+ All Categories
Home > Documents > Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from...

Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from...

Date post: 30-Apr-2020
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
194
Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri Martin Seletzky PhD thesis Biochemical Engineering, RWTH Aachen University
Transcript
Page 1: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms

Juri Martin Seletzky

PhD thesis Biochemical Engineering, RWTH Aachen University

Page 2: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri
Page 3: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms

Von der Fakultät für Maschinenwesen der Rheinisch-Westfälischen Technischen Hochschule

Aachen zur Erlangung des akademischen Grades eines Doktors der Ingenieurwissenschaften

genehmigte Dissertation

vorgelegt von

Juri Martin Seletzky

aus

Bonn

Berichter: Universitätsprofessor Dr.-Ing. Jochen Büchs

Universitätsprofessor Dr.-Ing. Peter Götz

Tag der mündlichen Prüfung: 13.06.2007

Diese Dissertation ist auf den Internetseiten der Hochschulbibliothek online verfügbar.

Page 4: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri
Page 5: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Abstract V

Abstract

The aim of this study was to develop and advance techniques required to carry out process

developments with suspended and immobilized aerobic microorganisms cultivated in batch or

continuous mode. The applicability is demonstrated with the biological model system

Corynebacterium glutamicum grown on lactic acid. Respiration measurement in shake flasks

is introduced as a screening method to characterize the behavior of microorganisms exposed

to stress factors such as carbon starvation, anaerobic conditions, organic acids, osmolarity and

pH. It is evaluated and compared with the standard measuring methods, exhaust gas analysis

and respirometry. A simple and inexpensive shake flask screening method to perform

immobilization studies with solid supports is introduced. This method is applied to study the

adhesion of dormant and growing cells to ceramic and glass support materials with different

porosities and surface modifications. A novel everyday scale-up strategy from shake flasks to

fermenters based on empirical correlations of the volumetric mass transfer coefficient and the

pH change of the medium is developed and applied for batch and continuous cultures. Mixing

time, gas-liquid mass transfer, the influence of antifoam agent, power input, gas hold-up,

productivity, and long term behavior of a novel external loop biofilm reactor with a

honeycomb ceramic monolith as immobilization carrier are determined. With modeling the

productivities of biofilm and standard reactors are compared for aerobic processes with

growth associated product formation. To identify optimal operating conditions, the influences

of dilution rate, mass transfer, carbon source concentration, and surface area on the

productivity is modeled. The experimental results are summarized in Chapter 9 (Summary of

conclusions and experimental results).

Page 6: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Zusammenfassung VI

Zusammenfassung

Ziel dieser Arbeit war die Entwicklung von Verfahren zur effizienten Prozessentwicklung mit

suspendierten und immobilisierten Mikroorganismen in ansatzweiser und kontinuierlicher

Prozessführung. Die Anwendung der Verfahren wurde mit dem biologischen Modellsystem

Corynebacterium glutamicum auf der Kohlenstoffquelle Milchsäure demonstriert. Atmungs-

messung im Schüttelkolben wird als Screeningverfahren zur Charakterisierung des Verhaltens

von Mikroorganismen angewendet, um die Auswirkungen von Stressfaktoren (Osmolarität,

pH-Wert, organische Säuren, Mangel an Nähstoffen und Sauerstoff) auf die metabolische

Aktivität zu untersuchen. Das Verfahren wird evaluiert und mit den Standardatmungsmess-

verfahren Abgasanalyse und Respirometrie verglichen. Ein einfaches und kostengünstiges

Screeningverfahren im Schüttelkolben zur Durchführung von Immobilisierungen auf Fest-

körpern wird eingeführt, um das Adhäsionsverhalten von ruhenden und wachsenden Zellen

auf oberflächemodifizierten porösen Glas- und Keramikwerkstoffen zu charakterisieren.

Aufbauend auf empirischen Korrelationen des volumetrischen Stoffaustauschkoeffizienten

und des pH-Werts wurde ein Verfahren zur Maßstabsvergrößerung vom Schüttelkolben zum

Fermenter entwickelt. Mischzeit, Stofftransport, Einfluss von Antischaummittel,

Leistungseintrag, Gashold-up, Produktivität und Langzeitverhalten eines neuartigen

Biofilmschlaufenreaktors mit einem keramischen Honeycombmonolithen als

Aufwuchsoberfläche wurden bestimmt. Durch Modellierung konnten die Produktivitäten von

Biofilm- und Standardreaktoren für aerobe Prozesse mit wachstumsgekoppelter

Produktbildung verglichen werden. Zur Identifikation optimaler Betriebsbedingungen wurden

der Einfluss der Durchflussrate, des Stofftransports, der Konzentration der Kohlenstoffquelle

und der Aufwuchsoberfläche auf die Produktivität modelliert. Die experimentellen Ergebnisse

sind in Kapitel 9 (Summary of conclusions and experimental results) zusammengefasst.

Page 7: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Vorwort VII

Vorwort

Die vorliegende Arbeit entstand während meiner Tätigkeit als wissenschaftlicher Mitarbeiter

am Lehrstuhl für Bioverfahrenstechnik der RWTH Aachen.

Mein Dank gebührt meinen Kooperationspartnern am Institut für keramische Komponenten

im Maschinenbau (IKKM) Karen Otten und Horst R. Maier, die immer bereitwillig und

fachkundig auf meine Fragen eingegangen sind und meine Wünsche und Anregungen zügig

und hilfsbereit umgesetzt haben.

Ohne den kontinuierlichen Einsatz, die wertvollen Anregungen und die Mühen meiner

Studenten Sebastian Hahn, Hendrik Schneider, Ute Noak, Silvia Denter, Irena Münz, Daniela

Dreymüller, Christian Pohlmann, Michael Stemmler, Eike Welk, Jens Fricke und Linn

Mehnert, die viele Nächte und Wochenenden am Lehrstuhl verbracht haben, wäre diese

Arbeit niemals möglich gewesen. Euch allen gilt mein besonderer Dank.

Für die finanzielle Unterstützung dieser Arbeit danke ich der deutschen Forschungs-

gemeinschaft (DFG).

Gut, dass Kaffeetrinken nicht nur der Nahrungsaufnahme dient. Für die abwechslungsreichen

Gespräche, Meetings, Krisentreffen und Verschnaufpausen am Kaffeeautomaten oder im

Grünen danke ich meinen Kollegen. Nur dank Eurer Hilfsbereitschaft, Kreativität und

Kooperationsbereitschaft aber auch Kritik ließ sich auch das scheinbar Unmögliche oft

verwirklichen.

Ohne die anregenden Gespräche am Abendbrotstisch wären viele wertvolle Ideen niemals in

diese Arbeit eingeflossen. Für diese spannende Lebensatmosphäre danke ich meiner Frau

Julie.

Mein herzlicher Dank gilt Jochen Büchs, der mir viel Vertrauen entgegengebracht hat und mir

den Freiraum gegeben hat, meine Ideen und Ziele zu verwirklichen.

Aachen, im Juni 2007 Juri Seletzky

Page 8: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Contents VIII

Contents

Abstract V

Zusammenfassung VI

Vorwort VII

Contents VIII

Nomenclature XIV

Roman symbols XIV

Greek symbols XVI

Subscripts, Superscripts XVII

Abbreviations XVIII

1 Introduction 1

2 Characterization of the biological model system Corynebacterium glutamicum with respiration measurement in shake flasks 5

2.1 Introduction 5

2.2 Materials and methods 8

2.2.1 Microorganism, media and culture conditions 8

2.2.2 Stress induced by carbon starvation 9

2.2.3 Stress induced by anaerobic conditions 9

2.2.4 Stress induced by lactic acid 10

2.2.5 Stress induced by osmolarity and pH 11

2.3 Results 11

2.3.1 Stress induced by carbon starvation 12

2.3.2 Stress induced by anaerobic conditions 13

2.3.3 Stress induced by lactic acid 15

2.3.4 Stress induced by osmolarity and pH 18

Page 9: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Contents IX

2.4 Discussion 21

2.4.1 Stress induced by carbon starvation 21

2.4.2 Stress induced by anaerobic conditions 22

2.4.3 Stress induced by lactic acid 22

2.4.4 Stress induced by osmolarity 23

2.4.5 Stress induced by pH 24

2.4.6 Evaluation of the general method 24

3 Evaluation of respiration measurement techniques 26

3.1 Introduction 26

3.2 Materials and Methods 28

3.2.1 Microorganism and Cultivation 28

3.2.2 Analytical procedures and culture conditions 28

3.3 Results and discussion 29

3.3.1 Accuracy 29

3.3.2 Precision 33

3.3.3 Quantitation limit and range 33

3.3.4 Summary and overall conclusion 36

4 Development of a scale-up approach for aerobic cultures 38

4.1 Introduction 38

4.2 Materials and methods 41

4.2.1 Microorganism and media 41

4.2.2 Fermentation devices 41

4.3 Determination of the culture conditions 42

4.3.1 Scale-up of the volumetric mass transfer coefficient and determination of non oxygen limited culture conditions 42

4.3.2 Scale-up of the pH 46

4.4 Model description 49

4.4.1 Description of the biological model 49

Page 10: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Contents X

4.4.2 Description of the pH model 51

4.5 Results and discussion 54

4.5.1 Gas-liquid mass-transfer 54

4.5.2 Scale-up from shake flasks to fermenters in batch and continuous mode 58

4.6 Conclusion 61

5 Development of a screening approach for immobilized biomass on solid supports 62

5.1 Introduction 62

5.2 Materials and methods 63

5.3 Results and discussion 67

6 The influence of material properties and surface modifications on the adhesion of dormant and growing cells 72

6.1 Introduction 72

6.2 Materials and methods 74

6.2.1 Microorganism, media, and culture conditions 74

6.2.2 Supports and surface modifications 75

6.2.3 Adhesion experiments 77

6.3 Results 79

6.3.1 Characterization of material properties and microorganisms 79

6.3.2 Adhesion experiments 83

6.4 Discussion 88

6.4.1 Characterization of material properties and microorganisms 88

6.4.2 Adhesion experiments 89

7 Characterization of a novel biofilm reactor with a honeycomb monolith as immobilization carrier 91

7.1 Introduction 91

7.2 Materials and methods 93

7.2.1 General description, mode of operation and design of the loop reactor 93

Page 11: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Contents XI

7.2.2 Measuring techniques 97

7.2.3 Flow regimes and Reynolds numbers in riser and downcomer 99

7.3 Results 100

7.3.1 Gas liquid mass transfer and antifoam agent 100

7.3.2 Specific power input 103

7.3.3 Pressure below single tube and monolith 104

7.3.4 Gas hold-up 104

7.3.5 Mixing 105

7.4 Discussion 106

7.4.1 Gas-liquid mass transfer and antifoam agent 106

7.4.2 Specific power input 108

7.4.3 Pressure below single tube and monolith 108

7.4.4 Gas hold-up 109

7.4.5 Mixing 109

7.4.6 The influence of the monolith on flow regime, mass transfer, and mixing time 110

8 Modeling aerobic biofilm reactors with growth associated product formation 112

8.1 Introduction 112

8.2 Model descriptions 114

8.2.1 Modeling tools 114

8.2.2 Model introduction 114

8.2.3 Model assumptions 116

8.2.4 Bulk phase 117

8.2.5 Gas phase and gas-liquid mass transfer 117

8.2.6 Biofilm phase 118

8.2.7 Biological conversion 120

8.2.8 Model description of the standard reactor 121

Page 12: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Contents XII

8.2.9 Model description of a simple not mechanistic biofilm model with constant immobilized biomass 121

8.3 Definition of the model parameters 122

8.4 Materials and methods biological experiments 125

8.4.1 Adaptation of the model parameters to the biological experiments 125

8.5 Results and discussion 126

8.5.1 General behavior of the biofilm reactor 126

8.5.2 Influence of mass transfer and carbon source concentration on the productivity 133

8.5.3 Influence of specific surface area and carbon source concentration on the productivity 136

8.5.4 Influence of diffusion coefficient, film density and Monod constant on the productivity 136

8.5.5 Comparison of experimental results and model predictions 139

8.5.6 Discussion of the comparison of experimental results and model predictions 141

9 Summary of conclusions and experimental results 143

9.1 Aim and introduction 143

9.2 Characterization of the biological model system C. glutamicum with respiration measurement in shake flasks 144

9.3 Evaluation of respiration measurement techniques 145

9.4 Development of a scale-up approach for aerobic cultures 145

9.5 Development of a screening approach for immobilized biomass on solid supports 146

9.6 The influence of material properties and surface modifications on the adhesion of dormant and growing cells 147

9.7 Characterization of a novel biofilm reactor with a honeycomb monolith as immobilization carrier 148

9.8 Modeling aerobic biofilm reactors with growth associated product formation 149

Appendix 151

General materials and methods 151

A.1 Microorganism and media 151

Page 13: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Contents XIII

A.1.1 Microorganism 151

A.1.2 Media, cultivation and culture conditions 151

A.1.3 Calculation of the media solubilities 152

A.1.4 Recording of titration curves of the media 153

A.2 Culture conditions and fermentation devices 153

A.2.1 Fermenters in batch mode 153

A.2.2 Fermenters in continuous mode 154

A.2.3 Shake flasks in batch mode 154

A.2.4 Shake flasks in continuous mode 154

A.3 Respiration measurement 156

A.3.1 Respiration measurement with exhaust gas analysis in fermenters 156

A.3.2 Respiration measurement with the RAMOS device in shake flasks 158

A.3.3 Respiration measurement with respirometry 158

A.4 Support materials for immobilization 159

A.5 Analytics 159

A.5.1 Biomass determination 159

A.5.2 HPLC chromatography 160

A.6 Modeling tools 160

Bibliography 161

Curriculum vitae 175

Page 14: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Nomenclature XIV

Nomenclature

Roman symbols

symbol name unit

A surface area of the support [m²]

a specific surface area [m²/m³]

AB amount titrant (acid or base) [mol/l]

ABtamount of titrant (acid or base) per time [mol/l/h]

AVµ average specific growth rate [h-1]

c concentration [mol/l], [g/l]

cf correlation factor between OTCmax, bio and OTCmax, sulfite

[-]

cO2 cumulative consumed oxygen [mol/l]

CTR carbon dioxide transfer rate [mol/l/h]

CV coefficient of variation [%]

D dilution rate [h-1]

d maximum inner shake flask diameter [cm]

d0 shaking diameter [cm] DD diffusion coefficient [cm²/s]

Dh hydrodynamic diameter [mm]

DO dissolved oxygen [%]

dt time interval [h]

h space coordinate [m], [µm]

H liquid level [mm]

J flux [mol/m²/h]

K Monod constant, dissociation constant [mol/l], [g/l]

K(H2O) dissociation constant of water [mol²/l²]

Page 15: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Nomenclature XV

kL gas-liquid mass transfer coefficient

[m/s]

kLa volumetric mass transfer coefficient

[s-1]

kLa/kLa0

ratio between volumetric mass transfer coefficient with antifoam agent and without antifoam agent

[-]

m maintenance coefficient [mol/g/h], [g/g/h]

n shaking frequency [rpm]

NH3 ammonium concentration [g/l]

OTCmax oxygen transfer capacity [mol/l/h]

OTR oxygen transfer rate [mol/l/h]

OTRmax maximum oxygen transfer rate [mol/l/h]

OUR oxygen uptake rate [mol/l/h]

p pressure [bar]

P power input [kW]

pO2 partial oxygen pressure [bar]

pR reactor pressure [bar]

q specific aeration rate [vvm]

Q aeration rate [l/min]

r rate [mol/l/h], [g/l/h]

R gas constant [bar·l/mol/K]

r² coefficient of determination [-]

Re Reynolds number [-]

rgas separator radius gas separator [mm]

S solubility, carbon source concentration [mol/l/bar], [g/l]

STY space time yield [g/l/h]

T temperature [K, °C]

t time [h]

tOTRmax fermentation time until [h]

Page 16: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Nomenclature XVI

OTRmax

V filling volume, volume [l], [ml]

lV& volume flow rate liquid [l/min]

v velocity [m/s]

Vggas volume headspace shake flask [ml]

Vm0molar gas volume at standard conditions [l]

X biomass concentration [g/l]

Y Yield coefficient [g/mol], [g/g]

yO2oxygen mole fraction gas phase [mol/mol]

YO2/S oxygen substrate coefficient [mol/g]

yO2loxygen mole fraction liquid phase [mol/mol]

YX/NH3 biomass ammonium yield [g/g]

YX/O2 biomass oxygen yield [g/mol]

YX/S biomass substrate yield [g/g]

Greek symbols

Symbol Name Unit

∆ difference

ε gas hold-up [-]

µ growth rate [h-1]

µmax maximum growth rate [h-1]

ρ biofilm density [g/l]

ν kinematic viscosity [cm²/s]

Page 17: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Nomenclature XVII

Subscripts, Superscripts

Symbol Name Unit

* equilibrium concentration

0 initial, original, beginning, standard condition

ba batch

bio biological conversion, biological medium

bulk variable of the bulk-phase

bulk↔film bulk-film mass transfer

c molar concentration [mol/l]

co continuous

CO2 carbon dioxide

F feed

feed concentration in the feed

film variable of the biofilm-phase

gas variable of the gas-phase

gas↔bulk gas-liquid mass transfer

i different species

imm immobilized biomass

in inlet, initial

g gas

l liquid

m maintenance

max maximum

min minimum

O2 oxygen

out outlet, harvest, exhaust

S carbon source, substrate

Page 18: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Nomenclature XVIII

sulfite sulfite system

sum sum of all species

p pump

sup support

X biomass

Abbreviations

Symbol Name

AV average

BFR biofilm reactor

COSBIOS Continuously Operated Shaking Bioreactor System

EGA exhaust gas analyzer

OD optical density

RAMOS Respiration Activity Monitoring System

SiC silicon carbide

STR stirred tank reactor

Page 19: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 1 Introduction 1

Chapter 1

Introduction

This study focuses on the process development and scale-up of aerobic microorganisms,

which may be suspended (planctonic, free) or immobilized on a support (Fig. 1.1). The aim of

this study was to develop novel and advance existing techniques and methods required to

efficiently perform process development of aerobic microorganisms with shake flasks and

fermenters. The applicability of the techniques was demonstrated with the biological model

system Corynebacterium glutamicum grown on lactic acid. As depicted in Fig. 1.1, the

process development comprised the following steps: characterization of the behavior of

suspended and immobilized Corynebacterium glutamicum, development and evaluation of

screening systems for suspended and immobilized microorganisms in shake flasks,

development of an everyday scale-up approach from shake flasks to fermenters for aerobic

cultures, development and characterization of a novel external loop biofilm reactor with a

honeycomb monolith as immobilization carrier, and finally the modeling and comparison of

biofilm fermenters and standard fermenters.

The process development strategy for aerobic microorganisms was based on two principles: I.

respiration and oxygen consumption are of fundamental importance for the behavior and

productivity of aerobic microorganisms, and II. to cope with the large numbers of experiments

required to develop and optimize an industrial production a simple and cost efficient

screening system is required. Shake flasks are suitable to serve this purpose.

Page 20: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 1 Introduction 2

Characterization biological model system and screening results

Screening system shake flask

Scale-up approach for aerobic cultures

Reactor development

Modeling and evaluation

Corynebacterium glutamicum carbon source lactic acid

Stress factors and growth conditions. Chapter 2

Stoichiometry for modeling. Chapter 2, Chapter 4

Adhesion of dormant and growing cells to surface modified glass and ceramic supports. Chapter 6

Evaluation of respiration measurement techniques in shake flasks and fermenters. Chapter 3

Development of a simple screening approach for immobilized biomass on solid supports. Chapter 5

Development of a scale-up approach for aerobic cultures from shake flasks to fermenters based on the volumetric mass transfer coefficient (kLa) and the pH. Chapter 4

Development and characterization of a novel external loop biofilm reactor with a honeycomb monolith as immobilization carrier. Chapter 7

Modeling aerobic biofilm reactors with growth associated product formation. Chapter 8

Modeling growth, respiration and pH in shake flasks and fermenters. Chapter 4

Fig. 1.1:Organisation and content of the present study.

I. According to [Anderlei and Büchs 2001, Stöckmann et al. 2003b], in aerobic cultures

almost every physiological activity is coupled to the respiratory uptake of oxygen, making the

oxygen transfer rate (OTR) a valuable parameter to monitor the metabolic activity of

biological cultures. The OTR reflects the physiological responses of the microorganisms to

different culture conditions such as temperature, pH, osmotic stress, nutrient limitation and

inhibition, product or by-product formation and inhibition. Various authors have

demonstrated this leading importance of the substrate oxygen. According to Gupta and Rao

[2003] changes in oxygen availability may lead to drastic effects on fermentation kinetics.

Clark et al. [1995] state, that oxygen limitation acts in an analogous manner to substrate

limitation. If the gas-liquid mass-transfer is the limiting step, efforts aiming at the

Page 21: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 1 Introduction 3

improvement of e.g. the enzymatic activity of the metabolic pathway leading to the desired

product, will not be of much benefit [Büchs 2001].

II. When designing an industrial fermentation, a large number of experiments are necessary to

select a strain, to improve the medium, and to optimize the fermentation procedure.

According to Kennedy et al. [1994], the only practical way of performing these experiments is

to use shake flasks. Probably more than 90 % of all culture experiments in biotechnology are

performed in shaking bioreactors [Büchs 2001].

Based on these two principals the characterization of the biological model system is

performed with the RAMOS technique, a novel device introduced by Anderlei and Büchs

[2001] and Anderlei et al. [2004] to measure the OTR in shake flasks. Respiration

measurement in shake flasks is introduced as a screening method to optimize the culture

conditions and to characterize the behavior of microorganisms exposed to stress factors such

as carbon starvation, anaerobic conditions, organic acids, osmolarity and pH (Chapter 2). The

accuracy, precision, quantitation limit and range of the novel measuring technique is verified

by comparison with the standard methods exhaust gas analysis and respirometry (Chapter 3).

Today, the formation and structure of biofilms and immobilized microorganisms on solid

supports is usually studied with complex, often especially designed reactors. This study

introduces a generic, simple, and inexpensive method to perform screening experiments with

solids in shake flasks. (Chapter 5). This screening method combined with the RAMOS

technology is applied to study the adhesion of dormant and growing cells to ceramic and glass

supports with different porosities and surface modifications (Chapter 6).

To transfer the results gained with the shake flask screening methods to fermenters, a scale-up

strategy is required. The scale-up strategy should make it possible to select the operating

conditions of aerobic cultures in shake flasks and fermenters. The scale-up strategy introduced

Page 22: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 1 Introduction 4

in this study is based on the volumetric mass transfer coefficient (kLa) and the pH change of

the medium. The effectiveness of the scale-up strategy is demonstrated by comparing the

behavior of Corynebacterium glutamicum on lactic acid in shake flasks and fermenters

operated in batch and continuous mode (Chapter 4).

A novel external loop biofilm reactor with a honeycomb ceramic monolith as immobilization

carrier is developed and characterized. The design of the reactor was suggested by an analogy

to the human body. The capillaries inside the monolith should deliver, similar to the blood

vessels, oxygen and nutrients to the cells, which are immobilized in and on the porous walls

of the monolith. The intention of the novel design is to develop an aerobic biofilmreactor

where the entire immobilized biomass can be supplied with oxygen (Chapter 7).

With modeling the productivities of biofilm and standard reactors are compared for aerobic

processes with growth associated product formation. The influence of dilution rate, mass

transfer, carbon source concentration, and surface area on the productivity is investigated to

identify conditions which favor the use of the standard reactor and conditions which favor the

biofilm reactor (Chapter 8).

Page 23: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 5

Chapter 2

Characterization of the biological model system Corynebacterium glutamicum with respiration measurement in shake flasks

Respiration measurement in shake flasks is introduced as a screening method to optimize the

culture conditions and to characterize the behavior of microorganisms. The screening method

should allow to identify optimum and critical growth conditions and characterize the behavior

of Corynebacterium glutamicum exposed to different stress factors. Most of the results

presented in this chapter have been described by Seletzky et al. [2006a].

2.1 Introduction

Stress tolerance and stress adaptation of microorganisms has been characterized to understand

the survival strategies of pathogenic organisms like Mycobacterium tuberculosis [Hu and

Coates 1999, Primm et al. 2000], Staphylococcus aureus [Watson et al. 1998], or food

spoiling bacteria like Listeria monocytogenes [Ferreira et al. 2001] and Listeria innocua

[Houtsma et al. 1996]. Escherichia coli was used by several researchers [Breidt jr et al. 2004,

Ihssen and Egli 2004, Mandel and Silhavy 2005, Presser et al. 1997, 1998, Roe et al. 2002,

Ross et al. 2003] to study the stress tolerance and stress responds of gram-negative organisms.

Microorganisms have developed complex stress-survival and stress-response strategies to

respond to various stress-factors like starvation, oxygen, osmolarity or pH. The knowledge

about these strategies has increased fundamentally in the last decade due to the advances of

molecular biological methods. However, the standard methods to determine the viability and

Page 24: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 6

activity after and during stress have remained unchanged. Viable cell counts with petri dishes

and activity determination with respirometry are still the most common methods to determine

the interaction between microorganisms and stress.

In this chapter a new method is introduced, to continuously monitor the oxygen transfer rate

(OTR) of microorganisms in shake flasks during and after exposure to stress without the

necessity to change the culture vessel or the medium. The OTR reflects the metabolic activity

and in most cases also the growth rate of microorganisms, due to the fact that almost every

physiological activity in aerobic cultures, is coupled to the respiratory uptake of oxygen

[Anderlei et al. 2004, Anderlei and Büchs 2000, Losen et al. 2004, Stöckmann et al.

2003a, b]. The new method has three major advantages over the standard methods. I.

Cultivation and metabolic activity determination are performed in the same culture vessel and

in the same medium, excluding stress-factors which may be induced by transferring the

microorganisms to plates or respirometers. II. Continuous online monitoring enables the

determination of the metabolic activity independent of the sampling interval. III. The

sampling times for molecular stress response analysis can be chosen in a flexible way

according to the on-line monitored behavior of the culture and not according to a fixed

predetermined sampling schedule. The mentioned advantages are all part of the technical

nature of the measuring technique. The broad applicability of the method is demonstrated by

analyzing the behavior of Corynebacterium glutamicum on minimal medium with lactic acid

as sole carbon source under the stress-factors carbon starvation, anaerobic conditions, lactic

acid, osmolarity, and pH.

The physiological behavior of biological cultures can be derived from the OTR by analyzing

the slope and the form of the OTR curve [Chapter 3.3.1, Anderlei et al. 2004, Anderlei and

Büchs 2001, Losen et al. 2004, Seletzky et al. 2006a, c, Stöckmann et al. 2003b], additionally,

Page 25: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 7

the integral of the OTR curve can be used to control the stoichiometry and the exhaustion of

essential nutrients [Chapter 3.3.2, Seletzky et al. 2006a, c, Stöckmann et al. 2003].

Corynebacterium glutamicum is widely used for the industrial production of amino acids

[Kircher and Pfefferle 2001]. It is a high-G+C gram-positive soil bacterium and closely

related to the Brevibacteria and Mycobacteria [Gruber and Bryant 1997, Kinoshita et al.

1957]. Stress behavior and amino acid production are linked because external triggers like

biotin limitation or heat stress are often required to start amino acid production [Barreiro et al.

2005, 2004, Gutmann et al. 1992, Stansen et al. 2005]. To improve amino acid production the

responds of C. glutamicum to substrate limitations/starvations have been investigated with a

strong focus on nitrogen [Burkovski 2003, Nolden et al. 2001, Silberbach et al. 2005], and

phosphate [Ishige et al. 2003, Kočan et al. 2006]. Further interest in the stress response of

C. glutamicum comes from medical research. C. glutamicum as a nonhazardous organism is

considered an ideal model to understand the stress response of closely related pathogenic

species such as Corynebacterium diphtheriae, Mycobacterium leprae, and Mycobacterium

tuberculosis. Physiological observations of Corynebacterium glutamicum on lactate under

standard batch and continuous conditions have been reported by Cocaign et al. [1993],

Cocaign and Lindley [1995], and Gourdon et al. [2000]. Studies of L- and D-lactate

metabolism and gene expression have been mainly aimed at lactate dehydrogenase,

anaplerotic reactions, co-utilization of lactate and other carbon sources, and pyruvate

accumulation [Cocaign et al. 1993, Gourdon et al. 2000, Koffas et al. 2003, and Stansen et al.

2005].

Page 26: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 8

2.2 Materials and methods

2.2.1 Microorganism, media and culture conditions

All experiments were carried out with the wild type of Corynebacterium glutamicum

ATCC 13032 [Appendix A.1.1, Kinoshita et al. 1957]. Medium composition and preparation

of the precultures are described in appendix A.1.2.

Cultures were cultivated on an orbital shaker (Lab-Shaker LS-W, Adolf Kühner AG,

Birsfelden, Switzerland) with a filling volume of 10 ml (oxygen unlimited), 25 ml, 30 ml or

60 ml (oxygen limited conditions), with a 50 mm shaking diameter and a shaking frequency

of 300 rpm. Cultivations were carried out in unbaffled 250 ml shake flasks with cotton plugs

or in 250 ml measuring flasks of the RAMOS device [Appendix A.3.2, Fig. A.2, Anderlei et

al. 2004, Anderlei and Büchs 2001]. The maximum oxygen transfer capacity (OTCmax) of the

different culture conditions was estimated according to the scale-up approach presented in

chapter 4. A single inoculated medium was prepared for each experiment. This inoculated

medium was then distributed to each shake flask. Only the flasks with different initial lactic

acid concentrations and the flasks with MOPS buffer were inoculated separately. None of the

experiments required a change of measuring device, measuring vessel or the medium during

the experiment, thus, avoiding metabolic activity losses produced by transferring the

microorganisms to e.g. respirometers or plates.

The final biomass concentration was determined as described in appendix A.5.1-. The

biomass concentration of samples taken during the experiment was calculated using an optical

density (OD600) dry weight correlation. Titration curves were recorded and normalized as

described in appendix A.1.4. The OTR of the shake flask cultures was determined with the

RAMOS device as described in appendix A.3.2.

Page 27: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 9

If during the cultivation the shaker had to be stopped, e.g. to take samples, the necessary

manipulation time was recorded. The manipulation time includes the sampling time for all 8

RAMOS flasks or 10 standard flasks and the time, which the shaker needs to decelerate and

reaccelerate to the set shaking frequency.

2.2.2 Stress induced by carbon starvation

To investigate the influence of carbon starvation on the metabolic activity of C. glutamicum,

the microorganisms were grown in batch fermentations with 10 g/l lactic acid and then

starved. The metabolic activity was tested by adding a lactic acid pulse after 28 h. After the

pulse the medium contained 5 g/l lactic acid. At the time of the pulse, the OTR of all cultures

was below 0.001 mol/l/h. The pulse was added to the cultures during the rinsing phase of the

RAMOS device 10 min prior to the start of the next measuring phase. Oxygen unlimited and

limited culture conditions were set by using filling volumes of 10 ml, 30 ml and 60 ml. The

initial pH was 7. The only manipulation necessary during the experiment was to stop the

shaker for ~10 min to add the lactic acid pulse.

2.2.3 Stress induced by anaerobic conditions

The behavior of C. glutamicum on lactic acid under anaerobic conditions was investigated by

comparing standard batch cultures (aerated with air) with batch cultures aerated twice for one

hour with nitrogen. After starting the nitrogen aeration, it was verified with the oxygen sensor

of the RAMOS device that the oxygen concentration in the flasks reached zero in less than

5 min. The anaerobic conditions were terminated by flushing the flasks with compressed air.

Standard oxygen concentrations were reached again in less than 5 min. The nitrogen aeration

was connected to the flask during the rinsing phase 10 min prior to the start of the next

measuring phase and disconnected directly after the previous measuring phase. Oxygen

unlimited and oxygen limited (not anaerobic) culture conditions were set by using filling

Page 28: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 10

volumes of 10 ml, 30 ml and 60 ml. The lactic acid concentration was 10 g/l and the initial

pH 7. The only manipulation necessary during the experiment was stopping the shaker for

~5 min to connect and disconnect the nitrogen aeration.

The influence of delayed agitation and aeration after inoculation on the growth of

C. glutamicum was investigated by comparing cultures, which were directly shaken after

inoculation with cultures, which were not shaken after inoculation for 1 h, 2 h, and 3 h. To

assure equal culture conditions of shaken and non shaken cultures the non shaken flasks were

placed directly next to the shaker in the same thermo-constant room. The filling volume was

25 ml, the lactic acid concentration 10 g/l, and the initial pH 7. The only manipulation

necessary during the experiments was stopping the shaker for ~8 min to connect the unshaken

flasks to the RAMOS device.

2.2.4 Stress induced by lactic acid

The stress induced by lactic acid was investigated by measuring the influence of different

initial lactic acid concentrations 1 g/l (0.51 osmol/l), 2.1 g/l (0.53 osmol/l), 4.7 g/l

(0.55 osmol/l), 9.6 g/l (0.72 osmol/l), 12.4 g/l (0.78 osmol/l), 13.7 g/l (0.81 osmol/l), 18.9 g/l

(0.93 osmol/l), 29.5 g/l (1.16 osmol/l) on respiration and growth of C. glutamicum. The

numbers in brackets are the calculated osmolarities of the entire medium. The lactic acid

concentrations were determined with HPLC chromatography as described in appendix A.5.2.

To avoid oxygen limitation even at higher lactic acid concentrations, the RAMOS device was

operated in the one flow mode with a filling volume of 10 ml. After terminating the

experiment, the final lactic acid concentrations (HPLC) (Chapter A.5.2), the final biomass

concentrations (dry weight, OD) (Chapter A.5.1) and the final pH were determined. During

the fermentation no manipulation was necessary.

Page 29: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 11

2.2.5 Stress induced by osmolarity and pH

The stress induced by osmolarity and pH was investigated by comparing the respiration of

two oxygen unlimited batch cultures, one without buffering and the other with a MOPS buffer

concentration of 42 g/l (0.2 M). The medium without buffer had an osmolarity of

0.73 osmol/l, which was increased by adding the buffer to 1 osmol/l. The filling volume was

10 ml, the lactic acid concentration 10 g/l, and the initial pH 7. The RAMOS device was

operated in the two phase mode. During the fermentation, no manipulation was necessary.

The growth characteristics of C. glutamicum at different initial pH was investigated by

comparing respiration and biomass formation of unbuffered batch cultures with different

initial pH 4, 4.5, 5, 5.5, 7.1, 7.4, 8.5, 9, 9.5. The biomass and the pH measurements were

performed in separate shake flasks with cotton plugs so as not to interrupt the oxygen transfer

rate measurements. The filling volume was 10 ml and the lactic acid concentration 10 g/l. The

RAMOS device was operated in the two phase mode and stopped only twice for 10 min to

take control samples. The shaker with the standard shake flasks was stopped 14 times for

~10 min to take samples (sample size 0.15 ml).

2.3 Results

The applicability of respiration measurements in shake flasks to analyze the behavior of

microorganisms during and after stress exposure was tested with the model system

Corynebacterium glutamicum grown on the carbon source lactic acid. The organism was

exposed to the stress factors, carbon starvation, anaerobic conditions, lactic acid, osmolarity,

and pH. Representative results are depicted in Fig. 2.1-2.7. All experiments were performed at

least twice and as minimum triple evidence was considered necessary to draw a conclusion.

Page 30: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 12

2.3.1 Stress induced by carbon starvation

The experiment was performed in three steps: first the microorganisms were raised in a

standard batch fermentation, second they were exclusively carbon starved (oxygen was

continuously supplied) for 12-19 h depending on the culture conditions, third the metabolic

activity of the microorganisms was tested by adding a lactic acid pulse. Fig. 2.1 depicts the

oxygen transfer rate (OTR) over fermentation time of the entire experiment. The OTR curves

reflect the experimental procedure. Initially, after inoculation, all cultures grew exponentially.

The exponential growth continued until it was impaired by the maximum oxygen transfer

capacity (OTCmax). The OTCmax depends only on the culture conditions (filling volume 10 ml

oxygen unlimited growth, 30 ml and 60 ml oxygen limited growth). After reaching OTCmax,

the OTR stayed at a plateau until the carbon source was exhausted and then dropped sharply.

With decreasing OTCmax, the fermentation time increased. Then, after the drop, the OTR

remained close to zero and the microorganisms were starving. Finally, after lactic acid was

added, the OTR of the oxygen limited cultures jumped to the level of the plateau before

starvation. The OTR of the oxygen unlimited culture exceeded even the maximum OTR

reached in the initial batch, indicating the growth of the microorganisms after the pulse was

added. After the renewed exhaustion of the carbon source the OTR dropped again. In

conclusion, a carbon starvation of up to 19 h does not influence the metabolic activity of

C. glutamicum.

Page 31: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 13

0 5 10 15 20 25 30 35 40 450.00

0.01

0.02

0.03

0.04

0.05

pulse 5 g/llactate

Oxy

gen

trans

fer r

ate

[m

ol/l/

h]

Fermentation time [h]

Fig. 2.1: Influence of carbon starvation

on the metabolic activity of

Corynebacterium glutamicum. After an

initial batch fermentation with a lactic

acid concentration of 10 g/l, a pulse of

5 g/l lactic acid was added after 28 h.

