, Bu, 62 and CC>2
were related to the specific rates as,
rj = qj X

The specific rate q^2 of reaction R2 was obtained from \\ as


8.9 Membrane and Cell Retention Reactors 459

V 2 , X MG X

where V2,x is the stoichiometric coefficient, and MGx is the C-mol mass of
biomass.
The rate for the respiration reaction, q^i, was found to be influenced by the
dissolved oxygen, and was described as follows:

qxi =

The yield coefficients for the complex components GI and C2 , were assumed
to be 1 g of each component for 1 g biomass, and the molecular weights of the
components were assumed to be the same as for the biomass. The initial
amounts of these components in the molasses medium were adjusted in the
simulation. The corresponding rates were proportioned according to the
growth rates as

rri = - rv and rr2 =

Pervaporation Model

The mass transfer in the pervaporation module was described as an equilibrium


process using constant enrichment factors. Thus the concentrations of product
in the reactor, Ac and Bu, were related to the concentration in the permeate,
Acpe and Bupe as,
Acpe = PAC Ac and Bupe = PRU Bu

Dynamic Reactor Mass Balances

Considering volume-specific flow rates,

_ FO _ Fpu _
D = y- , Dpu = -^r > Dpe =

where D, Dpu, Dpe were the flows for the feed, purge, and permeate, respectively
intr 1 .
460 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

The mass balances were as follows:

Total mass
D = Dpu + Dpe
Biomass
dx
"3T = rx - X Dpu
Sugar
dS
df = S0 D - S Dpu - rs
Acetoin
dAc
~dT = rAc ~ Ac D
pu ~ Acpe Dpe
Butanediol
dBu
~dt~ = rBu
~ Bu D U
P " BuPe DPe
Complex components
dCi
~dT = Ci>0 D - Ci Dpu - rCi
where Q = 1 and 2.

Oxygen transfer from the gas phase and the oxygen uptake rate determine the
DO in percent saturation as,

dDO 100
-HT~
m
= KLa (100 - DO) ^r OUR
2.34xlO"4
Here the equation is in terms of percent saturation using the DO saturation
value at 30 C of 2.34 x 10''4 mol/ L.

Nomenclature

Symbols

ACPE Acetoin concentration in permeate g/L


Ac Acetoin g/L, mol/L
ATP Adenosintriphosphate
Bu Butanediol g/L, mol/L
Ci Complex components, where i = 1 or 2 g/L
CPR Carbon dioxide production rate mol/L h
D Dilution rate 1/h
DO Dissolved oxygen concentration % saturation
8.9 Membrane and Cell Retention Reactors 461

UcBu Acetoin/butanediol ratio


lAcBu.max Max. Ac/Bu ratio
F Flow rate L/h
FM/P Permeate / purge flow rate
kAcBu Kinetic const. Ac/Bu mol/g(acetoin) h
kci Monod-const., growth on component i g/L
Monod-const., growth depending on DO % saturation
Inhibition constant product i g/L
KLa Gas -liquid transfer coefficient for O2 1/h
Monod-const. for respiration depending on
DO % saturation
Repression const., complex substrate 1 on
component 2 g/L
KS Monod-const., Growth on substrate g/L
Kprod Empirical const, for Ac/Bu-equilibrium
MGX Mol mass for biomass (1 C-mol) = 24.6 g/mol
OUR Oxygen uptake rate mol/L h
P Product, sum Acetoin + Butanediol g/L, mol/L
Specific rate of component j mol/g(biomass) h
Specific rate reaction i, where i =lto 4 mol/g(biomass) h
r
J Formation or uptake rates, components j
(Ac, Bu, O2, CO2, X and S) mol/L h
rpm Stirrer speed 1/min
Ri Chemical reaction i, where i = Ito 4
RQ Respiration quotient qcO2/(lO2
S Substrate concentration (e.g., sugar) g/L, mol/L
vvm Gas rate per volume liquid 1/min
V Reaction volume L
X Biomass concentration g/L, mol/L
Yi/j Yield coefficient (i formed/j used)

Greek symbols

ai Exponent inhibition kinetic product i (Ac or Bu)


pi Enrichment factor of pervaporation membrane for
component i (Ac or Bu)
\i Specific growth rate 1/h
R,max Maximum growth rate for component i 1/h
Vi j Stoichiometric coeff. of component j in reaction i
462 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Indices

Ac Acetoin
Bu Butanediol
Q Component i (unknown components 1 or 2)
CO2 Carbon dioxide
i Refers to reactions, i (1 to 4)
j Refers to components j (Ac, Bu, C>2, CC>2, X and S)
inh Refers to inhibition
02 Oxygen uptake
pe Permeate
pu Purge
P Product
Ri Chemical reaction i (1 to 4)
S Sugar
X Biomass
0 Feed or initial concentration
Ai Refers to reactions 1 to 4

Reference

Dettwiler, B. Dunn I. J., Heinzle E., and Prenosil J. E. "A Simulation Model
for the Continuous Production of Acetoin and Butanediol Using B. subtilis with
Integrated Product Separation by Pervaporation" Biotechnol Bioeng. 41, 791
(1993).

Program

The program is found on the CD-ROM.

Results

The results of Fig. 3 show the influence of the FM/P ratio, which corresponds
to the membrane area per unit reactor volume, on the biomass concentrations.
8.9 Membrane and Cell Retention Reactors 463

Here the enrichment factor was kept constant (pAc = 2.0), corresponding to the
values found for one of the membranes. In a second set of runs the enrichment
factor was varied for constant FM/P ratio = 1.0.

Run 14: 2013 steps in 0.317 seconds


110

100

60

50
6 8 10 12 14 16 18 20

Figure 3. Influence of permeate/purge flow ratio on the biomass concentration. (D=l,


BETAAC=2, FMP=0.4, 1.0, 2.0).

Run 19: 2013 steps in 0.333 seconds


0.25

0.15

..p.-'*"" j- *~ "*"
0.05

Figure 4. Influence of the enrichment factor on the acetoin and butanediol in the permeate.
(D=l, FMP=1 BETAAC=1, 3, 5).
464 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

8.9.3 Two-Stage Fermentor With Cell Recycle For


Continuous Production Of Lactic Acid
(LACMEMRECYC)

System

This example is based on a paper (1) in which a two stage fermentor-membrane


system for the continuous production of lactic acid was modelled. Membrane
retention of the active biomass can be expected to increase the productivity, but
the biomass concentration must be controlled in each reactor with a bleed
stream. The kinetics for this process, in which glucose is converted to lactic acid
by the bacteria Streptrococcus faecalls is described in the literature (2). Since
the rates are inhibited by product, it can be expected that a multistage system
will be advantageous.

(1-B2XB1F1+I?)

X 2 ,S 2 ,P 2

Figure 1. Two-stage membrane fermenter system for lactic acid production


8.9 Membrane and Cell Retention Reactors 465

Model

The model assumes completely-mixed stages and complete cell separation. The
operating parameters are feed flow rates or dilution rates and the bleed stream
fractions. The bleed stream from the first fermenter is led to the second
fermenter. The model equations are developed neglecting the volume of the
lines and separators. Note that the retentate streams are returned to their
respective reactors and can be considered as part of the well-mixed system.