Culture conditions: 250 ml shake flask,

filling volume before the pulse 10 ml, 30 ml, 60 ml, shaking diameter 50 mm, shaking

frequency 300 rpm, initial pH 7.

2.3.2 Stress induced by anaerobic conditions

The stress tolerance of C. glutamicum under anaerobic conditions was analyzed by comparing

constantly aerobic batch cultures with temporarily anaerobic batch cultures. Fig. 2.2a (values

are identical with the initial batch of Fig. 2.1) depicts the OTR of the constantly aerobic batch

cultures over fermentation time. Fig. 2.2c depicts the OTR of the twice nitrogen aerated batch

cultures over fermentation time. The general culture characteristics were similar to the

constantly aerobic batch. However, the fermentation time was prolonged and no growth

occurred during the anaerobic periods. In a separate experiment (results not shown) it could

be shown that the optical density of the culture also didn't increase during anaerobic

conditions. A more detailed insight in the growth behavior of C. glutamicum after anaerobic

stress is given in Fig 2.2b (oxygen unlimited growth conditions, filling volume 10 ml) and

Fig 2.2d (oxygen limited growth conditions, filling volume 30 ml and 60 ml). In both figures,

the time of anaerobic fermentation (twice 1 h) is cut out from the data set shown in Fig. 2.2c

to allow for a better comparison of the OTR curves. The cumulative consumed oxygen

(integral of the OTR curve) (Chapter 3.3.2) was identical for oxygen limited and not limited

conditions, providing an oxygen substrate yield (YO2/S) of 0.017 mol/g with a standard

Page 32: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 14

deviation of 0.0016 mol/g. The slope of the OTR and the whole course of the culture were not

affected by the anaerobic conditions, proving that the metabolic activity of C. glutamicum is

not affected by short periods of anaerobic conditions.

0.00

0.01

0.02

0.03

0.04

0.05B

DC

A

Oxy

gen

trans

fer r

ate

(OTR

) [m

ol/l/

h]

0 2 4 6 8 10 12 14 16 18 20 220.00

0.01

0.02

0.03

0.04 aeration withnitogen

Fermentation time [h]0 2 4 6 8 10 12 14 16 18 20 22 24

Fig. 2.2: Influence of anaerobic conditions on the metabolic activity of Corynebacterium glutamicum.

A. , , Continuous aeration with air. B. Comparison of oxygen unlimited cultures with and without

nitrogen aeration. C. , , After 4 h and 7 h aeration for 1 h with nitrogen. D. Comparison of oxygen

limited cultures with and without nitrogen aeration. A., C. Actual fermentation time. B., D. Nitrogen

aeration time subtracted from actual fermentation time. Culture conditions: 250 ml shake flask, filling

volume , 10 ml; , 30 ml; , 60 ml, shaking diameter 50 mm, shaking frequency 300 rpm,

lactic acid concentration 10 g/l, initial pH 7.

Similar observations were made with a second set of experiments. Anaerobic conditions were

produced by delaying the shaking of the cultures after inoculation. Fig. 2.3 depicts the OTR

over time after inoculation. The general course of the OTR curves was independent of the

shaking conditions reflecting the tolerance of C. glutamicum to anaerobic conditions.

However, the exponential growth phases and the exhaustion of the carbon sources were

Page 33: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 15

shifted for 1 h, 2 h and 3 h reflecting the delay of aeration after inoculation. If the OTR curves

of Fig. 2.3 are superposed by cutting out the non shaken periods, the OTR curves are nearly

identical (Figure not shown). In conclusion, during short anaerobic periods Corynebacterium

glutamicum falls in a state of dormancy. Similar to sleep, dormancy is defined as a state

where the microorganisms have no or only very reduced metabolic activity without growth,

however, the ability to be metabolically active and grow is not significantly reduced.

0 2 4 6 8 10 12 14 16 18 200.00

0.01

0.02

0.03

0.04

Oxy

gen

trans

fer r

ate

[m

ol/l/

h]

Time after inoculation [h]

Fig. 2.3: Influence of a delayed aeration on the

culture characteristics of C. glutamicum. After

inoculation the cultures were directly shaken, X

shaken after 1 h, shaken after 2 h, + shaken

after 3 h. Culture conditions: 250 ml shake flask,

filling volume 25 ml, shaking diameter 50 mm,

shaking frequency 300 rpm, lactic acid

concentration 10 g/l, initial pH 7.

2.3.3 Stress induced by lactic acid

The stress induced by varying acid concentrations (Fig. 2.4) was investigated by comparing

the oxygen transfer rate (OTR) of batch cultures with initial lactic acid concentrations ranging

from 1 g/l (0.01 mol/l) to 29.5 g/l (0.33 mol/l) at a constant initial pH of 7. Because the pH

was always above 7 (Fig. 2.4d), lactic acid (pKa 3.86) was entirely present in the dissociated

form. Up to a concentration of 10 g/l (Fig. 2.4a) the lactic acid concentration did not influence

the slope of the OTR, at higher concentrations (Fig. 2.4c) the slope decreased due to lactic

acid inhibition.

In aerobic cultures almost every physiological activity is reflected by the OTR [Anderlei et al.

2004, Anderlei and Büchs 2001, Losen et al. 2004, Stöckmann et al. 2003b]. Thus, in most

Page 34: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 16

cases it is possible to derive the growth rate of a biological culture from the slope of the OTR

as described in chapter 3.31. Using the slopes of the OTR curves depicted in Fig. 2.4a, c it

was possible to identify three growth phases. At fermentation times of less than 2.5 h

C. glutamicum showed an acceleration phase. As depicted in Fig. 2.5 at fermentation times

between 2.5 h and 6 h, pH values below 8, and initial lactic acid concentrations of less than

10 g/l the microorganisms grew with the maximum growth rate of 0.3 h-1-0.32 h-1. At higher

initial lactic acid concentrations (iC3H6O3 [g/l]) the growth rate (µ [h-1]) decreased linearly

with increasing initial lactic acid concentration [µ=0.38-0.0084·iC3H6O6, R²=0.92,

10 g/l<iC3H6O6<30 g/l]. During the late batch phase, when the lactic acid concentration was

low, and the pH value above 8, the growth rate (µ=0.235 h-1, standard deviation σ=0.016 h-1,

iC3H6O3>10 g/l) was independent of the initial lactic acid concentration.

Fig. 2.4b depicts the biomass substrate yield coefficient (YX/S) and the cumulative consumed

oxygen (Chapter 3.3.2) over the initial lactic acid concentration. The biomass substrate yield

decreased with increasing lactic acid concentration. The cumulative consumed oxygen,

however, increased linearly with the initial lactic acid concentration giving a constant oxygen

substrate yield (YO2/S=0.017 mol/g, standard deviation σ=0.0015 mol/g).

The consumption of the acid lactic acid leads to an increase of the pH value. Fig. 2.4d depicts

the final pH (symbols) reached by cultures with different initial lactic acid concentrations.

The measured increase of the pH values depends on the initial lactic acid concentration and on

the buffering capacity of the medium (Chapter 4.4.2). As described in chapter 4.3.2 titration

curves of the medium can be used as a simple tool to describe the final pH of biological

cultures and to identify the buffering effect of the different medium components.

Page 35: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 17

0 2 4 6 8 10 12 140.00

0.01

0.02

0.03

0.04

0.05

0.06

D

B

C

A

Fermentation time [h]

Oxy

gen

trans

fer r

ate

(OTR

) [m

ol/l/

h]

5 10 15 20 25 300.0

0.1

0.2

0.3

0.4

0.5

0.6

Inital lactic acid concentration [g/l]

Biom

ass

yiel

d y X/

S [g

/g]

Cum

ulat

ive

OTR

[m

ol/l]

0 4 8 12 16 20 24 280.00

0.02

0.04

0.06

0.08

0.10

0.12

Fermentation time [h]0 5 10 15 20 25 30

7.0

7.5

8.0

8.5

9.0

9.5

Fina

l pH,

pH

[-]

Initial lactic acid concentration [g/l]

0 5 10 15 20

0.5 M NaOH [ml]

Fig. 2.4: Influence of different initial lactic acid concentrations on respiration and growth of

Corynebacterium glutamicum. A. Oxygen transfer rate (OTR) over fermentation time, low initial lactic

acid concentrations▼ 1 g/l 2.1 g/l, ▲ 4.7 g/l, ● 9.6 g/l, ■ 12.4 g/l. B. + cumulative consumed oxygen

over initial lactic acid concentration, biomass substrate yield over initial lactic acid concentration.

C. OTR over fermentation time, high initial lactic acid concentrations ■ 12.4 g/l, 13.7 g/l, 18.9 g/l,

29.5 g/l. D. X final pH over initial lactic acid concentration. Titration curves normalized to pH 7 as

described in appendix A.1.4 over added amount titrant NaOH, —— standard medium, ········· medium

without lactic acid, - - - - medium without ammonium sulfate. Culture conditions: 250 ml shake flask,

filling volume 10 ml, shaking diameter 50 mm, shaking frequency 300 rpm, initial pH 7.

Fig. 2.4d depicts the titration curves (lines) of the standard medium (appendix A.1.2), of a

medium without lactic acid, and of a medium without (NH4)2SO4. The titration curves of the

media with and without lactic acid are nearly identical, affirming that the consumption of

lactic acid (pKa 3.86) can not influence the buffering capacity of the medium at pH above 7.

The titration curve of the medium without (NH4)2SO4 is much steeper, showing the buffering

effect of ammonium sulfate (pKa 9.25). Titration curves of a medium without phosphates

Page 36: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 18

(results not shown) demonstrated, that the phosphates (pKa 7.2) have only a minor influence,

proving, that the excess amount of ammonium present in the medium is responsible for the

relatively moderate pH increase at high initial lactate concentrations.

0 5 10 15 20 25 300.1

0.2

0.3

0.4

Initi

al µ

max

[h

-1]

Initial lactic acid concentration [g/l]

Fig. 2.5: Initial maximum growth rate (µmax) at

different initial lactic acid concentrations. The initial

µmax were derived from the slopes of the oxygen

transfer rate depicted in Fig. 2.4a, c. OTR values

with a fermentation time between 2.5 h and 6 h

were used to calculate the initial µmax.

In conclusion, initial lactic acid concentrations up to 10 g/l do not influence the growth of

C. glutamicum. At higher concentrations up to 30 g/l, the growth rate decreases linearly with

the lactic acid concentration. With increasing lactic acid concentration the biomass substrate

yield decreases, without affecting the oxygen substrate yield. The final pH reached by batch

cultures with different initial lactic acid concentrations followed the titration curve of the

medium.

2.3.4 Stress induced by osmolarity and pH

Fig. 2.6 depicts the OTR over fermentation time of oxygen unlimited, unbuffered and

buffered batch cultures. Both cultures had an initial pH of 7. The pH of the unbuffered culture

increased to 8.4, that of the buffered culture to 7.6. Up to a fermentation time of 8 h both

cultures had the same slope, demonstrating that the increased osmolarity does not influence

the growth of C. glutamicum. At fermentation times exceeding 8 h the growth rate of the

unbuffered culture decreased. The buffered culture continued to grow with the maximum

growth rate (see also the comparison between the Figs. 3.2a, b and Fig. 3.2c).

Page 37: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 19

20 2 4 6 8 10 10.00

0.01

0.02

0.03

0.04

0.05

0.06O

xyge

n tra

nsfe

r rat

e [

mol

/l/h]

Fermentation time [h]

Fig. 2.6: Influence of buffer on the growth of

Corynebacterium glutamicum. medium

without buffer, medium with 42 g/l MOPS

buffer. Culture conditions: 250 ml shake flask,

filling volume 10 ml, shaking diameter 50 mm,

shaking frequency 300 rpm, lactic acid

concentration 10 g/l, initial pH 7, final pH

unbuffered 8.4, buffered 7.6.

The effect of different initial pH on growth and respiration of C. glutamicum was investigated

by comparing OTR and growth curves of unbuffered batch cultures with initial pH ranging

from 4 to 9.5. Fig. 2.7 depicts the OTR (Fig. 2.7a), the biomass concentration (Fig. 2.7b) and

the pH (Fig. 2.7c) over fermentation time of cultures with different initial pH. The response of

C. glutamicum to different initial pH values can be classified in three groups. The first group

includes the initial pH values of 7.1 (6.4, 7.4 results not shown). The OTR and growth curves

of this group were nearly identical, they increased exponentially. The pH changed from its

initial value to 8.5. The second group showed a reduced metabolic activity and includes the

initial pH values of 5.5 and 8.5. Due to pH inhibition, the cultures of this group showed a

prolonged fermentation time and a reduced growth rate. The culture with an initial pH of 5.5

had a reduced OTR and growth rate at the beginning. However, with increasing pH the slope

became equal to the first group. In contrast, the OTR curve with an initial pH of 8.5 had an

equal OTR at the beginning, the slope, however, stayed low for the entire fermentation time.

The third group includes the initial pH values of 5, 9, (4, 4.5, 9.5 results not shown). None of

the cultures of this group showed any considerable metabolic activity. The OTRs remained

close to zero, the biomass concentrations did not increase or increased only slightly, and the

pH values stayed at their initial values. However, from the final pH values reached by the

cultures with a high initial lactic acid concentration (Fig. 2.4d) it can be concluded, that C.

Page 38: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 20

glutamicum can grow at high growth rates, at pH up to 9 if the microorganisms have time to

adapt. Small decreases of the pH at high initial pH (Fig. 2.7c) are probably caused by the

gassing out of NH3 at pH close to the pKa (9.25). The OTR curves (Fig. 2.7a) reached the

stationary phase earlier than the biomass (Fig. 2.7b) and pH (Fig. 2.7c) curves. During the

respiration measurements the shaker was stopped only twice for 10 min to take samples,

however, the shaker for the biomass and pH determination was stopped 14 times for 10 min to

take samples. If the interruptions of the shaker are subtracted from the fermentation time

(Fig. 2.7b, c only demonstrated for the culture with a pH value of 5.5), the fermentation time

of the OTR curves (Fig. 2.7a) match the fermentation time of the biomass and pH curves

(Fig. 2.7b, c). Probably stopping the shaker to take samples leads to short anaerobic periods,

which freeze growth and metabolic activity. The observed influence of the interruptions of the

shaker on the fermentation time confirms the conclusions drawn before about the stress

induced by anaerobic conditions. It can be concluded that the increased osmolarity of up to

1 osmol/l caused by the buffer does not effect the growth or respiration of C. glutamicum.

Initial pH ranging from 6.4 to 7.4 lead to uninhibited growth. Without adaptation the

organism did not grow at pH<5 or pH>9, but with adaptation growth was observed up to a pH

of 9.

Page 39: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 21

0 5 10 15 20 255

6

7

8

Fermentation time [h]

pH

[-]

0.000.010.020.03

0.040.050.06

C

B

AO

TR

[mol

/l/h]

0.00.51.01.52.02.53.03.5

Biom

ass

conc

entra

tion

[g/

l]Fig. 2.7: Influence of different initial pH 5,

5.5, 7.1, 8.5, 9 on respiration and

growth of C. glutamicum. A. Oxygen transfer

rate (OTR), B. Biomass concentration, C.

pH, B., C. + initial pH 5.5 the time of

interruptions of the shaker are subtracted

from the fermentation time. Culture

conditions: 250 ml shake flask, filling volume

10 ml, shaking diameter 50 mm, shaking

frequency 300 rpm, lactic acid concentration

10 g/l.

2.4 Discussion

2.4.1 Stress induced by carbon starvation

With respiration measurement in shake flasks, it was shown that a starvation time of 19 h

does not reduce the metabolic activity of C. glutamicum grown on lactic acid. The starvation

time used in this study can be sufficient to reduce the viability of gram-positive organisms.

E.g. Ferreira et al. [2001] observed for Listeria monocytogenes (WT) a 25 % decrease in

viability over a starvation time of 12 h. Other gram-positive organisms like Mycobacterium

tubeculosis show a high resistance to carbon starvation. With plate counts, Primm et al.

[2000] found no loss in viability of M. tuberculosis for 106 days. The high resistance against

carbon starvation of C. glutamicum observed in this study probably result from the complex

starvation-survival strategies of C. glutamicum [Wehmeier et al. 2001, 1998], which are

similar to the ones of M. tuberculosis [Primm et al. 2000].

Page 40: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 22

2.4.2 Stress induced by anaerobic conditions

In this study it was found that during short anaerobic periods Corynebacterium glutamicum

on lactic acid falls in a state of dormancy, without growth and without a loss in metabolic

activity. The behavior of C. glutamicum under anaerobic conditions on lactic acid has so far

not been investigated. On glucose Inui et al. [2004] observed no growth under oxygen

deprived conditions, however, a continuous formation of lactate and succinate. For

M. tuberculosis on complex medium Primm et al. [2000] found only a slight decrease in

viability over a 23 week anaerobic period. Kim et al. [2005] describe genes involved in

oxidative stress protection of C. glutamicum.

2.4.3 Stress induced by lactic acid

In this study, it was found that lactic acid present in the dissociated form (Fig. 2.4, 2.5) had no

influence on the growth of C. glutamicum up to an initial concentration of 10 g/l (0.11 M). At

higher concentrations of up to 30 g/l (0.33 M), the growth rate decreased linearly with the

lactic acid concentration. Cocaign et al. [1993] observed for C. glutamicum no influence of

initial lactate concentrations on the growth rate between 0.035 M and 0.125 M. According to

Houtsma et al. [1996] the dissociated form of lactate leads to a linear decrease of the growth

rate of Listeria innocua at concentrations between 0.25 M and 1.1 M. Above 1.25 M the

growth of L. innocua is completely inhibited. According to the review of Presser et al. [1997]

growth of various organisms is completely inhibited by the dissociated form of organic acids

at concentrations between 0.1 M and 0.8 M.

In this study, it was observed that with increasing lactic acid concentration the biomass

substrate yield (YX/S) decreases. For C. glutamicum decreasing growth rates combined with

decreasing biomass substrate yields under stress conditions were observed by Guillouet and

Engasser [1995]. YX/S between 0.36 g/g and 0.62 g/g depending on the growth rate and mode

Page 41: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 23

of operation were observed by Cocaign-Bousquet and Lindley [1995] and Cocaign et al.

[1993], however, these were not attributed to stress conditions. The oxygen substrate yield

(YO2/S) of 0.017 mol/g observed in this study is identical to the one reported by Cocaign-

Bousquet and Lindley [1995] (YO2/S was calculated from the data depicted in the figures), for

continuous cultures under conditions, where the substrate is entirely consumed. Why the

biomass substrate yield decreases with increasing lactic acid concentration, while the oxygen

substrate yield stays constant, could not yet be explained. However, it may be caused by the

formation of an undetected byproduct, or by a significant change in biomass composition.

2.4.4 Stress induced by osmolarity

In this study, an osmolarity of 1 osmol/l had no influence on the growth of C. glutamicum.

This agrees with Rönsch [2000], who found the same results for osmolalities up to

2.0 osmol/kg for C. glutamicum on sucrose, but noticed a decreased growth rate at

2.5 osmol/kg. By increasing the osmolality from 0.4 osmol/kg to 2 osmol/kg for

C. glutamicum on glucose, Guillouet and Engasser [1995] observed a reduction of the growth

rate from 0.7 h-1 to 0.2 h-1 and a reduction of the biomass substrate yield coefficient from

0.6 g/g to 0.3 g/g. The high tolerance of C. glutamicum against osmotic stress makes it

unlikely, that the observed reduction of the growth rate with increased initial lactic acid

concentration (10 g/l 0.73 osmol/l, 30 g/l 1.17 osmol/l) is caused by the increased osmolarity.

The high tolerance of C. glutamicum against osmotic stress is caused by the possibilities of

the organism to sense osmolarity and to take up and produce compatible solutes for protection

[Guillouet and Engasser 1995, 1996, Peter et al. 1998, Skjerdal et al. 1996, Varela et al. 2003,

Wolf et al. 2003].

Page 42: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 24

2.4.5 Stress induced by pH

In this study, initial pH ranging from 6.4 to 7.4 leats to uninhibited growth of C. glutamicum.

Without adaptation, the organism did not grow at pH<5 or pH>9. With adaptation however,

growth was observed up to a pH of 9. Similar results were reported by Shah et al. [2002], who

found lysine production of C. glutamicum from pH 6-9 with a maximum at pH 7.5. Bröer and

Krämer [1991] analyzed the influence of external pH on lysine efflux and found a maximum

at pH 7.8. The importance of adaptation for acid tolerance of gram-positive organisms has

been reported [Bodmer et al. 2000, Cotter and Hill 2003, O’Driscoll et al. 1996, Saklani-

Jusforgues et al. 2000].

With the chosen measuring set-up it is not possible to distinguish whether the observed

growth inhibition of C. glutamicum at low pH is caused by the pH itself or by the

undissociated form of the lactic acid. The undissociated form of lactic acid is believed to have

a higher inhibitory effect [Cotter and Hill 2003, Salmond et al. 1984, Shelef 1994] than the

dissociated form. At a pH of 5.5 about 0.0025 M and at a pH of 5 about 0.0075 M of the lactic

acid (total concentration 0.11 M) is present in the undissociated form. Presser et al. [1997,

1998] observed for Escherichia coli a complete inhibition of growth at undissociated lactate

concentrations exceeding 0.0083 M regardless of the pH or the total lactate concentration.

Houtsma et al. [1996] observed a total inhibition of Listeria innocua at pH 5.5 with a lactate

concentration of 0.2 M of which 0.0045 M were undissociated.

2.4.6 Evaluation of the general method

The broad applicability of respiration measurements in shake flasks to study the behavior of

C. glutamicum during and after exposure to stress was demonstrated. Experimental

procedures were developed to study the following stress factors: carbon starvation, anaerobic

conditions, organic acids, osmolarity and pH. Similar procedures may be applied to study the

Page 43: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 2 Characterization of the biological model system Corynebacterium glutamicum 25

responds of microorganisms to stress factors like heat, oxygen, antibiotics, hormones,

antibodies, or starvation of nitrogen, phosphorus, amino acids, or trace elements. The

importance to run stress response analysis without interferences during the cultivation was

demonstrated by showing the influence, which inevitable interuptions of the shaker had when

samples were taken. The short periods without agitation, necessary to transfer samples to new

culture vessels or measuring devices, like respirometers, may lead to detrimental anaerobic

conditions. Further, if evidence exists that the microorganisms show a different growth

behavior in liquid cultures than on plates, respiration measurement in shake flasks is

recommended.

The duration of the measuring cycle of the RAMOS device may limit the applicability of the

method. The standard duration of the measuring cycle is 30 min, which can be reduced for

strongly respiring cultures to 10 min. This is sufficient for most stress response analyses

because activity losses after exposure to stress mainly occur within hours or days. The new

possibility to study the stress response of microorganisms in online monitored shake flasks

makes it possible to study several cultures simultaneously, and increases the possible number

of experiments. The high throughput combined with the possibility to perform the entire

experiment in the same culture vessel and in the same medium, makes the respiration

measurement in shake flasks an important tool for research and industrial process

development.

Page 44: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 3 Evaluation of respiration measurement techniques 26

Chapter 3

Evaluation of respiration measurement techniques

To successfully perform the scale-up from shake flask to fermenter it is essential that the

measuring techniques applied for shake flask cultures and the measuring techniques applied

for fermenter cultures produce comparable results. The focus of this chapter is to compare and

evaluate three different respiration measurement techniques: exhaust gas analyzer, RAMOS

device, and respirometer. Most of the results presented in this chapter have been described by

Seletzky et al. [2006c].

3.1 Introduction

In aerobic cultures, almost every physiological activity is coupled to the respiratory uptake of

oxygen, making the oxygen transfer rate (OTR) a valuable parameter to monitor the metabolic

activity of biological cultures [Anderlei and Büchs 2001, Stöckmann et al. 2003b]. The OTR

reflects the physiological responses of the microorganisms to different culture conditions such

as temperature, pH, osmotic stress, nutrient limitation and inhibition, product or by-product

formation and inhibition (see also chapter 2). Furthermore, it can be used to guide sampling

times or the induction time for gene expression.

The comparison and evaluation of different respiration measuring techniques is rendered

difficult by the vast variety of devices and analytical procedures commonly in use. The

respiration can be measured in reaction vessels as different as fermenters or shake flasks with

devices as different as exhaust gas analyzers or respirometers and with sensors as diverse as

Page 45: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 3 Evaluation of respiration measurement techniques 27

magneto-mechanical, electro-chemical, or optical. This study introduces a method to compare

and evaluate different respiration measurement devices and analytical procedures considering

accuracy, precision, quantitation limit, and range. It focuses on the evaluation of the

measuring techniques, their advantages and limitations and their possible areas of application.

The respiration was measured with three commercially available devices: An exhaust gas

analyzer (EGA) coupled to a fermenter, a RAMOS (Respiration Activity Monitoring System)

device, which is a novel technique to monitor the respiration in shake flasks online, and a

respirometer.

Measuring the respiration of bioreactors with exhaust gas analyzers (EGA) has been state of

the art for years. They are an important tool for process optimization in research and industrial

process control. Shake flasks are inexpensive and can be operated simultaneously in large

numbers. In large industrial companies, up to several hundred thousand individual

experiments may be performed in shake flasks every year [Büchs 2001]. Nevertheless, most

experiments are still conducted without any online monitoring, hampering focused screening

and complicating scale-up by neglecting the effect of oxygen supply and fermentation time on

growth and product formation. Only in recent years has the monitoring of the OTR in shake

flasks [Anderlei and Büchs 2001, Anderlei et al. 2004, Raval et al. 2003, Silberbach et al.

2003, Stöckmann et al. 2003a, b, Wittmann et al. 2003] become more common. A

commercially available technique is the RAMOS device, which enables simultaneous online

monitoring of the OTR in several biological cultures under sterile conditions. Respirometry is

a wide spread technique to measure the oxygen uptake of various cell suspensions such as

bacteria [Boyles 1978, Ridgway 1977, Wittmann et al. 2001], microcrustacean [Montagnolli

et al. 2004], or macrophages [Frost et al. 2005]. It is mainly used to determine the effect of

different environmental conditions or of the addition of toxic substances or of growth factors

on the viability. The respiration is usually measured with electro-chemical oxygen electrodes

Page 46: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 3 Evaluation of respiration measurement techniques 28

[Lee and Tsao 1979], or more recently with optical sensors [Kohls and Scheper 2000,

Wittmann et al. 2003].

3.2 Materials and Methods

3.2.1 Microorganism and Cultivation

All experiments were carried out with the wild type of Corynebacterium glutamicum ATCC

13032 [Appendix A.1.1, Kinoshita et al. 1957]. The organism was cultivated with two

different media. A complex medium with glucose as carbon source and a defined minimal

medium with lactic acid as carbon source to avoid the formation of anaerobic or overflow

metabolites. The culture conditions and media are as described in appendix A.1.2.

3.2.2 Analytical procedures and culture conditions

Fermentations were carried out in a laboratory fermenter and in a 50 l fermenter as specified

in appendix A.2.1. The dissolved oxygen (DO) was maintained above 30 % by adjusting the

stirrer speed. In case of excessive foam formation the antifoam agent (Plurafac LF 1300,

BASF, Ludwigshafen, Germany) was added.

The OTR of shake flask cultures was measured with a RAMOS device (Hitec Zang,

Herzogenrath, Germany) as specified in appendix A.3.2. To avoid oxygen limited culture

conditions the operating conditions were selected with the scale-up strategy described in

chapter 4. All shaken cultures were cultivated on an orbital shaker as described in appendix

A.2.3.

A respirometer with an electro-chemical oxygen electrode (Rank Brothers, Cambridge,

England) was used to measure the OTR of complex medium samples each drawn from an

individual 250 ml standard shake flask with cotton plug or of minimal medium samples drawn

Page 47: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 3 Evaluation of respiration measurement techniques 29

from a fermenter. The measurement was performed as described in a appendix A.3.3. The

solubility of the media were calculated as described in appendix A.1.3.

3.3 Results and discussion

Biological experiments were conducted to evaluate the ability of the EGA, the RAMOS

device and the respirometer to determine the OTR. With a first set of experiments the

accuracy of each individual measuring technique was determined independently, using

defined minimal medium cultures. In a second set of experiments the measuring techniques

were compared using minimal and complex medium cultures. Finally, based on theoretical

considerations and literature data quantitation limit and range of the devices were compared.

All validation characteristics in this chapter agree with the definitions of the Q2B validation

of analytical procedures [ICH].

3.3.1 Accuracy

The accuracy of the measuring techniques was evaluated by determining the specific growth

rate (µ), which reflects the effect of environmental conditions on the activity of

microorganisms (Chapter 2.3). µ of an exponentially growing culture can be deduced from the

OTR using the slope of a regression function (Eq. 3.1), t0 being the start of the exponential

growth phase. If the culture growth is assumed to be ideally exponential, according to Eq. 3.1,

the coefficient of determination (r²) reflects the accuracy of the measuring set-up.

t0tt eOTROTR ⋅µ⋅= (3.1)

The specific growth rate of each experiment was determined by fitting the measuring data

with Eq. 3.1, using the least square method. For each fit the coefficient of determination (r²)

was calculated. Replicates were compared by calculating the average specific growth rate

(AVµ) and its coefficient of variation (CVµ).

Page 48: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 3 Evaluation of respiration measurement techniques 30

0.00

0.01

0.02

0.03

0.04

0.05

0.06

CVµ=2.2 %

CVµ=2.2 %

9.5 - 11 h

4 - 9.5 h

r2>0.99

r2>0.99

AVµ=0.21 h-1

AVµ=0.27 h-1

A

oxyg

en tr

ansf

er ra

te (O

TR)

[mol

/l/h]

0 2 4 6 8 10 12 140.00

0.01

0.02

0.03

0.04

0.05

r2>0.99

CVµ=6.5 %

AVµ=0.35 h-14 - 10.5 h

CVµ=6.7 %

B

fermentation time [h]

0.00

0.01

0.02

0.03

0.04

0.05

CVµ=2.9 %

9.5 - 11 h

4 - 9.5 h

r2>0.99

r2>0.99AVµ=0.23 h-1

AVµ=0.31 h-1

C

oxyg

en tr

ansf

er ra

te (O

TR)

[mol

/l/h]

Fig. 3.2: Accuracy of C. glutamicum minimal

medium fermenter and shake flask cultures with

exhaust gas analyzer (EGA) and RAMOS

device. A. laboratory fermenter with EGA, pH not

controlled , , . B. shake flasks with

RAMOS device, pH not controlled , , , +,

X, . C. laboratory fermenter with EGA, pH

controlled ; 50 l fermenter, pH controlled , .

Exponential fits ⎯, · · · ·. Culture conditions

laboratory fermenter: working volume 1 l, specific

aeration rate 2 vvm, Culture conditions 50 l

fermenter: working volume 15 l, specific aeration

rate 0.5 vvm. The dissolved oxygen (DO) of the

fermenter cultures was maintained above 30 %

by adjusting the stirrer speed. Culture conditions

shake flasks: unbaffled 250 ml measuring flasks,

shaking diameter 50 mm, shaking frequency

300 rpm, filling volume 10 ml.

Fig. 3.2a depicts the OTR over the fermentation time of independent not pH controlled

minimal medium laboratory fermenter cultures with EGA (3 replicates). Two different

exponential growth phases could be described. At fermentation times between 4 h and 9.5 h

the organism grew with an AVµ of 0.27 h-1 (solid line), which decreased to 0.21 h-1 (dotted

line) between 9.5 h and 11 h. Fig. 3.2b depicts independent not pH controlled minimal

medium shake flask cultures with RAMOS device (6 replicates). These cultures also show

two different exponential growth phases. The average growth rates of the shake flask cultures

are similar to the not pH controlled fermenter cultures (4-9.5 h, AVµ=0.31 h-1, solid line;

9.5-11 h, AVµ=0.23 h-1, doted line). Fig. 3.2c depicts the OTR over fermentation time of pH

Page 49: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 3 Evaluation of respiration measurement techniques 31

controlled minimal medium laboratory fermenter and 50 l fermenter cultures with EGA

(3 replicates). The pH controlled cultures showed only a single exponential growth phase. The

average growth rate of the pH controlled cultures (4 h-10.5 h, AVµ=0.35 h-1, solid line) was

higher than AVµ of the not pH controlled cultures. The different fermenter scales had no

influence on the culture characteristics. For all experiments depicted in Fig. 3.2 the coefficient

of variation of the average growth rate (CVµ) was below 7 % and the coefficient of

determination (r²) always above 0.99.

The comparison of the three measuring techniques is depicted in Fig. 3.3 (a. minimal medium,

b. complex medium). To set equal culture conditions in fermenters and shake flasks, the pH

was not controlled and a single inoculated medium was prepared, which was distributed to the

fermenters and shake flasks. With all three measuring techniques it was possible to observe

the general culture characteristics (Fig. 3.3). The slope of the OTR curves, the maximum OTR

and the cultivation time were comparable. The OTR curves of the minimal medium cultures

(Fig. 3a) recorded with EGA and RAMOS are very similar to the ones depicted in Fig. 3.2a,

b. With the average of the growth rates of Fig. 3.2a and Fig 3.2b (µ=0.29 h-1, dashed line,

µ=0.22 h-1, dotted line) the measuring data of both devices can be fitted with a r² higher than

0.99. The complex medium cultures (Fig. 3.3b) have a higher growth rate than the minimal

medium cultures.

Page 50: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 3 Evaluation of respiration measurement techniques 32

0 2 4 6 8 10 120.00

0.01

0.02

0.03

0.04

0.05

0.06

respirometer r2 not determinableEGA, RAMOS r2>0.99

EGA, RAMOS r2>0.99

respirometer r2>0.96

9.5 - 11 hµ=0.22 h-1

µ=0.29 h-1

4 - 9.5 hA

oxy

gen

tran

sfer

rate

[mol

/l/h]

fermentation time [h]0 2 4 6 8 10

EGA, RAMOS r2>0.97

respirometer r2>0.94

0 - 4.5 hµ=0.66 h-1

B

fermentation time [h]

Fig. 3.3: Comparison of different respiration measurement techniques. Oxygen transfer rate of

C. glutamicum over fermentation time. A. minimal medium with lactic acid , B. complex medium with

glucose. Fermenter with exhaust gas analyzer (EGA) ⎯; shake flask with RAMOS device ; fermenter

with respirometer ; shake flask with respirometer . Exponential fits - - -, · · · ·. Culture conditions

laboratory fermenter: working volume 1 l, specific aeration rate 2 vvm. Culture conditions shake flasks:

unbaffled 250 ml standard flasks, or RAMOS measuring flasks, shaking diameter 50 mm, shaking

frequency 300 rpm, filling volume 10 ml.

The measuring values of EGA and RAMOS device can be fitted with a µ of 0.66 h-1 (0-4.5 h,

dashed line), resulting for both devices in a r² higher than 0.97. The accuracy of the

respirometer was found to be noticeably lower than the accuracy of the EGA and the RAMOS

device. This observation did not depend on the type of bioreactor (fermenter, shake flask).

The minimal medium samples (Fig. 3.3a) were taken from a single fermenter culture (2

replicates per fermentation time). Each complex medium sample was taken from an individual

shake flask (2 replicates per fermentation time). With the respirometer it was not possible to

differentiate the two exponential growth phases on minimal medium. The lower accuracy is

reflected by the, in comparison to the EGA and the RAMOS device, lower coefficient of

determination (minimal medium 4-9.5 h r²=0.96, complex medium 0-4.5 h r²=0.94). The

specific growth rates observed in this study are in good agreement with the data of Cocaign et

al. [1993], and Cocaign-Bousquet et al. [1996], who found for Corynebacterium glutamicum

ATCC 17965 batch cultures a maximum growth rate of 0.35 h-1 on lactate and of 0.6 h-1 on

Page 51: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 3 Evaluation of respiration measurement techniques 33

glucose. In conclusion, the high accuracy of EGA and RAMOS device allows to describe the

respiration and culture characteristics of microorganisms in detail. Thus, both are methods to

be applied for scale-up purposes. With the respirometer it is possible to describe the general

culture characteristics of a biological culture. However, the accuracy is too low for a detailed

analysis of the culture characteristics.

3.3.2 Precision

The precision of the measuring devices was compared with the cumulative consumed oxygen

(cO2), which is independent of the biological kinetics, and depends on the amount of limiting

substrate. cO2 is calculated by integrating the OTR according to Eq. 3.2. The fermentation

time until OTRmax (tOTRmax) was used as upper boundary (see also chapter 6.3.1.2, Fig. 6.5).

∫=maxOTRt

02O dtOTR)t(c (3.2)

The cO2 values were derived from the data presented in Fig. 3.3. The use of a single

inoculated medium assured identical substrate concentrations for all measuring devices. cO2 of

the minimal media cultures was 0.22 mol/l with a deviation of less then 2 % between EGA

and RAMOS. The scatter of the respirometer did not allow a determination of cO2. The

complex media cultures had a cO2 of 0.11 mol/l with a deviation of less then 5 % EGA and

RAMOS device. The precision of EGA and RAMOS has been found to be equivalently high.

The measuring values of the respirometer are largely dependant on personal experience of the

operator resulting in a low precision and robustness.

3.3.3 Quantitation limit and range

EGA, RAMOS device and respirometer were compared considering the minimal (Vmin) and

maximal (Vmax) reaction volume, minimal OTRmin, and maximal OTCmax oxygen transfer

Page 52: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 3 Evaluation of respiration measurement techniques 34

capacity, and online measurement possibilities. All results are summarized in Tab. 3.1. The

measuring set-ups consist of three components: an oxygen measuring device (EGA, RAMOS,

respirometer), a fermentation vessel (fermenter, shake flask), and a fermentation environment

(eg. temperature and pH control, thermo-constant room). All three components and their

interaction influence the quantitation limit and range of the measuring set-ups, making it

difficult to precisely determine quantitation limits and ranges. Particularly, the characteristics

of stirred tank fermentations depend more on the fermenter size, which can range from 250 ml

to 500 m², and the mode of operation than on the EGA. The measurement of low OTR values

can be influenced by the quality of the temperature control. Considering these constrains the

values in Tab. 3.1 are only intended to give a general orientation.

3.3.3.1 Exhaust gas analyzer

The minimal volumetric flow of ~0.5 l/min to keep the magneto-mechanical oxygen sensor

working determines Vmin. A standard fermenter q (~2 vvm) therefore has a Vmin of ~0.25 l. It

can be reduced by changing to electro-chemical oxygen electrodes or to mass spectrometry.

Vmax is determined by the size of the fermenter. Power input and the size of the fermenter

determine OTCmax. A standard stirred tank fermenter has an OTCmax of 0.2-0.6 mol/l/h

[Enfors and Mattiasson 1983]. If the minimal specific aeration rate is considered to be

0.01 vvm, a standard value for cell cultures, and the minimal oxygen concentration difference

1 10-4 mol/mol, OTRmin is 1 10-4 mol/l/h. With multiplexing up to 5 fermenters can be

operated with one EGA if the measuring interval is increased to 30 min. Continuous online

monitoring is possible.

3.3.3.2 RAMOS (Respiration Activity Monitoring System)

To reduce the measuring error due to evaporation, Vmin should be higher than 5 ml. Vmax

depends on the oxygen requirements of the microorganisms and the size of the shake flask. A

bacteria culture in a 250 ml shake flask requires a Vmax of less then 25 ml [Maier and Büchs

Page 53: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 3 Evaluation of respiration measurement techniques 35

2001]. For low respiring cell cultures Vmin can reach up to 150 ml in 250 ml shake flasks.

OTRmin of the standard RAMOS device is ~1 10-3 mol/l/h and can be decreased to

1 10-4 mol/l/h by further reducing the influence of ambient conditions especially of

temperature fluctuations. OTCmax is determined by the culture conditions [Chapter 4.3.1,

Maier and Büchs 2001, Seletzky et al. 2007]. For a standard shaker (nmax=350 rpm) and a

reaction volume of 10 ml, OTCmax is ~0.065 mol/l/h and can be raised to 0.1 mol/l/h by

increasing the shaking frequency up to 500 rpm or using baffled shake flasks. The OTR is

measured intermittently with a measuring interval of 10-30 min.

3.3.3.3 Respirometer

The minimal sample size (~1 ml) needed to wet the oxygen electrode determines Vmin.

Smaller samples can be processed in respirometers with needle type electro-chemical or

optical oxygen sensors. The air tightness of the measuring chamber and the oxygen

consumption of the electrochemical oxygen electrode determine OTRmin. Theoretically, even

the respiration of single organisms can be measured. With an optical oxygen sensor we could

determine OTRs lower than 1 10-5 mol/l/h (data not shown). OTCmax is ~0.05 mol/l/h, at

higher respiration rates the oxygen consumption is faster than the mass transfer of the

aeration. The mass transfer of the culture vessel has no influence on the measurement,

impeding the detection of oxygen limitations due to insufficient culture conditions. Online

monitoring is not possible (manual sample injection).

Page 54: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 3 Evaluation of respiration measurement techniques 36

Tab. 3.1: Summary of

the different oxygen

transfer rate measuring

devices: analytical pro-

cedure, quantitation

limit, range, precision.

The values given in this

table are intended to

give a general orien-

tation. However, values

vary with the mea-

suring set-up and the

device used.

Name Exhaust gas analyzer

RAMOS Respirometer

reaction vessel fermenter shake flask any

common O2 sensor type

magneto-mechanical

electro-chemical electro-chemical

number of parallel fermentation vessels

1-(5) 6-12 (1)

online monitoring yes yes no

measuring interval continuous 10-30 min limited by manual handling

reaction volume [l] >0.25 0.005-0.1 for the device used 0.001-0.007

OTC, OURmax [mol/l/h] 0.2-0.6 0.08 0.05

OTR, OURmin [mol/l/h] 1 10-4 1 10-4 >1 10-5

3.3.4 Summary and overall conclusion

To design an industrial fermentation a large number of experiments, according to Büchs

[2001] in large companies up to several hundred thousand individual experiments per year,

are necessary, to select a strain, to improve the medium and to characterize the productivity at

different culture conditions. These screening tasks are normally performed in shake flasks or

laboratory fermenters. Because of their simplicity and inexpensiveness, according to

[Kennedy et al. 1994], the only practical way is to perform the major part of these

experiments in shake flasks. However, shake flasks lack the possibility to monitor the culture

during the experiment. This limits their application to simple standard tasks and may result in

unexpected scale-up problems. To overcome these limitations Anderlei et al. [2004], and

Page 55: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 3 Evaluation of respiration measurement techniques 37

Anderlei and Büchs [2001] introduced the RAMOS device, which allows the online

monitoring of the respiration of microbial cultures in shake flasks.

The focus of this chapter was to compare and evaluate three different respiration measurement

techniques exhaust gas analyzer, RAMOS device, and respirometer. By choosing an

appropriate biological model system it was possible to compare the different techniques

independent of their analytical procedure or the reaction vessel used for cultivation. Exhaust

gas analyzer and RAMOS device resulted in very similar culture characteristics. Accuracy,

and precision of both devices were high. Respirometry is a simple and cost efficient tool to

check the activity of a biological culture. But accuracy and precision have been found too low

for screening purposes. The results show that the respiration measured in shake flasks and

fermenters can be very similar. This makes it possible to increase the productivity by

performing screening experiments in online monitored shake flasks, which traditionally are

performed in fermenters. Additionally, the online monitoring increases the knowledge gained

from a single shake flask experiment and facilitates the identification of parameters critical for

scale-up.

Page 56: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 38

Chapter 4

Development of a scale-up approach for aerobic cultures

To transfer the results gained with the shake flask screening methods to fermenters, a scale-up

strategy is required. The scale-up strategy should make it possible to select the operating

conditions of aerobic cultures in shake flasks and fermenters. The scale-up strategy introduced

in this study is based on the volumetric mass transfer coefficient (kLa) and the pH change of

the medium. The effectiveness of the scale-up strategy is demonstrated by comparing the

behavior of Corynebacterium glutamicum on lactic acid in shake flasks and fermenters

operated in batch and continuous mode. Most of the results presented in this chapter have

been described by Seletzky et al. [2007].

4.1 Introduction

When designing an industrial fermentation, a large number of experiments are necessary to

select a strain, to improve the medium, and to optimize the fermentation procedure.

According to Kennedy et al. [1994], the only practical way of performing these experiments is

to use shake flasks. Probably more than 90 % of all culture experiments in biotechnology are

performed in shaking bioreactors [Büchs 2001]. However, production takes place in

fermenters. Therefore a scale-up strategy is required to reproduce the culture characteristics

and the kinetic parameters from shake flasks in fermenters. According to Gupta and Rao

[2003] shake flasks have not been used to perform scale-up of fermentation processes due to

the lack of knowledge of the conditions under which shake flask fermentations are performed.

Page 57: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 39

Sumino et al. [1993] point out that a better understanding and a thorough utilization of shake

flasks should make a direct method for scaling-up a fermentation process possible. In recent

years, the knowledge about the conditions in shake flasks has increased. Maier and Büchs

[2001], Maier et al. [2004], Maier [2002] characterized the mass-transfer, Büchs et al. [2000 I,

II] and Peter et al. [2004] characterized the power input, Peter et al. [2006] the

hydromechanical stress, and Büchs et al. [2006], Maier and Büchs [2001], Peter et al. [2004]

the fluid movement. Until lately shake flask experiments could be performed only in the batch

mode of operation. The COSBIOS device (Continuously Operated Shaking Bioreactor

System) introduced by Akgün et al. [2004] broadens the field of application to continuous

cultures. In continuous cultures under steady-state conditions the environment is well defined

and the results obtained are generally more reliable and reproducible [Sipkema et al. 1998].

Various authors have demonstrated the leading importance of the substrate oxygen.

According to Gupta and Rao [2003] changes in oxygen availability may lead to drastic effects

on fermentation kinetics. Clark et al. [1995] state that oxygen limitation acts in an analogous

manner to substrate limitation, and Katzer et al. [2001] point out that the production of

secondary metabolites is influenced not only by factors like nitrogen or phosphate limitation

but may also be dependent on oxygen limitation. If the gas-liquid mass-transfer is the limiting

step, efforts aiming at the improvement of e.g., the enzymatic activity of the metabolic

pathway leading to the desired product, will not be of much benefit [Büchs 2001]. The high

importance of oxygen for microorganisms with a high oxygen demand like Corynebacterium

glutamicum makes a constant volumetric mass transfer coefficient (kLa) an efficient scale-up

strategy. The strategy was successfully applied based on experimental data for batch cultures

by Gupta and Rao [2003], and by Kim et al. [2003] for the scale-up from laboratory to

industrial fermenters. However, it has not yet been applied based on empirical correlations or

for continuous cultures.

Page 58: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 40

The pH value of the culture medium effects growth, respiration and product formation of

microorganisms. The influence of the pH on Corynebacterium glutamicum has been described

in the chapters 2.34, 2.4.5 and by Bröer and Krämer [1991], Seletzky et al. [2006a], Shah et

al. [2002]. Thus, the scale-up from shake flasks to fermenters can only succeed, if the pH shift

is similar. Shake flasks are usually operated without pH control, while in fermenters the pH

can be controlled by titrating with acids or bases. The pH of shake flask cultures is normally

stabilized by buffering the medium. However, the addition of buffer increases the ionic

strength and reduces the oxygen solubility of the medium. The reduced solubility may lead to

oxygen-limited conditions [Büchs 2001]. Moreover, if organic acids, like lactate or acetate,

are used as carbon source, the pH may shift even in strongly buffered medium. However, in

continuous cultures the use of buffers can be reduced or avoided by anticipating the pH

change of the culture by appropriately adjusting the pH of the feed solution. Thus, the feed

does not only add nutrients, but also controls the pH. The great difference in the way the pH is

controlled in shake flasks and fermenters makes it necessary to include in any scale-up

strategy a pH model. The model can be used to predict the pH shift of the shake flasks in

batch mode, to estimate the amount of buffer required to stabilize the pH, to predict the

amount of titrant required to control the pH of the fermenter cultures, and to calculate the pH

of the feed solution required to anticipate the pH change of continuous cultures.

Hydromechanical stress, which is often regarded as an important scale-up parameter [Katzer

et al. 2001, Amanullah et al. 1999, Sahoo et al. 2003, Cui et al. 1998], doesn't have to be

considered for the model organism Corynebacterium glutamicum. According to Chamsartra et

al. [2005] variations in agitation do not damage C. glutamicum or cause a significant change

in physiological response.

The focus of this chapter is to introduce a practical scale-up strategy from shake flask to

fermenter in batch and continuous mode for everyday use. The scale-up strategy should allow

Page 59: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 41

the operating conditions of aerobic cultures to be selected. The scale-up strategy is based on

the volumetric mass transfer coefficient (kLa) and the pH change of the medium. The

effectiveness of the scale-up strategy is demonstrated by comparing the behavior of

Corynebacterium glutamicum on lactic acid in shake flasks and fermenters operated in batch

and continuous mode. Additionally, a simple biological and a mechanistic pH model is used

to describe the behavior of Corynebacterium glutamicum in shake flasks and fermenters with

the same kinetic parameters.

4.2 Materials and methods

4.2.1 Microorganism and media

Experiments were carried out with the wild type of Corynebacterium glutamicum

ATCC 13032 [Kinoshita et al. 1957]. The preparation of the precultures and the media were

performed as described in appendix A.1.2. The minimal medium contained per liter: 12.5 g

lactic acid for batch cultures, 11 g for continuous fermenter cultures and 5 g for continuous

shake flask cultures. The biomass and lactic acid concentrations and the oxygen solubility of

the medium were determined as described in appendix A.5.1, A.5.2, A.1.3.

4.2.2 Fermentation devices

4.2.2.1 Batch cultures

Fermenter batch cultures were processed in a laboratory fermenter as described in appendix

A.2.1. To assure equal culture conditions in fermenters and shake flasks the pH of the batch

cultures was not controlled. The oxygen transfer rate of the fermenter cultures was measured

with an exhaust gas analyzer as described in appendix A.3.1.

The oxygen transfer rate (OTR) of shake flask cultures was measured with a RAMOS device

as described in appendix A.3.2. Cultivations for biomass, substrate and pH determination

Page 60: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 42

were carried out in unbaffled 250 ml shake flasks with cotton plugs. To reduce the sampling

time to a minimum, a separate shake flask was sacrificed for each sample and the sample flask

was not replaced on the shaker.

4.2.2.2 Continuous cultures

Continuous fermenter cultures were performed with a laboratory fermenter as described in

appendix A.2.2. Steady state conditions were assumed after at least 5 residence times. Each

dilution rate was determined by measuring the filling volume of the fermenter and the weight

difference of the feedstock. The density of the medium was measured to be 1.01 kg/l. The pH

was kept constant over time at pH 7 by titrating with 2M H2SO4. The amount of titrant added

was recorded with a balance. The density of 2 M H2SO4 at 20°C was 1123 g/l.

Continuous shake flask cultures were performed with the COSBIOS device (Continuously

Operated Shaking Bioreactor System) introduced by Akgün et al. [2004] as described in

appendix A.2.4. Eight flasks were operated simultaneously at different dilution rates making

it possible to record a biomass over dilution rate (X-D) diagram in less than one week.

4.3 Determination of the culture conditions

4.3.1 Scale-up of the volumetric mass transfer coefficient and determination of non oxygen limited culture conditions

In order to apply a constant volumetric mass transfer coefficient (kLa) as scale-up strategy, it

is necessary to characterize the influence of the operating conditions on the volumetric mass

transfer coefficients, in fermenters and shake flasks. The mass transfer in fermenters has been

reviewed by van't Riet [1979] and is not the focus of this work. This chapter describes a

practical approach to predict the kLa in shake flasks and a method to prevent oxygen-limited

culture conditions in shake flasks.

Page 61: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 43

Generally, it is difficult precisely reproduce the mass transfer characteristics of a fermenter

with a shake flask. However, the metabolic activity of many microorganisms (e.g. for

C. glutamicum [Chamsartra et al. 2005]) is not effected by the oxygen concentration of the

liquid phase as long as the concentration is above the critical value close to zero. Thus, in

most cases it is sufficient to distinguish oxygen-limited and non oxygen limited operating

conditions.

To assure non oxygen limited culture conditions, the maximum oxygen consumption of the

microorganisms (maximum oxygen uptake rate (OURmax)) should not exceed the maximum

amount of oxygen delivered by the gas-liquid mass transfer (maximum oxygen transfer

capacity (OTCmax)) (Eq. 4.1).

maxmax OTC OUR < (4.1)

If no product is formed, the maximum oxygen uptake rate (OURmax) of a biological culture

depends on the maximum growth rate (µmax), the biomass oxygen yield (YX/O2), and the

maximum biomass concentration (Xmax) (Eq. 4.2). Otherwise, the product formation should be

included in Eq. 4.2. If the initially added substrate is entirely consumed for biomass

formation, the maximum biomass concentration can be estimated with Eq. 4.3 by multiplying

the biomass substrate yield (YX/S) and the substrate concentration (S) of the limiting substrate.

maxmaxO/X

max XY

1OUR2

⋅µ⋅= (4.2)

SYX S/Xmax ⋅= (4.3)

The oxygen transfer rate (OTR) can be described with Eq. 4.4, where kLa is the volumetric

mass transfer coefficient, SO2 the oxygen solubility, pR the reactor pressure, yO2 the oxygen

mole fraction in the gas phase, and yO2l the oxygen mole fraction in the liquid phase. The

Page 62: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 44

maximum oxygen transfer capacity (OTCmax) at a given kLa is obtained under oxygen-limited

conditions, when yO2l is close to zero (Eq. 4.5).

( )lOOROL 222yypSakOTR −⋅⋅⋅= (4.4)

22 OROLmax ypSakOTC ⋅⋅⋅= (4.5)

For the chemical sulfite system, Maier [2002] found an empirical correlation (Eq. 4.6) for the

kLa in shake flasks, considering the operating parameters shaking frequency (n) [rpm], filling

volume (V) [ml], shaking diameter (d0) [cm], and maximum inner shake flask diameter (d)

[cm].

92.138.00

83.016.16sulfiteL ddVn1067.6ak ⋅⋅⋅⋅⋅= −− (4.6)

The correlation is valid for standard glass Erlenmeyer flasks according to [DIN 12380] with

hydrophilic walls, shaking frequencies of 50-500 rpm, relative filling volumes of 4-20 % (the

relative filling volume is defined as the filling volume divided by the nominal flask volume),

shaking diameters of 1.25-10 cm, and nominal flask volume between 50-1000 ml. The

accuracy of the kLa correlation for shake flasks is 30 %, putting it in the same range as the kLa

correlation for fermenters, which lies, according to van't Riet [1979], between 20 % and 40 %.

The maximum oxygen transfer capacity of the sulfite system OTCmax, sulfite can be estimated,

for a given set of operating conditions, with Eqs. 4.5 and 4.6. The oxygen solubility (SO2) for

the 1 M sulfite system at 22.5° C used by Maier et al. [2004] and Maier [2002] is 0.00058

mol/l/bar. SO2 was calculated as described in appendix A.1.3.

According to Maier et al. [2004] the maximum oxygen transfer capacity for biological media

(OTCmax, bio) can be estimated for any operating condition by using a correlation factor (cf)

between OTCmax, bio and OTCmax, sulfite (Eq. 4.7). The correlation factor is determined for a

Page 63: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 45

single set of operating parameters by correlating OTRmax, sulfite calculated with Eqs. 4.5 and 4.6

with OTCmax, bio, experimentally determined with a RAMOS device from an oxygen-limited

biological culture.

sulfitemax,biomax, OTCcfOTC ⋅= (4.7)

Alternatively, if no RAMOS device is available, OTCmax, bio can be estimated with Eq. 4.8

using the volumetric mass transfer coefficient of the sulfite system according to Eq. 4.6.

22 ORbio,OsulfiteLLsulfite

Lbiobiomax, ypSak

kkOTC ⋅⋅⋅⋅= (4.8)

The oxygen solubility of the biological medium SO2, bio can be estimated as described in

appendix A.1.3. The specific surface area (a) of a shake flask depends only on the operating

conditions and not on the medium composition [Büchs 2001]. Thus, the specific surface area

of differently composed solutions with waterlike viscosities can be assumed equal. With this

assumption, differences between the kLa of the sulfite and the biological system result only

from the relation between the gas-liquid mass transfer coefficient of the biological (kLbio) and

the sulfite system (kLsulfite).

For defined aqueous electrolyte solutions, kL can be derived from the diffusion coefficient of

oxygen using the correlations of Akita [1981], however, biological solutions often contain

undefined complex media components and the microorganisms produce biological

surfactants. The influence of complex media components and biological surfactants on kL

cannot yet be determined. Thus, only general concepts, describing the interaction between

surfactants and kL can be used to estimate kLbio. So far no such concepts exist for the gas-

liquid interface of shake flasks. However, for bubble aerated systems the gas-liquid mass

transfer coefficient is considered to depend on the mobility of the gas-liquid interface. The

mobility is reduced through the accumulation of surfactants until the interface can be

Page 64: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 46

considered rigid. If the interface is mobile, the addition of surfactants leads to a strong

decrease of kL. If, however, the interface is almost rigid, the addition of more surfactants leads

only to a small change of kL. For completely mobile interfaces in purified water Alves et al.

[2005] found kL values for single bubbles ranging from 3.5·10-4 m/s to 5.5·10-4 m/s, which

decrease to 0.75·10-4-1·10-4 m/s for rigid interfaces. For single bubbles with a rigid surface,

Vasconcelos et al. [2002] gives a minimum kL of 1·10-4 m/s. Alves et al. [2004] show that

sulphate reduces the mobility of the interface and reports a kL of 0.87·10-4 m/s for bubbles

made completely rigid with polyethylene glycol (PEG). For 250 ml flasks kLsulfite values of

0.4·10-4-1.1·10-4 m/s can be derived from the data of Maier et al. [2004] and Maier [2002],

using the model of Maier and Büchs [2001] to calculate the specific surface area (a). Thus,

kLsulfite of the 1 M sulfite system corresponds to kL values of a bubble aerated system with a

completely rigid interface. For a biological medium with Cnaetomium cellulolythicum the

review of Kawase et al. [1992] reports kLbio values of 0.75·10-4-1·10-4 m/s. These values also

correspond to kL values of a bubble aerated system with a completely rigid interface. If the

general dependencies, which govern kL in a bubble aerated system, can be transferred to shake

flasks, the interface of the sulfite system and the biological system can be considered rigid,

and thus variations in medium composition and biological surfactants should have only a

small influence on kL. In conclusion, until the dependency between ionic strength, biological

surfactants, surface contamination and gas-liquid mass transfer coefficient is not clarified, it is

reasonable to assume that the gas-liquid mass transfer coefficient of the sulfite (kLsulfite) and

the biological (kLbio) system is equal.

4.3.2 Scale-up of the pH

The scale-up from shake flasks to fermenters can only succeed if the pH shift is similar. A

practical laboratory method is introduced to determine the pH of the feed medium for

continuous cultures, which makes it possible to operate continuous cultures at the desired pH

Page 65: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 47

without titration. This simple method can be performed with standard laboratory equipment.

The practical execution of the method is depicted in Fig. 4.1. The first step is to record the

titration curves of the continuous and batch medium. In our case, the continuous medium

contained MOPS buffer, which was absent in the batch medium. Since the substrate, which is

nearly entirely consumed in continuous cultures, may have a buffering effect, the titration

curves should be determined without the limiting substrate. To increase the comparability, the

titration curves depicted in Fig. 4.1 were normalized to pH 7 by subtracting the amount of

titrant necessary to adjust the pH to 7 from the actually titrated amount. This yields positive

amounts for a pH greater than 7 and negative amounts for a pH smaller than 7. The second

step is to record the pH shift during a complete batch culture. In our case the pH of a batch

culture with a lactic acid concentration of 12.5 g/l shifted from 7 to 8.6. The third step is to

read from the titration curve of the batch medium the amount of titrant (acid or base) required

to compensate the pH shift of the batch culture (∆ABbatch). In our case ~0.054 mol/l NaOH

were required. The fourth step is to calculate the amount of titrant required to compensate the

pH shift of the continuous culture (∆ABconti) by considering the different substrate

concentrations of the batch (Sbatch 12.5 g/l) and the continuous medium (Sconti=5 g/l) according

to Eq. 4.9.

batch

contibatchconti S

SABAB ⋅∆=∆ (4.9)

Page 66: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 48

The final step is to read the pH of the continuous medium from the titration curve of the

continuous medium. As shown in Fig. 4.1, in our case a pH of 6.5 resulted in a stable pH of 7.

Fig. 4.1: Practical laboratory method to

determine the pH of the feed solution of

not titrated continuous cultures.

titration curve of shake flask medium

for continuous cultures with 20 g/l MOPS

buffer. titration curve of shake flask

medium for batch shake flask cultures

without MOPS buffer. All titration curves

were normalized to pH 7, by subtracting

the amount of the titrant NaOH necessary to adjust the pH to 7 from the actually titrated amount,

giving positive amounts for pH greater than 7 and negative amounts for pH smaller than 7. Batch

cultures with the batch medium (lactic acid concentration 12.5 g/l, without MOPS buffer) reached a

final pH of 8.6. ∆ABbatch is the amount of titrant required to compensate this pH shift. To autostabilize

the pH of continuous cultures without titration at a pH of 7, the continuous medium (lactic acid

concentration 5 g/l, 20 g/l MOPS buffer) should have an initial pH of 6.5. ∆ABconti is the amount of

titrant required to anticipate this pH shift. The sections of the titration curves necessary to describe the

pH change caused by the biological activity (Fig. 4.3d) is described with the mechanistic pH model

(solid lines) as specified in chapter 4.4.2.

-0.04 -0.02 0.00 0.02 0.04 0.06 0.085

6

7

8

9

10

titration curvecontinuous medium

titration curvebatch medium∆ABconti

5 g/l lactic acid∆ABbatch

12.5 g/l lactic acid

pH

normalized NaOH [mol/l]

As described in the results section the continuous medium requires a lower lactic acid

concentration to avoid oxygen limited conditions. The lower lactic acid concentration

compensates the higher filling volumes of the COSBIOS device. All batch media did not

contain buffer to use the same medium for shake flask and fermenter cultures and fermenter

media usually do not contain buffer. This is possible, because C. glutamicum can grow

between pH 5 and pH 9 [Chapter 2.3.4, Seletzky et al. 2006a]. Furthermore, the large pH shift

of the batch cultures allowed a better demonstration of the possibilities of the pH model. In

Page 67: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 49

the medium for continuous shake flask cultures buffer was added, to avoid uncontrolled pH

shifts during the change from batch to continuous operation.

4.4 Model description

A single model is applied to describe the behavior of batch and continuous cultures in

fermenters, and shake flasks. The model consists of a biological and a pH model. The

biological model describes the metabolic activities: growth, substrate consumption, and

respiration of the microorganisms. The mechanistic pH model is based only on the medium

composition, which may change through the biological activity. The pH model can describe

the pH change caused by the consumption of the substrates, the amount of titrant necessary to

stabilize the pH or the pH of the feed solution for continuous cultures required to keep the pH

at its desired value. All simulations were carried out with the software package gPROMS

(general PROcess Modelling System), (PSE, London, Great Britain). Model equations

describing microbial growth and microbial substrate consumption are denominated in mass

concentrations [g/l]. Model equations describing the change of ion concentrations and pH are

denominated in molar concentrations [mol/l] (symbols of the concentrations start with a [c]).

4.4.1 Description of the biological model

The increase of the biomass concentration (X), the consumption of lactic acid (S) and

ammonia (NH3) are modeled using the differential Eqs. 4.10-4.12. For batch cultures, the

dilution rate (D) is set to zero. The consumption of trace elements was neglected.

XDXdtdX

⋅−⋅µ= (4.10)

( SSDXY

1dtdS

FS/X

−⋅+⋅µ⋅−= ) (4.11)

Page 68: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 50

( 3F3NH/X

3 NHNHDXY

1dt

dNH

3

−⋅+⋅µ⋅−= ) (4.12)

The growth of C. glutamicum is described with a Monod-kinetics for the limiting substrate

lactic acid (Eq. 4.13). For ammonia and oxygen, no Monod kinetics were set-up, since

ammonia is present in the medium in excess and all cultivations are performed without

oxygen limitation. The Monod constant KS was set to 0.0045 g/l, an average value of the

Monod constants for bacteria measured by Owens and Legan [1987].

Smax KS

S+

⋅µ=µ (4.13)

The oxygen consumption of the microorganisms (OUR) is described by the growth dependant

algebraic Eq. 4.14.

XY

1OUR2O/X

⋅µ⋅−= (4.14)

The biomass substrate yield (YX/S), oxygen biomass yield (YX/O2), and maximum growth rate

(µmax) are derived from the experimental data of this work; the nitrogen biomass yield (YNH3)

was determined using the method described by Stöckmann et al. [2003b] (results not shown).

All kinetic parameters are summed in Tab. 4.1, all initial values in Tab. 4.2.

Page 69: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 51

Tab. 4.1: Kinetic parameters of

the biological model.

kinetic parameter value

KS 0.00 g/l45

YX/NH3 6.25 g/g

YX/O2 batch 20.7 g/mol

YX/O2 conti 25.9 g/mol

YX/S batch 0.36 g/g

YX/S conti 0.45 g/g

µmax batch 0.23 h-1

µmax conti 0.32 h-1

Tab. 4.2: Initial values for the

biological and the pH model. initial value value

c(NH4)2SO4 0.15 mol/l

S0 batch cultures 12.5 g/l

S0=SF conti shake flask cultures 5 g/l

S0=SF conti fermenter cultures 11 g/l

X0 0.32 g/l

4.4.2 Description of the pH model

The change of the pH is described with a static pH model. For biological cultures, a static pH

model can be used because the acid-base reactions are much faster than the metabolic

reactions of the microorganisms. The model is based on the dissociation equations using

Bronsted's acid-base concept, material balances, and the electroneutrality condition. The

medium was described as an infinitely diluted solution; a simplification successfully used by

Gustafson et al. [1995], Gustafson and Waller [1992], Nicolas-Simonnot et al. [1995], Støle-

Hansen and Foss [1997], and Westerlund et al. [1995]. The pH changes result through the

Page 70: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 52

consumption of lactic acid (Eq. 4.11) and ammonia (Eq. 4.12). While the consumption of

lactic acid leads to an increase, the consumption of ammonia leads to a decrease of the pH.

The pH change is buffered by the phosphates (pKa=7.21) and by the ammonia (pKa=9.25) in

the medium.

The pH is defined as the negative logarithm of the hydronium ion concentration (cH+). The

(4.15)

Considering the focus on simplicity and practicability of the model, the influence of trace

ability of water to release and accept a proton is expressed by the ionic product of water

Eq. 4.15.

( ) +− ⋅= cHcOHOHK 2

elements, creation of complexes and formation of carbon dioxide on the pH was neglected.

Their concentrations are small in comparison to the concentrations of the strongly pH active

medium components lactate, ammonium, phosphates, and MOPS buffer. Eqs. 4.16-4.19

describe the dissociation of these strongly pH active components. The relevant dissociation

constants are summed up in Tab. 4.3.

( )363

353363 OHcC

cHOHcCOHCK

+− ⋅= (4.16)

( )+

++ ⋅

=4

34 cNH

cHcNHNHK (4.17)

( )−

+−− ⋅

=42

24

42 POcHcHcHPO

POHK (4.18)

The medium of the continuous shake flask cultures additionally contained the buffer MOPS.

( )cMOPS

cHcMOPSMOPSK+− ⋅

= (4.19)

Page 71: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 53

Tab. 4.3: Dissociation constants

according to Lide [1999] and

Dawson [1986].

able [Chapter 2.3.4, Seletzky et al.

are regarded as completely dissociated

(4.20)

(4.21)

(4.22)

The mass balances (Eqs. 4.23-4.25) assure the conservation of mass with cisum being the sum

of all species of the medium component (i).

(4.23)

(4.24)

424242 cHPOPOcHPOcKHHPOcK (4.25)

When a solution is formed by mixing electrically neutral aqueous solutions of acids, bases

and salts, the resulting solution should also be

dissociation constant value

K(C3H6O3) 10-3.25 mol/l

K(H2O) at 30° C 10-13.833 mol²/l²

K(H2PO4-) 10-7.21 mol/l

K(MOPS) 10-7.2 mol/l

K(NH4+) 10-9.25 mol/l

At a pH between 5 and 10, where C. glutamicum may be vi

2006a], salts, strong acids, and strong bases

(Eqs. 4.20-4.22). Acids and bases are considered as strong, when having a dissociation

constant smaller than 10-3 or larger than 10-12.

=+ cNaOHcNa

4242 POcKHHPOcK2cK +⋅=+

( ) 4242

4 SONHccSO =−

363353sum33 OHcCOHcCHOcC += −

34424 cNHcNHSO)NH(c2 +=⋅ +

−− +=+ 24

electrically neutral. This is assured by the

Page 72: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 54

electroneutrality condition (Eq. 4.26), which sets the sum of the positively charged ions equal

to the sum of the negatively charged ions.

−−−−4

−++++ ⋅++⋅++=+++ 2442

24353 cHPO2POcHSO2OHcCcOHcKcNHcNacH (4.26)

The pH of the continuous fermenter cultures is controlled by titrating with sulfuric acid

(H2SO4). The strong acid is always assumed to be completely dissociated, therefore, only the

concentration of the sulfate ions SO42- was modeled. The variation of the sulfate ion

concentration is described by Eq. 4.27, which substitutes Eq. 4.22. The sulfate ion

concentration depends on the concentration of ammonium sulfate (c(NH4)2SO4) in the feed

medium, and the amount of titrant H2SO4 added per time (ABtSO42-). Numerically the amount

of titrant was adjusted in the way that the pH was kept constant at 7.

( ) −−−−

+−⋅= 24t

24F

24

4 SOABcSOcSODd

2

tdcSO (4.27)

4.5 Results and discussion

4.5.1.1 Evaluation of the predicted oxygen transfer capacity (OTCmax)

volumetric mass transfer coefficient (kLa), it is

ating conditions. Fig.

4.5.1 Gas-liquid mass-transfer

To apply a scale-up strategy based on the

necessary to predict the interaction of gas-liquid mass transfer and oper

4.2 compares the OTCmax, bio predicted with Eqs. 4.6 and 4.8 (concave lines) with the

OTCmax, bio of C. glutamicum cultures experimentally determined (crosses) in the RAMOS

device (OTR values of the plateaus in the Figs. 2.1, 2.3, 5.3a). The maximum deviation

between the predicted and measured OTCmax, bio is 15 %.

Page 73: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 55

For the relation between the oxygen transfer capacity of the biological system (OTRmax, bio)

and the oxygen transfer capacity of the sulfite system (OTCmax, sulfite) an average correlation

factor (cf) according to Eq. 4.7 of 1.8 can be calculated from the experimental data shown in

Fig. 4.2 (crosses). This value is close to the relation of 1.72 between the oxygen solubility of

the biological medium and the sulfite system (SO2bio/SO2sulfite), proving that the gas-liquid mass

transfer coefficient of the biological system kLbio is approximately equal to the gas-liquid mass

transfer coefficient of the sulfite system kLsulfite.

4.5.1.2 Determination of oxygen non limited culture conditions

To distinguish oxygen-limited and non limited conditions, Fig. 4.2 compares the maximum

oxygen uptake rate (OURmax) (straight line) of Corynebacterium glutamicum at varying lactic

acid concentrations with the maximum oxygen transfer capacity (OTCmax, bio) (concave lines)

at different operating conditions. For the calculation of OURmax (Eqs. 4.2 and 4.3) the kinetic

parameters for the continuous cultures (YX/S=0.45 g/g, YX/O2=25.9 g/mol, µmax=0.32 h-1), as

defined in the following section, were used. The OTCmax, bio, according to Eqs. 4.6 and 4.8 is

depicted for a shaking diameter of 5 cm at shaking frequencies of 350 rpm, 300 rpm, 250 rpm,

200 rpm, and 150 rpm over the filling volume. The sample (gray lines) depicted in Fig. 4.2

demonstrates that not oxygen-limited culture conditions for a Corynebacterium glutamicum

continuous culture with a lactic acid concentration of 10 g/l, using a shaking diameter of 5 cm

and a shaking frequency of 300 rpm, require a filling volumes of less than 11 ml.

Page 74: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 56

Fig. 4.2: Determination of oxygen-

limited and non limited culture

conditions, by comparing the

maximum oxygen uptake rate

(OURmax) of a Corynebacterium

glutamicum culture at different lactic

acid concentrations with the modeled

and measured maximum oxygen

transfer capacity (OTCmax, bio) at

different operating conditions. OURmax

(straight line). OTCmax, bio calculated (concave lines): 250 ml shake flask, shaking frequency — — —

350 rpm, ——— 300 rpm, • • • • • 250 rpm , — • — • —200 rpm, • • —• • — 150 rpm, shaking diameter

50 mm. X OTCmax, bio measured with the RAMOS device, culture conditions: 250 ml shake flask,

shaking frequency 300 rpm, shaking diameter 5 cm, filling volume 20 ml, 25 ml, 30 ml, 60 ml.