The kinetics is given by a product inhibition

ii ii S KLP

and a death rate of the bacteria caused by product is also included.

The specific glucose uptake rate is given by


1
Ys
=+
qs M-
The lactic acid production rate is given by

Mass Balances for the First Stage


Defining the dilution rate as DI=FI/VI

For the total cells


l l
~dT~~
For the living cells
466 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

For the glucose

For the lactic acid


l fl l l l l
dt

Mass Balances for the Second Stage


The dilution rate for the second stage depends on the bleed ratio from the first
stage and the ratio of feed rates.

For the total cell mass

= BjDjXn - (8! + f )D!B2X2 +

For the active cells

^ = IBAXJ - I(B! + 00^2X2 + (ji2 - k d2 )x


For the glucose

fD!Sf2 + BAS1 - ( 8 ! +f)D 1 S 2 -q 2 X 2

For the lactic acid product

1 1 . >
2+V2X2
Clt UC (JC (JC

Productivity of Lactic Acid


It is assumed that all streams containing product can be recovered from all of
the streams. The productivities are calculated as follows:

From the first stage


fti=D1(P1-Pfl)
8.9 Membrane and Cell Retention Reactors 467

From the second stage

Pr 2 =-D 1 [(B 1 +f)P 2 -B 1 P 1 -fP f2 ]


VAi

The total productivity

Pr = -i-D 1 [(l-B 1 )P 1 +(B 1 +f)P 2 -P f ,-fP f2 ]


JL ~t~ UC

Substrate Conversion

From stage 1

From stage 2
(B 1+ f)S 2
Xco =1 --
82
B^+fSfz
Total glucose conversion

_ (l-B 1 )S 1 +(B 1 +f)S 2


Xg I --
S
f 1 + fSf 2

Program

In the simulation product-free feed streams are assumed with flow rates between
10-30 kg m 3 . The program is on the CD-ROM.

Nomenclature

a Death constant m 3 kg"1


B Bleed ratio
D Dilution rate s'1
f Flowrate ratio, F2/F1
F Volumetric flowrate m3 s'1
kd Specific death rate s"1
Basis specific death rate s"1
468 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

KP Product inhibition constant kg nr3


KS Saturation constant kg nr3
P Lactic acid concentration kg nr3
Pr Productivity kg nrV1
qs Specific substrate uptake kg kg-V1
s Glucose concentration kg nr3
t Time s
V Reactor volume m3
*s Substrate conversion -
X Concentration of living cells kgnr 3
xt Total cell concentration kgm~ 3
a Reactor volume ratio -
YP,YS Kinetic constants kg kg-1
8P,8s Kinetic constants kg kg-1
H Specific growth rate s-1
V Spec, product formation rate kg kg'1 s-1

Exercises
8.9 Membrane and Cell Retention Reactors 469

References

A. Nishiwaki and I. J. Dunn, "Performance of a two stage fermentor with cell


recyle for continous production of lactic acid", Bioprocess Engineering 21;
299-305, 1999.

H. Ohara, K. Hiyama, T. Yoshida, "Kinetics of growth and lactic acid


production in continuous and batch culture" Appl.Microbiol.Biotechnol. 38,
403-407, 1992.

Results

Run 1: 405 steps in 0.0167 seconds


-30

^Xt1:1
...Xt2:1 25
20-
... S2:1
20
P2:1
rsj 15-
X

gioH
10 CO

v
1
10 15 20 30 35 40
TIME

Figure 2. Results showing the steady state for biomass substrate and product in both stages.
470 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Run 4: 8320 steps in 1.53 seconds


160 -1

140 i 0.9

| 120-I
0.8
100
B
\ -0.7 rf
X 80
-
0.6 tfT
60 X
1
-0.5
flf 40
A.. 0.4

0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
B1

Figure 3. Results showing that the productivity changes slightly with bleed ratio but that the
biomass concentration and the substrate conversion depend highly on bleed ratio.

8.9.4 Tubular Hollow Fiber Enzyme Reactor


Module for Lactose Hydrolysis (LACREACT)

System

This tubular reactor- radial diffusion model assumes a series of nine well-mixed
tanks to describe a single hollow fiber module. Flow of lactose substrate passes
axially through the inner region of the fiber lumen. By diffusion the substrate
is transported radially outward from the lumen through the membrane and into
the cylindrical porous support surrounding the membrane. The reaction takes
place in this support region, where the immobilized enzyme is located. The
products of hydrolysis are glucose and galactose, which diffuse back toward the
liquid phase in the lumen. The parameter N can be used to adjust the number
of shells required. The module is assumed to consist of a large number of
identical fibers*
8.9 Membrane and Cell Retention Reactors 471

Model

The model is developed by finite-differencing both the axial and radial


directions. Thus there are axial stages in series, here nine plus the recycle tank,
and there are multiple cylindrical sections of porous support in the radial
direction. As depicted in the figures below, there is a convective liquid flow
from each stage to another. Diffusional flows carry substrate and product
radially from one cylindrical section of the porous support to another. Figures
1, 2 and 3 give the geometrical details.

FRfiber LAfiber

LAtank

FRfiber

Figure 1. Hollow-fiber module showing only a single fiber as modelled by nine stages.
472 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Figure 2. Details of a single well-mixed axial section.

Figure 3. Details of a finite-differencing of the porous support in the radial direction.

Fig. 3 shows a cross-section of the hollow fiber membrane showing the inner
hollow fiber region (white) and the outer porous support (shaded). The finite-
difference shell of volume V2 (white) is shown with diffusion fluxes of lactate
JLAI entering and leaving JLAI- It is important to account for the radial
variation of volumes and diffusional areas. Note that the segments are
numbered from outside to inside, 1 to N.

For each tank the component balances account for the accumulation, the flow
in and out and the diffusion in or out from segment N of the porous
membrane. For lactose in tank 1
8.9 Membrane and Cell Retention Reactors 473

LA FR
lumenl _ fiber */ T A T A x , JlAltN]* Aj[N]
H
ai
- v
v
v^^tank ' L/Mumenl ) + 77
v
lumen lumen

For the external recycle tank, the flow leaving all of the fibers enters it and also
the feed stream enters it. The total flow rate leaving must equal the sum of these
rates. Thus for lactate

Number * FRfiher T A X Ffeed / A


T T A N
. LAtank ) + -Jeed_(LAfeed - LA tank )
V
tank tank

Taking the component balances for each segment in the enzyme zone account
for accumulation, diffusion in and reaction.
Thus for lactose in the enzyme regions of the first tank

where Aj is the area available to diffusion and TLAI!^] is the reaction rate for
lactose in the ith enzyme segment of the first tank. Note that the above
equations do not include the balance for segment 1. The wall condition
requires that this balance contains only the rate of diffusion from segment 1 to
segment 2, as seen in the program.

The reaction rate is assumed, as confirmed by experiment, to have the form of

*!__ LAt[i]

Here the units are mole lactose per cm3 of porous support volume per second.

Program

Repeated here for the first tank section are the lactose balances, fluxes and rates
as given in the program on the CD-ROM.