0 1 2 3 4 5 6 7 8 9 10 11 12

lactic acid concentration [g/l]

0 5 10 15 20 25 30 35 40 45 50 55 600.000.010.020.030.040.050.060.07

OTC

max

, bio, O

UR

max

[m

ol/l/

h]

filling volume [ml]

In most cases, non oxygen-limited culture conditions can be obtained by adjusting the

operating conditions. However, if this is technically not possible or not sufficient, the

substrate concentration of the medium has to be reduced. This was the case for the continuous

shake flask cultures, where the in respect to the batch cultures higher filling volumes of

~25 ml reduced OTCmax, bio. To avoid oxygen-limited conditions, the substrate concentration

had to be reduced to 5 g/l.

Page 75: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 57

Fig. 4.3: Culture charac-

teristics of C. glutamicum

in shake flasks and fer-

menters in batch (ba, left)

and continuous (co, right)

mode of operation. Solid

symbols: fermenter cultur-

es, open symbols: shake

flask cultures, solid lines:

simulations. A. ,

biomass concentration.

B. , oxygen transfer

rate. C. , lactic acid

concentration. D. batch

cultures pH, continu-

ous ferm cultures

amount of titrant H

enter

a

2SO4

added, continuous sh ke

flask cultures pH. For

dilution rates exceeding

the washout point steady-state conditions have not completely been reached. Culture conditions

fermenter: filling volume 1 l, stirrer speed max. 1200 rpm, specific aeration rate 2 vvm, lactic acid

concentration batch cultures 12.5 g/l, continuous cultures 11 g/l. The pH of the batch cultures was not

controlled. The pH of the continuous cultures was controlled with 2M H2SO4. The initial pH of the

media was 7. Culture conditions shake flasks: shaking frequency 300 rpm, batch cultures filling

volume 10 ml, continuous cultures filling volume ~25 ml, shaking diameter 5 cm, lactic acid

concentration batch cultures 12.5 g/l, continuous cultures 5 g/l. The pH of the shake flask cultures was

not controlled, the initial pH of the batch media was 7 and of the continuous medium 6.5.

0

1

2

3

4

5

6

Bba

AcoAba

biom

ass

conc

entra

tion

[g/

l]

0 2 4 6 8 10 126.0

6.5

7.0

7.5

8.0

8.5Dba

pH

[-]

fermentation time [h]

0.00

0.01

0.02

0.03

0.04

0.05

oxyg

en tr

ansf

er ra

te

[mol

/l/h]

Cco

Bco

0.0 0.1 0.2 0.3 0.4 0.5

Dco

dilution rate [1/h]

0

2

4

6

8

10

12

Cba

lact

ic a

cid

conc

entra

tion

[g/

l]

la

ctic

aci

d co

ncen

tratio

n [

g/l]

0.000

0.002

0.004

0.006

0.008

0.010

titra

nt 1

M H

2SO

4 [m

ol/l/

h]0

1

2

3

4

5

6

7

0.0

0.5

1.0

1.5

2.0

2.5

3.0

bio

mas

s co

ncen

tratio

n [

g/l]

Page 76: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 58

4.5.2 Scale-up from shake flasks to fermenters in batch and continuous mode

The scale-up strategy introduced is evaluated by comparing the behavior of Corynebacterium

glutamicum on lactic acid in shake flasks and fermenters in batch and continuous mode of

operation. The experimental data is depicted in Fig. 4.3. The shake flask (open symbols) and

fermenter (solid symbols) batch cultures are depicted on the left (Fig 4.3aba-4.3dba). The shake

flask and fermenter continuous cultures are depicted on the right (Fig 4.3aco-4.3dco). The

general behavior of C. glutamicum is independent of the scale and of the type of bioreactor.

The biomass formation (Fig. 4.3a), the respiration (Fig. 4.3b), the substrate consumption (Fig.

4.3c), and the pH change (Fig. 4.3d) are very similar for shake flask and fermenter cultures.

The behavior of the organism can be described independent of the scale and type of bioreactor

with the same biological and pH model (solid lines).

4.5.2.1 Growth rate and maximum growth rate

The maximum growth rate of a batch culture can be derived from the slope of the biomass

concentration (Fig. 4.3aba) or from the slope of the oxygen transfer rate (OTR) (Fig. 4.3bba)

[Chapter 3.3.1, Seletzky et al. 2006c, Stöckmann et al. 2003b]. Independent of the scale, three

growth phases can be identified from the OTR curves of the batch cultures (Fig. 4.3bba): an

acceleration phase at fermentation times shorter than 4 h with growth rates lower than

0.15 h-1, an exponential growth phase where the microorganism growth with the maximum

growth rate (µmax) of 0.31 h-1, and a deceleration phase with a growth rate of 0.23 h-1 at pH

above 8, where the organism is pH inhibited [Chapter 2.3.4, Seletzky et al. 2006a]. The µmax

of 0.26 h-1 derived from the biomass concentrations is lower, because the sampling interval

did not allow a distinction of the last two growth phases. In order to focus on a simple and

practical scale-up method and not on modeling the biological behavior in detail, the simple

biological model described by Eqs. 4.10-4.14 was used, applying for the entire batch culture

an average growth rate of 0.23 h-1.

Page 77: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 59

The highest dilution rate (D) at which the cultures do not wash out is considered as the

maximum growth rate of the continuous cultures. The washout of the fermenter cultures

occurred between 0.318 h-1 and 0.322 h-1, that of the continuous shake flask cultures between

0.296 h-1 and 0.322 h-1. Therefore, a µmax of 0.32 h-1 was used as model parameter for the

continuous cultures. In conclusion, the maximum growth rate of Corynebacterium

glutamicum on lactic acid is independent of the mode of operation or the culture vessel. The

growth rates observed in this study are similar to the data of Cocaign et al. [1993], and

Cocaign-Bousquet and Lindley [1995] who found a maximum growth rate of 0.35 h-1 for

Corynebacterium glutamicum ATCC 17965 batch cultures on lactate.

4.5.2.2 Biomass formation and substrate consumption

The biomass substrate yield (YX/S) of the batch cultures, derived from the data of Fig. 4.3aba

and Fig. 4.3cba, lies between 0.28 g/g and 0.37 g/g, being generally higher at lower lactic acid

concentrations. For the model, a YX/S of 0.36 g/g was used.

The YX/S of the continuous cultures (Fig. 4.3aco, Fig. 4.3cco) is higher than that of the batch

cultures and lies between 0.41 g/g and 0.48 g/g. For the model, an average YX/S of 0.45 g/g

was used.

The differences of YX/S for Corynebacterium glutamicum on lactate has also been observed by

[Cocaign et al. 1993], who found an average YX/S of 0.36 g/g for C. glutamicum ATCC 17965

batch cultures, which increased to 0.49 g/g during the exponential growth phase. Cocaign-

Bousquet and Lindley [1995] observed a YX/S between 0.36 g/g and 0.42 g/g for continuous

cultures. The lower YX/S of the batch cultures compared to the continuous cultures probably

results from the higher mean lactic acid concentration. As described in chapter 2.3.3 [Seletzky

et al. 2006a], in batch cultures YX/S decreases with increasing initial lactic acid

concentrations.

Page 78: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 60

4.5.2.3 Oxygen consumption

The oxygen substrate coefficient YO2/S of 0.0174 mol/g was equal for batch and continuous

cultures and for shake flask and fermenter cultures. Thus, the batch cultures with a YX/S of

0.36 g/g have a biomass oxygen yield (YX/O2) of 20.7 g/mol. The continuous cultures with a

YX/S of 0.45 g/g have a YX/O2 of 25.9 g/mol.

The oxygen substrate yield observed in this study is identical to the one reported by Cocaign-

Bousquet and Lindley [1995] for continuous cultures under conditions where the substrate is

entirely consumed (calculated from the data depicted in the figures). A constant YO2/S

combined with different YX/S was also observed in chapter 2.3.3 [Seletzky et al. 2006a] for

batch cultures with different initial lactic acid concentrations.

4.5.2.4 pH and titration

The consumption of lactic acid leads to an increase of the pH. The pH of the batch cultures

increases from 7 to 8.6 (Fig. 4.3dba). To stabilize the pH of the continuous fermenter culture

about 0.27 mol H2SO4 per mol of metabolized lactic acid were added (Fig. 4.3dco). By using a

feed solution with a pH of 6.5, the pH of the continuous shake flask cultures was stabilized at

7.1 (Fig. 4.3dco). With the same pH model it was possible to describe the pH increase of the

batch cultures (Fig. 4.3dba), the amount of titrant required to stabilize the pH of the continuous

fermenter cultures (Fig. 4.3dco) and the pH of the feed solution necessary to stabilize the pH

of the continuous shake flask cultures (Fig. 4.3dco). In the pH range important for biological

growth [Chapter 2.3.4, Seletzky et al. 2006a], simulated titration curves of the batch and the

continuous medium are in good agreement with the experimentally determined titration

curves (Fig. 4.1) (see also Fig. 2.4d).

Page 79: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 4 Development of a scale-up approach for aerobic cultures 61

4.6 Conclusion

Commonly, the scale-up from shake flasks to fermenters is considered difficult, since shake

flasks are not well characterized. Based on recently published empirical models, which

describe the interaction between operating conditions and culture conditions for shake flasks,

it was possible to develop a scale-up strategy for aerobic cultures, which focuses on the

volumetric mass transfer coefficient (kLa) and the pH. The empirical model reported for the

chemical sulfite system was evaluated and adjusted to the conditions of the biological system.

The scale-up from shake flasks to fermenters of Corynebacterium glutamicum on lactic acid

was successfully accomplished for batch and continuous mode of operation. The behavior

could be described with a simple biological model combined with a mechanistic pH model.

The scale-up strategy developed in this chapter is applicable for most aerobic cultures with

low viscosities.

Page 80: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 5 A screening approach for immobilized biomass on solid supports 62

Chapter 5

Development of a screening approach for immobilized biomass on solid supports

This chapter introduces and characterizes a simple and inexpensive method to perform

immobilization studies with solid supports using standard histology and microbiology

laboratory equipment. The new procedure should provide reproducible conditions

independent of the properties of the support. Most of the results presented in this chapter have

been described by Seletzky et al. [2006b].

5.1 Introduction

The variety of solid supports used for immobilization studies range from synthetic materials

over metals, ceramics, and glass to bones. Many supports cannot be placed directly in the

reaction vessel due to the following possible problems: the supports may float on the surface

or simply rest on the bottom of the reaction vessel. The support can erode through the power

input necessary to mix the solid and the liquid phase as well as to assure the required mass-

transfer. In case of very hard supports their movement may damage the reaction vessel. To

protect support, and reaction vessel, and to standardize the hydrodynamic conditions, the

support has to be placed in a carrier. In this work, histology-cassettes also referred to as

embedding-cassettes were used as carriers. Histology-cassettes have been used for many years

in histopathology laboratories to prepare biological specimens for histological observations.

They are commonly available, inexpensive, solvent and acid resistant, automatable, and the

slots in the cassette walls allow the liquid to circulate freely through the carrier.

Page 81: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 5 A screening approach for immobilized biomass on solid supports 63

As standard reaction vessels shake flasks were chosen because they are easy to handle,

inexpensive, well established, characterized (Chapter 4.1), and several experiments can be

performed simultaneously. The procedure introduced in this chapter focuses on

immobilization of biomass; however, it is also adaptable for enzymic or inorganic catalysts.

Cultivations in the batch mode can be used to study the initial adhesion of microorganisms to

supports. However, batch cultures cannot provide the long and constant operating conditions

required to study the formation of biofilms. Such operating conditions can be obtained by

continuous cultivations in the COSBIOS device (Continuously Operated Shaking Bioreactor

System) introduced by Akgün et al. [2004] [Appendix A.2.4]. The respiratory activity of the

shake flask cultures can be monitored using the RAMOS device (Respiration Activity

Monitoring System) [Chapters 3, A.3.2, Anderlei and Büchs 2001, Anderlei et al. 2004,

Seletzky et al. 2006c].

The focus of this chapter is to establish standard reaction conditions and to characterize the

hydrodynamic and mass transfer characteristics of histology-cassettes with supports in shake

flasks.

5.2 Materials and methods

At high shaking frequencies, it is difficult to observe the fluid motion, especially in small

shake flasks. Therefore, a novel rotating camera (Fig. 5.1) [Lotter 2003] has been applied to

visualize the movement of the shaken fluid. The rotating camera produces a stable image of

the liquid in the shake flask, enabling an analysis of the movement and the position of the

histology-cassette and the bulk liquid at different shaking frequencies, shaking diameters and

filling volumes. Fig. 5.1a shows a photograph of the apparatus, Fig. 5.1b a schematic

depiction. Fig. 5.1c depicts the movements of flask, camera and bulk liquid. To visualize the

different motions, the position of a sticker, in the form of an arrow, placed on flask and

Page 82: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 5 A screening approach for immobilized biomass on solid supports 64

camera is shown. A flask on an orbital shaker, performs a translation (tip of the arrow is

always pointed in the same direction), while the centrifugal force causes the bulk liquid to

rotate. Therefore, a camera which rotates (tip of the arrow pointed towards the center) around

the flask, pointing always in the direction of the centrifugal force produces a stable image. To

generate the translation of the flask and the rotation of the camera with the same motor, flask

and camera have to rotate with the same frequency in opposite directions. The inversion of the

rotation is achieved with a gear unit. With a wireless transmission (2.4 GHz video transmitter,

Conrad Electronic GmbH, Hirschau, Germany) the video stream taken by the camera

(VCAM-110, Phytec Technologie Holding AG, Mainz, Germany) is continuously sent to a

receiver. To increase the contrast of the liquid, a solution of 100 µM fluorescein and 0.2 M

phosphate buffer (pH 8.3) was used.

Standard histology-cassettes (UNI-Safe with lid, Medizinische Diagnostik-Methoden GmbH,

Germany) were applied. As support material non-porous glass sheets and porous

recrystallized silicon carbide sheets were used (Appendix A.4). Biological experiments were

carried out with Corynebacterium glutamicum. The preparation of medium and preculture are

as described in appendix A.1.2. Conditions for batch cultures: initial lactate concentration

8 g/l, initial pH 7. Conditions for continuous cultures: dilution rates 0.1-0.8 h-1, lactate

concentration 5 g/l, medium pH 6.5, this resulted in a pH of 7 (Chapter 4.3.2) under steady

state conditions. The continuous cultures were performed with a COSBIOS device as

described in appendix A.2.4.

Page 83: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 5 A screening approach for immobilized biomass on solid supports 65

Fig. 5.1: The rotating camera produces a stable image

of the shake flask, allowing continuous observations of

the bulk liquid and the position of the histology-

cassette. A. Photograph, B. Schematic depiction side

view, C. Schematic depiction top view. The movement

of shake flask and camera are symbolized by an arrow

shaped sticker.

A

translation of flask

C

rotation of liquid

movement of flask

movementof camera

rotation of camera

300min-1

wireless transmission

camera

shaking radius

shake flask

weight

table

motorcomputer

gear unit

B

M coupling

Page 84: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 5 A screening approach for immobilized biomass on solid supports 66

To determine the influence of the histology-cassette on the oxygen transfer capacity

(OTCmax), the maximum oxygen transfer rate was determined with a set of shaking

conditions, applying an oxygen limited biological culture and the chemical sulfite-oxidation

method. The oxygen transfer rate (OTR) of the biological cultures was determined with a

RAMOS device as described in appendix A.3.2.

The optical sulfite-oxidation method described by Hermann et al. [2001] was applied. The

reaction is catalyzed by a number of metal ions such as Co2+, Cu+, Fe+, and Mn+ [Linek and

Vacek 1981], which may be present as impurities in the support material. To verify that the

reaction rate of the sulfite-oxidation method is not influenced by the material of the support,

100 mg of ceramic powder [SiC (Heimbach GmbH, Düren, Germany), SiC (Ibiden Europe

BV, Paris, France); cordierite (Rauschert, Kloster Veilsdorf, Veilsdorf, Germany)] or 100 mg

of glass powder [quartz (2.5 m²/g BCR 172, Fluka, Seelze, Germany); silicic acid (288772

particle size 20 µm, Sigma-Aldrich Chemie GmbH, Munich, Germany)] were added to the

sulfite solution. Powder and sulfite solution form dispersion, creating a huge contact area

without disturbing the hydrodynamics of the liquid.

The influence of the histology-cassette on the oxygen transfer capacity (OTCmax) was

determined with the oxygen sulfite method by comparing the OTCmax of flasks without

histology-cassette with flasks with a glass or silicon carbide sheet. The conditions of the

experiment were: 0.5 M Na2SO3, temperature 26.5°C, 250 ml shake flask without a cotton

plug, shaking frequency 300 rpm, shaking diameter 50 mm, filling volume 25 ml. The

evaporation was prevented by wrapping the entire shaker in a plastic bag. The bag was

aerated with humidified air.

Page 85: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 5 A screening approach for immobilized biomass on solid supports 67

5.3 Results and discussion

To characterize the behavior of the liquid and the histology-cassette in shake flasks and

beakers at different shaking frequencies, filling volumes, flask sizes, and weights of the

support, pictures taken with the rotating camera (Fig. 5.1) were analyzed. Fig. 5.2 depicts

position and movement of the liquid (25 ml) and the histology-cassette in a 250 ml shake

flask at different shaking frequencies. At a shaking frequency of 50 rpm, the liquid

completely covers the flask bottom. The histology-cassette tumbles over the bottom of the

flask performing a random movement independent of the liquid. At a frequency of 100 rpm,

the flask bottom is still completely covered. The histology-cassette starts to follow the

movement of the liquid. At a shaking frequency of 150 rpm, the center of the flask bottom is

no longer covered by the liquid. The liquid forms a sickle in the shape of a crescent at the

flask wall; the histology-cassette rotates with the liquid and leans in a flat angle between the

flask bottom and the flask wall. At frequencies between 150 rpm and 250 rpm, an increase of

the shaking frequency causes the bottom of the flasks to be less and less covered by the liquid.

The maximum liquid height of the sickle increases and the histology-cassette rotates with the

liquid. At all shaking frequencies between 50 rpm and 250 rpm, the histology-cassette is only

partially covered with liquid. Between a shaking frequency of 250 rpm and 300 rpm, the

histology-cassette changes into a vertical position. The histology-cassette moves along the

flask wall in the center of the sickle leaning only against the flask wall. In this position, it is

completely covered by the liquid. At shaking frequencies exceeding 300 rpm, only the

maximum liquid height of the sickle increases further. The general position and movement of

the liquid and the histology-cassette remain unchanged. If the shaking frequency is reduced,

the histology-cassette stays in the upright position up to a frequency of ~150 rpm, and then

changes into a horizontal position. Filling volumes of less than 25 ml are not sufficient to

cover the cassette with liquid. To assure that the histology-cassette is entirely covered with

liquid in a 250 ml shake flask at 300 rpm, the filling volume should exceed 25 ml. At higher

Page 86: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 5 A screening approach for immobilized biomass on solid supports 68

filling volumes, however, the oxygen transfer capacity (OTCmax) may not be sufficient to

support non oxygen limited growth of aerobic bacteria [Chapter 4.5.1.2, Seletzky et al. 2007].

For reactions, which do not consume oxygen, a shaking frequency of 150 rpm and a filling

volume of 40 ml are sufficient to mix the liquid and to wet the histology-cassette.

Up to a load of ~5 g the weight of the support does not influence the movement of the

histology-cassette. There is also no minimum loading, even without support the histology-

cassette rotates with the liquid and changes into the vertical position at shaking frequencies

between 250 rpm and 300 rpm. Heavier loads up to 12 g (cassette filled with lead) rotate

freely, however, the histology-cassette doesn’t change into the vertical position. It stays in the

position shown for 250 rpm in Fig. 5.2a.

In shake flasks smaller than 250 ml, the histology-cassette cannot move freely. It can neither

change into the vertical position, nor rotate with the liquid. Shake flasks larger than 250 ml

may be used, although in order to wet the histology-cassette, the filling volume has to be

increased.

In vessels with vertical walls, such as beakers or laboratory bottles, the histology-cassette

shows varying behavior, sometimes the histology-cassette changes into the vertical position,

sometimes it doesn’t (Fig. 5.2b). If the histology-cassette does not change into the vertical

position, it is hardly wetted by the liquid. The liquid forms a film along the wall and the

cassette rotates on the bottom.

It is possible to place several histology-cassettes in one large shake flask. The cassettes rotate

with the liquid and align one behind the other. However, depending on the position of the

histology-cassette the liquid covers the cassettes differently (Fig. 5.2c), limiting the

comparability of the experiments.

Page 87: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 5 A screening approach for immobilized biomass on solid supports 69

Fig. 5.2: Observations with the rotating camera: Position and movement of histology-cassettes and

bulk liquid in shake flasks and beakers. A. Histology-cassette with a glass sheet in a 250 ml shake

flask, shaking frequency from 50 rpm to 300 rpm, shaking diameter 50 mm, filling volume 25 ml.

B. Histology-cassette in a 400 ml beaker with an inner diameter of 77 mm, both pictures had equal

shaking conditions, shaking frequency 300 rpm, shaking diameter 50 mm, filling volume 30 ml.

C. Three histology-cassettes in a 1000 ml shake flask, shaking frequency 200 rpm, shaking diameter

50 mm, filling volume 100 ml. The center of the flask or beaker bottom is zero of the scale. Liquid

0.2 M phosphate buffer with 100 µM fluorescein, pH 8.3.

Without modifying the standard method of the COSBIOS device, it was possible to

continuously cultivate C. glutamicum with histology-cassette for two weeks. The suspended

biomass could be described with a biomass dilution rate (X-D) diagram and with SEM

microscopy a biofilm, which covered the support, was observed (results not shown).

Page 88: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 5 A screening approach for immobilized biomass on solid supports 70

To determine the influence of the histology-cassette on the gas-liquid mass transfer in shake

flasks, the oxygen transfer capacity (OTCmax) obtainable with a constant set of shaking

conditions, was determined by measuring the oxygen transfer rate OTR of a biological culture

under oxygen limited conditions (Fig. 5.3a) and by determining OTCmax with the chemical

sulfite-oxidation method (Fig. 5.3b). Fig. 5.3a depicts the OTR over fermentation time of

Corynebacterium glutamicum cultures on lactate with and without histology-cassette. After

inoculation, the OTR increases exponentially until OTCmax is reached. Subsequently the OTR

stays at a plateau until the carbon source is exhausted and then drops sharply. The histology-

cassette neither influences the height of the plateau, nor the fermentation time until the

depletion of the carbon source. Fig. 5.3b compares the OTCmax of shake flasks with and

without histology-cassette at different filling volumes, using the sulfite-oxidation method.

With increasing filling volume the OTRmax decreases, however, it is not influenced by the

presence of a histology-cassette. An influence of the support on the reaction rate of the sulfite-

oxidation method can be excluded. The OTCmax of experiments without histology-cassette

was not influenced by the addition of various ceramic and glass powders to the sulfite

solution (results not shown). In conclusion, the addition of histology-cassettes to liquid

cultures in shake flasks does not influence the gas-liquid mass transfer of biological cultures

or the sulfite-oxidation method.

Page 89: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 5 A screening approach for immobilized biomass on solid supports 71

Fig. 5.3: The influence of the histology-cassette on

the oxygen transfer capacity. A. Oxygen transfer

rate in a biological culture: without a histology-

cassette, X with a histology-cassette filled with a

porous silicon carbide sheet. Culture conditions:

organism C. glutamicum, carbon source 8 g/l lactic

acid, initial pH 7, temperature 30° C, shaking

frequency 300 rpm, shaking diameter 50 mm, filling

volume 25 ml. B. Oxygen transfer capacity

determined with the optical sulfite oxidation method

[Hermann et al. 2001]: without an histology-

cassette, + with an histology-cassette filled with a

glass sheet, with an histology-cassette filled with

a porous silicon carbide sheet. The total consumption of sulfite was visually determined by evaluating

the color change of the pH indicator on recorded CCD camera images. Conditions: 0.5 M Na2SO3,

temperature 26.5° C, shake flask 250 ml without cotton plug, shaking frequency 300 rpm, shaking

diameter 50 mm, filling volume 20 ml, 30 ml, 40 ml.

10 20 30 40 500.01

0.02

0.03

0.04

oxyg

en tr

ansf

er c

apac

ity

[mol

/l/h]

filling volume [ml]

0 2 4 6 8 10 12 14 16 18 200.00

0.01

0.02

0.03

0.04

B

A

oxyg

en tr

ansf

er ra

te

[mol

/l/h]

fermentation time [h]

Page 90: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 72

Chapter 6

The influence of material properties and surface modifications on the adhesion of dormant and growing cells

This chapter compares the adhesion properties of dormant and growing cells on glass and

ceramic supports to analyze whether the screening results gained with dormant cells in

artificial salt solutions can be used to predict the adhesion behavior of growing cells in

medium.

6.1 Introduction

Adhesion experiments are mostly performed in artificial salt or buffer solutions [Bendinger et

al. 1989, Li and Logan 2004, Sternström 1989, van Loosdrecht et al. 1987], in which the

dormant cells have to survive without nutrients. However, biofilm reactors are operated with

growing cells in medium. Adhesion is influenced by the properties of the microorganisms as

well as by the characteristics of the support material. The adhesion properties vary between

different species, even when they are closely related [Bendinger et al. 1993, Rijnaarts et al.

1993, van Loosdrecht et al. 1987]. They are influenced by the medium composition [Büchs et

al. 1988] and the growth phase [Chavant et al. 2002, Walker et al. 2005]. For coryneform

bacteria, Bendinger [1993] correlates adhesion with the chain length of the mycolic acids at

the cell surface. He observed that the cells get more hydrophobic with increasing chain length.

Generally, hydrophobic bacteria adhere to a greater extend than hydrophilic bacteria [An and

Friedman 1998, Rijnaarts et al. 1993, Sternström 1989, van Loosdrecht et al. 1987], and

Page 91: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 73

adhesion is increased by hydrophobic support materials [An and Friedman 1998, Li and

Logan [2004], van Loosdrecht et al. 1990].

The adhesion properties of metal oxides, an important element of glass and ceramics, has only

been peripherally in the focus of research. Li and Logan [2004] observed, that iron (III) oxide

(Fe2O3) and aluminum oxide (Al2O3) had better adhesion properties than silicon oxide (SiO2)

the main compound of glass. However, he could not definitely exclude effects resulting from

different surface roughnesses.

Adhesion is not only influenced by the chemical properties of the support material, but also

by structure, porosity and pore size. According to theoretical considerations of Wang and

Wang [1989] a pore size of two to five fold the cell diameter gives maximum adhesion. This

was experimentally confirmed by Samonin and Elikova [2004]. Messing et al. [1979] found

that a material, which has the major part of the pores between 1 µm and 30 µm, is best for the

adhesion of E. coli (D= 1-6 µm).

Adhesion increases with the specific surface area of the support. Support materials with a high

porosity and a small pore size should therefore increase adhesion. However, the pore size has

to be large enough to accommodate the cells. Thus, not the total surface area of a support

material is important for adhesion, but only the surface area formed by pores lager than the

minimum pore diameter accessible to cells [Samonin and Elikova 2004]. If this is not

considered, the specific surface area of a support material will not correspond to the adhesion

properties [Kida et al. 1990, Samonin and Elikova 2004]. For organisms which multiply by

fission, pore sizes larger than ~2-3 fold the cell size should allow the cells to reproduce inside

the pores and to move freely in and out of the pores.

To speed up biofilm formation, which can take several weeks [Léon Ohl et al. 2004], the

initial adhesion can be increased by modifying the surface of the support materials with

Page 92: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 74

polymer-mediated interactions [Rutter 1980]. As polymers polyethylenimin applied by Melo

and D'Souza [1999] and DEAE dextran applied by Büchs et al. [1988] were used.

The influence of a solid surface on activity, substrate conversion, and growth rate of

microorganisms is still disputed. According to the review of van Loosdrecht et al. [1990], the

presence of a solid surface may increase, decrease or not affect the activity of

microorganisms. The review revealed, that most activity changes attributed to surfaces can be

explained with indirect effects like different media composition for attached and suspended

cells (cells not attached to the support) or with the experimental set-up.

In this study, the commercially important amino acid producer [Kircher and Pfefferle 2001]

Corynebacterium glutamicum was used as model organism. It has been previously used as

model organisms for biofilm reactors and for adhesion studies by Büchs et al. [1988], Henkel

et al. [1990], Kim and Ryu [1982], and Park and Jeong [2001].

6.2 Materials and methods

6.2.1 Microorganism, media, and culture conditions

Experiments were carried out with the wild type of Corynebacterium glutamicum

ATCC 13032. Culture conditions media and preculture preparation were performed as

described in appendix A.1.2. The respiration of the microorganisms in shake flasks was

measured with a RAMOS device as described in appendix A.3.2. The biomass concentration

was determined as described in appendix A.5.1.

The influence of the residence time in a NaCl solution on the activity of Corynebacterium

glutamicum was tested with a RAMOS device by inoculating the main culture with cells from

a twice washed preculture, which was left, after washing and before inoculation under

continuous stirring for 0 h (~1 h in NaCl solution during the washing procedure of the

Page 93: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 75

preculture), for 1 h, 3 h, 6 h, 15 h in a 9 g/l NaCl solution. 8 ml minimal medium (lactic acid

concentration 10 g/l) was inoculated with 2 ml NaCl solution resulting in a lactic acid

concentration of 8 g/l. The NaCl solution had an OD600 of 3 equal to the OD of the adhesion

experiments with dormant cells. The operating conditions of the orbital shaker (Lab-Shaker

LS-W, Adolf Kühner AG, Birsfelden, Switzerland) used for the activity test were a shaking

diameter of 50 mm, a shaking frequency of 300 rpm, and a filling volume of 10 ml.

The hydrophobicity of Corynebacterium glutamicum was determined by measuring the

contact angel with a goniometer microscope (G402, Krüss GmbH, Hamburg, Germany) as

described by van Loosdrecht et al. [1987]. Cells were taken from precultures washed once

with 1 M NaCl then collected on a cellulose acetate membranes with a pore size of 0.2 µm

(Sartorius AG, Göttingen, Germany). The reported contact angel for a 0.1 M (5.85 g/l) NaCl

solution on a bacterial layer is a mean of 10 independent measurements.

6.2.2 Supports and surface modifications

6.2.2.1 Support materials

As glass support materials for adhesion experiments non-porous glass microscope slices and

porous glass filter discs of borosilicate glass (porosities 2, 3, 4) were used. As ceramic

supports porous silicon carbide and porous cordierite was used. Further properties of the

support materials are given in Tab. 6.1 and in appendix A.4.

6.2.2.2 Surface modifications

Prior to the surface modification, the surfaces were activated by placing the supports in nitric

acid (50 g/l). The supports were evacuated for 15 min and then cooked for 4 h. The activated

supports were rinsed with deionized water for 12 h, dried at 100° C and stored in an airtight

container.

Page 94: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 76

The activated supports were modified with polyethylenimin with an average molecular weight

of ~750,000 (500 g/l aqueous solution, Sigma-Aldrich Chemie GmbH, Munich, Germany)

and with DEAE dextran hydrochloride with an average molecular weight of ~500,000 and a

nitrogen content of ~3% (personal communication of the supplier) (Fluka, Seelze, Germany).

The surface modifications with polyethylenimin were carried out with a 5 g/l polyethylenimin

solution. The supports were placed in a beaker with stirrer evacuated for 15 min and then left

for 12 h with continuous stirring. Erosion of the supports was inhibited by placing them on a

sieve. The modified supports were rinsed with deionized water for 12 h, dried at 75° C and

stored in an airtight container. The surface modifications with DEAE dextran were carried out

as described by Büchs et al. [1988].

6.2.2.3 Characterization of support materials and surface modifications

The influence of the surface modifications on the hydrophobicity was determined by

measuring on microscope slices the contact angel of 0.1 M (5.85 g/l) NaCl solution with a

goniometer microscope (G402, Krüss GmbH, Hamburg, Germany). The contact angles given

are a mean of 10 measurements.

Density, porosity, median pore diameter, and pore diameter distribution were determined with

a mercury high pressure intrusion porosimeter (Autopore II 9220, Micromeritics GmbH,

Mönchengladbach, Germany) according to ISO 15901-1. For all materials a mercury contact

angle of 141.3° and a mercury surface tension of 0.485 N/m was considered. A cylindrical

pore model was assumed to calculate porosities, and pore diameters using the Washburn

equation [Garboczi and Bentz 1991, Ilavsky et al. 1997]. Scanning electron microscope

images of platinum sputtered supports were taken with a (JSM 6400, Jeol GmbH, Munich,

Germany, measurements IKKM RWTH-Aachen) using an acceleration voltage of 15 kV.

Page 95: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 77

6.2.3 Adhesion experiments

6.2.3.1 Adhesion of dormant cells

The experimental procedure used to determine the adhesion of dormant cells is depicted in

Fig. 6.1. A single stock solution with an optical density (OD600) of 3 was prepared by diluting

stationary phase cells (cultivation time 12 h) harvested from complex medium cultures in a

NaCl solution (9 g/l). In preliminary experiments, (results not shown) an OD of 3 has proven

best to observe adhesion. The OD did neither fall to zero, nor was it too high to observe

adhesion. 30 ml of the stock solution are distributed to 50 ml beakers, which were covered

with aluminum foil to prevent evaporation. The beakers equipped with a stir bar (20 mm

length, 0.6 mm diameter, VWR international, Wien, Austria) were placed on a multistirrer

(stirrer speed 300 rpm) (Variomag Poly 15, H+P Labortechnik AG, Oberschleißheim,

Germany). To protect the supports from the movement of the stir bar, they were placed on a

rubber ring (41 mm diameter, 27 mm inner diameter, 3 mm thickness). The adhesion was

measured by recording the decrease of the OD over time. For each measuring point, the

average of two samples with a sample volume of 1 ml was calculated. The OD was converted

to biomass concentration using an OD dry weight correlation. To determine for each

experiment the optical density dry weight correlation, the biomass concentration of the stock

solution was determined with 30 ml as described in appendix A.5.1. In a final step the specific

adhered biomass (X/A) was calculated by multiplying the difference between the initial

biomass concentration of the stock solution and the biomass concentration of the sample with

the liquid volume and by dividing it by the external surface area (not the pore surface area) of

the support. With reference experiments the adhesion of beaker walls, rubber ring, and stir bar

were quantified. To simplify the comparison of reference and supports, the specific adhered

biomass of the supports (X/A) depicted in Fig. 6.6 still include the adhesion of the reference

(beaker walls, rubber ring, stir bar) and the specific adhered biomass of the reference is

Page 96: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 78

related to the surface area of a standard support (A=0.0014 m²) and not to the surface area of

the beaker walls. Thus, if the specific adhered biomass of a support and the specific adhered

biomasses of the reference are in the same range the observed decrease of the OD can be

mainly attributed to the adhesion of the reference. All adhesion experiments presented in this

work were at least double determined.

time

OD

sample OD determination

cultivation Corynebacterium

glutamicum

10 x parallel

time

result IIspecific adhered biomass

(X/A) over time

stock solution OD ~3

9 g/l NaCl

X/A

distributing

washing centrifuging

diluting OD/X

correlation

9 x parallel1 x

surface modified supports

support surface area (A)

result Ioptical density

(OD) over time

adhesionsupport on

protection-ring inbeaker on multi-

stirrer

pre-culture

Fig. 6.1: Schematic depiction of the operating procedure to determine the adhesion of dormant cells.

6.2.3.2 Adhesion of growing cells

The experimental procedure used to determine the adhesion of growing cells is depicted in

Fig. 6.2. Standard 250 ml shake flasks and RAMOS measuring flasks [Anderlei and Büchs

2001, appendix A.3.2], some equipped with an histology-cassette (UNI-Safe with lid,

Medizinische Diagnostik-Methoden GmbH, Germany), which contained a glass or ceramic

support, were autoclaved. This screening method and the behavior of histology-cassettes in

shake flasks is described in detail in chapter 5. The sterile flasks were filled from a single

inoculated minimal medium with a lactic acid concentration of 8 g/l, which was prepared by

mixing the pellets of several precultures. Respiration and biomass growth are determined by

measuring the oxygen transfer rate with a RAMOS device and the optical density (OD). For

each OD measuring point the average of two samples with a sample volume of 0.1 ml was

Page 97: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 79

calculated. The biomass concentration was calculated from the OD using a previously

determined OD dry weight correlation. The specific adhered biomass (X/A) was calculated by

multiplying the biomass concentration difference between the culture without support

(reference) and a culture with support with the filling volume of the flask and by dividing by

the external surface area of the support. The operating conditions of the orbital shaker were a

shaking diameter of 50 mm, a shaking frequency of 300 rpm, a filling volume of 25 ml.

Fig. 6.2: Schematic depiction of the operating procedure to determine the adhesion of growing cells.

6.3 Results

6.3.1 Characterization of material properties and microorganisms

perties of the glass and ceramic materials. The properties of

sample OD determination

inoculated medium

Corynebacterium glutamicum

time

result II specific adhered

biomass (X/A) over time

X/A

difference biomass with and without

support

modified supports

surface support surface area (A)

result Ioptical

density (OD) over time

adhesion& cultivation

with and without supports

pre-culture

1 x time

OD

distributing

8 xparallel

Cultivation conditions: shake flask 250 ml, shaking diameter 50 mm, shaking frequency 300 rpm, filling

volume 25 ml, lactic acid concentration 8 g/l.

6.3.1.1 Material properties

Tab. 6.1 lists the material pro

glass materials and cordierite showed a high repeatability with variations of less than 10 %.

The properties of silicon carbide (SiC) showed noticeably higher differences. All

Page 98: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 80

measurements were performed at least three times. The porosity of the porous glass materials

lies around 30 % (Tab. 6.1). It decreases with increasing median pore diameter. Cordierite has

with 22 % the lowest, SiC with 35-55 % the highest porosity. The median pore diameter of

the porous glass materials lies between 15 µm and 85 µm. Cordierite has a small median pore

diameter of 1.7 µm. The median pore diameter of SiC (43-55 µm) lies between glass porosity

2 and glass porosity 3. The bulk density, which includes the pores, lies between 1.4 kg/dm³

and 2.4 kg/dm³. It decreases with increasing porosity.

Fig. 6.3 depicts the distribution of the pore diameters. The measured pore size distributions

ies of the support materials used for adhesion experiments. For all porous materials,

microscope

porosity 4 porosity 3 porosity 2 silicon

carbide

(square brackets) for the porous glass materials are similar to the specifications of the

manufacturer (round brackets): porosity 4 [10-16 µm] (10-16 µm), porosity 3 [25-32µm]

(16-40 µm), porosity 2 [50-80µm] (40-100 µm). The pores of cordierite are between 1 µm

and 2 µm. SiC had pores between 32-65 µm and between 190-310 µm. This wide spectrum of

pore sizes for SiC could be an explanation for the observed differences in porosity and mean

pore diameter.

Tab. 6.1: Propert

porosity, median pore diameter, and density have been determined with a mercury high pressure

intrusion porosimeter (Autopore II 9220, Micromeritics GmbH, Mönchengladbach, Germany), assuming

cylindrical pores. Scanning electron microscope images of platinum sputtered materials were taken with

a (JSM 6400, Jeol GmbH, Munich, Germany) using an acceleration voltage of 15 kV.

glass glass glass glass cordierite

slice

100 µm

SEM images

porosity 0 % 36 % 32 % 27 % 22 % 35-55 %

median

0 µm 15 µm 28 µm 85 µm 1.7 µm 43-5 µm pore diameter

5

bulk density

2.4 kg/dm³ 1.4 kg/dm³ 1.5 kg/dm³ 1.6 kg/dm³ 2.0 kg/dm³ 1.2-1.4 kg/dm³

100 µm

100 µm 100 µm 100 µm

100 µm

Page 99: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 81

Fig. 6.3:

different support materials determined with a

o

layers of Corynebacterium glutamicum. The cont

Fig. 6.4: Hydrophobicity of surface

non porous soda lime glass microscope

eglutamicum

Fig. 6.5 shows the influence of the residence time in a NaCl solution on the activity of

Corynebacterium glutamicum. With increasing residence time in the NaCl solution prior to

0.5 1 5 10 50 100 5000.00

0.02

0.04

0.06

0.08

0.10

0.14 Distribution of the pore diameter for

mercury high pressure intrusion porosimeter

(Autopore II 9220, Micromeritics GmbH,

Mönchengladbach, Germany). cordierite,

glass porosity 4, glass porosity 3,

glass porosity 2, silicon carbide.

lution on microscope slices and bacterial

act angle of microscope slices without

Fig. 6.4 depicts the contact angle of 0.1 M NaCl s

0.12minimumdiameter

foradhesiom

incr

emen

tal v

olum

e [

ml

/g]

pore diameter [µm]

Hg

surface modification is 24.1° with a standard deviation of (σ=2.8°); it increases through the

surface modification with DEAE dextran to 31.7° (σ=5.2°) and to 45.3° (σ=2.7°) through the

modification with polyethylenimin. Contact angle and standard deviation of C. glutamicum is

with 42.5° (σ=2.4°) in the same magnitude as the surface modified microscope slices. Glass

surfaces as well as the microorganisms are hydrophilic. The glass surfaces get more

hydrophobic through the modification with the polymers DEAE dextran and polyethylenimin.

modified (DEAE dextran, polyethylenimin)

without DEAE Ethylenimin Coryne0

10

20

30

40

50

cont

act a

ngel

]st

anda

rd d

evia

tion

[°]

slices and Corynebacterium glutamicum.

Contact angels were determined with

0.1 M NaCl solution as described by van

Loosdrecht et al. [1987].

activity of Corynebacterium 6.3.1.2 Influence of the residence time in NaCl on th

Page 100: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 82

inoculation, the oxygen transfer rate (OTR) and its slope decrease directly after inoculation.

The maximum oxygen transfer rate (OTR ) reached by the cultures and the slope during the

last three hours before OTR were not influenced by the residence time in the NaCl

solution, allowing the fermentation time until OTR (t ) to be used as a measure for the

activity loss of the precultures. t is the transition point to the stationary phase at which

ide

0-15 h). Thus, the activity of the cells in the NaCl solution decreases linearly with the

residence time in the NaCl solution. If the same linear reduction of the activity also occurs

during the preparation (washing and centrifuging with NaCl) of the preculture, the preparation

would increase the fermentation by about 0.5 h.

of Corynebacterium glutamicum. Residence

time in a NaCl solution after washing and

,

the inoculated medium 8 g/l, initial pH 7. Nomenclature

fermentation time until OTRmax.

0 4 8 12 16 2

max

max

max OTRmax

OTRmax

the organism has completely consumed the carbon source [Chapter 2.3.1, Seletzky et al.

2006a]. A linear correlation between tOTRmax and the res nce time in the NaCl solution

(tNaCl) was observed (tOTRmax=0.5·tNaCl+11.6, coefficient of determination r²>0.99, tNaCl

Fig. 6.5: Influence of the residence time cells

were left in a 9 g/l NaCl solution on the activity

00.00

0.01

0.02

0.03

0.04OTR

OTRmaxmaxt

oxyg

en tr

ansf

er ra

te

[mol

/l/h]

fermentation time [h]

before inoculation of a minimal medium culture:

0 h (~1 h in NaCl solution during the

washing procedure of the preculture), 1 h,

3 h, 6 h, 15 h. Culture conditions:

filling volume 10 ml, lactic acid concentration of

: OTR

shaking frequency 300 rpm, shaking diameter 50 mm

max maximum oxygen transfer rate, tOTRmax

Page 101: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 83

6.3.2 Adhesion experiments

6.3.2.1 Adhesion of dormant cells

Fig. 6.6 depicts the influence of different glass and ceramic materials with various surface

modifications on the adhesion of dormant cells dissolved in a NaCl solution. Generally,

independent of material or surface modification the adhesion slows down with increasing

adhesion time. Adhesion was strong during the first 3 hours and came to a halt after 6 to 7

hours. The adhesion of the reference (results not shown) had a good repeatability. About

1±0.2 g/m² of the observed specific adhesion of the supports can be attributed to the adhesion

of the reference. Fig. 6.6a compares the adhesion of the reference (bold solid line, fit curve of

the experimental data) with the adhesion of the non porous microscope slices (symbols). The

adhesion of the microscope slices is only slightly higher than the adhesion of the reference,

thus the observed adhesion can be attributed mainly to the adhesion of the reference. The

adhesion of glass supports with different porosities and surface modifications is depicted in

Figs. 6.6b, c, e. To demonstrate the repeatability of the method, the depicted data points for

one type of support are taken from two independent measurements. Independent of the

porosity it can be observed that the adhesion on porous glass supports is noticeably higher

than the adhesion on the reference. The adhesion of DEAE dextran modified supports is two

to three times higher than the adhesion of supports without surface modification.

Polyethylenimin gives an intermediate adhesion. The adhesion on glass with porosity 4

(Fig. 6.6b) and on glass with porosity 3 (Fig. 6.6c) is for all surface modifications similar. The

adhesion on glass with porosity 2 (Fig. 6.6e) is lower, however, it is still much higher than the

adhesion of non porous glass supports (Fig. 6.6a). The adhesion of the ceramics is depicted in

Figs. 6.6d, f. The observed adhesion of cordierite (Fig. 6.6d) is only a little higher than the

adhesion of the reference, thus, the adhesion can be mainly attributed to the adhesion of the

reference. However, the stronger adhesion of DEAE modified supports can still be observed.

Page 102: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 84

The adhesion of silicon carbide (SiC) (Fig. 6.6f) is in the same range as the adhesion of glass

with porosity 2, also the influence of the surface modifications is similar to the glass supports.

In contrast to all other materials, the adhesion of SiC showed strong variations (results not

shown). In some experiments, the final specific adhered biomass for DEAE dextran modified

supports was only 3 g/m² or the adhesion of polyethylenimin modified support exceeded the

adhesion of DEAE dextran modified supports.

Page 103: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 85

0

1

2

3

4

5D

EB

C F

A

adhe

red

biom

ass

[g/m

²]

0 1 2 3 4 5 6 7 8

adhesion time [h]0 1 2 3 4 5 6 7

0

2

4

6

8

adhesion time [h]

adhe

red

biom

ass

[g/m

²]

0

2

4

6

8

adhe

red

biom

ass

[g/m

²]

Fig. 6.6: The influence of surface modifications and porosities on the adhesion of dormant cells on

glass and ceramic supports. A. glass supports without porosity (microscope slices), B. glass supports

with porosity 4, C. glass supports with porosity 3, D. cordierite supports, E. glass supports with

porosity 2, F. silicon carbide supports. without surface modification, modified with

polyethylenimin, modified with DEAE dextran. A.-F. (bold solid line) adhesion of the reference

(beaker walls, rubber ring, stir bar). The adhesions of the supports includes the adhesion of the

reference. Conditions: continuously stirred 50 ml beaker with 30 ml NaCl solution with an initial OD

of 3. Experimental procedure as depicted in Fig. 6.1.

6.3.2.2 Adhesion of growing cells

Fig. 6.7 depicts the behavior of growing cells in cultures without supports (reference open

symbols) and of cultures with supports (solid symbols) for silicon carbide (left) and cordierite

Page 104: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 86

(right). The respiration of the cultures (Figs. 6.7a, d) (results not shown for the glass

materials) did not depend on the support material. Small differences between the oxygen

transfer rate of the different cultures showed no correlation to the properties of the supports.

The biomass concentration of the batch cultures increases exponentially (Figs. 6.7b, e). A

detailed analysis of the slopes and yield coefficients of Corynebacterium glutamicum batch

cultures is given in chapter 2.3.3 and chapter 3.3.1. Generally, for all materials (results not

shown for the glass supports) the adhesion of cells reduced the slope of the biomass

concentration, the reference always reached the highest final biomass concentration, the

biomass concentration differences between cultures with and without support were small (less

than 0.5 g/l) and the suspended biomass concentration was always much higher than the

amount of biomass adhered on the supports.

Fig. 6.8 depicts the specific adhered biomass over fermentation for glass supports with

different porosities and surface modifications. The specific adhered biomass (Figs. 6.7c, f and

Fig. 6.8) generally increased slowly for the first 6 h, then increased more rapidly for 3 to 4 h

and finally reached a decreasing plateau. The influence of surface modifications on the

specific adhered biomass of the ceramic supports (Figs. 7c, f) and of the porous glass supports

(Fig. 6.8) was similar to the observations made with dormant cells. DEAE dextran modified

supports had the highest adhesion, non modified the lowest and polyethylenimin modified

supports showed an intermediate adhesion. For all surface modifications glass supports

(Fig. 6.8) with porosity 4 adhered better than glass supports with porosity 3, which adhered

better than glass supports with porosity 2. The adhesion of SiC lay between the glasses of

porosity 2 and 3 (Fig. 6.7c, Fig. 6.8). Cordierite had the lowest adhesion (Fig. 6.7f).

Page 105: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 87

0

1

2

3

C

B

biom

ass

conc

entr

atio

n [g

/l]

0 2 4 6 8 10 12 140

1

2

3

4

5

6

fermentation time [h]

adhe

red

biom

ass

[g/m

²]

0 2 4 6 8 10 12 14

fermentation time [h]

F

E

D

0.00

0.01

0.02

0.03

0.04A

oxyg

en tr

ansf

er ra

te

[mol

/l/h]

Fig. 6.7: The influence of surface modifications on the adhesion of growing cells on silicon carbide

(left) and cordierite (right). reference without support, support without surface modification,

support modified with polyethylenimin, support modified with DEAE dextran. A., D. oxygen

transfer rate B., E. biomass concentrations C., F. specific adhered biomass. Culture conditions:

shaking frequency 300 rpm, shaking diameter 50 mm, filling volume 25 ml, lactic acid concentration

8 g/l, initial pH 7. Experimental procedure as depicted in Fig. 6.2.

Page 106: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 88

Fig. 6.8: The influence of surface modifications and

porosities on the adhesion of growing cells on glass

supports. A. without surface modification, B. modified

with polyethylenimin, C. modified with DEAE dextran.

porosity 2, porosity 3, porosity 4. Culture

conditions: shaking frequency 300 rpm, shaking

diameter 50 mm, filling volume 25 ml, lactic acid

concentration 8 g/l, initial pH 7. Experimental

procedure as depicted in Fig. 6.2.

1

2

3

4

5

6

7

B

C

A

adhe

red

biom

ass

[g/m

²]

0

2

4

6

8

10

12

14

adhe

red

biom

ass

[g/m

²]

0 2 4 6 8 10 120

2

4

6

8

10

12

14

fermentation time [h]

adhe

red

biom

ass

[g/m

²]

6.4 Discussion

6.4.1 Characterization of material properties and microorganisms

The observed contact angle of 42.5° for C. glutamicum is similar to contact angels reported by

Büchs et al. [1988] (49° for phosphate saturated C. glutamicum, 28° for phosphate depleted

cells). The contact angle of the not surface modified microscope slices (24.1°) is comparable

to the contact angle (22±3°) reported for standard float glass by Li and Logan [2004].

Page 107: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 89

6.4.1.1 Influence of the residence time in NaCl on the activity of Corynebacterium glutamicum

In this study, it was observed, that the activity of C. glutamicum in a NaCl solution decreases

with increasing residence time in a NaCl solution. The decreasing activity could affect the

adhesion properties of the cells comparable to the growth rate [Chavant et al. 2002, Walker et

al. 2005]. The reason for the activity loss could not be specified, however, several stress

factors, which might cause the activity loss can be excluded. As described in chapter 2.3.1

[Seletzky et al. 2006a] carbon starvation in fermentation medium of up to 19 h does not

influence the activity of C. glutamicum. Primm et al. [2000] found no loss in viability during

carbon starvation for the related organism M. tuberculosis for 106 days. It is improbable that

the osmolality of the solution causes the activity loss. As described in chapter 2.3.4 [Guillouet

and Engasser 1995, Seletzky et al. 2006a] C. glutamicum has a high tolerance against osmotic

stress. As described in chapter 2.3.2 [Seletzky et al. 2006a] anaerobic conditions of up to 1 h

do not affect the activity of C. glutamicum. According to Chamsartra et al. [2005], mechanical

stress does not affect C. glutamicum even at high stirrer speeds.

6.4.2 Adhesion experiments

The chemical properties, which may influence adhesion, are similar for all investigated

materials. All have a SiO2 surface layer with a greater or lesser degree of impurities. Thus,

geometrical differences between the support materials like porosity and pore diameter should

influence the adhesion more strongly than the chemical properties. Little adhesion was

observed for support materials with no porosity (microscope slices), or with pore sizes too

small to accommodate the microorganisms (cordierite) (Fig. 6.3). Small pores create a large

surface area, however, if the pores are to small to accommodate the cells, this surface area

cannot be used. Glass with porosities 3 and 4 showed the highest adhesion. These type of

glass have pore sizes between 10 µm and 32 µm. An intermediate adhesion had materials with

Page 108: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 6 Influence of material properties on the adhesion of dormant and growing cells 90

large pores (glass porosity 2, SiC) in comparison to the size of the microorganisms. The

adhesion of SiC showed large variations, probably this reflects the observed differences in the

material properties. According to the theoretical considerations of Wang and Wang [1989], a

pore size of twice to five times the cell diameter gives maximum adhesion. For E. coli, an

organism, which has a similar size as C. glutamicum, Messing et al. [1979] found that to

accumulate biomass a support material should have at least 70 % of the pores between 1 µm

and 30 µm. A material having the major portion of its pores smaller than 1 µm would exclude

the cells, while a material with the predominant number of pores above 30 µm does not have a

high usable surface area. Superimposed on the material properties is the effect of the chemical

surface modifications. It is unlikely that the higher hydrophobicity of the DEAE dextran and

polyethylenimin modified supports is mainly responsible for the increase in adhesion. Even so

according to An and Friedman [1998], Li and Logan [2004], and van Loosdrecht et al. [1990]

a higher hydrophobicity increases adhesion. Probably, the higher amount of binding sites

leads to the increased adhesion. Even so, the adhesion properties of dormant cells in artificial

salt solutions and of growing cells in medium are different, resulting in different adhesion

profiles over time. Materials or surface modifications which showed a high adhesion with

dormant cells, showed also a high adhesion with growing cells. In conclusion, for an initial

screening it can be reasonable to assume that results gained with dormant cells can predict the

behavior of growing cells.

Page 109: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 91

Chapter 7

Characterization of a novel biofilm reactor with a honeycomb monolith as immobilization carrier

This chapter describes the hydromechanical and mass transfer characteristics of a novel

external loop bioreactor. The riser of the reactor consists of a ceramic honeycomb monolith,

which serves as immobilization carrier. To investigate the influence of monoliths on the

performance of loop bioreactors, the reactor was operated with a monolith as riser or

alternatively with a standard single tube.

7.1 Introduction

Monolith reactors [Kawakami et al. 1989, Tata et al. 1999, Wojnowska-Baryla et al. 1999,

Yongming et al. 2002] are proposed as an alternative to conventional packed-bed reactors, for

the immobilization of biomass. The design of the monolith loop reactors can be compared

with the human body. The capillaries inside the monolith deliver, similar to blood vessels,

oxygen and nutrients to the cells, which are immobilized in and on the porous walls of the

monolith. Similar to the blood circulation the liquid is constantly circulated through pumps, or

the difference in density between riser and downcomer (airlift). The intention is to develop an

aerobic biofilm reactor where the entire immobilized biomass can be supplied with oxygen.

Microorganisms, which are immobilized on solid carriers, form a biofilm. The mass transport

inside biofilms is predominantly by diffusion [Stewart 2003, 1998]. Due to mass transport

limitation, the active and aerobic part of the biofilm is relatively thin. A literature analysis

Page 110: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 92

[Bishop et al. 1995, Boessmann et al. 2003, Horn and Hempel 1998, Howell and Atkinson

1976, Mori 1985, Zhang and Bishop 1994] revealed that for air aerated systems the active

aerobic biofilm thickness is nearly always less then 200 µm and often below 100 µm. To

supply oxygen to the entire immobilized biomass, the size of the porous immobilization

carrier should be adapted to the active biofilm thickness. Honeycomb monoliths with a wall

thickness of 200-400 µm are commercially available.

In contrast to packed bed reactors, liquid and air bubbles flow through the monolith in well-

defined capillaries. This design reduces the pressure drop and prevents the forming of inactive

and dry spots. Inactive and dry spots reduce the productivity of the biofilm reactor and may

lead to cooling problems. They are formed in packed bed reactors through the blocking of

channels with biomass and through the accumulation of gas. In summary monolith reactors

may have the following advantages especially if compared to conventional packed bed

reactors [Kawakami et al. 1989, Sattlerfield and Özel 1977, Tata et al. 1999, Wojnowska-

Baryla et al. 1999, Yongming et al. 2002]: (1) high mechanical strength, (2) low pressure

drop, (3) high chemical resistance, (4) better liquid distribution at low liquid flow rates, (5)

even mass transfer inside the monolith, (6) good heat transfer, (7) thin walls, (8) large pore

size and large specific surface area, (9) ideal porous shape allowing for high biomass load,

(10) no dry spots with inactive biomass.

The riser of conventional loop reactors usually consists of a single tube with a diameter of

several centimeters. A monolith consists of many capillaries. For most commercially available

monoliths, the diameter of each capillary is in the range of a millimeter. The change from

laminar to turbulent flow depends on the Reynolds number. The laws, which govern the

transition from laminar to turbulent flow of small diameter tubes or capillaries [Maynes and

Webb 2002], are comparable to the ones in large diameter tubes [Idelchik 1994]. However,

even when the total cross section is comparable, laminar flow is much more likely to occur in

Page 111: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 93

monolith reactors than in conventional reactors with a single tube because the Reynolds

number decreases with smaller tube diameters.

A fermentation broth is usually non-coalescing due to the presence of certain types of surface-

active materials, which tend to stabilize foams when they are formed [Kawase and Moo-

Young 1990]. To control the foam of biological systems antifoam agent is added to the

medium. The addition of antifoam agents causes a significant decrease of the volumetric mass

transfer coefficient (kLa) [Adler et al. 1980, Kawase and Moo-Young 1990, 1987, Koide et al.

1985, Takashi and Yoshida 1979, Yagi and Yoshida 1974]. The antifoam agent reduces the

mass transfer in the following two ways [Kawase and Moo-Young 1990]: (1) interfacial

blockage (increase in mass transfer resistance), (2) hydrodynamic change (suppression of the

mobility of the surface).

7.2 Materials and methods

7.2.1 General description, mode of operation and design of the loop reactor

Fig. 7.1 schematically depicts the design and mode of operation of the novel loop bioreactor.

The air enters the reactor through a porous glass plate. Bubbles and liquid are homogeneously

mixed and flow upwards through a silicon carbide (SiC) honeycomb monolith or alternatively

through a single rectangular tube. In a gas separator the liquid and the gas bubbles are

separated. For continuous biological cultures, the gas separator is additionally used, to add

feed medium, antifoam agent, titrant, and to remove the harvest. Inside the gas separator, the

pH and dissolved oxygen (DO) are measured. From the gas separator the liquid flows into the

downcomer where the temperature is controlled with a heat exchanger. The liquid is then

pumped by two gear pumps into the riser where it is mixed again with the gas bubbles. To

achieve a homogeneous inflow over the entire cross section the liquid flows over a ring into

the riser (Fig. 7.1 bottom). The distribution of the liquid volume inside the reactor is listed in

Page 112: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 94

Tab. 7.1. In contrast to airlift reactors the pumps decouple the liquid flow rate and aeration

rate. At low liquid flow rates the reactor is similar to a bubble column, at intermediate liquid

flow rates it should be comparable to airlift reactors, while at high liquid flow rates it should

behave similarly to jet loop reactors.

The sterile inlet gas stream (0-5 l/min) is controlled by a thermal mass flow controller

(5850TR, Brooks, Hatfield, USA). The aeration of the reactor is performed through a porous

glass filter plate (porosity 0, pore diameter 160-250 µm, Robu Glasfiltergeräte GmbH,

Hattert, Germany). The exhaust gas is cooled with the exhaust gas cooler of a laboratory

fermenter (Biostat M, Braun Biotech, Melsungen, Germany). The composition of the exhaust

gas is measured with an exhaust gas analyzer (Advance Optima, Magnos 106, ABB

Automation, Frankfurt, Germany). The liquid is circulated with two gear pumps

(8200 WM020R, Scherzinger GmbH, Furtwangen, Germany). The volume flow of the gear

pumps (0-10.5 l/min) was tested by measuring the time required to fill a 50 l bucket with

buffer solution (results not shown). During continuous operation the addition of the feed

medium, titrant (2M H2SO4), antifoam agent (Plurafac LF 1300, BASF, Ludwigshafen,

Germany) and the harvest are performed with peristaltic pumps (Reglo-Analog, Ismatec,

Wertheim-Mondfeld, Germany). The temperature of the reactor is controlled at 30° C with a

circulating water bath (U12+R400, Lauda, Lauda-Königshofen, Germany). The temperature

sensor is placed in the gas separator. The pH is controlled and the DO recorded with the

control unit of a laboratory fermenter (Biostat M, Braun Biotech, Melsungen, Germany). The

DO is measured with DO probes (Ingold-Mettler Toledo, Greifensee, Switzerland).

Porous reaction bonded recrystallized silicon carbide (SiC) honeycomb monoliths (Heimbach

GmbH, Düren, Germany) were used. Properties of the honeycomb monolith: outer geometry

330 mm x 40 mm x 40 mm, quadratic capillaries, number of capillaries 529, capillary inner

diameter 1.3 mm, wall thickness 400 µm, channels per cm² 33, open cross section (sum of the

Page 113: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 95

inner cross sections of the capillaries) 894 mm², further properties are described in appendix

A.4, Tab. 6.1, and Fig. 6.3. The outside of the porous monolith was sealed with epoxy resin

(Bylapox E+H, Byla GmbH, Runkel, Germany). Before sealing the monolith, the viscosity of

the resin was increased, by leaving the resin for 0.5 h after mixing epoxy resin and curing

agent. With scanning electron microscopy (SEM), it was confirmed, that this procedure

prevented the epoxy resin to be sucked into the porous SiC through capillary forces (results

not shown). The biocompatibility of the epoxy resin was assured with respiration

measurement in shake flasks (Chapter 2) (results not shown). As single tube, a rectangular

stainless steel tube (330 mm x 29 mm x 29 mm, cross section 841 mm²) was used.

The wild type of Corynebacterium glutamicum ATCC 13032 [Chapter A.1.1, Kinoshita et al.

1957] was used as biological model system. The composition of the minimal medium (carbon

source lactic acid 5 g/l, pH 7), preparation of the preculture, and culture conditions are

described in chapter A.1.2. Alternatively, a buffer solution (17 g/l NaH2PO4·2H2O) with the

same ionic strength (0.66 mol/l) as the biological medium was used. At this ionic strength,

medium and buffer solution are coalescence inhibited [Lessard and Zieminski 1971]. The

entire reactor was sterilized in an autoclave.

Page 114: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 96

∆p

70

30

125

269330

40

65

16

20

4

monolith

pump

aeration

exhaust gas

porous glass plate

silicontube

70

liquid

viewingglas

170

heatexchanger

gasseparator

DO topelectrode

pressuredrop pump

DObottom

19

6076

length ofboth tubes

124

exhaust gascooler

aeration

ring

from downcomer

from down-comer

feed, titrant, antifoamharvest

∆p

liquivolume

d

riser: liquid inside the SiC monolith

0.4 l

riser: liquid inside the single tube

0.3 l

riser: liquid below the monolith

0.45 l

downcomer and pumps

0.15 l

gas separator fill level 30 mm

0.7 l

Fig. 7.1: Schematic depiction, mode of

operation, and design of the monolith

loop reactor. The photograph at the

bottom shows how the liquid flows from

the downcomer into the riser. To in-

crease the readability of the figure, the

view of downcomer and gas separator

were rotated 90°. The downcomer is

actually behind the monolith.

Tab. 7.1: Liquid volumes of the reactor.

Page 115: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 97

7.2.2 Measuring techniques

This study characterizes the influence of honeycomb monoliths in loop reactors, with the

intention to design and optimize loop biofilm reactors with immobilized biofilms on

honeycomb monoliths. Therefore, if the measurement technique allowed a sterile operation,

experiments were performed with the biological model system C. glutamicum, otherwise

buffer solution was used.

7.2.2.1 Gas liquid mass-transfer

The volumetric mass transfer coefficient (kLa) was determined with the biological model

system, applying the global balance method [Sobotka et al. 1982, Knoll et al. 2005] according

to Eqs. 7.1, 7.2. During the measurement, the reactor was operated in continuous mode with a

dilution rate of 0.1 h-1. This guarantees stable and reproducible operating conditions. As

described by Takashi and Yoshida [1979] kLa may depend on the cell concentration.

⎟⎠⎞

⎜⎝⎛ ⋅−⋅⋅

=

in2out2RO

L

O100DOOpS

OTRak

2

(7.1)

⎟⎟⎠

⎞⎜⎜⎝

⎛⋅

−−−−

−⋅= out2out2out2

in2in2in2

0m

OCOO1COO1

OV

qOTR (7.2)

The solubility was calculated as described in appendix A.1.3. The DO was measured at two

positions of the reactor (Fig. 7.1): in the gas separator at the opposite side of the downcomer

(DO top) and below the monolith (DO bottom). For each DO probe a kLa value was

calculated according to Eqs. 7.1, 7.2.

7.2.2.2 Specific power input

The total specific power input (P/V) of the loop reactor is composed of the power input

generated by the pumps (PP) and the power input by aeration. The power input generated by

Page 116: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 98

the pumps depends on the liquid flow rate ( ), and the pressure difference before and after

the pump (∆P

V&

p) (Eq. 7.3). T-pieces were placed in the downcomer before and after the pump

to determine the change in pressure (∆pP) (Fig. 7.1). All pressure measurements were

performed in buffer solution with the pressure sensor (almemo system: sensor FDA612mr,

data logger Ma 2290-8, Ahlborn, Holzkirchen, Germany).

PlP pVP ∆⋅= & (7.3)

The power input by aeration was estimated as described by Christi [1989], and Liepe et al.

[1988], assuming isothermal gas expansion.

7.2.2.3 Pressure below monolith and single tube

The pressure below the monolith or single tube is measured with a pressure sensor placed

59 mm under the monolith (Fig. 7.1). The difference between the hydrostatic pressure of the

not operated reactor (liquid level of 30 mm, aeration and pumps switched off) and the

operated reactor (aeration and pumps switched on) is recorded.

7.2.2.4 Gas hold-up

The gas hold- up (ε) was determined in the biological system by determining the difference of

the liquid level in the gas separator (∆H) between the aerated and the non aerated reactor

(Eq. 7.4) [Blenke 1979, Christi 1989].

l2

seperatorgas

2seperatorgas

lg

g

VrHrH

VVV

+⋅π⋅∆

⋅π⋅∆=

+=ε (7.4)

7.2.2.5 Mixing time

The mixing time of the reactor is determined with the colorizing/decolorizing method [Stein

1987, Danckwerts 1957]. 0.02 g/l pH indicator phenolphthalein is added to the buffer

solution, and the pH set to 8.4. After injecting 1 ml 5 M NaOH with a syringe (inject solo

Page 117: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 99

5 ml with needle 0.8 x 120 mm, B. Braun Melsungen, Melsungen, Germany) into the inlet of

the downcomer (bottom of the gas separator) the mixing time is recorded visually with a

stopwatch. The phosphate buffer (pKa{ 2.16, 7.21, 12.32}) has almost no buffering effect in

the pH range (8.4-10) at which the phenolphthalein changes its color. After recording the

mixing time, the pH was set again to 8.4 with 5 M H2SO4. To avoid high salt concentrations,

the solution was renewed every 10 measuring points [Stein 1987]. Three different mixing

times were recorded: mixing time M1 is the time required until the colored solution flows out

of the monolith or single tube, mixing time M2 is the time required until the colored solution

arrives again at the bottom of the heat exchanger, mixing time M3 is the time required until

the liquid in the reactor is completely mixed.

7.2.3 Flow regimes and Reynolds numbers in riser and downcomer

The Reynolds number (Re) (Eq. 7.5) and the laminar-turbulent transition were calculated

according to Idelchik [1994]. The velocity (v) inside the downcomer was calculated by

dividing the volume flow of the liquid phase by the cross section. Velocities inside the riser

were calculated by dividing the combined volume flow of the liquid and gas phase by the

cross section. The hydrodynamic diameter (Dh) for cylindrical and rectangular cross sections

was calculated with the equations provided by Idelchik [1994]. The kinematic viscosity (ν) of

water (0.01 cm²/s) was used to calculate Re. Biological medium and buffer solution have

water like viscosities. The relative roughness required to estimate the laminar-turbulent

transition was assumed to be 0.007.

υ⋅

= hDvRe (7.5)

Page 118: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 100

7.3 Results

7.3.1 Gas liquid mass transfer and antifoam agent

Fig. 7.2 depicts the influence of a single tube (top row) and a monolith (bottom row) on the

volumetric mass transfer coefficient (kLa) at different liquid flow rates ( ) and aeration rates

(Q). The k

V&

La was calculated using a DO probe placed in the riser below the single tube or

monolith (kLaDO,bottom) or alternatively using a DO probe placed in the gas separator (kLaDO,top)

(Chapter 7.2.2.1, Fig. 7.1).

kLaDO,bottom determined with antifoam agent (Figs. 7.2a, d) are comparable for operations with

a single tube and a monolith. At liquid flow rates between 4 l/min and ~15 l/min the kLa is

independent of the liquid flow rate. It increases roughly linearly with the aeration rate from

0.01 s-1 (Q=1 l/min) to 0.06 s-1 (Q=5 l/min). At liquid flow rates between ~15 l/min and

21 l/min the kLa depends on the aeration and liquid flow rate. Higher liquid flow rates

increase the kLa by a factor of 2-3 compared to the constant kLa at low liquid flow rates.

kLaDO,top is strongly influenced by the monolith, while in an operation with the single tube kLa

did not depend on the position of the DO probe (Figs. 7.2a, b), the monolith leads to a reduced

kLaDO,top (Figs. 7.2d, e). For all conditions kLaDO,top measured with monolith (Fig. 7.2e)

remained below 0.05 s-1. In contrast to measurements with a single tube the liquid flow rate

had nearly no influence on kLaDO,top.

Figs. 7.2c, f depict kLaDO,bottom measured without antifoam agent. The general characteristics

are comparable to the measurements with antifoam agent. At low liquid flow rates, kLaDO,bottom

increases roughly linearly with the aeration rate and is nearly independent of the liquid flow

rate. At high liquid flow rates, kLaDO,bottom increases with the aeration rate and liquid flow rate.

kLaDO,bottom determined with monolith was slightly higher than with single tube. The

Page 119: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 101

comparison of the volumetric mass transfer coefficient with antifoam agent (kLa)

(Figs. 7.2a, d) and without antifoam agent (kLa0) (Figs. 7.2c, f) revealed that the addition of

antifoam agents causes a significant decrease of the volumetric mass transfer coefficient. The

ratio between kLa and kLa0 lay between 0.3 and 0.7. The influence of the antifoam agent on

the kLa decreased with higher liquid flow rates and increased with higher aeration rates.

0.00

0.05

0.10

0.15

with antifoam agent

single tube

with antifoam agent

A

k LaDO

,bot

tom

[s-1]

0.0

0.1

0.2

0.3single tube

without antifoam agent

C

k La

DO,b

otto

m

[s-1]

0 5 10 15 20 250.00

0.05

0.10

0.15SiC monolithD

liquid flow rate [l/min]

k LaDO

,bot

tom

[s-1]

0 5 10 15 20 250.0

0.1

0.2

0.3SiC monolithF

liquid flow rate [l/min]

k LaDO

,bot

tom

[s-1]

0 5 10 15 20 250.00

0.05

0.10

0.15SiC monolithE

k LaD

O,to

p [s

-1]

liquid flow rate [l/min]

0.00

0.05

0.10

0.15

single tubeB

k LaD

O,to

p [s

-1]

Fig. 7.2: Influence of monolith, single tube, and antifoam agent on the volumetric mass transfer

coefficient (kLa). kLa was determined in the biological model system Corynebacterium glutamicum in

continuous mode, applying the global balance method. The dissolve oxygen (DO) probes were placed

below the monolith / single tube (DO bottom) and in the gas separator (DO top), the gas concentrations

were measured with an exhaust gas analyzer. Conditions: aeration rate 1 l/min, 2 l/min, 3 l/min,

4 l/min, 5 l/min, temperature 30° C, the liquid flow rate was changed by varying the revolution of

the gear pumps, concentration of the antifoam agent 1000 ppm (Plurafac LF 1300, BASF,

Ludwigshafen, Germany), liquid volume monolith 1.7 l, liquid volume single tube 1.55 l.

Page 120: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 102

Fig. 7.3 depicts the influence of different antifoam agent concentrations on the foam structure

(Figs. 7.3a-c) and on the kLa (Fig. 7.3d). At concentrations between 0 ppm and 5 ppm the

foam consists of large bubbles Fig. 7.3a. If antifoam agent is added to the reactor, the foam is

only locally destroyed at the injection point. At higher antifoam agent concentrations between

5 ppm and 1000 ppm, the foam structure changes to smaller bubbles (Fig.7.3b). If antifoam

agent is added, the foam is at first destroyed, however, after a short period, it re-forms. Only

at concentrations above 1000 ppm (Fig. 7.3c), the foam is permanently destroyed. An addition

of as little as 1 ppm antifoam agent leads to a 25 % reduction of the kLa value (Fig. 7.2d).

Until a concentration of 50 ppm, kLa decreases with the antifoam agent concentration. At

higher concentrations, the kLa becomes independent of the antifoam agent concentration.

Fig. 7.3: Influence of antifoam agent

(Plurafac LF 1300, BASF, Ludwigshafen,

Germany) A.-C. on the foam structure in

the gas separator and on D. the volumetric

mass transfer coefficient (kLa). Conditions:

single tube, aeration rate 3 l/min, liquid flow

rate 17 l/min, all other conditions are equal

to Fig. 7.2.