JLA1[1. . (N-l) ]=-DLA* ( L A l E i + 1 ] -LAl[i] ) /DR


Flux of lactose flowing between segments i+1 and i

JLAl E N ] =DLA* ( L A I [N] -LAlumenl) /DR


Flux of lactose into inner lumen section
474 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

RATELA1 [1. . N] =-vmax*LAl [i] / (LAI [i]


+Km* (1+ (GA1 [i] / K i n h i b ) ) )
Reaction rate for lactose,mole/s cm3

d/dt (LAlumenl) = (FRf iber/Vlumen) * (Latank-LAlumenl )


+ ( JLA1[N] *AJ[N] ) /Vlumen
Dynamic balance for lactose in the inner lumen volume of the first tank

d/dt (LAI [2 . .N] )-(l/V[i] ) * (JLAl[i-l] *AJ[i-l] -


JLA1 [ i ] * AJ [ i ] ) +RATELA1 [ i ]
Dynamic balance for lactose in the segment 2 to N of the first tank

d/dt ( LAI [1] )=-(l/V[l] )*(JLA1[1] *AJ[1] )+RATELAl[l]


Dynamic lactose balance for the segment 1 (wall condition) of the first tank.

d / d t ( L A t a n k ) = ( Number * F R f i b e r / V t a n k ) * (LAlumen9-
L A t a n k ) + ( F f e e d / V t a n k ) * (LAf e e d - L A t a n k )
Dynamic lactose balance for the circulation tank

The geometry of the fiber is programmed such that the lumen radius and the
total fiber radius are given. The number of porous segments can be varied.

Nomenclature

Additional symbols for the geometrical factors are defined in the program on
the CD-ROM

Symbols

DGA Galactose diffusivity in porous


membrane cm2/h
DGL Glucose diffusivity in porous
membrane cm2/h
DLA Lactose diffusivity in porous
membrane cm2/h
EO Enzyme loading mg E in each fiber
FR Recycle flow cm3
GAfeed Galacose in feed mole/cm3
GLfeeci Glucose in feed mole/cm3
Ki n hib Kinetic inhibition constant mol/cm3
8.9 Membrane and Cell Retention Reactors 475

Kinetic constant mol/cm3

?
LA, GA, GL
Kinetic constant for vmax
Length of fiber,
Lactose, galactose and glucose
mol/mgE h cm3
cm
mol/cm3
Number Number of fibers
V Volumes cm
Maximum rate mol/cm3 h

Indices

fiber Refers to lumen


lumen Refers to lumen
Refers to tanks or segments
tank Refers to recycle tank

Exercises
476 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Results

Run 1:619 steps in 53.4 seconds

100 200 400 500 600


TIME

Figure 4. Approach to steady state for the tank concentrations.


Run 1: 619 steps in 53.4 seconds
9.4e-5

L
9.393e-5

Figure 5. Radial concentration gradients for the lactose and glucose. The left axis corresponds
to the outside of the fiber.
8.9 Membrane and Cell Retention Reactors 477

8.9.5 Immobilized Animal Cells in a Fluidi/ed Bed


Reactor (ANIMALIMMOB)

System

This example is based on experiments with immobilized BHK cells in a


fluidized bed of solid or porous carriers. The fluidized bed itself has an
expanded volume of 700 mL, The complete reactor system contains a volume
of 3.5 L. As seen in the figure below, the arrangement of the electrodes at the
inlet and outlet of the reactor allows an accurate difference measurement of the
oxygen uptake rate. Oxygen transfer takes place only in the conditioning
vessel, while oxygen consumption is only in the fluidized bed column, where
the cells are located.

GAS GAS INLET AIR


OUTLET FILTER OXYGEN
CARBON
OXIDE

OXYGEN
j (MEASUREMENT
CHAMBER

SAMPLE
PORTM .

THERMOSTAT
VESSEL

Figure 1. Fluidized bed for culturing animal cells on solid carriers.

The recirculation reactor is modelled by taking into account the separate


aeration tank and the geometry of the column reactor. It is assumed that the
reactor is not well mixed, but is described by a tanks-in-series model for the
column with immobilized cells and a separate well-mixed aeration tank reactor,
as shown below.
478 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

j3, FR

Spent medium

Figure 2. Schematic of the model structure, where Ci refers to any component concentration.

For the tanks-in-series description of the column, 3 (or more) tanks in series are
used. The mass balance for one component in tank 2 is then

Here V is the volume of one tank and r is the reaction rate of the component.
The circulation flowrate is FR. This balance equation form would apply to all
components, but not for the biomass since it is immobilized.
The kinetic model assumes the following:

1. Growth of cells is linked to the consumption of glucose and Yx/s = 0.28g


biomass produced per g glucose consumed.
2. Lactate is produced in proportion to the glucose uptake rate with
YiacG=2.0.
3. Oxygen, glutamine, lactate and glucose concentrations influence the rates.
4. Multiple Monod kinetics can be applied.
5. The medium is in contact with air and the solubility of oxygen is 8 mg/L.

From the data in the dissertation of Keller (1) can be calculated the yield of
lactate with respect to glucose, giving Yiacc = 2.0 mmol lactate/mmole glucose
or 1.1 g sodium lactate/g glucose. This can be used to calculate the production
rate of lactate. Also calculated from the dissertation is YgiutG=:36mg
glutamine/g glucose.
Thus the glucose uptake rate becomes

"OX K lac
w
sox K sglut 4-C lac
8.9 Membrane and Cell Retention Reactors 479

Note that the last term in the equation models an inhibition by lactate.

The growth rates are based on the specific substrate uptake rates. Other specific
rates are related by yield coefficients and biomass concentration.
For example for growth rate,
rx = qG X Yxg

Data of experimental values:


Maximal volume of the expanded fluidized bed; 0.6 L
Entire reactor system volume; 3.5 L
Medium throughput; 6.5 L/day
Feed glucose concentration; 3.9 g/L
Glucose consumption rate; 4.7 mmol/h
Oxygen uptake rate; 3.7 mmol/h
Max. cell density: nonporous carriers; 2*10 cells/ml expanded
bed volume
porous carriers; 4*10 cells/ml expanded bed volume
Ratio of inoculum cell number; approx. 5 % of final cell
number
Total biomass (porous carrier); 6.24 g (Approx. 2*10 cells per
g biomass)
Oxygen transfer coeff.; KLa 2.15 1/hr

Program

The program is on the CD-ROM.

Nomenclature

Feed cone, for glucose g/L


Cgiutf Feed glutamine concentration mg/L
CgiutF Feed glutamine concentration mg/L
Cox Dissolved oxygen mg/L
Coxsat Saturation for oxygen mg/L
F Flowrate L/h
FR Circulation flowrate L/h
GUR Glucose uptake rate mmol/h
480 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

KLa Mass transfer coeff. 1/h


Klac Inhibition constant for lactate g/L
K sg Saturation constant, for glucose g/L
KSglut Saturation constant for glutamine mg/L
KSOX Saturation constant for oxygen mg/L
OUR Oxygen uptake rate mmo 1/h
qCmax Glucose uptake rate max g/h g cells
QOxmax Oxygen uptake rate max mg/g cells h
V Volume of 3 tanks in the column L
V4 Aeration tank volume L
X Biomass concentration g/L
YoiutG Glutamine uptake rate mg glut./g glue.
YLacG Yield coefficient glucose g lactate/g
YoxG Yield coefficient glucose mg oxygen/g
YXG Yield coefficient, biomass to glucose g/g

Exercises
8.9 Membrane and Cell Retention Reactors 481

References

Keller, J. Dissertation No. 9373, ETH, 1991.

Keller, J., Dunn, I. J. and Heinzle E. "Improved Performance of the Fluidized


Bed Reactor for the Cultivation of Animal Cells" in Production of Biologicals
from Animal Cells in Culture, Ed. Spier, Griffiths, Meignier, Butterworth-
Heinemann, 10th ESACT Meeting, 513-515 (1991).

Keller, J. and Dunn, I. J. "A Fluidized Bed Reactor for the Cultivation of
Animal Cells", In: Advances in Bioprocess Engineering, Eds. E. Galindo and O.
T. Ramirez, Kluver, pp 115-122 (1994).

Results

Run 1:1618 steps in 1.03 seconds


11 '

10'

9' 6 O
O
8

? 7

5
O
4 iliii X[3]:1

3
20 40 80 100 140 160
TIME

Figure 3. Batch run showing the difference in dissolved oxygen between the tank sections.
482 8 Simulation Examples of Biological Reaction Processes Using Berkeley Madonna

Run 1:1618 steps in 1.13 seconds


8

20 60 80 100 120 140 160

Figure 4. In this run with F=0.05 L/h first oxygen limitation develops and later glucose
limitation. The immobilized biomass in the three sections increases at different rates due to the
glucose gradient in the column.
Appendix: Using the Berkeley
Madonna Language

9.1 A Short Guide to Berkeley Madonna

Computer Requirements

Two Berkeley Madonna versions are supplied with this book on a CD, one for
PC with Windows and one for the Power Macintosh. More information with
downloads can be found on the following website:
http://www.berkeleymadonna.com

Installation from CD

The files are compressed on the CD in the same form as they are available on
Internet. Information on registering Madonna is contained in the files.
Registration is optional since all the examples in the book can be run with the
unregistered version. Registration makes available a detailed manual and is
necessary for anyone who wants to develop his or her own programs.

Running Programs

To our knowledge, Madonna is by far the easiest simulation software to use, as


can be seen on the Screenshot Guide in this Appendix. Running an example
typically involves the following steps:

Start Berkeley Madonna and open a prepared program file.

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
484 9 Appendix: Using the Berkeley Madonna Language

Adjust the font and size to suit.

Go to Model/Equations on the menu and study the equations and program


logic.

Go to Parameters/Parameter Window on the menu and see how the values


are set. They may be different than on the program. Those with a * can
be reset to the original values. Also, if necessary, here the integration
method and its parameters (DT, Stoptime, DTmax, Dtmin, Tolerance, etc.)
values can be changed.

Decide which plot might be interesting, based on the discussion in the text.

Go to Graph/New Window and then Graph/Choose Variables to select data


for each axis. All calculated results on the left side of the equations are
available and can be selected.

Run the program by clicking on Run.

Adjust the graph by setting the legend with the legend button. Perhaps put
one of the variables on the right side of the graph with Graph/Choose
Variables.

Possibly select the range of the axes with Graph/Axis Settings. Choose
colors or line types with the buttons.

Decide on further runs. It is most common to want to compare runs for


different values of the parameters. This is usually done with Parameters
/Batch Runs and also with Parameters/Define Sliders. If the overlay button
is set then more than one set of runs can be graphed on top of the first set.
Sometimes more than one parameter needs to be set; this is best done with
changes done in the Parameters/Parameter Window, with an overlay graph
if desired.

As seen at the end of the Screenshot Guide, Parameter Plot runs are very
useful to display the steady-state values as a function of the values of one
parameter. For this, one needs to be sure that the Stoptime is sufficient to reach
steady-state for all the runs.
When running a program with arrays, as found in the finite-differenced
examples, the X axis can be set with [i] and the Y axes with the variables of
interest. The resulting graph is a plot of the variable values at the Stoptime in
all of the array sections. For equal-sized segments, this is the equivalent of a
plot of the variables versus distance. If the steady-state has been reached then
9.1 A Short Guide to Berkeley MADONNA 485

the graph gives the steady-state profile with distance. More on running
programs is found in Sec. 2 of the Appendix.

Special Programming Tips

Berkeley Madonna, like all programming languages, has certain functions and
characteristics that are worth noting and that do not appear elsewhere in this
book.

Editing text
The very convenient built-in editor is usually satisfactory. Also the program
can be written with a word processor and saved as a text-only file with the suffix
".mmd". Madonna can then open it.

Finding programming errors. Look at a table output of the variables


Sometimes programs do not run because of errors in the program that cause
integration problems. Some hint as to the location of the error can often be
found by making an output table of all the calculated variables. This is done
by going to Graph/New Window and then Graph/Choose Variables and
selecting all the variables. Then the program is run and the table button is
chosen. Inspection of all the values in the table during the first one or two time
intervals will usually lead to an isolation of the problem for those values that are
marked in red with NAN (not a number). Also, values going negative can be
found easily here and often indicate an integration error. Sometimes this can
be overcome with a limit function of the form, limit X>=0.

Is a bracket missing?
Madonna tests for bracket pairs, and a missing bracket will be indicated.

Setting the axes. Watch the range of values.


Remember that each Y axis can have only one range of values. This means that
you must choose the ranges so that similarly sized values are located on the
same axis.

Are there bugs or imperfections in Madonna?


Yes, there are some that we are aware of. You may find some or you may have
some special wishes for improvements. The Madonna developers in Berkeley,
California would be glad to receive your suggestions. See the homepage to
contact them

Making a pulse input to a process.


This can be done it two ways: To turn a stream on and off either use the pre-
programmed PULSE function or use an IF-THEN-ELSE statement.
486 9 Appendix: Using the Berkeley Madonna Language

Making a more complex conditional control of a program.


In general the IF-THEN-ELSE conditional statement form is used, combined
with logical expressions as found in the HELP. This can involve a switching
from one equation to another within this statement. Another way is to use flags
or constants that take values of 0 or 1 and are multiplied by terms in the
equations to achieve the desired results. Nesting of multiple IF statements is
possible:
V = I F ( D i s k < l AND P > 1 . 9 )
THEN 0.85*KV*P/SQRT(TR+273)
ELSE IF (Disk<l AND P< = 1.9 AND P>1.1)
THEN KV*P/SQRT(TR+273)*SQRT(1+(1/P)*(1/P))

Parameter estimation to fit parameters to data.


For fitting sets of data to one or more parameters the data can be imported as a
text file and fitted by going to Parameters/Curve Fit. The
Edit/Preferences/Graph Window provides the possibility of having the data as
open circles. The required data format can be found in the file KLADATA.

Optimisation of a variable.
There is optimization available under Parameters/Optimize, but if it is
something simple with one or two parameters, then sliders can also be
effectively used. If the value of a maximum is sought as a function of a single
parameter value, then the Parameter Plot for maximum value can be used.

Finding the influence of two parameters on the steady state?