D

0 0.1 1 10 100 10000.00

0.05

0.10

0.15

k LaD

O,b

otto

m

[s-1]

antifoam concentration [ppm]

>1000 ppm >5 ppm 5 ppm -0

C B A

Page 121: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 103

Operating the reactor without antifoam agent would be favorable. However, long-term

biological experiments in continuous mode over several weeks (Chapter 8.5.5) could only be

performed safely with antifoam agent. The reactor design with single tube showed

homogeneously distributed dissolved oxygen, while the design with monolith led to a reduced

mass transfer in the gas separator. The lower mass transfer can influence the productivity of

aerobic processes. Therefore, to optimize the performance of monolith loop reactors it is

important to clarify the mechanisms responsible for the inhomogeneous mass transfer

distribution. The following sections will address this problem.

7.3.2 Specific power input

Fig. 7.4 depicts the specific power input at different liquid flow rates. The power input (P/V)

increases parabolically with the liquid flow rate up to 4 kW/m³. While at low liquid flow rates

single tube and monolith have equal P/V, at high liquid flow rates, the monolith has a slightly

higher P/V. In Fig. 7.4, solely the power input caused by the pumps is considered. The P/V by

aeration was small and nearly independent of the liquid flow rate. It increased with the

aeration rate. At an aeration rate of 5 l/min P/V by aeration lay between 0.2 kW/m³ and

0.3 kW/m³.

0 5 10 15 20 250

1

2

3

4

5

liquid flow rate [l/min]

spec

ific

pow

er in

put (

P/V)

[k

W/m

³] Fig. 7.4: Influence of single tube, monolith and liquid

flow rate on the specific power input (P/V) of the loop

reactor P/V was determined by measuring the pressure

difference over the pump. ——— parabolic fit curve.

Conditions: buffer solution.

Page 122: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 104

7.3.3 Pressure below single tube and monolith

Fig. 7.5 depicts the pressure measured below the single tube and monolith. The pressure

under the single tube (Fig. 7.5a) decreases with the aeration rate and increases with the liquid

flow rate. Generally, the pressure differences generated by the single tube are small. At low

liquid flow rates the pressures below the single tube is smaller than the hydrostatic pressure of

the unaerated liquid. At high liquid flow rates the influences of aeration and liquid flow rate

compensate each other, resulting in a pressure under the single tube close to the hydrostatic

pressure.

Compared to the single tube the pressure below the monolith (Fig. 7.5b) is always higher and

shows a distinctly different behavior. The pressure is always higher than the hydrostatic

pressure. At low liquid flow rates, the pressure decreases with the aeration. At high liquid

flow rates, in contrast to an operation with single tube, it increases with the aeration.

Fig. 7.5: Influence of liquid flow rate, and

aeration rate on the pressure below

single tube and monolith. Conditions:

aeration rate 0 l/min, 2 l/min,

3 l/min, 4 l/min, buffer solution, the

pressure sensor is placed 59 mm below

the monolith, the hydrostatic pressure

(aeration and pumps switched off) is

defined as a pressure difference of 0.

0 5 10 15 20 25-6

-4

-2

0

2

4single tubeA

pres

sure

diff

eren

ce u

nder

sing

le tu

be

[mba

r]

liquid flow rate [l/min]0 5 10 15 20 25

0

5

10

15

20

25

liquid flow rate [l/min]

SiC monolithB

pres

sure

diff

eren

ce u

nder

mon

olith

[m

bar]

7.3.4 Gas hold-up

Fig. 7.6 depicts the gas hold-up of the loop reactor operated with single tube (Fig. 7.6a) and

operated with monolith (Fig. 7.6b). The gas hold-up increases with the aeration rate and

Page 123: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 105

decreases linearly with the liquid flow rate. Monolith and single tube result in comparable

gas-hold ups. The gas hold-up of the downcomer was below 0.005, and thus negligible.

0 5 10 15 200.00

0.01

0.02

0.03

0.04

0.05

0.06

BA single tube

gas

hold

-up

[-]

0 5 10 15 20 25 liquid flow rate [l/min] liquid flow rate [l/min]

SiC monolith Fig. 7.6: Influence of A. single tube,

B. monolith, aeration rate, and liquid flow

rate on the gas hold-up. The gas hold-up

was determined by measuring the difference

of the liquid level in the gas separator.

Aeration rate 1 l/min, 3 l/min,

5 l/min, ——— linear fit, all other

conditions are equal to Fig. 7.2.

7.3.5 Mixing

Fig. 7.7 depicts the influence of single tube and monolith on the mixing time. Generally, the

mixing time decreases with higher liquid flow rates and higher aeration rates. The time the

liquid needs to flow through downcomer and riser (mixing time M1) is equal for single tube

and monolith (Fig. 7.7c). At low liquid flow rates the mixing time M1 decreases with the

aeration rate, while at high liquid flow rates the aeration has only a negligible effect

(Fig. 7.7e). In operation with single tube and monolith, the liquid is completely mixed over

the entire cross section, when it exits the riser (Fig. 7.7a) (not shown for the single tube). The

mixing time until the pulse arrives again at the bottom of the heat exchanger (mixing time

M2) was roughly twice as long as M1 (results not shown). The mechanisms, which influence

M2 and M1, are comparable. M2 is equal for monolith and single tube. Equivalent to the

other mixing times, the time required to mix the entire reactor (mixing time M3) decreases

with higher liquid flow rates and higher aeration rates (Figs. 7.7d, f). While in operation with

single tube mixing times M2 and M3 are nearly identical, in operation with monolith mixing

time M3 is much longer, particularly at low liquid flow rates (Fig. 7.7d). The monolith

Page 124: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 106

reduces the mixing intensity in the outer regions of the gas separator, where the dissolved

oxygen (DO) probe is placed (Figs 7.7a, b).

Fig. 7.7: Influence of monolith,

liquid flow rate and aeration rate on

the mixing time. Left: mixing time

M1, right: mixing time M3. A., B.

photographs of the gas separator at

mixing times M1, M3, SiC monolith,

liquid flow rate 21 l/min, aeration

rate 3 l/min. C., D. single tube,

SiC monolith, aera ion rate

3 l/min. E., F. SiC monolith,

aeration rate 1 l/min, 3 l/min,

5 l/min. All fit functi s (solid

lines) are of first order exponential

decay. Measurement technique:

colorizing / decolorizing method,

buffer solution 17 g/l NaH

t

on

7.4 Discussion

7.4.1 Gas-liquid mass transfer and antifoam agent

The volumetric mass transfer coefficient was found independent of the liquid flow rate below

15 l/min. This corresponds to a liquid velocity in the riser of 0.3 m/s. The porous glass filter

2PO4

2H2O, pH indicator 0.02 g/l phenol-

phthalein, 1 ml 5 M NaOH is in-

jected with a syringe into the inlet of

the downcomer. Definition of the mixing times: M1 time required until the colored solution flows out of

the monolith or single tube, M3 time required until the entire reactor is mixed.

A B

DO probe

downcomer

0.0

2.5

5.0

7.5

10.0 C

mix

ing

time

M1

[s]

0

10

20

30

40

singletube

SiC monolith

D

mix

ing

time

M3

[s]

0 5 10 15 20 250.0

2.5

5.0

7.5

10.0 E

mix

ing

time

M1

[s]

liquid flow rate [l/min]0 5 10 15 20 25

0

10

20

30

40 F

mix

ing

time

M3

[s]

liquid flow rate [l/min]

Page 125: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 107

plate used for aeration produces bubbles with a diameter of ~2 mm. The bubble diameter was

calculated according to Voit et al. [1987] and Weuster-Botz et al. [2001]. It may be more then

locity of 0.3 m/s. For bubble

The addition of antifoam agent leads to a reduction of the kLa. In this study the ratio between

L

(kLa0) lay between 0.3 and 0.7. The influence of antifoam agent on the kLa decreased with

higher liquid flow rates and increased with higher aeration rates. For airlift and bubble column

reactors Koch et al. [1995] and Koide et al. [1985] report kLa/(kLa)o of 0.4. Most of the

antifoam agent concentration at concentrations much lower then were required for an efficient

L

concentrations above 50 ppm, however, more than 1000 ppm were required to control

foaming. The observation, that the kLa is independent of the antifoam agent concentration at

a coincidence that bubbles of this diameter rise with a ve

diameters between 2 mm and 15 mm, the bubble velocity is often considered roughly

independent (constant at 0.3 m/s) of the bubble diameter [Christi 1989, Deckwer 1992,

Fleischer 2000, García Calvo 1989, Heijen JJ and van Riet 1984]. Comparable to the results

presented by Choi, [2001] for an external loop airlift reactor at liquid flow rates below

15 l/min, kLa increases roughly linearly with the aeration rate from 0.01 s-1 to 0.06 s-1.

the volumetric mass transfer coefficient with antifoam agent (k a) and without antifoam agent

kLa/(kLa)o reported in the review of Kawase and Moo-Young [1990] were between 0.2 and

0.4. The reported kLa/(kLa)o for stirred tank reactors are higher. Benedek and Heideger [1971]

report kLa/(kLa)o between 0.7 and 0.8. Most of the data in the review of Kawase and Moo-

Young [1990] were between 0.5 and 0.9. It seems, that at low liquid flow rates and power

inputs the reactor behaves more like an airlift reactor, while at high power inputs it behaves

more like a stirred tank or jet loop reactor.

Koch et al. [1995] observed that the surface tension of the liquid became independent of the

foam control. In this study the k a was independent of the antifoam agent concentration at

Page 126: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 108

high concentrations, was also reported by Koide et al. [1985] and Takashi and Yoshida

[1979].

7.4.2 Specific power input

The influence of the monolith on the mass transfer is discussed in the last section of this

chapter.

In this study it was observed, that with increasing liquid flow rates the specific power input

of the kLa, a decrease in gas hold-up leads to a reduction of the kLa [Christi 1989,

Deckwer 1992]. Maybe these two effects compensate each other, resulting in a constant kLa at

low liquid flow rates. While the power input increases parabolically, the gas hold-up

d a lower hydrostatic pressure [Christi 1989, Deckwer 1992]. The pressure

drop of tubes increases with the fluid velocity [Idelchik 1994], resulting in a rise of the

the monolith corresponds to the

increases, while the gas hold-up decreases. While an increase in power input leads to an

increase

decreases linearly with the liquid flow rate. Thus, at high liquid flow rates the decreasing gas-

hold up cannot compensate the increasing power input anymore, resulting in an increase of

the kLa. The differences in power input between single tube and monolith were small. It is

unlikely, that they are responsible for the observed lower gas liquid mass transfer in operation

with monolith.

7.4.3 Pressure below single tube and monolith

The pressure below the single tube decreases with the aeration rate and increases with the

liquid flow rate. With the aeration rate, the gas hold-up increases, resulting in a lower density

of the liquid an

pressure with the liquid velocity. The higher pressure below

increase of the pressure drop with decreasing tube diameter [Idelchik 1994]. The increase of

the pressure below the monolith with the aeration rate at high liquid flow rates may result

Page 127: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 109

from the energy required to press the air bubbles, which are larger than the capillary diameter,

into the monolith. The influence of the overpressure under the monolith of maximal 20 mbar

on the oxygen transfer capacity of the reactor (Eq. 4.5) is negligible.

7.4.4 Gas hold-up

The gas hold-up increased with the aeration rate and decreased with the liquid flow rate. As

described in the literature [Christi 1989, Deckwer 1992, Choi 2001], in aerated reactors the

gas hold-up generally increases with the aeration rate. In contrast to airlift reactors, in the

reactor design used in this study, aeration rate and liquid flow rate are decoupled. Thus, if the

, the air bubbles are transported faster out of the reactor resulting in a

ate influences mixing in all

sections of the reactor (downcomer, riser, gas separator), the aeration can exclusively

influence mixing in the riser. Thus, at high liquid flow rates the higher fluid velocity produced

liquid flow rate increases

reduced gas hold-up. Decreasing gas hold-ups at high power inputs were also observed by

[Blenke 1979] with a jet loop reactor. The gas hold-up of the downcomer was found to be

negligible. This observation agrees with [Al-Masry 2004]. He showed for external loop airlift

reactors that the gas hold-up of the downcomer is negligible if the volume ratio between the

liquid volume of the gas separator and the liquid volume of the entire reactor is above 30 %.

The reactor used in this study has a ratio of ~40 % (Tab. 7.1). Single tube and monolith had

comparable gas-hold-ups, thus differences in the gas hold-up are most likely not responsible

for the lower gas liquid mass transfer in operation with monolith.

7.4.5 Mixing

The mixing time decreases with the liquid flow rate. With increasing liquid flow rates, the

circulation time decreases, resulting in a reduced mixing time. The aeration influenced the

mixing time only at low liquid flow rates. While the liquid flow r

Page 128: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 110

by the aeration in the riser is small relative to the circulation of the liquid. The influence of the

fferences in specific power

input, pressure, and gas hold-up between an operation with single tube and monolith were

lower k a results most likely from different mixing conditions in the gas separator.

monolith on the mixing time is discussed in the following section.

7.4.6 The influence of the monolith on flow regime, mass transfer, and mixing time

The gas liquid mass transfer in the gas separator was found to be distinctly lower in operation

with monolith than in operation with single tube. To clarify the mechanisms responsible for

this phenomenon, the influence of the monolith was analyzed on specific power input,

pressure below the monolith, gas hold-up, and mixing time. The di

small. As discussed previously, it is unlikely that they are responsible for the reduced kLa in

the gas separator. The mixing times M1 and M2 were equal for single tube and monolith. M3

was longer for the monolith. It was observed that the monolith reduces the mixing intensity in

the outer regions of the gas separator, where the dissolved oxygen (DO) probe is placed. The

L

Page 129: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 7 Biofilm reactor with a honeycomb monolith as immobilization carrier 111

Fig. 7.8: Hydrodynamic conditions in the

gas separator and flow regimes in single

tube, monolith, and downcomer. A., B.

photographs of the gas separator

A. single tube, B. monolith. Conditions:

buffer solution, liquid flow rate 21 l/min,

aeration rate 3 l/min. C. calculated

Reynolds numbers, aeration rate

0 l/min, 3 l/min, 5 l/min, down-

comer 0 l/min. The Reynolds numbers

and the laminar-turbulent transition were

calculated according to Idelchik [1994].

0 5 10 15 20 2510

100

1000

10000

50000downcomer

lam

inar

turb

ulen

t

SiC monolith

single tube

Rey

nold

s nu

mbe

rin

side

cha

nnel

s of

rise

r [-]

liquid flow rate [l/min]

A

B

single tube

SiC monolith

C

Figs. 7.8a, b depict the movement of the liquid in the gas separator. In operation with single

tube, a highly agitated fountain is present (notice also the splashes on the of wall Fig. 7.8a)

above the riser. The fountain creates waves, which mix the liquid in the gas separator. In

contrast, the fountain formed in operation with monolith is calm and uniform (Fig. 7.8b),

resulting in a prolonged mixing time and a lower volumetric mass transfer coefficient. The

different behavior of single tube and monolith correspond to different flow regimes in the

riser. The flow inside the single tube is turbulent (Fig. 7.8c), while the flow in the capillaries

of the monolith is laminar. The laminar flow in the monolith is most likely responsible for the

longer mixing time and the reduced volumetric mass transfer coefficient.

Page 130: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 112

Chapter 8

Modeling aerobic biofilm reactors with growth associated product formation

This chapter compares the productivities of biofilm reactors and standard reactors at different

operating conditions, focusing on aerobic biological systems with growth associated product

formation. The aim is, to predict the behavior of biofilm reactors at different operating

conditions, and to distinguish between operating conditions, which favor the use of biofilm

reactors, and operating conditions, which favor standard reactors.

8.1 Introduction

Continuous cultures performed in biofilm reactors are generally expected to have a higher

productivity than continuous cultures performed in standard reactors without immobilized

microorganisms. This expectation is based on the assumption, that immobilization enables

reactor operation at high biomass concentrations [Gjaltema et al. 1997], because immobilized

cells do not wash out. The reactor can be operated at high dilution rates [Mori 1985]. This

study considers the entire biofilm reactor: varying model parameters, which could be

manipulated by an operator or design engineer, such as volumetric mass transfer coefficient

(kLa), the carbon source concentration of the feed solution (cS,bulk,feed), and the specific surface

area of the support (A/V). This is different to the more common biofilm focused approach,

which mainly studies the influence of model parameters at the bulk-biofilm interface such as

dissolved oxygen (DO), carbon source concentration of the bulk phase (cS,bulk), and fluid

Page 131: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 113

velocity on the behavior and structure of the biofilm itself [Howell and Atkinson 1976,

Morgenroth et al. 2004, Rittmann and McCarty 1980, van Loosdrecht 2002].

Biofilms are complex heterogeneous structures, which consist of variable distributed cellular

aggregates with void spaces or water channels [Costerton et al. 1995]. Depending on the

microorganism, the substrate concentration and the shear stress, this heterogeneous "biofilm

architecture" can change its porosity, biofilm density and diffusion coefficient [Bishop et al.

1995, Picioreanu et al. 2000]. Until a biofilm reaches a steady state, it can take several weeks

[Léon Ohl et al. 2004] making modeling the only convenient approach to study the behavior

of biofilm reactors in a vast variety of different conditions.

The mass transport inside biofilms and the development of the "biofilm architecture" has been

studied using complex 3-D biofilm models [Chang et al. 2003, Morgenroth et al. 2004,

Noguera et al. 1999, Picioreanu et al. 2000, van Loosdrecht et al. 2002]. It was found that the

mass transport inside biofilms is predominantly by diffusion. Only if the biofilm colonies are

isolated, and the fluid velocity high, may convective transport occur in the channels of the

biofilm [Picioreanu et al. 2000, Stewart 2003, 1998]. Further, it was observed, that in systems

with high shear forces smooth and flat biofilms are formed [van Loosdrecht et al. 2002]. High

shear forces (high fluid velocities, high power inputs) are required in most aerobic processes

to achieve a volumetric mass transfer coefficient (kLa) sufficiently high to supply the

microorganisms with oxygen. Morgenroth et al. [2004] compared complex 3-D with simpler

1-D biofilm models and observed that the 1-D models provide very similar results to the more

complex numerical solutions if the biofilm is flat. According to Bonomo et al. [2001], 1-D

models are the only ones, that can reasonably support process engineers in biofilm reactor

design, due to their intrinsic simplicity and the need for small sets of input parameters.

Page 132: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 114

Biofilms usually consist of an active film thickness, which is supplied with substrates, and an

anaerobic or inactive film thickness, where at least one substrate is exhausted. For oxygen the

active film thickness and the decrease of the dissolved oxygen inside the biofilm has been

experimentally determined with microelectrodes by Boessmann et al. [2003], Horn et al.

[2002], and Zhang and Bishop [1994]. According to the simulations performed with a 1-D

model by Atkinson and Davies [1972] and Howell and Atkinson [1976], the "ideal" film

thickness for the use in biofilm fermenters is equal to the penetration depth of the substrates

(active film thickness). Howell and Atkinson [1976] further postulated, that the "ideal"

thickness of a biofilm limited by both oxygen and substrates should be equal to the smaller of

the two penetration depths.

8.2 Model descriptions

8.2.1 Modeling tools

All simulations were carried out on standard PC's with the software package gPROMS

(general PROcess Modelling System), (PSE, London, Great Britain).

8.2.2 Model introduction

The principal components of an aerobic biofilm reactor are depicted in Fig. 8.1a. Carbon

source and oxygen are fed to the reactor and then consumed and converted to biomass, carbon

dioxide and products, which leave the reactor together with the not consumed carbon source

and oxygen. The biofilm reactor additionally contains a support with a surface (A) on which

the biofilm can grow. As a sample reactor a monolith loop reactor is depicted in Fig. 8.1a,

which contains as support a honeycomb monolith. This type of biofilm reactor has been used

by Kawakami et al. [1989], Tata et al. [1999], Wojnowska-Baryla et al. [1999], and

Yongming et al.[2002]. The pump depicted in Fig. 8.1a stands for a high shear stress (fluid

Page 133: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 115

velocity), and a high power input which is required to produce a flat biofilm and a high gas-

liquid mass transfer. As depicted in Fig. 8.1 b for the model the biofilm reactor is divided into

three phases: bulk-phase, gas-phase and biofilm-phase. Gas- and bulk-phase are assumed to

be ideally mixed whereas in the biofilm-phase mass transport occurs only by diffusion

(Fig. 8.1c).

harvest

cO2,gas,feed

cS,bulk,feed

A

surface area of support (A)

cO2,gas; cCO2,gas

filling volume (V)

substrate feed exhaust gas

pump aeration

cX,bulk; cS,bulk

ordinary differential equations ideally mixed reactor system

partial differential equations only diffusion

film bulk

C

concen-tration (ci)

space (h)

XS

O2

ri,bulk,bio ri,film,bio

CO2 ci,bulk ci,film

O2, CO2 O2B

S

bulk

D film

ci,bulk,feed

ri,bulk film

ri,bulk film X, CO2

q

X, S

ci,bulk S, O2

ci,gas,feedci,gas

O2 (OTR)

ri,gas bulk CO2 (CTR)

gas

ri,gas bulk

Fig. 8.1: Model structure of an aerobic biofilm

reactor. A. Schematic depiction of a biofilm reactor.

As sample reactor a monolith loop reactor is

depicted. B. Mass transport between the three

reactor phases as defined by assumption 1.

Denominations below arrows represent the variable

names used in the model description. C. Schematic

depiction of the spatial concentrations in the ideally

mixed bulk-phase (assumption 2) and the biofilm-

phase, where mass transport occurs only by

diffusion (assumption 2).

Page 134: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 116

8.2.3 Model assumptions

The 1-D biofilm model applied in this study is based on the following fundamental

assumptions:

1. The biofilm reactor (Fig. 8.1a) consists of three phases a gas-phase, a bulk-phase, and a

biofilm-phase (Fig. 8.1b).

2. Gas-phase and bulk-phase are considered as ideally mixed (Fig. 8.1c). The mass transport

of carbon source, oxygen, and carbon dioxide in the biofilm phase occurs only by

diffusion (Fig. 8.1c) according to Fick's law. Thus, concentrations in gas- and bulk-phase

are only functions of time, while concentrations in the biofilm-phase are functions of time

and space.

3. At the biofilm surface (bulk-biofilm interface) all concentrations, except for the biomass

are equal in the bulk- and the biofilm phase (Fig. 8.1c). The effect of the liquid film

diffusion in the boundary layer between bulk-phase and biofilm-phase is neglected.

4. The biofilm can grow as long as both substrates oxygen and carbon source are present,

and the biofilm density is below the maximum biofilm density (ρfilm,max). If the biofilm

density exceeds ρfilm,max, additionally formed biomass is detached, increasing the biomass

concentration of the bulk-phase.

5. The biological conversion in the bulk- and biofilm-phase can be described with Monod

and maintenance kinetics. The kinetic parameters, yield coefficients, Monod constants,

and maintenance constants are considered of being equal for bulk- and biofilm-phase.

Page 135: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 117

8.2.4 Bulk phase

The ordinary differential equation (Eq. 8.1) is the core component of the model. As depicted

in Fig. 1b it combines feed (D·ci,bulk,feed) and harvest (-D·ci,bulk) of the dissolved components

with the gas-liquid mass transfer (ri,gas↔bulk) (Eq. 8.3), the mass transfer between bulk- and

biofilm-phase (ri,bulk↔film) (Eqs. 8.12, 8.13), and the rate of the biological conversion of the

bulk-phase (ri,bulk,bio) (Eqs. 8.14-8.17). Eq. 8.1 exists four times, once for each simulated

concentration (X, O2, CO2, S). As a measure for the growth associated product formation of

the biofilm reactor, the space time yield (STY) (Eq. 8.2) of the biomass is used. A product

was not explicitly considered, because the STY of the biomass and the STY of an exclusively

growth associated product are linearly linked. Both depend only on the substrate feed rate, the

biomass concentration and the growth rate.

{ S,CO,O,XirrrcDcDdt

dc22bio,bulk,ifilmbulk,ibulkgas,ibulk,ifeed,bulk,i

bulk,i ∈+++⋅−⋅= ↔↔ } (8.1)

bulk,XcDSTY ⋅= (8.2)

8.2.5 Gas phase and gas-liquid mass transfer

The gas-phase is considered ideally mixed (assumption 2). Eqs. 8.3, 8.4 describe the gas-

liquid mass transfer. Eq. 8.5 defines the dissolved oxygen (DO). Eq. 8.6 states, that the

amount of oxygen consumed is equal to the amount of carbon dioxide produced, guaranteeing

a constant volume of the reactor.

( ) ( ) { }22gas,ifeed,gas,ibulk,i*

bulk,iLbulkgas,i CO,Oiccqccakr ∈−⋅=−⋅=↔ (8.3)

{ 22gas,iRi*

bulk,i CO,OicpSc ∈⋅⋅= } (8.4)

Page 136: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 118

100cc

DO *0,bulk,O

bulk,O

2

2 ⋅= (8.5)

CTROTRrr bulkgas,CObulkgas,O 22=−= ↔↔ (8.6)

8.2.6 Biofilm phase

Assumption 2 defines, that mass transport inside the biofilm-phase occurs only by diffusion

(with exception of the biomass). Diffusion according to Fick's 2nd law is described by Eq. 8.7.

The behavior of the biomass in the film is described by Eq. 8.8. rX,film,bio (Eqs. 8.14-8.17)

describes the biological conversion of the biofilm, rX,bulk↔film the bulk-biofilm interaction

(assumption 4) (Eq. 8.13), and rX,film,inoculum, rX,film,death the dynamic adaptation of the biofilm.

( ) ( ) { S,CO,Oirdh

t,hcdD

dtt,hdc

22bio,film,i2film,i

2

film,iDfilm,i ∈+⋅= } (8.7)

( ) ( ) ( t,hrrt,hrrdt

t,hdcdeath,film,Xinoculum,film,Xfilmbulk,Xbio,film,X

film,X −+−= ↔ ) (8.8)

The immobilized biomass concentration in the reactor (cX,bulk,imm) is calculated with Eq. 8.9 by

integrating the biomass concentration of the film (cX,film) from the bulk-film interface (h=0) to

the support surface (h=hsup) and by multiplying with the specific surface area of the support

(A/V). This procedure converts the distributed (1-D) biofilm-phase to the lumped (0-D) bulk-

phase.

dh)t,h(cVA)t(c

suphh

0hfilm,Ximm,bulk,X ∫

=

=

⋅= (8.9)

For the partial differential equations, two boundary conditions were defined. The assumption

made by assumption 3, that at the biofilm surface (bulk-biofilm interface) the concentrations

Page 137: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 119

of the bulk- and the biofilm phase are equal, is described by Eq. 8.10. Eq. 8.11 defines, that

the support is not penetrable.

( ) ( ) { S,CO,Oi0htct,hdc 22bulk,ifilm,i ∈== } (8.10)

( ) { S,CO,Oihh0dh

t,hdc22sup

film,i ∈== } (8.11)

8.2.6.1 Mass transfer between bulk- and biofilm-phase

The mass transfer between bulk- and biofilm-phase (ri,bulk↔film) is described by

Eqs. 8.12, 8.13. The flux across the bulk-phase film-phase interface (h=0) for O2, CO2, S is

governed by Fick's 1st law (Eq. 8.12). These substances can diffuse into as well as out of the

biofilm. The multiplication of the fluxes (Ji,bulk↔film) with the specific surface area of the

support converts the distributed (1-D) biofilm-phase to the lumped (0-D) bulk-phase.

( ) { S,CO,Oi0hdh

t,hdcD

VAJ

VAr 22

film,ii

Dfilmbulk,ifilmbulk,i ∈=⋅⋅=⋅= ↔↔ } (8.12)

Eq. 8.13 describes the mass transfer of the biomass between bulk- and biofilm-phase.

According to assumption 4, if the biomass concentration of the biofilm (cX,film) exceeds the

maximum film density (ρfilm,max), additionally formed biomass is detached. kdetach is a not

mechanistic time constant.

dh)0,ρ)t,h(cmax(kV/AJA/Vr max,film

suph

0hfilm,XachdetfilmbulkX,filmbulkX, −⋅⋅=⋅= ∫

=

=↔↔ (8.13)

8.2.6.2 Dynamic adaptation of the biofilm

To allow the biofilm to adapt to parameter changes, the film thickness can increase and

decrease. A constant space above the support (e.g. 100 µm) is always slightly inoculated with

the rate (rX,film,inoculum) (equations not shown). If both substrates are present, the biofilm grows

Page 138: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 120

and increases its thickness. If at least one of the substrates is limiting, the biofilm dies with

the rate rX,film,death (equations not shown) and reduces its thickness. This implementation may

lead to inconsistencies of the mass balance during the dynamic adaptation, however does not

influence the steady state conditions.

8.2.7 Biological conversion

Model assumption 5 assumes, that the kinetic parameters of the biological conversion are

equal for bulk- and biofilm-phase. Eqs. 8.14-8.17 describe the reaction rates of the biological

conversion (ri,bio) necessary for the bulk-phase (Eq. 8.1) and the biofilm-phase (Eqs. 8.7, 8.8).

Xbio,X cr ⋅µ= (8.14)

XSXS/X

bio,S cmcY

1r ⋅−⋅µ⋅−= (8.15

XOXO/X

bio,O cmcY

1r2

2

2⋅−⋅µ⋅−= (8.16)

XCOXCO/X

bio,CO cmcY

1r2

2

2⋅+⋅µ⋅= (8.17)

The biological conversion in bulk- and biofilm-phase is described with Monod and

maintenance kinetics. To avoid the numerical difficulties associated with double Monod

kinetics [Kovárová-Kovar and Egli 1998, Bader 1978] double limitations were excluded.

According to Eq. 8.18 always the stronger limiting substrate determines the growth rate (µ)

[Berber et al. 1998]. This approach is possible because, as described in the results section,

double limitations occur only in a small transition region.

Page 139: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 121

⎥⎥⎦

⎢⎢⎣

++⋅µ=µ

22

2

OO

O

SS

Smax Kc

c,

Kcc

min (8.18)

The maintenance reaction produces no biomass and the carbon source lactic acid is

metabolized according to Eq. 8.19. The maintenance coefficients (in mol) of oxygen (mO2),

carbon source (mS), and carbon dioxide (mCO2) are linked by Eq. 8.20.

OH3CO3O3OHC 222363 ⋅+⋅⎯→⎯⋅+ (8.19)

SCOO mm3m322

=⋅=⋅ (8.20)

8.2.8 Model description of the standard reactor

The behavior of a chemostat culture without immobilized biomass is expressed by Eq. 8.21.

The biological conversion is described by Eqs. 8.14-8.18, and the gas-liquid mass transfer by

Eqs. 8.3-8.6.

{ S,CO,O,XirrcDcDdt

dc22bulkgas,ibio,bulk,ibulk,ifeed,bulk,i

bulk,i ∈++⋅−⋅= ↔ } (8.21)

8.2.9 Model description of a simple not mechanistic biofilm model with constant immobilized biomass

A simple not mechanistic biofilm model is described by Eqs. 8.22, 8.23. In contrast to the

mechanistic biofilm model, the concentration of the immobilized biomass (cXimm,bulk) is a

constant independent of the culture conditions. cXimm,bulk can not be washed-out and grows

with the same growth rate as the bulk-phase. The biological conversion except for the biomass

is described by Eqs. 8.15-8.18, the growth of suspended and immobilized biomass is

described by Eq. 8.23, the gas-liquid mass transfer by Eqs. 8.3-8.6.

Page 140: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 122

{ S,CO,O,XirrcDcDdt

dc22bulkgas,ibio,bulk,ibulk,ifeed,bulk,i

bulk,i ∈++⋅−⋅= ↔ } (8.22)

( ) .constcwithccr bulk,Ximmbulk,Ximmbulk,Xbio,X =+⋅µ= (8.23)

8.3 Definition of the model parameters

For all simulations, if not mentioned differently in the text, the standard values specified in

Tab. 8.1 were used. As biological model system the commercially important amino acid

producer [Kircher and Pfefferle 2001] Corynebacterium glutamicum was used. The organism

has been previously used for immobilization studies by Büchs et al. [1988], Henkel et al.

[1990], Kim and Ryu [1982], and Park and Jeong [2001]. All kinetic parameters of the

biological conversion required for Eqs. 8.15-8.20 except for the Monod constants were

derived from the continuous stirred tank fermentations described in chapter 4.5.2. Differences

in the model parameters between chapter 4 and chapter 8 result through the consideration of

the maintenance kinetics. Lactic acid as carbon source has the advantage that it can be

consumed only aerobically by C. glutamicum [Chapter 2.3.2, Seletzky et al. 2006a], thus,

corresponding to the model description (assumption 4) only aerobic growth is possible.

For the Monod kinetics (Eq. 8.18) average values of the Monod constants for bacteria

measured by Owens and Legan [1987] were used. The Monod constant of lactic acid KS was

set to 50 µM (0.0045 g/l) the Monod constant of oxygen to 1 µM. The oxygen solubility of

the minimal medium was calculated as described in appendix A.1.3.

Biofilms consist mostly of water, the starting point in evaluating the diffusion coefficient of

the biofilm (DDi,film) is an estimate of the diffusion coefficient of the solute (i) in water

(DDi,H2O). The presence of microbial cells, extracellular polymeric substances (EPS), and

inorganic material reduces the diffusion coefficient [Stewart 1998]. DDi,H2O were estimated

Page 141: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 123

using the values and the temperature correlations given by Perry et al. [1997], and Lide

[1999]. Based on the data for bacterial biofilms provided by Stewart [2003, 1998] average

effective diffusive permeabilities (DDi,film/DDi,H2O) were calculated. The effective diffusive

permeabilities and the resulting diffusion coefficients DDi,film are summed in Tab. 8.1.

According to the review of Stewart [1998], the maximum film density (ρfilm,max) (Eq. 8.13,

assumption 4) of microbial biofilms can range from 8 g/l to 200 g/l and can reach up to

400 g/l in flocs or pellets. However, if phenol degrading and artificial biofilms are excluded

most of the reported biofilms have film densities between 35 g/l and 85 g/l. In this study a

film density of 70 g/l was used as the standard value.

Page 142: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 124

Tab. 8.1: Standard values of the

parameters: reactor geometry,

mass transfer, biofilm character-

istics, and kinetics of the

biological conversion.

parameter value, unit

A 0.91 m²

cO2,gas,feed 0.21 mol/mol

cS,bulk,feed 5 g/l

DDCO2,film 1.38·10-5 cm²/s

DDCO2,film/ DD CO2,H2O 0.634

DDO2,film 1.79·10-5 cm²/s

DDO2,film/ DD O2,H2O 0.634

DDS,film 0.465·10-5 cm²/s

DDS,film/ DDS,H2O 0.33

kLa 0.1 s-1

KO2 1·10-6 mol/l

KS 0.0045 g/l

mO2 0.00113 mol/g/h

mS 0.03 g/g/h4

pR 1 bar

q 2 vvm

SCO2 at 30°C 0.029 mol/l/bar

SO2 at 30°C 0.0011 mol/l/bar

V 1.5 l

YX/O2, YX/CO2 28.5 g/mol

YX/S 0.47 g/g

ρfilm,max 70 g/l

µmax 0.32 h-1

Page 143: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 125

8.4 Materials and methods biological experiments

Continuous fermentations were performed with the external loop biofilm reactor described in

chapter 7.2.1. As immobilization carrier a silicon carbide (SiC) honeycomb monolith was

used (Chapter 7.2.1, Appendix A.4). The filling volume of the reactor was 1.7 l, the aeration

rate 3 l/min, and the liquid flow rate 19 l/min. To speed up the biofilm formation (Chapter

6.3.2.2), the monolith was surface modified with DEAE dextran as described in chapter

6.2.2.2. The fermentations were carried out with Corynebacterium glutamicum (Appendix

A.1.1), using the minimal medium described in appendix A.1.2 with a lactic acid

concentration of 5 g/l. Lactic acid as carbon source has the advantage that as assumed by the

model assumptions, it can be metabolized only aerobically (Chapter 2.3.2). Oxygen transfer

rates, biomass concentrations, and lactic acid concentrations were determined as described in

appendix A.3.1, A.5.1, and A.5.2. The dissolved oxygen was measured below the monolith as

described in chapter 7.2.1. To build up the biofilm, the reactor was left at a low dilution rate

of 0.23 h-1 for four days before increasing the dilution rate. Steady state conditions were

assumed, when the measuring values were stable for at least 2 residence times. To reach these

conditions at high dilution rates, up to 10 residence times were required.

The continuous stirred tank fermentions presented in this chapter are identical to the ones in

chapter 4.5.2 (Fig. 4.3).

8.4.1 Adaptation of the model parameters to the biological experiments

Two different biofilm models were used to describe the performance of the loop biofilm

reactor. A mechanistic biofilm model as described in the chapters 8.2.2-8.2.7 (Eqs. 8.1-8.20)

and a not mechanistic biofilm model as described in chapter 8.2.9 (Eqs. 8.22, 8.23). In the

mechanistic model, biofilm thickness and growth rate are functions of the carbon source and

Page 144: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 126

oxygen concentrations. In the not mechanistic model, the immobilized biomass concentration

(cXimm,bulk) is constant and independent of the culture conditions.