A Parameter Plot choosing the "final" value can be used to find the influence
of one variable on the steady state. The second parameter can be changed in
the Parameter Window and additional parametric runs made and plotted with an
overlay plot. Thus it is possible to obtain a sort of contour plot with a series of
curves for values of the second parameter. Unfortunately, a contour plot is not
yet possible.

Nice looking results are not always correct.


A warning! It is possible to obtain results from a program that at first glance
seems OK. Always make sure that the same results are obtained when DT is
reduced by a factor of 10 or when a different integration method is used.
Plotting all the variables may reveal oscillations that indicate integration errors.
These may not be detectable on plots of a few variables.

Setting the integration method and its parameters?


It is recommended to choose the automatic step-size method AUTO and to set
equal values of DT and DTMAX. Run the integration once and reduce both
parameters by one-half and run again. If the results are good, try to improve
the speed by increasing both parameters. Finally it should be possible to set
9.1 A Short Guide to Berkeley MADONNA 487

DTMAX higher than DT, but sometimes the resulting curves are not smooth if
DTMAX is too high. In most cases, good results are obtained with AUTO and
DT set to about 1/1000 of the smallest time constant.
If no success is found with AUTO, then try STIFF and adjust by the same
procedure. Oscillations can sometimes be seen by zooming in on a graph; often
these are a sign of integration problems. Sometimes some variables look OK
but others oscillate, so look at all of them if problems arise. Unfortunately
there is not a perfect recipe, but fortunately Madonna is very fast so the trial-
and-error method usually works out.

Checking results by mass balance


For continuous processes, checking the steady-state results is very useful.
Algebraic equations for this can be added to the program, such that both sides
became equal at steady state. For batch systems, all the initial mass must equal
all the final mass, not always in mols but in kg. Expressed in mols the
stoichiometry must be satisfied.

What is a "Floating point exception"?


This error message comes up when something does not calculate correctly, such
as dividing by zero. This is a common error that occurs when equations
contain a variable in the denominator that is initially zero. Often it is possible
to add a very small number to it, so that the denominator is never exactly zero.
These cases can usually be located by outputting a table of all the variables.

Plotting variables with distance and time.


Stagewise and finite-differenced models involve changes with time and
distance. When the model is written in array form the variable can be plotted as
a function of the array index. This is done by choosing an index variable for
the Y axis and the [ ] symbol for the X-axis. The last value calculated is used in
the plot, which means that if the steady-state has been reached then it is a
steady-state profile with distance. An example is given in the"Screenshot
Guide" in Sec. 2 of the Appendix and in the example CELLDIFF.

Notation for the differential.


In this book the differential form d/dt(x) is usual. However the x' form has the
advantage that it appears in the Choose Variables menu and can be plotted. It
can also be used directly in another equation.

Writing your own plug-in functions or integration methods.


Information on using C or C++ for this can be obtained by making contact
through the BerkeleyMadonna homepage.
488 9 Appendix: Using the Berkeley Madonna Language

9.2 Screenshot Guide to Berkeley Madonna

This guide is intended as a supplementary introduction to Berkeley Madonna,


Version 8.0.1.

(CHEMQSTATST
(FILE/CHEMCr1)

{Constants}
UM=OJKS=OJK1
8F*10 D1=0.25
Stoptime=80

{Conditional equation for 0}

!nrfX=1

lnt P=0

{Mass Balances}
; 81OMASS BALANCE EQUATION
; SUBSTRATE Q^ANGE EQUATION
; PRODUCT BALANCE EQUATION
9.2 Screenshot Guide to Berkeley MADONNA 489

Figure 1. The example CHEMO has been opened and the Menu (From left: File, Edit, Flowchart
active only for flowchart programs, Model, Compute, Graph, Parameters, Window and Help) and
Graph Buttons (From left: Run, Lock, Overlay, Table, Legend, Parameters, Colors, Dashed Lines,
Data Points, Grid, Value Output and Zoom).

Figure 2. The Berkeley Madonna menus are shown above.


490 9 Appendix: Using the Berkeley MADONNA Language

UP AND QPEFIAT1GW

3T6RTTIME
iTOPTIMI
Kl=Q.03K2=O.OBY=fl.B DT 0,02
BF=10 D1=OJS MTOUT 0
LJM 0.3
KS 0.1
{Conditional equation for D} Kl 0,03
D=iftime=tstirttnenDl o.oa
0.8
10
01 0,34
toil 8=10 5
initP=G HITX 1
NITS 10
{Mass Balances} N1TP 0
; EQ
; BALANCE

Figure 3. The Model/Equations was chosen. Seen here is also the Parameter Window.

Figure 4. If a new graph is chosen under Graph/New Window then the data must be selected
under Graph/Choose Variables.
9.2 Screenshot Guide to Berkeley MADONNA 491

4J-,

4- -9,5

3,5 9

3 8.5
2.5 -8
2- -7.5

1.5 7

f- -8.5
0,5-
0 5.5
20 40 80 SO 108 120 140 ISO 130 200
TIME

Figure 5. A graph window for variables on the left and right-side Y axis with Legend Button
selected.

Run 4:10000 steps m 0.167 seconds


..........................10

7.

6-
-7
5- 6

X 4- 5 01
-4
3-
-3
4-
-2
1- 1
0
20 40 80 100 120 140 180 !80 200

Figure 6. An Overlay Graph for three values of Dl as selected in the Parameter Window.
492 9 Appendix: Using the Berkeley MADONNA Language

Figure 7. Part of the window to define the Sliders.

Figure 8. A graph of two slider runs, showing the Parameters Menu pulled down.
9.2 Screenshot Guide to Berkeley MADONNA 493

Figure 9. The Batch Runs window for 5 values of SF.

Figure 10. A Parametric Plot was chosen for 40 runs changing values of Dl to give the
final, steady-state values. The Data Button was pressed to give the points for each run.
494 9 Appendix: Using the Berkeley MADONNA Language

Figure 11. The Optimize Window, with the value of Dl being selected to minimize the
expression -D*X. The value found was 0.27, corresponding to the Parameter Plot results.

Run 11 KOODOO steps ir, 2.53 seconds

Figure 12. Two Parameter Plots overlaid showing the effect of reducing Y from 0.8 to 0.6.
9.2 Screenshot Guide to Berkeley MADONNA 495

Figure 13. A program written in an array form allows plotting all the values versus time by
choosing the variable vector, here S[ ] versus TIME for the program CELLDIFF.

Figure 14. From the same program as Fig. 13, radial profiles of three runs are plotted in an
overlay plot. The [i] values can be selected in the Choose Variables. Here the parameter
Radius has been changed to demonstrate the large influence of diffusion length.
496 9 Appendix: Using the Berkeley MADONNA Language

Figure 15. Here the program file KLAFTT is run and fitted to data in the text-file KLADATA.
The data consists of two columns: time and CE at equal intervals as seen by the open circles on
the plot. Note that the fit variable is CE and the parameter varied to minimize the difference in
least squares is KLA.
10 Alphabetical List of Examples

ACTNITR 267 NITRIF 327


ANEAMEAS 433 OLIGO 275
ANIMALIMMOB 476 OXDYN 337
BATFERM 193 OXENZ 378
BIOFILM 372 PENFERM 230
BIOFILTDYN 342 PENOXY 253
CELLDIFF 388 PHB 279
CHEMO 799 PHBTWO 374
CHEMOSTA 258 PLASMID 477
COMMENSA 400 REPFED 245
COMPASM 406 REPLCUL 249
CONINHIB 261 STAGED 290
CONTCON 367 SUBTILIS 455
DCMDEG 280 TEMPCONT 354
DEACTENZ 308 TITERBIO 349
DUAL 275 TITERDYN 349
ELECTFIT 335 TURBCON 363
ENZCON 305 TWOONE 422
ENZDYN 383 TWOSTAGE 286
ENZSPLIT 377 VARVOL 224
ENZTUBE 272 VARVOLD 224
ETHFERM 240 YEASTOSC 447
FBR 295
FEDBAT 204
FERMTEMP 358
FILMPOP 425
INHIB 327
KLADYN 335
KLAFIT 335
LACMEMRECYC 464
LACREACT 470
LINEWEAV 272
MAMMCELLCYCLE 445
MEMINH 450
MIXPOP 477
MMKINET 209
NITBED 299
NITBEDFILM 393

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
11 Index

Absorption, 117 Analogy, 114


Absorption tank, 136 Analytical solution, 19
Accumulation, 125 Analytically, 108
Accumulation terms, 135 Animal cell culture, 128
Acetogenic step, 89 Apparent reaction rate, 133
Acid-base equilibria, 47 Approximation, 113, 141
Acidogenic step, 89 Aqueous phase, 118
Active, 110 Arithmetic-mean, 141
Adaptive Control, 174 Automatic process control, 161
Adaptive tuning, 178 Automatic reset, 165
Aerated tank, 130 Auxiliary variable, 174
Aerated tank with oxygen electrode, 336 Axial, 114
Aeration, 117, 137 Axial profiles, 114
Aeration efficiency, 123 Axial segments, 115
Aeration rates, 126 Backmixing, 137
Aeration systems, 126 Backmixing flow contribution, 142
Aerobic sewage treatment, 128 Backmixing stream, 140
Agitation, 137 Balance region, 23
Air or oxygen sparging, 134 Balances, 101
Air saturation, 126 Batch, 57, 64
Air supply, 128 Batch aeration, 126
Airlift bioreactor, 139 Batch fermentation, 11, 103
Algebraic loop, 49 Batch reactor periods, 57
Alginate, 149 Batchwise, 134
Alginate bead, 149 Bed, 134
Allosteric kinetics, 74 Biocatalysis, 118
Ammonia, 134 Biocatalyst diffusion model, 153
Ammonium, 300 Biocatalytic reaction, 147
Ammonium ion, 133 Biofilm, 154
Anaerobic degradation, 89 Biofilm, 134, 145
Analogous, 114 Biofilm nitrification, 160

Biological Reaction Engineering, Second Edition. I. J. Dunn, E. Heinzle, J. Ingham, J. E. Pfenosil


Copyright 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
ISBN: 3-527-30759-1
500 Index

Biofilm Reactor, 133 Cells, 117


Biofilter, 343 Centrifugation, 110
Biofloc, 145 Chemical reaction, 126
Biofloc, 149 Chemostats, 104
Biological activity, 123 Circulating liquid supply, 129
Biological film, 149 Circulation time, 55
Biological floes and films, 388 Closed-loop response, 170
Biological oxygen uptake, 125 Coalescence, 126
Biological reaction, 145 Cocurrently, 139
Biological reactors, 101 Cohen-Coon controller settings, 170
Biological systems, 117 Column systems, 135
Biomass, 10, 103 Commensalism, 87
Biomass recycle, 110 Comparator, 162
Biomass retention, 111, 299, 393 Competition, 150
Biomass separation, 111 Competitive, 74
Bioprocess control, 175 Completely mixed gas or liquid phases, 137
Bioreaction, 117 Complex diffusion-reaction processes, 157
Bioreactor, 112 Complex kinetics, 115
Bioreactor modelling, 101 Complex models, 123, 138
Bio-trickling filter, 343 Complexity, 138
Boundary conditions, 153 Component, 122
Briggs-Haldane mechanism, 68 Component Balances, 22
Broth, 117 Components, 113
Bubble, 117, 137 Computer Solution, 19
Bubble coalescence, 137 Concentration driving force, 122
Bubble column, 139 Concentration gradient approximation, 151
Bubble size, 140 Concentration gradients, 119, 137
Bulk liquid, 145 Concentration inhomogeneities, 137
Bulk reactor concentration, 153 Concentration profile, 114
Buoyancy, 137 Conical sand bed, 134
Carbon, 134 Continuous, 64, 109, 118
Carbon dioxide, 122 Continuous Baker's Yeast Culture, 94
Carbon dioxide production rate (CPR), 10 Continuous feed and effluent stream, 134
Carbonate, 134 Continuous Operation, 60
Carrier, 145 Continuous phase, 118
Carrier matrix, 145 Continuous-cycling, 171
Cascade, 138 Contois Equation, 82
Cascade control, 172 Control, 111
Cell concentration, 101 Control point, 165
Cell productivity, 55 Control region, 23
Cell recycle, 110 Control strategy, 176
Index 501

Controlled variable, 161 Diffusional limitation, 146


Controller, 162 Diffusional mass transfer, 145
Controller action, 163 Diffusional mass transfer coefficient, 147
Controller equations, 33 Diffusion-reaction parameter, 155
Controller output, 166 Diffusion-reaction phenomena, 151
Controller tuning, 169 Diffusion-reaction systems, 151
Convection currents, 137 Diffusive mass transfer, 138
Convective flow, 26, 125 Digital simulation, 14
Convective streams, 26 Digital simulation languages, 14
Conversions, 113 Dilution rate, 105
Cross-sectional area, 114 Dimensionless, 108
Cyclic fed batch, 109 Dimensionless form, 391
Damkohler number, 155 Dimensionless group, 151
Data fitting, 339 Dimensionless parameter, 150, 154
Datafile, 341 Dimensionless variables, 155, 159, 337
Dead zone, 163 Discontinuous control, 163
Death rate, 78 Disks of liquid, 113
Degree of backmixing, 140 Dispersed, 117, 118
Density, 104 Dispersed phase, 118
Deoxygenated, 126 Dissociation equilibrium constant, 47
Deoxygenated feed method, 130 Dissolved oxygen concentration, 123, 335
Depth, 137 Dissolved oxygen electrode, 126
Derivative control, 163 Distance, 116
Deviations, 161 Distance coordinate, 151
Deviations from ideal stage mixing, 140 Double Michaelis-Menten Kinetics, 73
Difference form, 116 Double-Monod kinetics, 83, 158
Difference segment, 115 Driving forces, 119
Differential, 107 Droplet, 118
Differential control constant, 355 D-value, 77
Differential equation, 127 Dynamic, 115
Differential equations, 140 Dynamic component balances, 104
Differential gap, 163 Dynamic kla, 126-127
Differential-difference equations, 159 Dynamic Method, 335
Diffusion, 118, 119, 145 Dynamic simulation, 49
Diffusion and reaction, 388 Dynamics of measurement, 127
Diffusion control, 133 Dynamics of the liquid phase, 128
Diffusion film, 146, 336 Effective diffusivity, 30, 121
Diffusion layer, 149 Effective rate, 148
Diffusion path, 145 Effective reaction rate, 146
Diffusion rate, 149 Effectiveness Factor, 155, 391
Diffusional flux, 26, 149, 389 Efficiency, 110
502 Index