At the set operating conditions (aeration rate 3 l/min, liquid flow rate 19 l/min), the

volumetric mass transfer coefficient (kLa) of the loop biofilm reactor with antifoam agent lies

between kLaDO,bottom 0.035 s-1 and kLaDO,top 0.075 s-1 (Fig. 7.2d, e). For both models, a median

kLa of 0.045 s-1 was used. The maximum film density (σfilm,max) was chosen as fit parameter

for the mechanistic model. The best accordance between measuring values and model was

achieved with a σfilm,max of 7 g/l. The fit parameter of the not mechanistic biofilm model is the

constant immobilized biomass concentration (cXimm,bulk). A cXimm,bulk of 0.42 g/g described the

results best. All other model parameters are left at the standard values depicted in Tab. 8.1.

8.5 Results and discussion

8.5.1 General behavior of the biofilm reactor

Fig. 8.2 depicts the influence of the dilution rate on the behavior of the biofilm reactor under

the standard conditions specified in Tab. 8.1. The variables depicted in Fig. 8.2 can all (with

some restraints for the immobilized biomass) be measured under sterile conditions with

standard equipment (dry weight, exhaust gas analyzer, dissolved oxygen electrode, HPLC,

weighing of the support before and after immobilization). Thus, Fig. 8.2 reflects the behavior

of the biofilm reactor likely to be observed by most operators. In an aerobic system with

growth associated product formation, space time yield (STY), oxygen transfer rate (OTR) and

dissolved oxygen (DO) are linked. With increasing dilution rate, STY and OTR increase

linearly, and DO decreases linearly until the activity maximum of the reactor is reached. At

the activity maximum STY and OTR reach a maximum (STYmax, OTRmax) and the DO a

minimum (DOmin).

Page 145: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 127

0.0 0.2 0.4 0.6 0.8 1.00.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

STY

OTR

cS

cX

cXimm

DO

c X,bu

lk

[g/l]

c S,

bulk

[g/l]

c Xim

m,b

ulk

[g/l]

S

TY [

g/l/h

]

dilution rate (D) [h-1]

0

20

40

60

80

100DCBAFig. 8.3

STYmax

DO

[%

], O

TR

[mm

ol/l/

h]

Fig. 8.2: The performance of an aerobic biofilm reactor at different dilution rates. Arrows on top of the

figure refer to the dilution rates, at which Figs. 8.3a-d were created. Model parameters: kLa=0.1 s-1,

cS,bulk,feed=5 g/l, A/V=0.6 m²/l, ρfilm,max=70 g/l, pR=1 bar. Nomenclature: A/V specific surface area of the

support, cS,bulk carbon source concentration bulk phase, cS,bulk,feed carbon source concentration in the

feed medium, cX,bulk biomass concentration bulk phase, cX,bulk,imm immobilized biomass concentration

related to the liquid volume of the bulk phase, DO dissolved oxygen, OTR oxygen transfer rate, pR

reactor pressure, ρfilm,max biofilm density.

At dilution rates exceeding the activity maximum STY and OTR decrease and DO increases.

The carbon source concentration in the bulk phase (cS,bulk) is close to zero until the activity

maximum is reached and then increases sharply. Due to the maintenance, the biomass

concentration in the bulk-phase (suspended biomass) (cX,bulk) increases sharply at low dilution

rates (D). With a further increase of D, cX,bulk reaches a slightly increasing plateau. At D

above the activity maximum cX,bulk decreases, however the reactor is never completely

Page 146: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 128

washed-out due to the growth of the biofilm. Similar to the suspended biomass the

immobilized biomass concentration in the reactor (cX,bulk,imm) increases sharply at low D. With

a further increase of D, cX,bulk,imm increases linearly until it reaches a peak (cX,bulk,imm,max) then

decreases sharply to a local minimum and with a further increase of D slightly increases

again. The local minimum corresponds to the activity maximum of the reactor. The sharp

peak of the immobilized biomass may be one of the reasons for the sometimes observed

oscillating behavior of biofilm reactors. The data shown in Fig. 8.2 is not sufficient to

mechanistically identify the dilution rate with the maximum productivity (STYmax), or to

explain the sharp decrease of the immobilized biomass. A mechanistic explanation can only

be derived, if the behavior of the biofilm is additionally analyzed at different dilution rates.

Fig. 8.3 depicts growth rates and concentrations inside the biofilm at four characteristic

dilution rates. The variables depicted in Fig. 8.3 are usually measurable only in especially

designed biofilm reactors with particular equipment (micro electrodes, confocal laser

scanning microscopy). Thus, for engineering purposes, they are only accessible with

modeling. Fig. 8.2 and Fig. 8.3 are derived from the same simulation, allowing comparisons

of the behavior of the entire biofilm reactor and the behavior of the biofilm. At the bulk-

biofilm interface, the solute concentrations (hence also the growth rate) are equal in the bulk-

and the biofilm phase (assumption 3, Fig. 8.1c). That means, the values depicted in Fig. 8.2

match the values with a film thickness of 0 µm (Y-axis, h=0) depicted in Fig. 8.3. Generally,

in a biofilm reactor (Fig. 8.3) the growth rates in the bulk-phase and in the biofilm are always

smaller than the dilution rate (µbulk<D, µfilm<D), different to a standard continuous reactor in

which the growth rate is equal to the dilution rate (µbulk=D). The biomass concentration jumps

at the bulk-biofilm interface from the low concentration in the bulk-phase (cX,bulk ~2 g/l) to the

higher concentration in the biofilm (cX,film), which is according to assumption 4 under steady-

state conditions equal to the maximum biofilm density (cX,film=ρfilm,max=70 g/l).

Page 147: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 129

Fig. 8.3a depicts the characteristics of the biofilm at a low dilution rate of 0.3 h-1. The carbon

source concentration is always lower than 5 times the Monod constant (cS<5 KS) (see also

Fig. 8.4), thus bulk- and biofilm phase are substrate limited. Carbon source concentration and

growth rate gradually decrease inside the biofilm to zero. Oxygen does not limit growth

(cO2>>5 KO2) (see also Fig. 8.4). Fig. 3b depicts the biofilm at the dilution rate with the

maximum immobilized biomass concentration (cX,bulk,imm,max) (D=0.425 h-1). Together with

the dilution rate, carbon source concentration and growth rate increase in the bulk-phase.

Bulk- and biofilm phase are still substrate limited (cS<5 KS) and carbon source concentration

and growth rate still gradually decrease inside the biofilm to zero. The higher carbon source

concentration increases the biofilm thickness. Oxygen does not limit growth in the bulk-phase

(cO2>5 KO2), however, is nearly totally consumed inside the biofilm (see also Fig. 8.4).

Fig. 8.3c depicts the biofilm at the dilution rate with the maximum space time yield

(D=0.525 h-1). The biomass in the bulk-phase grows with the maximum growth rate, it is

neither carbon source nor oxygen-limited (cS>>5 KS, cO2>5 KO2). Also inside the biofilm

substrate does not limit growth (cS>>5 KS). The oxygen concentration gradually decreases to

zero inside the biofilm. In contrast to the substrate limited cases this gradual decrease is not

reflected by the growth rate. The growth rate remains close to the maximum growth rate and

then, with the exhaustion of oxygen, drops sharply. The change from a carbon source limited

biofilm to an oxygen-limited biofilm is accompanied by a decrease of the film thickness,

reflecting the sharp decrease of the immobilized biomass observed in Fig. 8.2. Fig. 8.3d

depicts the biofilm at a high dilution rate (D=0.6 h-1). The general behavior of the biofilm is

similar to the characteristics at D=0.525 h-1. Only the film thickness is slightly higher. The

higher carbon source concentration does not influence growth, because already at D=0.525 h-1

growth was solely oxygen-limited.

Page 148: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 130

h=hsup h=0

0.00

0.05

0.10

0.15

0.20

0.25

0.30

0.35

0.40

c S,fil

m

[mm

ol/l]

, µfil

m

[h-1]

support

µ

cO2,film

c O2,f

ilm

[mm

ol/l]

B

0 20 40 60 800.00

0.05

0.10

0.15

0.20

0.25

0.30

0.35

c X,fil

m

[g/l]

C

film thickness [µm]0 20 40 60 80 100

D0

10

20

30

40

50

60

70

80cX,film A

c X,fil

m

[g/l]

cS,film

c S,fil

m

[mm

ol/l]

0

10

20

30

40

50

60

70

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

c O2,f

ilm

[mm

ol/l]

, µfil

m

[h-1]

0

5

10

15

20

25

30

35

40

c S,fil

m

[mm

ol/l]

Fig. 8.3: Concentrations and growth rates inside the biofilm at different dilution rates (arrows on top of

Fig. 8.2). A. dilution rate (D) below the maximum growth rate of the microorganism D=0.3 h-1.

B. dilution rate with maximum immobilized biomass concentration D=0.425 h-1, C. dilution rate with

maximum space time yield (STYmax) D=0.525 h-1, D. dilution rate higher than STYmax D=0.6 h-1. At the

x-axis=0, according to the boundary condition, the concentrations of the bulk-phase are equal to the

concentrations in the biofilm-phase. The model parameters are equal to the simulation of Fig. 8.2.

Nomenclature: cO2,film oxygen concentration in the biofilm, cS,film carbon source concentration in the

biofilm, cX,film biomass concentration in the biofilm, µfilm growth rate in the biofilm.

The analysis of the biofilm at different dilution rates was attempted to mechanistically explain

the relationship between dilution rate and productivity, to allow the identification of the

dilution rate with the maximum productivity (STYmax), and to explain the sharp peak of the

immobilized biomass observed in Fig. 8.2. The mechanistic explanation can be gained, by

combining three arguments:

Page 149: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 131

1. The productivity of the biofilm, that is the biomass transfer between bulk- and

biofilm-phase (rX,bulk↔film), which according to Eqs. 8.1 and 8.2 directly influences the

productivity of the biofilm reactor (STY), is determined by the maximum film

thickness, that is the total concentration of immobilized biomass (cXbulk,imm), and by

the growth rate of the immobilized biomass.

2. The amount of carbon source fed per time (D·cS,bulk,feed) depends on the dilution rate,

however, the amount of oxygen fed per time, that is the gas-liquid mass transfer

(rO2,gas↔bulk), is independent of the dilution rate (kLa=const.).

3. The Monod constant of the carbon source is higher than the Monod constant of

oxygen (KS>KO2).

Similar to a standard continuous reactor at low dilution rates bulk- and biofilm-phase are

carbon source limited (Fig. 8.3a). An increase of D increases the amount of carbon source fed

per time. As long as this additional amount of carbon source can be consumed by the

suspended and/or the immobilized biomass the productivity of the reactor increases as

depicted in Fig. 8.2. The additional carbon source can be consumed through an increase of the

biomass concentration (increase of the film thickness, Fig. 8.3a to Fig. 8.3b) and/or by an

increase of the growth rate (Fig. 8.3a to 8.3b to 8.3c), however, this increased productivity

leads to an increased oxygen consumption as described by the OTR curve in Fig. 8.2. Thus,

because the gas-liquid mass transfer is constant and independent of D, at a certain dilution

rate the reactor has to become oxygen-limited.

Fig. 8.4 depicts the concentrations of oxygen and carbon source at the support surface

(h=hsup). To highlight their influence on the growth rate, the concentrations were divided by

their respective Monod constants (cS,film/KS, cO2,film/KO2). The figure is divided into two parts,

with a small transition region between them. At low dilution rates the concentration of the

Page 150: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 132

carbon source is nearly zero at the support surface and the oxygen concentration is high. At

high dilution rates the situation is reversed, the oxygen concentration is low and the carbon

source concentration high. The transition region coincides with the steep decrease of the

biofilm thickness. The sharp decrease of cX,bulk,imm results from the difference of the Monod

constants (KS>KO2). If the Monod constant of the carbon source (KS) is set equal to the

Monod constant of oxygen (KO2) the peak disappears (diagram not shown) (see also chapter

8.4.4, Fig. 8.7d). The course of the growth rate inside the biofilm corresponds to the Monod

kinetics of the limiting substrate. The high Monod constant of the carbon source leads to a

gradual decrease of the growth rate inside the biofilm (Figs. 8.3a, b) under carbon source

limited conditions. The low Monod constant of oxygen leads to a nearly constant growth rate

in the biofilm with a sharp decrease near the support surface (Figs. 8.3c, d) under oxygen-

limited conditions. During the transition from carbon source to oxygen-limited conditions the

thick and slow growing biofilm is transformed in a slim but fast growing film, resulting in the

observed sharp decreases of the immobilized biomass concentration.

In conclusion, the dilution rate with the maximum productivity is reached when the biofilm is

completely transformed from a carbon source to an oxygen-limited state. The sharp peak of

the immobilized biomass results from the different Monod constants of carbon source and

oxygen.

Page 151: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 133

Fig. 8.4: The transition from a carbon

source limited to an oxygen-limited

biofilm. Comparison of the immo-

bilized biomass concentration in the

reactor (cX,bulk,imm) with the concen-

trations of oxygen (cO2,film), and car-

bon source (cS,film) at the support

surface (h=hsup, Fig. 8.3a). Oxygen

and carbon source concentrations

have been divided by their Monod constants (cO2,film/KO2, cS,film/KS). The model parameters are equal to

the simulation of Fig. 8.2.

0.35 0.40 0.45 0.50 0.550.01

0.1

1

510

100

cO2,film/KO2

cS,film/KS

at

the

supp

ort s

urfa

ce h

=hsu

pc S,

film/K

S; c

O2,f

ilm/K

O2

[-]

dilution rate (D) [h-1]

0

1

2

3

cX,bulk,imm,max STYmax

cX,bulk,imm

c X,bu

lk,im

m

[g/l]

8.5.2 Influence of mass transfer and carbon source concentration on the productivity

Fig. 8.5 depicts the influence of the two generally most easily manipulable operating

parameters, carbon source concentration of the feed medium (cS,bulk,feed) and volumetric mass

transfer coefficient (kLa) on the productivity. It compares at a constant kLa value the

maximum space time yield (STYmax) of standard reactor and biofilm reactor at different

cS,bulk,feed. While cS,bulk,feed is, except for applications in waste water treatment, usually freely

adjustable over a wide range (solubility of glucose monohydrate in water at 25° C 820 g/l) the

kLa value can usually only be changed in a much smaller range by varying e.g. the aeration

rate or the stirrer speed. The two kLa values selected for the simulations in Fig. 8.5 represent

the kLa range for loop reactors at a specific power input of 1 kW/m³ [Blenke 1979]. STYmax

was derived from simulations of the type depicted in Fig. 8.2. The dilution rate, at which

STYmax occurs, is not fixed, it is different for each set of operating conditions. If cS,bulk,feed is

low, the productivity of the biofilm reactor is distinctly higher than the productivity of the

standard reactor (Fig. 8.5a). At low cS,bulk,feed a tripling of the kLa value does not increase

Page 152: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 134

STYmax of the standard reactor and leads only to a small increase of STYmax of the biofilm

reactor. With an increase of cS,bulk,feed the STYmax of standard and biofilm reactor increase

linearly, with the slope of the biofilm reactor being flatter than the slope of the standard

reactor. STYmax increases with cS,bulk,feed until it is equal for both reactors. At even higher

cS,bulk,feed the STYmax of both reactors remains at the same constant value. A higher kLa value

leads to a higher constant value, however, it does not change the general characteristics. The

reason for this behavior can be derived from Fig. 8.5b, which depicts the dissolved oxygen

(DO) at STYmax. At low cS,bulk,feed the DO is high, therefore, a thick and productive biofilm

can develop, which leads to an advantage of the biofilm reactor. With increasing cS,bulk,feed and

STYmax the DO decreases, resulting in a thinner and less productive biofilm. At high cS,bulk,feed

the DO of both reactors is close to zero, thus no aerobic biofilm can be formed.

Consequentially biofilm and standard reactor behave equally. In other words, both reactors

have the same and the highest productivity, if the mass transfer capacity of the reactor is fully

used. The interaction of productivity and mass transfer capacity of standard reactors has been

described by Hollander [1993].

0

1

2

3

4

5

kLa=0.1 s-1

kLa=0.3 s-1A

STY m

ax

[g/l/

h]

0 10 20 30 40 50

20

40

60

80

0

Fig. 8.5: Comparison of ——— biofilm reactor and

• • • • • standard reactor without immobilized biomass.

The influence of the carbon source concentration of

the feed medium (cS,bulk,feed) and the volumetric mass

transfer coefficient (kLa) on: A. the productivity

(STYmax, maximum space time yield) and B. the

dissolved oxygen (DO) at STYmax. Model parameters:

kLa=0.1 s-1 and kLa=0.3 s-1, cS,bulk,feed=5-50 g/l, biofilm

reactor A/V=0.6 m²/l, standard reactor A/V=0 m²/l,

ρfilm,max=70 g/l, pR=1 bar.

cS,bulk,feed [g/l]

kLa=0.1 s-1

kLa=0.3 s-1

B

DO

at S

TYm

ax

[%]

Page 153: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 135

Fig. 8.6a compares the productivity of the two reactor types at a constant DO. The DO was

numerically kept constant by adjusting the reactor pressure (pR). Under these conditions, the

mass transfer cannot constrain the development of the biofilm. At a constant DO the biofilm

reactor has a constant absolute advantage to the standard reactor, however, with increasing

carbon source concentration of the feed the relative advantage of the biofilm reactor

decreases. Considering the more complicated operating procedure of the biofilm reactors, they

have an advantage to standard reactors only at low carbon source concentrations, even if the

mass transfer is not limiting the productivity of the biofilm.

In conclusion, for aerobic processes with growth associated product formation the simpler and

well established standard reactor is preferable for most applications in industrial

biotechnology, where the carbon source concentration of the feed can be adjusted without

restraints. For applications in waste water treatment, where the carbon source concentration in

the feed is typically low and normally can not be adjusted, a biofilm reactor can be preferable.

0.0 0.5 1.0 1.5 2.0 2.5 3.00

2

4

6

8

10

cS,bulk,feed=5 g/l

cS,bulk,feed=20 g/l

cS,bulk,feed=50 g/lB

STY m

ax

[g/l/

h]

A/V [m2/l]

0 10 20 30 40 50 600

2

4

6

8

standard reactor

biofilm reactor

A

STY m

ax

[g/l/

h]

cS,bulk,feed [g/l]

1

2

3

4

5

reac

tor p

ress

ure

(pR)

[b

ar]

1

2

3

4

5

6

reac

tor p

ress

ure

(pR)

[b

ar]

Fig. 8.6: The influence of the carbon source

concentration of the feed medium (cS,bulk,feed) and

the specific surface area of the support (A/V) at a

constant dissolved oxygen (DO) of 60% on the

productivity (STYmax, maximum space time yield).

DO was kept constant by adjusting the reactor

pressure (pR). Model parameters: DO=60 %,

kLa=0.1 s-1, cS,bulk,feed=5-50 g/l, ρfilm,max=70 g/l, A.:

biofilm reactor A/V=0.6 m²/l, standard reactor

A/V=0 m²/l.

Page 154: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 136

8.5.3 Influence of specific surface area and carbon source concentration on the productivity

Monolith loop biofilm reactors, which have moved in the focus of research in recent years

[Chapter 7, Kawakami et al. 1989, Tata et al. 1999, Wojnowska-Baryla et al. 1999, Yongming

et al. 2002] generally, use commercially available honeycomb monoliths. To study the

influence of reactor design and specific surface area on the productivity, the commercially

available SiC honeycomb monolith as described in appendix A.4 was selected. Fig. 8.6b

depicts the influence of the specific surface area (A/V) at a constant dissolved oxygen of 60 %

on the productivity. The standard reactor is depicted at an A/V of zero. The theoretical

maximum of A/V=2.9 m²/l is reached, if only the liquid volume inside the capillaries of the

monolith is considered as V. From Fig. 8.6b it can be concluded, that independent of the

substrate concentration of the feed (cS,bulk,feed) the productivity (STYmax) increases linearly

with the specific surface area. However, as depicted in Fig. 6a the relative advantage of the

biofilm reactor decreases with increasing cS,bulk,feed.

In conclusion, an increase of the specific surface area leads to a linear increase of the biofilm

reactor productivity if the dissolved oxygen is kept constant.

8.5.4 Influence of diffusion coefficient, film density and Monod constant on the productivity

As described previously, the biofilm is oxygen-limited at STYmax and the productivity of the

biofilm is determined by the immobilized biomass concentration in the reactor (cX,bulk,imm),

and by the growth rate of the immobilized biomass. Under these conditions the model

parameters diffusion coefficient of oxygen in the biofilm (DDO2,film), Monod constant of

oxygen (KO2) and maximum film density (σfilm,max) should have a high impact on the

productivity of the biofilm reactor. Fig. 8.7a depicts the influence of these three model

parameters on STYmax. Figs. 8.7b-d show the growth rate inside the biofilm and the film

Page 155: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 137

thickness when one of the parameters (DDO2,film, KO2, σfilm,max) is varied. DO2,film/DDO2,H2O was

varied between 0.15-1 and ρfilm,max between 5 g/l and 90 g/l, thereby covering the commonly

observed range of microbial biofilms [Stewart 1998]. KO2 was varied between 1 µM and

15 µM covering the main range of bacterial cultures [Owens and Legan 1987]. For the

simulations depicted in Fig 8.7 a single parameter was always varied, the others were left at

the standard value specified in Tab. 8.1.

It may be expected that a higher biofilm density, an easier penetrable biofilm and a lower

Monod constant would increase the productivity of the biofilm reactor. However, even great

parameter changes lead only to moderate changes of STYmax (Fig. 8.7a), e.g an increase of

ρfilm,max by a factor of 18 from 5 g/l to 90 g/l leads only to an increase of STYmax by a factor of

1.4 from 0.81 g/l/h to 1.14 g/l/h. Two self-stabilizing factors are mainly responsible for the

moderate changes of the productivity: first, any increase of the productivity leads to an

increase of the oxygen consumption, resulting in a decrease of the dissolved oxygen (kLa=

const.). Consequently, the productivity of the biofilm can increase only slightly. Second, the

variation of a biofilm property is compensated by a change of the film thickness, leaving the

total productivity of the biofilm nearly unchanged. E.g. the increase of ρfilm,max from 5 g/l to

90 g/l leads to a decrease of the film thickness from ~200 µm to ~40 µm (Fig. 8.7c).

Variations of the Monod constant change the characteristics of the growth rate in the biofilm.

At low KO2 the growth rate remains at a plateau and then decreases sharply (Fig. 8.7d,

Figs. 8.3c, d). An increase of KO2 leads to a more gradual decrease of the growth rate

comparable to the biofilm under carbon source limitation (Fig. 8.7d, Figs. 8.3a, b). In

conclusion, inaccuracies of the model parameters DDO2,film, KO2 and σfilm,max have only a minor

influence on the maximum productivity of a biofilm reactor with a growth associated product

formation, because variations of the parameters are compensated by the biofilm properties

Page 156: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 138

film thickness and film growth rate. Or vice versa, even relatively large changes of the biofilm

properties are difficult to detect with a measuring set-up, which determines only growth

associated variables of the entire biofilm reactor like suspended biomass, oxygen transfer rate

or dissolved oxygen.

0 20 40 60 80 100

0.8

0.9

1.0

1.1

1.2

1.3A

ρfilm,max

STY m

ax

[g/l/

h]

maximum film density (ρfilm,max) [g/l]

0 50 100 150 2000.0

0.1

0.2

0.3

0.4

70 [g/l]

50 [g/l] 20 [g/l]

5 [g/l]

90 [g/l]

D KO2C ρfilm,max

film thickness [µm]

µ film

[h-1]

0 2 4 6 8 10 12 14 16

DDO2

KO2

Monod constant oxygen (KO2) [µM]

0.0 0.2 0.4 0.6 0.8 1.0DDO2,film

/DDH2O [-]

0 25 50 750.0

0.1

0.2

0.3

0.4

1 [µM]2.5 [µM]

15 [µM]

µfil

m

[h-1]

film thickness [µm]

0 25 500.0

0.1

0.2

0.3

0.4

75

0.15 [-] 0.4 [-] 0.634 1 [-]

µfil

m

[h-1]

B DDO2,film/DDO2,H2O

film thickness [µm]

Fig. 8.7: The influence of the oxygen diffusion coefficient (DDO2,film), the maximum film density

(σfilm,max), and the Monod constant of oxygen (KO2) on the biofilm. The diffusion of oxygen is graphed

as DDO2,film/DDO2,H2O (relation between the diffusion coefficient in the biofilm and the diffusion coefficient

in water). A. The influence of DDO2,film/DDO2,H2O, σfilm,max and KO2 on the maximum space time yield

(STYmax) of the biofilm reactor. The influence of B. DDO2,film/DDO2,H2O, C. σfilm,max, and D. KO2 on the

growth rate inside the biofilm (µfilm) and the film thickness. Model parameters: kLa=0.1 s-1,

cS,bulk,feed=5 g/l, A/V=0.6 m²/l, pR=1 bar standard values: (in B.-D. depicted with a solid line)

DDO2,film/DDO2,H2O=0.634, ρfilm,max=70 g/l, KO2=1 µM. For each simulation one parameter was varied, the

others left at the standard value.

Page 157: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 139

8.5.5 Comparison of experimental results and model predictions

Fig. 8.8 compares the culture characteristics of a biofilm reactor (BFR) and a stirred tank

reactor (STR) at different dilution rates (D). To facilitate the comparison the ratio between the

Y-axes is equivalent to the ratio of the carbon source concentrations in the feed medium. At

dilution rates lower than the maximum growth rate of 0.32 h-1 (Chapter 4.5.2.1), the culture

characteristics of BFR and STR are comparable and both can be described with the same

kinetic parameters (yield and maintenance coefficients). The culture characteristics of the

STR have been described in detail in chapter 4.5.2. At dilution rates higher than the maximum

growth rate the STR washes out, all concentrations become zero or the feed or inlet

concentration (the values of the STR higher than µmax depicted in Fig. 8.8 have not reached

the steady state). The activity of the biofilm prevents the wash-out of the BFR, suspended

biomass concentration (cX,bulk), space time yield (STY) and oxygen transfer rate (OTR) slowly

decrease with increasing dilution rates while the dissolved oxygen (DO), and the lactic acid

concentration (cS,bulk) increase towards the inlet or feed concentration. Even at a dilution rate

twice the maximum growth rate, none of the concentrations has reached zero or the feed

concentration.

It is possible to describe the culture characteristics of the BFR with the mechanistic (biofilm

thickness and growth are functions of the substrate concentrations) and with the non

mechanistic (immobilized biomass concentration always constant) biofilm reactor models

(Fig. 8.8). The non mechanistic model predicts that the BFR has a slightly higher maximum

productivity (STYmax) than the STR. The mechanistic model predicts an equal STYmax for

BFR and STR.

Page 158: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 140

Fig. 8.8: Comparison of biofilm

reactor (BFR, left Y-axis, solid

symbols) and stirred tank reactor

(STR, right Y-axis, open symbols).

Comparison of measured (symbols)

and modeled (lines) reactor behavior.

——— mechanistic biofilm model

(concentrations and activities of the

immobilized biomass are functions of

carbon source and oxygen concen-

trations), - - - - non mechanistic

biofilm model (the immobilized bio-

mass concentration is a constant

independent of the culture con-

ditions). Culture conditions: BFR;

filling volume 1.7 l, specific aeration

rate 1.75 vvm, liquid flow rate

19 l/min, DEAE dextran surface

modified SiC monolith, lactic acid

concentration feed medium 5 g/l;

STR; filling volume 1 l, stirrer speed max. 1200 rpm, specific aeration rate 2 vvm, lactic acid

concentration feed medium 11 g/l, for dilution rates of the STR which exceed the washout point

steady-state conditions have not completely been reached. Model parameters: mechanistic model;

volumetric mass transfer coefficient (kLa) 0.045 s-1, maximum film density (ρfilm,max) 7 g/l; non

mechanistic model kLa 0.045 s-1, concentration of the immobilized biomass (cXimm,bulk) 0.42 g/l; all

other parameters as depicted in Tab. 8.1. Nomenclature: cX,bulk suspended biomass concentration,

STY space time yield, OTR oxygen transfer rate, DO bottom dissolved oxygen measured below the

monolith (Fig. 7.1), cS,bulk lactic acid concentration.

0.00.51.01.52.02.53.0

A

c X,bu

lk,B

FR

[g/l]

0.00

0.01

0.02

0.03C

OTR

BFR

[mol

/l/h]

0.0 0.1 0.2 0.3 0.4 0.5 0.6012345

E

dilution rate (D) [h-1]

c S,bu

lk,B

FR

[g/l]

0

20

40

60

80D

DO b

otto

m BF

R [%

] 0.00

0.02

0.04

0.06

OTR

STR

[mol

/l/h]

0123456

c X,bu

lk,S

TR

[g/l]

024681012

c S,bu

lk,S

TR

[g/l]

0.0

0.2

0.4

0.6

0.8B

STY BF

R [g

/l/h]

0.0

0.4

0.8

1.2

1.6

STY ST

R [g

/l/h]

Page 159: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 141

8.5.6 Discussion of the comparison of experimental results and model predictions

BFR and STR could be described with the same kinetic parameters. This is in good agreement

with the result presented in chapter 6.3.2.2, that suspended and immobilized biomass have the

same respiration. It further confirms that the assumptions made by model assumption 5

(Chapter 8.2.3) are reasonable.

The non mechanistic model predicts a higher maximum productivity (STYmax) than the

mechanistic model. The non mechanistic model does not consider the influence of the low

dissolved oxygen at dilution rates close to STYmax on the activity and thickness of the biofilm

(Figs. 8.8d, 8.5).

With both models, it is possible to describe the culture characteristics of the BFR. Several

reasons could be responsible for this observation. The biofilm could not be limited by the

oxygen or carbon source as assumed by the mechanistic biofilm model (Chapter 8.2.3,

assumption 4). It could instead be limited by shear stress, resulting in a constant immobilized

biomass concentration. The biofilm was maybe not fully developed and steady state

conditions had not been reached. Biofilms often need weeks to reach a steady state [Léon Ohl

et al. 2004].

The immobilized biomass concentration is low, and the predicted maximum productivities of

BFR and STR are similar. As described in chapter 8.5.2 BFR and STR have equal

productivities, if the mass transfer capacity of the reactor is entirely used. Therefore, to

improve the productivity of the BFR it is necessary to increase the gas-liquid mass transfer

[Henkel et al. 1990]. A higher gas-liquid mass transfer would increase the advantage of the

BFR, and allow operating the BFR at higher carbon source concentrations. High volumetric

mass transfer coefficients (kLa) are difficult to achieve because biological fermentations

require antifoam agent. As described in chapter 7.3.1, antifoam agent leads to a strong

Page 160: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 8 Modeling aerobic biofilm reactors with growth associated product formation 142

reduction of kLa. Further, the flow inside the capillaries is laminar even at high liquid flow

rates [Chapter 7.4.6]. In some cases, it could be a good alternative to increase the reactor

pressure instead of kLa [Knoll et al. 2005].

Page 161: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 9 Summary of conclusions and experimental results 143

Chapter 9

Summary of conclusions and experimental results

This chapter gives a summary of the most important conclusions and experimental results

presented in this work.

9.1 Aim and introduction

This study focuses on the process development and scale-up of aerobic microorganisms,

which may be suspended (planctonic, free) or immobilized on a support. The aim was to

develop novel and advance existing techniques and methods required to perform process

developments of aerobic microorganisms with shake flasks and fermenters efficiently. The

applicability of the techniques was demonstrated with the biological model system

Corynebacterium glutamicum grown on lactic acid. The process development strategy for

aerobic microorganisms was based on two principels: I. respiration and oxygen consumption

are of fundamental importance for the behavior and productivity of aerobic microorganisms,

and II. to cope with the large numbers of experiments required to develop and optimize an

industrial production, a simple and cost efficient screening system is required, shake flasks are

suitable to serve this purpose.

Page 162: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 9 Summary of conclusions and experimental results 144

9.2 Characterization of the biological model system C. glutamicum with respiration measurement in shake flasks

Respiration measurement in shake flasks was used to characterize the metabolic activity of

microorganisms. The method can be applied for media and strain development, and to

characterize the behavior of microorganisms during and after stress exposure. The new

method combines the simplicity, low resource requirements, and possibility to perform

several experiments simultaneously of shake flask cultures with the possibility to online

monitor the culture characteristics of fermenter cultures with exhaust-gas analysis. For stress

analysis, which is usually performed by determining the viable cell number with petri dishes

or the activity with respirometry, the new method has the additional advantages that the

metabolic activity can be determined independent from manual sampling and without the

necessity to change the culture vessel or the cultivation medium. This excludes stress-factors,

which may be induced by transferring the microorganisms to plates or respirometers. These

advantages allow the productivity to be increased by performing screening experiments in

online monitored shake flasks, which traditionally are performed in fermenters. Additionally,

the online monitoring increases the knowledge gained from a single shake flask experiment

and facilitates the identification of parameters critical for process development.

The applicability of the method was verified by characterizing the behavior of

Corynebacterium glutamicum grown on the carbon source lactic acid under the stress factors

carbon starvation, anaerobic conditions, lactic acid, osmolarity and pH. The following

conditions had no effect on the metabolic activity of C. glutamicum: a carbon starvation of up

to 19 h, anaerobic conditions, lactic acid concentrations up to 10 g/l, MOPS buffer

concentrations up to 42 g/l, or pH from 6.4 to 7.4. Lactic acid concentrations from 10 g/l to

30 g/l lead to a decrease of the growth rate and the biomass substrate yield without effecting

the oxygen substrate conversion. Without adaptation, the organism did not grow at pH<5 or

Page 163: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 9 Summary of conclusions and experimental results 145

pH>9. The negative influence which interruptions of the shaker during sampling may have on

the growth of microorganisms was demonstrated.

9.3 Evaluation of respiration measurement techniques

The accuracy, precision, quantitation limit and range of the novel measuring technique

respiration measurement in shake flasks are verified by comparison with the standard methods

exhaust gas analysis and respirometry. Respiration measurement techniques are difficult to

compare, due to the vast variety of devices and analytical procedures commonly in use.

Corynebacterium glutamicum complex and minimal medium cultures were used to compare

and evaluate the devices. With all measuring devices, it was possible to determine the general

culture characteristics. The exhaust gas analyzer and the RAMOS device produced almost

identical results. The scatter of the respirometer was noticeably higher. The exhaust gas

analyzer is the technique of choice, if the reaction volume is high or a short reaction time is

required. The possibility to monitor cultures simultaneously makes the RAMOS device an

important tool for media and strain development. If online monitoring is not compulsive, the

respiration of the investigated microbial system extremely low, or the sample size small, a

respirometer is recommended.

9.4 Development of a scale-up approach for aerobic cultures

To transfer the results gained with shake flask screening methods to fermenters, a scale-up

strategy is required. The scale-up strategy should allow to select the operating conditions of

aerobic cultures in shake flasks and fermenters. A practical scale-up strategy for everyday use

is introduced for the scale-up of aerobic cultures from shake flasks to fermenters in batch and

continuous mode. The strategy is based on empirical correlations of the volumetric mass

transfer coefficient (kLa) and the pH. It makes it possible, to set oxygen not limited culture

Page 164: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 9 Summary of conclusions and experimental results 146

conditions by adjusting the gas-liquid mass transfer to the respiration of the microorganisms.

It predicts the influence of the shake flask operating conditions: flask size, filling volume,

shaking frequency, and shaking diameter on the gas-liquid mass transfer, and the influence of

growth rate, biomass concentration, and limiting substrate concentration on the respiration.

The accuracy of the empirical kLa correlations and the assumptions required to use these

correlations for an arbitrary biological medium were discussed. To determine the optimal pH

of the culture medium, a simple laboratory method based on titration curves of the medium

and a mechanistic pH model, which is solely based on the medium composition, was applied.

The effectiveness of the scale-up strategy was demonstrated by comparing the behavior of

Corynebacterium glutamicum on lactic acid in shake flasks and fermenters in batch and

continuous mode. The maximum growth rate (µmax=0.32 h-1) and the oxygen substrate

coefficient (YO2/S=0.0174 mol/l) of C. glutamicum on lactic acid were equal for shake flask,

fermenter, batch, and continuous cultures. The biomass substrate yield was independent of the

scale, but it was lower in batch cultures (YX/S=0.36 g/g) than in continuous cultures

(YX/S=0.45 g/g). The experimental data of shake flask and fermenter cultures in batch and

continuous mode could be described with the same biological and mechanistic pH model.

9.5 Development of a screening approach for immobilized biomass on solid supports

A generic, simple, and inexpensive screening method to perform immobilization studies with

solid supports in shake flasks was introduced. The screening method uses solely standard

histology and microbiology laboratory equipment. Histology cassettes were used to

standardize hydrodynamic conditions and to protect immobilized biomass and solid support.