Electrode measurement dynamics, 127, 129 First order lag equation, 127, 337
Electrode membrane, 337 First order lag model, 127
Electrode response characteristic, 336 First-order, 132
Electrode time constant, 127, 129 First-order time lag, 354
Elemental balances, 23 Flocculant cell mass, 145
Energy balances, 49 Flotation, 110
Entrance, 113 Flow interaction, 140
Enzymatic, 112 Flow velocity, 114, 148
Enzyme, 112, 118 Fluid, 120
Enzyme loading, 148, 149 Fluid elements, 113
Enzyme reactor, 115 Fluidized bed, 133, 299, 149
Enzyme-substrate complex, 69 Flux, 120
Equations, 113 Food/biomass ratio, 112
Equilibrium, 10, 122 Fractional conversion, 126
Equilibrium oxygen concentration, 337 Fractional response, 127
Equilibrium relationships, 46 Free rise velocity, 137
Equilibrium value, 132 Functional modes of control, 163
Errors, 161 Gas, 117
Exit, 113 Gas absorber, 136
Experimental reactor, 130 Gas Absorption, 117
Exponential, 108 Gas and liquid films, 121
Exponential and limiting growth phases, 103 Gas balance, 125
External film, 145 Gas balance method, 128, 126
External mass transfer, 145 Gas bubbles, 117
External transport rate, 153 Gas concentrations, 125
Extraction, 118 Gas flow rates, 125
Fed Batch, 58, 64 Gas holdup, 55, 124
Feed Forward Control, 173 Gas inlet, 137
Feedback, 161 Gas phase, 117
Fermentation, 101 Gas-liquid, 117, 120
Fermentation media, 137 Gas-liquid systems, 122
Pick's Law, 29, 120 Gas-liquid transfer, 336
Film coefficients, 122 Gel, 145
Final, 109 Growth, 110
Final control element, 162 Growth rate, 32, 103
Finite difference, 30 Heat of agitation, 52
Finite difference Model, 151 Heat of fermentation, 49
Finite differencing, 388 Heat Transfer, 51
Finite differencing technique, 159 Henry coefficient, 340
Finite-differencing, 115, 153 Henry's law, 33, 122, 337
First order, 65 Henry's law constant, 122
Index 503

Heterogeneous reaction systems, 155 Interphase transfer, 26, 125


Hill Kinetics, 74 Intraparticle transfer, 145
Hydrostatic pressure, 137, 140 Intrinsic reaction rate, 150
Ideal Gas Law, 39 Ion charge balance, 47
Ideal gas law, 133, 337 Ion exchange resins, 149
Idealized flow conditions, 137 Kinetic, 106
Idealized plug flow, 137 Kinetic control, 147
Ideally mixed, 136 Kinetic model, 136
Immiscible, 118 Kinetic rate constant, 147
Immobilization, 145 Kinetic regime, 148
Immobilization matrix, 145 Kinetic relationship, 65
Immobilized, 129 Kinetics control, 149
Immobilized biocatalyst systems, 145 Kla, 335
Immobilized enzyme and cell systems, 118 Lag phase, 103
Impermeable solid, 149 Lag time, 170
Incomplete oxygen penetration, 160 Laplace transformation, 338
Increments, 151 Large bioreactors, 137
Industrial fermenters, 137 Large scale, 137
Information flow diagram, 33 Length, 115
Inhibition, 73 Length of diffusion path, 151
Inhibitory Substrate, 367 Limiting, 140, 159
Initial conditions, 103 Limiting substrate, 103, 106
Initial value, 20 Limiting substrate concentration, 78
Inlet, 106 Linear gradients, 120
Inoculum, 103 Linear growth, 108
Input rate, 102 Lineweaver-Burk diagram, 72
Integral, 163 Liquid, 117
Integral control constant, 355 Liquid balance, 125
Integral time constant, 166 Liquid balance equation, 128
Integrated, 108, 115 Liquid film, 337
Integration procedure, 20 Liquid film control, 123
Integration step length, 20 Liquid flow terms, 135
Integration time interval, 20 Liquid medium, 117
Intensity of mixing, 55 Liquid phase, 117
Intensity of mixing, 140 Liquid recycle stream, 134
Interconnected, 138 Liquid surface, 137
Interface, 117, 118, 120 Liquid velocities, 137
Interfacial concentrations, 122 Liquid-liquid, 118
Internal mass transfer, 145 Liquid-phase impeller zones, 141
Internal structure, 119 Logistic Equation, 82
Interphase, 119 Luedeking-Piret model, 85
504 Index

Maintenance coefficient, 10 Molecular diffusion, 29


Maintenance factor, 84 Molecular diffusion, 120
Mammalian Cell Cycle Control, 445 Molecular diffusion coefficient, 121
Manipulated variable, 161 Monod equation, 67
Mass balance equation, 16 Monod kinetics, 10
Mass Transfer, 117, 119 Monod-type equation, 390
Mass transfer capacity coefficient, 122, 123 Monod-type rate expressions, 105
Mass transfer coefficients, 121 Multiphase reaction, 117
Mass transfer control, 147 Multiple impeller, 140
Mass transfer resistance, 145 Multiple-organism populations, 86
Material balance equations, 101 Multiple-substrate Monod kinetics, 82
Mathematical, 150 Multi-stage, 138
Mathematical model, 12, 137 Mutual inhibition, 343
Mathematical modelling, 151 Mutualism, 87
Matrix elements, 151 Natural logarithmic, 127
Maximum, 109 Nernst-diffusion film, 146
Maximum observed rate, 148 Nitrate, 300
Maximum rates, 158 Nitrate ion, 133
Maximum reaction rate, 70 Nitrification, 133, 299
Measurement dynamics, 126, 128, 336 Nitrification reactions, 157
Measurement signal, 127 Nitrite, 300
Measurements, 106 Nitrite ion, 133
Measuring element, 162 Nitrobacter, 134
Mechanical agitation, 137 Nitrogen, 127
Mechanical energy, 137 Nitrosomonas, 134
Medium, 106 Non-competitive, 74
Membrane, 112, 127 Non-porous carrier, 146
Membrane filtration, 110 Numerical solution, 19
Methanogenic step, 89 Objective function, 178
Michaelis-Menten constant, 70, 148 Offset, 163
Michaelis-Menten kinetics, 148 Oil phase, 118
Microbial interaction, 86 One-dimensional diffusion, 158
Microbial physiology, 106 On-line adaptive optimizing control, 178
Microbiological, 133 On-line monitoring, 126
Mixing, 113 On-Off Control, 163
Mixing zones, 140 Open-loop tuning technique, 169
Mode of control, 168 Operation, 110
Model, 109 Order of magnitude analysis, 155
Modelling, 113, 388 Organism balance, 102
Molar flow rate of air, 133 Oscillations, 355
Molar reaction rates, 136 Oscillations of continuous culture, 94
Index 505

Outlet, 106 Plant cell culture, 128


Output rate, 102 Plug flow, 113, 141
Overall mass transfer capacity coeff., 123 Poly-6-hydroxybutyric Acid (PHB), 93
Overall mass transfer rate equation, 123 Porous, 119
Overall order of reaction, 148 Porous biocatalyst, 119
Overall rate of reaction, 62 Porous solid, 145
Overall resistance to mass transfer, 123 Power inputs, 137
Oxidation steps, 134 Predator-Prey Kinetics, 86
Oxygen, 117, 122, 388 Pressure, 137
Oxygen balances, 123 Process control, 10, 56, 161
Oxygen depletion, 132 Process dynamics, 127
Oxygen diffusion effects, 157 Process optimization, 12, 56
Oxygen electrode, 335 Process reaction curve, 169
Oxygen electrode dynamic, 337 Process response, 128
Oxygen gas phase concentrations, 128 Product, 118
Oxygen gradients, 139 Product inhibition, 113
Oxygen limitation, 388 Product inhibition kinetics, 63
Oxygen requirements, 158 Production rate, 102
Oxygen transfer, 125 Productivity, 105, 110
Oxygen transfer coefficient, 136, 335 Programmed, 174
Oxygen transfer rate (OTR), 10 Proportional, 143, 163
Oxygen uptake rate, 38, 125, 128, 136, 390 Proportional control constant, 355
OUR, 38 Proportional-Derivative (PD) Controller, 166
Oxygen-enriched air, 131 Proportional-Integral controller, 355
Oxygen-sensitive culture, 97 Proportional-Integral-Derivative
Packed, 149 Controller, 167
Parameter, 111 Proportlonal-Reset-Rate-Control, 167
Parameter estimation, 131 Pulse, 113
Partial differential equation, 116, 153 Quadratic equation, 148
Partial pressure, 122 Quasi-homogeneous, 135
Penetration, 149 Quasi-homogeneous reaction, 158
Penetration distance, 159, 388 Quasi-steady state, 107
Penetration-limiting, 159 Radial variations, 140
Perfect mixing, 137 Rate expressions, 103
Perfect plug flow, 137 Rate of accumulation, 21, 102
Performance, 113 Rate of fermentation, 126
pH control, 49 Rate of mass transfer, 121
Phase interface, 120 Rate of oxygen transfer, 123
Phases, 117 Rate of oxygen uptake, 123
Physical model, 12, 137 Rate of substrate uptake, 83
Physical properties, 122 Rate of supply, 148
506 Index

Reactants, 117 Saturation constant, 10


Reaction, 114, 117, 118 Scheduled adaptive control, 174
Reaction capability, 149 Second-order response lag, 337
Reaction control, 133 Sections, 138
Reaction Heat, 51 Sedimentation, 110, 111
Reaction parameter, 155 Semi-Continuous Reactor, 314, 349
Reaction rate constant, 132 Separation, 110
Reaction rate control, 133 Separator, 111
Reaction site, 118, 145 Series of tank reactors, 62
Reaction surface, 147 Set point, 161
Reaction-rate limitation, 136 Shear, 55
Reactor, 101, 138 Simulation, 107
Reactor cascade, 62 Simulation example, 104
Reactor column, 135 Simulation methods, 140
Reactor efficiency, 111 Simulation programming, 153
Reactor modes, 63 Simulation programs, 14
Reactor operating conditions, 159 Simulation results, 159
Reactor outlet, 136 Simulation software, 15
Re-aerated, 126 Simultaneous diffusion and reaction, 150
Recycle loop, 134 Single stage, 138
Recycle loop configuration, 134 Single-pass conversion, 135
Recycle rates, 134 Slab, 151
Recycle ratio, 111 Slope, 18, 127
Regimes, 133 Solid, 118
Research, 106 Solid biocatalyst, 119
Reset time, 166 Solid carrier, 145
Residence time, 105, 116 Solid phase, 135
Residence time distribution, 137 Solid-liquid interfacial area, 135
Residual error, 167 Solubility, 122
Resistance to mass transfer, 120, 123 Solution, 143
Respiration quotient (RQ), 10 Spatial variations, 123
Response, 127 Specific area for mass transfer, 121
Response curve, 126 Specific carbon dioxide production rate, 11
Retention, 110, 145 Specific carbon dioxide uptake rate, 84
Riser, 139 Specific death rate, 78
Sample, 128 Specific growth rate, 10, 78, 105
Sampled data control, 174 Specific interfacial area, 122
Sampling frequency, 175 Specific oxygen, 11
Sampling interval, 174 Specific oxygen uptake rate, 84
Sand, 134 Specific product production rate, 84
Saturation, 127 Specific substrate uptake rate, 84
Index 507

Spherical bead, 388 System, 110


Spherical shell, 389 Tank, 101
Stages, 138 Tank sizes, 113
Stagewise model, 139 Tanks-in-series, 112
Stagewise modelling, 139, 140 Teisser Equation, 81
Stagnant, 120 Temperature, 148
Stagnant film, 145 Temperature measurement, 354
Stagnant flow, 120 Theoretical basis, 112
Starting, 109 Thiele modulus, 155
Startup, 61 Thin film, 120
Startup period, 61, 62 Time, 109
Steady state, 104 Time constant for heater, 355
Steady state conditions, 104 Time constant for measurement, 355
Steady state tubular reactor design, 115 Time constant measurement, 340
Steady state values, 140 Time constants, 128
Steady-state, 105 Time constants for transfer, 340
Steady-state balances, 21 Time-varying, 114
Steady-state position, 165 Titration, 126
Step change, 127 Total interfacial area, 121
Sterile, 104 Total mass balance, 102
Stirring power, 137 Total system, 134
Stirring speed, 148 Toxicity, 134
Stirring speed, 149 Tracer, 113
Stoichiometric coefficients, 40 Tracer experiment, 140
Stoichiometric oxygen requirements, 159 Tracer techniques, 137
Stoichiometric relations, 134 Transfer control, 133, 149
Stoichiometry, 40, 126 Transfer parameters, 138
Structured kinetic model, 11 Transmission lines, 162
Substrate balance, 102 Transport of material, 117
Substrate concentration, 101 Transport streams, 26
Substrate gradients, 151 Transport-reaction process, 147
Substrate inhibition, 65, 108 Trial and error method, 169
Substrate uptake rate, 55 Tubular, 113
Sulfite method, 126 Tubular reactor, 62, 113
Sulfite oxidation, 126 Turbine impeller, 134
Support, 119 Turbulence, 120
Surface, 118 Turbulent flow, 120
Surface concentration, 147 Two position action, 163
Surface reaction, 145 Two-Film Theory, 120
Sustained oscillations, 94 Ultimate gain, 169
Symmetry, 151 Ultimate period, 171
508 Index

Uncompetitive, 74 Well-mixed, 24, 101, 123


Uptake rate, 111 Well-mixed gas phase, 133
Variable, 106 Well-mixed liquid zones, 140
Viscosity, 149 Well-mixed tank, 25
Volumetric flow rate, 101 Whole cells, 149
Washout, 105 Yield coefficient, 10, 32, 102
Wastage, 110 Zero order, 65, 147, 148
Waste water, 111 Zero-order kinetics, 391
Water, 122 Ziegler-Nichols Method, 169

S-ar putea să vă placă și