Histology cassettes have the following advantages: they are readily available, inexpensive,

solvent and acid resistant, automatable, and the slots in the cassette walls allow liquid to

circulate freely. A new rotating camera, which produces a stable image of the rotating liquid

Page 165: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 9 Summary of conclusions and experimental results 147

in the shake flask, was used, to observe movement and position of histology cassettes in shake

flasks. This visualization method revealed that in a 250 ml shake flask, stable operating

conditions are achieved at a shaking frequency of 300 rpm, a filling volume of 30 ml, and a

shaking diameter of 50 mm. In vessels with vertical walls, such as beakers or laboratory

bottles, the movement of the histology cassette is not reproducible. Mass transfer

characterization with the biological model system Corynebacterium glutamicum and the

chemical sulfite-oxidation method revealed that histology cassettes do not influence the gas-

liquid mass transfer of shake flasks.

9.6 The influence of material properties and surface modifications on the adhesion of dormant and growing cells

A newly developed screening method for immobilized biomass on solid supports was applied

to study the adhesion of dormant (artificial salt solution), and growing (biological medium)

Corynebacterium glutamicum on glass, silicon carbide and cordierite supports with different

porosities and surface modifications (polyethylenimin and DEAE dextran). To relate material

properties and adhesion, the hydrophobicity of the microorganism (42.5°) and support

materials (not modified microscope slices 24.1°, polyethylenimin modified 45.3°, DEAE

dextran modified 31.7°) was determined with contact angel measurements. Density, porosity,

and pore size distribution of the support materials were measured with mercury porosimetry.

The activity of dormant C. glutamicum in a NaCl solution decreased linearly over time. With

respiration measurement in shake flasks the respiration of suspended cells and immobilized

cells was found equal. Compared to not modified materials, surfaces modified with DEAE

dextran improved adhesion more than surfaces modified with polyethylenimin. Materials with

pore diameters between 10 µm and 30 µm had the highest adhesion. None porous materials or

materials with pores smaller or equal to the cell size showed only little adhesion. Even so, the

adhesion profiles over time of dormant cells in artificial salt solutions were different to the

Page 166: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 9 Summary of conclusions and experimental results 148

adhesion profiles of growing cells in media. For an initial screening it is reasonable to assume

that results gained with dormant cells in artificial salt solutions can predict the behavior of

growing cells in medium.

9.7 Characterization of a novel biofilm reactor with a honeycomb monolith as immobilization carrier

A novel external loop biofilm reactor is introduced. The riser of the reactor consists of a

ceramic honeycomb monolith, which serves as immobilization carrier. The riser of the reactor

was equipped with a monolith or alternatively with a standard single tube. The influence of

the monolith on mass transfer, specific power input, pressure drop, gas hold-up, and mixing

time was analyzed. Experiments were performed with the biological model system

Corynebacterium glutamicum or with a buffer solution with the same ionic strength as the

biological medium.

The relation between the volumetric mass transfer coefficient with antifoam agent (kLa) and

without antifoam agent lay between 0.3 and 0.7. The kLa was independent of the antifoam

agent concentration at concentrations above 50 ppm, however, more than 1000 ppm were

required to control foaming. Below 15 l/min the kLa was independent of the liquid flow rate,

higher flow rates up to 21 l/min led to an increase by a factor of 2-3. This behavior probably

results from the interaction of power input and gas hold-up. The power input increased

parabolically, while the gas hold-up decreased linearly with the liquid flow rate.

Monolith and single tube had equal mixing times in downcomer and riser, however, the

monolith drastically reduced mixing and mass transfer in the gas separator. These phenomena

can most likely be explained with different flow regimes in the riser. While the flow inside the

single tube is turbulent, the flow in the capillaries of the monolith is laminar.

Page 167: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 9 Summary of conclusions and experimental results 149

9.8 Modeling aerobic biofilm reactors with growth associated product formation

It can take several weeks until a biofilm reaches a steady state, making modeling the only

convenient approach to studying the behavior of biofilm reactors at a vast variety of different

conditions. The productivities of biofilm and standard reactors were compared for aerobic

processes with growth associated product formation. The behavior of the biofilm reactor was

modeled assuming ideally mixed bulk- and gas-phases, and a 1-D biofilm phase with mass

transport solely by diffusion. The influence of dilution rate, kLa value, carbon source

concentration, and surface area on the productivity was investigated. At a constant kLa value

three sections can be differentiated: carbon source limited at low dilution rates, a very small

transition section, and oxygen-limited at high dilution rates. If the Monod constant of the

carbon source is higher than the Monod constant of oxygen, the biofilm is thick and slow

growing under carbon limitation and slim and fast growing under oxygen limitation, resulting

in a sharp decrease of the biofilm thickness during the transition section. The maximum

productivity of the biofilm reactor occurs at the end of the transition from carbon source to

oxygen limitation. Inaccuracies of the model parameters diffusion coefficient, Monod

constant and biofilm density have only a small influence on the maximum productivity of the

reactor because variations are compensated by an adaptation of film thickness and film

growth rate. The model was used, to describe the experimental results of an external loop

biofilm reactor with a honeycomb monolith as immobilization carrier. The model predicts that

standard and biofilm reactors have similar productivities at high carbon source concentrations.

At low concentrations, the biofilm reactor is favorable. Thus, for aerobic processes with

growth associated product formation the simpler and well established standard reactor is

preferable for most applications in industrial biotechnology, where the carbon source

concentration of the feed can be adjusted without restraints. For applications in wastewater

Page 168: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Chapter 9 Summary of conclusions and experimental results 150

treatment, where the carbon source concentration in the feed is typically low and normally

cannot be adjusted, a biofilm reactor can be preferable.

Page 169: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Appendix General material and methods 151

Appendix

General materials and methods

This appendix describes the standard materials and methods used in most experiments

presented in this work. Variations of these standard procedures are described in each chapter

and in the figure legends.

A.1 Microorganism and media

A.1.1 Microorganism

All experiments in this work were carried out with the wild type of Corynebacterium

glutamicum ATCC 13032 (Kinoshita et al. 1957). Corynebacterium glutamicum is widely

used for the industrial production of amino acids [Kircher and Pfefferle 2001]. It is a high-

G+C gram-positive soil bacterium and closely related to the Brevibacteria and Mycobacteria

[Gruber and Bryant 1997, Kinoshita et al. 1957].

A.1.2 Media, cultivation and culture conditions

For all cultivations, a complex medium with the carbon source glucose or a minimal medium

with the carbon source lactic acid was used.

The complex medium contained per liter (letters in brackets refer to the supplier): 20 g

glucose monohydrate (R), 10 g yeast extract (R), 10 g peptone (M), 2.5 g NaCl (M), 0.25 g

MgSO4 (R). The pH was adjusted to 7.2, the glucose was sterilized separately.

Page 170: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Appendix General material and methods 152

The minimal medium contained per liter (letters in brackets refer to the supplier): 10 g lactic

acid (R), 20 g (NH4)2SO4 (R), 1 g KH2PO4 (R), 2 g K2HPO4 (F), 0.25 g MgSO4·7H2O (R) ,

30 mg (HO)2C6H3COOH (R), 10 mg CaCl2·H2O (A), 10 mg MnSO4·H2O (M), 10 mg

FeSO4·7H2O (S), 1 mg ZnSO4·7H2O (F), 0.2 mg CuSO4 (M), 0.2 mg biotin (R), 0.02 mg

NiCl2·7H2O (R). The pH was adjusted to 7 with NaOH (R). The trace elements, the biotin, the

3,4-Dihydroxybenzoic acid ((HO)2C6H3COOH) were sterile filtered, the lactic acid was

autoclaved separately. The lactic contained ≥950 g/l L(+) lactic acid.

The chemicals were supplied by the following companies (underlined letters refer to the

abbreviations used in the media descriptions): AppliChem GmbH, Darmstadt, Germany;

Fluka, St. Gallen, Switzerland; Merck KGaA, Darmstadt, Germany; Roth, Karlsruhe,

Germany; Sigma-Aldrich , St. Louis, MO, USA.

Complex medium plate cultures were used to inoculate complex medium precultures. The

precultures were harvested after 10-12 h. The precultures reached a final pH of ~7.4. Complex

medium main cultures were directly inoculated from the precultures with a biomass

concentration of 0.3 g/l.

For minimal medium main cultures, the precultures were washed twice in 9 g/l NaCl. The

pellet was resuspended in minimal medium and the main culture inoculated with a biomass

concentration of 0.42 g/l. The temperature for all cultivations was 30° C.

A.1.3 Calculation of the media solubilities

The solubilities (complex medium 0.0011 mol/l/bar, minimal medium 0.001 mol/l/bar) were

calculated according to [Rischbieter and Schumpe 1996, Schumpe and Deckwer 1979,

Schumpe at al. 1982]. The effect of lactic acid on SO2 is assumed equal to that of acetate. For

Page 171: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Appendix General material and methods 153

lactic acid, no data was available. However, the error can be assumed small because organic

acids have only a small influence on the solubility.

A.1.4 Recording of titration curves of the media

Titration curves were recorded by stepwise adding 0.5 M NaOH to the medium, never

increasing the pH by more then 0.1. To increase the comparability, the titration curves were

normalized to pH 7. The normalization was done by subtracting the volume of 0.5 M NaOH

necessary to adjust the pH to 7 from the actually titrated volume giving positive volumes for

pH greater than 7 and negative volumes for pH smaller than 7.

A.2 Culture conditions and fermentation devices

A.2.1 Fermenters in batch mode

Fig. A.1a depicts the experimental set-up for the batch fermenter cultures. Fermentations were

carried out in a laboratory fermenter (Biostat M, Braun Biotech, Melsungen, Germany), total

capacity 1.5 l, working volume 1 l, specific aeration rate 2 vvm, rushton turbine (4 blades,

diameter 47 mm, blade height 9 mm). The inlet gas stream of the fermenter was controlled

with a thermal mass flow controller (5850TR, Brooks, Hatfield, PA, USA).

Additional fermentations were carried out in a 50 l fermenter (LP351, Bioengineering AG,

Wald, Switzerland) total capacity 50 l, working volume 15 l, specific aeration rate 0.5 vvm,

3 rushton turbines of those 1 submersed (6 blades, diameter 120 mm, blade height 25 mm), 4

baffles (height 600 mm, width 30 mm). A more detailed description of the fermenter is given

by [Maier et al. 2001].

Page 172: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Appendix General material and methods 154

The dissolved oxygen (DO) was maintained above 30 % by adjusting the stirrer speed. In case

of excessive foam, the antifoam agent (Plurafac LF 1300, BASF, Ludwigshafen, Germany)

was added.

A.2.2 Fermenters in continuous mode

Fig. A.1c depicts the experimental set-up for the continuous fermenter cultures. The fermenter

(Biostat M, Braun Biotech, Melsungen, Germany) used for the batch cultures was equipped

with feed and harvest. Steady state conditions were assumed after at least 5 residence times.

Each dilution rate was determined by measuring the filling volume of the fermenter and the

weight difference of the feedstock. The density of the medium was measured to be 1.01 kg/l.

The pH was kept constant over time at pH 7 by titrating with 2M H2SO4. The amount of

titrant added was recorded with a balance. The density of 2 M H2SO4 at 20° C was 1123 g/l.

A.2.3 Shake flasks in batch mode

Fig. A.1b depicts the fermentation devices used for batch shake flask cultures. Cultures were

cultivated on an orbital shaker (Lab-Shaker LS-W, Adolf Kühner AG, Birsfelden,

Switzerland) with a filling volume of 10 ml, a 50 mm shaking diameter and a shaking

frequency of 300 rpm. Cultivations were carried out in unbaffled 250 ml shake flasks with

cotton plugs or in 250 ml measuring flasks of the RAMOS device (Chapter A.3.2). Orbital

shaker and RAMOS device were placed in a thermo constant room.

A.2.4 Shake flasks in continuous mode

Fig. A.1d depicts the fermentation device used for continuous shake flask cultures. The

cultivations were carried out in a COSBIOS device (Continuously Operated Shaking

Bioreactor System) introduced by Akgün et al. [2004]. An advanced version is commercially

Page 173: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Appendix General material and methods 155

available from Adolf Kühner AG (Birsfelden, Switzerland). The COSBIOS device used in

this study consists of 250 ml shake flasks with inlets in the headspace for aeration (1 vvm)

and medium supply, and an outlet on the side of the flask for exhaust gas and culture broth.

The surplus culture broth leaves the flask if the maximum liquid height of the rotating liquid

reaches the edge of the outlet. Through the centrifugal force and the aeration, the liquid

accelerates into the outlet. Different substrate feed rates were adjusted with a single

multichannel peristaltic pump (IPC-N, Ismatec, Glattbrug, Switzerland) by using tubes

(Ismaprene, Ismatec GmbH, Wertheim-Mondfeld, Germany) with internal diameters of 0.38,

0.51, 0.64, 0.76, 0.89, 1.02, 1.14, and 1.3 mm. The orbital shaker was operated with a shaking

diameter of 5 cm and a shaking frequency of 300 rpm. The filling volume was determined for

each flask after the experiment was terminated. It varied between 24-26 ml. To avoid oxygen-

limited culture conditions due to the, compared with the shake flasks in batch mode (10 ml),

higher filling volume of the COSBIOS, the lactic acid concentration was reduced to 5 g/l. To

achieve a pH of 7 during the cultivation, the pH of the feed medium was adjusted to 6.5

(Chapter 4.3.2). Eight flasks were operated simultaneously at different dilution rates, making

it possible to record a biomass over dilution rate (X-D) diagram in less than one week.

Page 174: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Appendix General material and methods 156

Fig. A.1: Schematic depiction of

the culture devices for batch and

continuous cultivation in shake

flasks and fermenters. A. stirred

tank fermenter with exhaust gas

analyzer in batch mode, B. stan-

dard shake flask with a cotton

plug, and respiration measurement

with RAMOS device, C. stirred

tank fermenter with exhaust gas

analyzer in continuous mode, D.

continuous fermentation in special

shake flasks.

A exhaust gas analyzer stirrer

aeration shaker

aeration

B

O2

oxygen sensor

& 8 flasks

D

harvest

multi-channel peristaltic pump

feed shaker

aeration

8 flasks

harvest

acid

feed

C exhaust gas analyzer

A.3 Respiration measurement

A.3.1 Respiration measurement with exhaust gas analysis in fermenters

Fig. A.2a schematically depicts the analytical procedure to determine the OTR with an

difference between the inlet gas stream (O

exhaust gas analyzer (EGA). The OTR is calculated by specifying the oxygen concentration

Magnos 106, ABB Automation, Frankfurt, Germany). The outlet gas stream is dried by a

mass flow controller (5850TR, Brooks, Hatfield, PA, USA), thus, variations of the aeration

2in) and the outlet gas stream (O2out). The OTR of

fermenter cultures was measured with a magneto-mechanical EGA (Advance Optima,

cooler and the volume flow rate (0.5 l/min) through the EGA is kept constant with a thermal

rate cannot influence the EGA. It was calibrated before each experiment with nitrogen and a

test gas (25 % O2, 5 % CO2 and 70 % N2). The OTR was recorded every two minutes.

Page 175: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Appendix General material and methods 157

t [min]

pO2

∆pO2

∆t

I III II

I rinsing phase II measuring phase

End of aeration

t [s]

DO

∆DO

∆t

O2

t [h]

O2in

O2out

Exhaust gas analyser Continous on-line

RAMOS device Intermittent on-line

Respirometry Off-line

magnetic stirrer

Clark electrode

O

sample

initaly aerated culture

l

g2

VTRV

tpOOTR

⋅⋅⋅

∆∆

=

( )out2in20m

OOVqOTR −⋅=

2OSt

DOOTR ⋅∆

=

A

B

C

O2in

O2out

fermenterexhaustgas analyzer

stirrer

pO2 valves

O2 oxygen sensor

Fig. A.2: Schematic depiction of different analytical procedures to determine the oxygen transfer rate

(OTR). Nomenclature: A. q specific aeration rate, Vm0 molar gas volume at standard conditions, O2in

oxygen concentration of the inlet gas stream, O2out oxygen concentration of the outlet gas stream; B. pO2

partial oxygen pressure in the headspace of the shake flask, ∆t time interval of the measuring phase, Vg

gas volume of the headspace of the shake flask, R gas constant, T temperature, Vl liquid filling volume;

C. DO dissolved oxygen, ∆t time interval of the measurement, SO2 oxygen solubility. Simplified equations

are depicted.

Page 176: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Appendix General material and methods 158

A.3.2 Respiration measurement with the RAMOS device in shake flasks

The OTR was measured with a RAMOS (respiration activity monitoring system) [Anderlei et

al. 2004, Anderlei and Büchs 2001, Seletzky et al. 2006c] (Hitec Zang, Herzogenrath,

Germany) utilizing an electro-chemical oxygen sensor (Fig. A.2b). The OTR is measured by

periodically repeating an automated measuring cycle. A measuring cycle is composed of a

measuring phase (flask closed air-tight, 10 min) and a rinsing phase (continuous air flow,

20 min). Thus, recreating the average oxygen supply of a standard flask with cotton plug

[Anderlei et al. 2004]. The OTR is calculated from the decrease of the partial oxygen pressure

(∆pO2) in the headspace of the shake flask (gas volume Vg). To allow the monitoring of

weakly respiring cell cultures accuracy and precision of the measurement is increased by

recalibrating the oxygen sensors before each measuring phase using the steady state gas

composition at the end of the rinsing phase [Anderlei et al. 2004, Anderlei and Büchs 2001].

The variation of the partial oxygen pressure in the headspace of the shake flask caused by the

measuring cycle is smaller than differences in partial oxygen pressure produced by the

variability of manually produced cotton plugs [Anderlei et al. 2004, Mrotzek et al. 2001].

Thus, the stresses imposed by standard flasks and the RAMOS device are comparable.

A.3.3 Respiration measurement with respirometry

Fig. A.2c schematically depicts the analytical procedure used to determine the OTR with a

respirometer. A respirometer with an electro-chemical oxygen electrode (Rank Brothers,

Cambridge, England) was used to measure the OTR. A sample taken from an arbitrary culture

vessel is aerated and afterwards the decrease of the dissolved oxygen (DO) over time (∆t) is

measured. The OTR can be calculated, using the oxygen solubility (SO2). A sample of 5 ml

culture broth was directly transferred to the measuring chamber. The aeration was done with a

silicone tube (ID. 3mm), which produced large bubbles to avoid the forming of micro

Page 177: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Appendix General material and methods 159

bubbles, which might act as an oxygen source during the measurement. DO gradients were

avoided by stirring the measuring chamber with a magnetic stirrer. After aeration, the

measuring chamber was quickly closed. The time interval between a DO of 80 % and 20 %

was recorded with a stopwatch. All manipulations were carried out in a thermo-constant room

at 30° C. The respirometer was calibrated prior to the measurement with air and nitrogen. The

solubilities of the media were calculated as described in appendix A.1.3.

A.4 Support materials for immobilization

As glass support materials non-porous glass microscope slices of soda lime glass ISO 8037/I

(Roth, Karlsruhe, Germany) (26 mm length, 26 mm width, 1 mm thickness) and porous glass

filter discs of borosilicate glass 3.3 DIN ISO 3585 (porosities 2, 3, 4, Robu Glasfiltergeräte

GmbH, Hattert, Germany) were used. Glass sheets were cut from the filter discs and then

ground to obtain a uniform size (length 31 mm, width 20 mm, thickness 2 mm).

As ceramic supports, porous reaction bonded recrystallized silicon carbide (Heimbach GmbH,

Düren, Germany), and porous cordierite (Mg2Al4Si5O11) (Rauschert, Kloster Veilsdorf,

Veilsdorf, Germany) were used. Ceramic sheets were cut from honeycomb catalyst monoliths

and then ground to obtain a uniform size (length 30 mm, width 20 mm, thickness 1.5 mm).

Additional properties of the support materials are depicted in Tab. 6.1 and Figs. 6.3, 6.4.

A.5 Analytics

A.5.1 Biomass determination

The biomass concentration of fermenter cultures was determined by centrifuging 5 ml culture

broth at 5000 rpm in dried (24 h, 105° C, cooled down in a dessicator) and weighed centrifuge

tubes. After removing the supernatant, the pellet was dried for 24 h at 105° C, then cooled

Page 178: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Appendix General material and methods 160

down in a dessicator and weighed. For biomass measurements from shake flask cultures,

1.5 ml culture broth was used. All biomass concentrations were determined at least twice.

A.5.2 HPLC chromatography

The lactic acid concentrations were determined with HPLC chromatography (Dionex with RI

and UV-VIS detectors, Dionex Sunnyvale, CA, USA). As stationary phase an organic acid

resin (CS Chromatography, Langerwehe, Germany) and as eluent 1 mM H2SO4 with a flow

rate of 0.6 ml/min was used. The column temperature was maintained at 30° C. All substrate

concentrations were determined at least twice.

A.6 Modeling tools

All simulations were carried out with the software package gPROMS (general PROcess

Modelling System), (PSE, London, Great Britain).

Page 179: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Bibliography 161

Bibliography

Adler I, Diekmann J, Hartke W, Hecht V, Rohn F, Schügerl K (1980) Bubble coalescence behaviour in

biological media. Eur. J. Appl. Microbiol. Biotechnol. 10: 171-186

Akgün A, Maier B, Preis D, Roth B, Klingelhöfer R, Büchs J (2004) A novel parallel shaken bioreactor system

for continuous operation. Biotechnol. Prog. 20: 1718-1724

Akita K (1981) Diffusivities of gases in aqueous solutions. Ind. Eng. Chem. Fundam. 20: 89-94

Al-Masry WA (2004) Influence of gas separator and scale-up on the hydrodynamics of external loop circulating

bubble columns. Chem. Res. Des. 82: 381-389

Alves SS, Maia CI, Vasconcelos JMT (2004) Gas-liquid mass transfer coefficient in stirred tanks interpreted

through contamination kinetics. Chem. Eng. Proc. 43: 823-830

Alves SS, Orvalho SP, Vasconcelos JMT (2005) Effect of bubble contamination on rise velocity and mass

transfer. Chem. Eng. Sci. 60: 1-9

Amanullah A, Blair R, Nienow AW, Thomas CR (1999) Effects of agitation intensity on mycelial morphology

and protein production in chemostat cultures of recombinant Aspergillus oryzae. Biotechnol. Bioeng.

62: 434-446

An YH, Friedman RJ (1998) Concise review of mechanisms of bacterial adhesion to biomaterial surfaces. J.

Biomed. M. Res. 43: 338–348

Anderlei T, Büchs J (2001) Device for sterile online measurement of the oxygen transfer rate in shaking flasks.

Biochem. Eng. J. 7: 157-162

Anderlei T, Zang W, Papaspyrou M, Büchs J (2004) Online respiration activity measurement (OTR, CTR, RQ)

in shake flasks. Biochem. Eng. J. 17: 187-194

Atkinson B, Davies IJ (1972) The completely mixed microbial film fermenter. Trans. Instn. Chem. Engrs.

50: 208-216

Bader FG (1978) Analysis of double-substrate limited growth. Biotechnol. Bioeng. 20: 183-202

Barreiro C, González-Lavado E, Brand S, Tauch A, Martín J-F (2005) Heat shock proteome analysis of wild-

type Corynebacterium glutamicum ATCC 13032 and a spontaneous mutant lacking groEL1, a dispensable

chaperone. J. Bacteriol. 187: 884-889

Barreiro C, González-Lavado E, Pátek M, Martín J-F (2004) Transcriptional analysis of the groES-groEL1,

groEL2, and dnaK genes in Corynebacterium glutamicum: characterization of heat shock-induced promoters. J.

Bacteriol. 186: 4813-4817

Page 180: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Bibliography 162

Bendinger B, Rijnaarts HHM, Altendorf K, Zehnder AJB (1993) Physiochemial cell surface and adhesive

properties of coryneform bacteria related to the presence and chain length of mycolic acids. Appl. Environ.

Microbiol. 59: 3973-3977

Benedek A, Heideger WJ (1971) Effect of additives on mass transfer in turbine aeration. Biotechnol. Bioeng.

13: 663-684

Berber R, Pertev C, Türker M (1998) Optimization of feeding profile for baker's yeast production by dynamic

programming. Biopro. Eng. 20: 263-269

Bishop PL, Zhang TC, Fu Y-C (1995) Effects of biofilm structure, microbial distributions and mass transport on

biodegradation processes. Wat. Sci. Tech. 31: 143-152

Blenke H (1979) Loop reactors. Adv. Biochem. Eng. Biotechnol. 13: 121-214

Bodmer T, Miltner E, Bermudez LE (2000) Mycobacterium avium resists exposure to the acidic conditions of

the stomach. FEMS Microbiol. L. 182: 45-49

Boessmann M, Staudt C, Neu TR, Horn H, Hempel DC (2003) Investigation and modeling of growth, structure

and oxygen penetration in particle supported biofilms. Chem. Eng. Technol. 26: 219-222

Bonomo L, Pastorelli G, Quinto E (2001) Simplified and Monod kinetics in one-dimensional biofilm reactor

modeling: a comparison. Wat. Sci. Tech. 43: 295-302

Boyles DT (1978) Specific growth rate measurement in an oxygen electrode chamber. Biotechnol. Bioeng.

20: 1101-1104

Breidt jr F, Hayes JS, McFeeters RF (2004) Independent effects of acetic acid and pH on survival of Escherichia

coli in simulated acidified pickle products. J. Food Protection 67: 12-18

Bröer S, Krämer R (1991) Lysine excretion by Corynebacterium glutamicum 1. Identification of a specific

secretion carrier system. Eur. J. Biochem. 202: 131-135

Büchs J (2001) Introduction to advantages and problems of shaken cultures. Biochem. Eng. J. 7: 91-98

Büchs J, Maier U, Lotter S, Peter CP (2006) Calculating liquid distribution in shake flasks on rotary shakers at

water like viscosities. Biochem. Eng. J. accepted

Büchs J, Maier U, Milbradt C, Zoels B (2000) Power consumption in shaking flasks on rotary shaking machines:

Part I: power consumption measurement in unbaffled flasks at low liquid viscosity. Biotechnol. Bioeng.

68: 589-593.

Büchs J, Maier U, Milbradt C, Zoels B (2000) Power consumption in shaking flasks on rotary shaking machines:

Part II: nondimensional description of specific power consumption and flow regimes in unbaffled flasks at

elevated liquid viscosity. Biotechnol. Bioeng. 68: 594-601

Büchs J, Mozes N, Wandrey C, Rouxhet PG (1988) Cell adsorption control by culture conditions. Influence of

Page 181: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Bibliography 163

phosphate on surface properties, flocculation and adsorption behaviour of Corynebacterium glutamicum. Appl.

Microbiol. Biotechnol. 29: 119-128

Burkovski A (2003) I do it my way: regulation of ammonium uptake and ammonium assimilation in

Corynebacterium glutamicum. Arch. Microbiol. 179: 83-88

Chamsartra S, Hewitt CJ, Nienow AW (2005) The impact of fluid mechanical stress on Corynebacterium

glutamicum during continuous cultivation in an agitated bioreactor. Biotechnol. L. 27: 693-700

Chang I, Gilbert ES, Eliashberg N, Keasling JD (2003) A three-dimensional, stochastic simulation of biofilm

growth and transport-related factors that affect structure. Microbiol. 149: 2859-2871

Chavant P, Martinie B, Meylheuc T, Bellon-Fontaine M, Hebraud M (2002) Listeria monocytogenes LO28:

surface physicochemical properties and ability to form biofilms at different temperatures and growth phases.

Appl. Environ. Microbiol. 68: 728-737

Choi KH (2001) Hydrodynamic and mass transfer characteristics of external-loop airlift reactors without

extension tube above the downcomer. Korean J. Chem. Eng. 18: 240-246

Christi MY (1989) Airlift biorectors. Elsevier, Barking New York

Clark GJ, Langley D, Bushell ME (1995) Oxygen limitation can induce microbial secondary metabolite

formation: investigations with miniature electrodes in shaker and bioreactor culture. Microbiol. 141: 663-669

Cocaign M, Monnet C, Lindley ND (1993) Batch kinetics of Corynebacterium glutamicum during growth on

various carbon substrates: use of mixtures to localize metabolic bottlenecks. Appl. Microbiol. Biotechnol.

40: 526-530

Cocaign-Bousquet M, Guyonvarch A, Lindley ND (1996) Growth rate-dependent modulation of carbon flux

through central metabolism and the kinetic consequences for glucose-limited chemostat cultures of

Corynebacterium glutamicum. Appl. Environ. Microbiol. 62: 429-436

Cocaign-Bousquet M, Lindley ND (1995) Pyruvate overflow and carbon flux within the central metabolic

pathways of Corynebacterium glutamicum during growth on lactate. Enzyme Microb. Technol. 17: 260-267

Costerton JW, Lewandowski Z, Caldwell DE, Korber DR (1995) Microbial biofilms. Annu. Rev. Microbiol.

49: 711-745

Cotter PD, Hill C (2003) Surviving the acid test: responses of gram-positive bacteria to low pH. Microbiol. Mol.

Bio. Rev. 67: 429-453

Cui YQ, van der Lans RGJM, Luyben KChAM (1998) Effects of dissolved oxygen tension and mechanical

forces on fungal morphology in submerged fermentation. Biotechnol. Bioeng. 57: 409-419

Danckwerts PV (1957) Measurement of molecular homogeneity in a mixture. Chem. Eng. Sci. 7: 116-117

Dawson RMC (1986) Data for Biochemical Research. Clarendon Press, Oxford

Page 182: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Bibliography 164

de Hollander JA (1993) Kinetics of microbial product formation and its consequences for the optimization of

fermentation processes. A. v. Leeuwenhoek 63: 375-381

Deckwer W-D (1992) Bubble column reactors. John Wiley & Sons, Chichester

Enfors SO, Mattiasson B (1983) Oxygenation of processes involving immobilized cells. In: Mattiasson B (ed)

Immobilized cells and organelles. CRC Press, Boca Raton, Fl, USA, pp 41-60

Ferreira A, O’Byrne CP, Boor KJ (2001) Role of σB in heat, ethanol, acid, and oxidative stress resistance and

during carbon starvation in Listeria monocytogenes. Appl. Environ. Microbiol. 67: 4454–4457

Fleischer C (2000) Detaillierte Modellierung von Gas-Flüssigkeits-Reaktoren. PhD thesis university Stuttgart.

VDI Verlag, Düsseldorf

Frost MT, Wang Q, Moncada S, Singer M (2005) Hypoxia accelerates nitric oxide-dependent inhibition of

mitochondrial complex I. in activated macrophages. Am. J. Physiol. Regul. Integr. Comp. Physiol. 288: 394-400

Garboczi EJ, Bentz DP (1991) Digitized simulation of mercury intrusion porosimetry. Advances cementitious

materials, Ceramic Trans. 16: 365-380

García Calvo E (1989) A fluid dynamic model for airlift reactors. Chem. Eng. Sci. 44: 321-323

Gjaltema A, van der Marel N, van Loosdrecht MCM, Heijnen JJ (1997) Adhesion and biofilm development on

suspended carriers in airlift reactors: hydrodynamic conditions versus surface characteristics. Biotechnol.

Bioeng. 55: 880-889

Gourdon P, Baucher M-F, Lindley ND, Guyonvarch A (2000) Cloning the malic enzyme gene from

Corynebacterium glutamicum and role of the enzyme in lactate metabolism. Appl. Environ. Microbiol.

66: 2981-2987

Gruber TM, Bryant DA (1997) Molecular systematic studies of eubacteria, using σ70-type sigma factors of group

1 and group 2. J. Bacteriol. 179: 1734–1747

Guillouet S, Engasser JM (1995) Growth of Corynebacterium glutamicum in glucose-limited continuous cultures

under high osmotic pressure. Influence of growth rate on the intracellular accumulation of proline, glutamate,

and trehalose. Appl. Microbiol. Biotechnol. 44: 496-500

Guillouet S, Engasser JM (1996) Growth of Corynebacterium glutamicum in ammonium- and potassium-limited

continuous cultures under high osmotic pressure. Appl. Microbiol. Biotechnol. 46: 291-296

Gupta A, Rao G (2003) A study of oxygen transfer in shake flasks using a non-invasive oxygen sensor.

Biotechnol. Bioeng. 84: 351-358

Gustafsson TK, Skrifvars BO, Sandström KV, Waller KV (1995) Modelling pH for Control. Ind. Eng. Chem.

Res. 34: 820-827

Page 183: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Bibliography 165

Gustafsson TK, Waller KV (1992) Nonlinear and adaptive control of pH. Ind. Eng. Chem. Res. 31: 2681-2698

Gutmann M, Hoischen C, Krämer R (1992) Carrier-mediated glutamate secretion by Corynebacterium

glutamicum under biotin limitation. Biochim. Biophys. Acta 1112: 115-123

Heijen JJ, van't Riet (1984) Mass transfer, mixing and heat transfer phenomena in low viscosity bubble column

reactors. Chem. Eng. J. 28: 21-42

Henkel HJ, Johl HJ, Trösch W, Chmiel H (1990) Continuous production of glutamic acid in a three phase

fluidised bed with immobilized Corynebacterium glutamicum. Food Biotechnol. 4: 149-154

Hermann R, Walther N, Maier U, Büchs J (2001) Optical method for the determination of oxygen-transfer

capacity of small bioreactors based on sulfite oxidation. Biotechnol. Bioeng. 74: 355-363

Horn H, Hempel DC (1998) Modeling mass transfer and substrate utilization in the boundary layer of biofilm

systems. Wat. Sci. Tech. 37: 139-147

Horn H, Wäsche S, Hempel DC (2002) Simulation of biofilm growth, substrate conversion and mass transfer

under different hydrodynamic conditions. Wat. Sci. Tech. 46: 249-252

Houtsma PC, Kant-Muermans ML, Rombouts FM, Zwietering M (1996) Model for the combined effects of

temperature, pH, and sodium lactate on growth rates of Listeria innocua in broth and bologna-type sausages.

Appl. Environ. Microbiol. 62: 1616–1622

Howell JA, Atkinson B (1976) Influence of oxygen and substrate concentrations on the ideal film thickness and

the maximum overall substrate uptake rate in microbial film fermenters. Biotechnol. Bioeng. 18: 15-35

Howell JA, Atkinson B (1976) Influence of oxygen and substrate concentrations on the ideal film thickness and

the maximum overall substrate uptake rate in microbial film fermenters. Biotechnol. Bioeng. 18: 15-35

Hu Y, Coates ARM (1999) Transcription of two sigma 70 homologue genes, sigA and sigB, in stationary-phase

Mycobacterium tuberculosis. J. Bacteriol. 81: 469–476

ICH Int. Conf. Harmonization tech. Requirements reg. Pharmaceuticals Human use. (1996) Guidance for

industry Q2B Validation of analytical procedures.

Idelchik IE (1994) Handbook of hydraulic resistance. CRC Press, Boca Raton Ann Arbor London Tokyo

Ihssen J, Egli T (2004) Specific growth rate and not cell density controls the general stress response in

Escherichia coli. Microbiol. 150: 1637–1648

Ilavsky J, Berndt CC, Karthikeyan J (1997) Mercury intrusion porosimetry of plasma-spayed ceramic. J. Mat.

Sci. 32: 3925-3932

Page 184: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Bibliography 166

Inui M, Muratami S, Okino S, Kawaguchi H, Vertès AA, Yukawa H (2004) Metabolic analysis of

Corynebacterium glutamicum during lactate and succinate productions under oxygen deprivation conditions. J.

Mol. Microbiol. Biotechnol. 7: 182-196

Ishige T, Krause M, Bott M, Wendisch VF, Sahm H (2003) The phosphate starvation stimulon of

Corynebacterium glutamicum determined by DNA microarray analyses. J. Bacteriol. 185: 4519-4529

Katzer W, Blackburn M, Charman K, Martin S, Penn J, Wrigley S. (2001) Scale-up of filamentous organisms

from tubes and shake-flasks into stirred vessels. Biochem. Eng. J. 7: 127-134

Kawakami K, Kawasaki K, Shiraishi F, Kusunoki K (1989) Performance of a honeycomb monolith bioreactor in

a gas-liquid-solid three-phase system. Ind. Eng. Chem. Res. 28: 394-400

Kawase Y, Harlard B, Moo-Young M (1992) Liquid-phase mass transfer coefficients in bioreactors. Biotechnol.

Bioeng. 39: 1133-1140

Kawase Y, Moo-Young M (1987) Influence of antifoam agents on gas hold-up and mass transfer in bubble

columns with non-Newtonian fluids. Appl. Microbiol. Biotechnol. 27: 159-167

Kawase Y, Moo-Young M (1990) The effect of antifoam agents on mass transfer in bioreactors. Bioprocess Eng.

5: 169-173

Kennedy MJ, Reader SL, Davies J, Rhoades DA, Silby HW (1994) The scale up of myceleal shake flask

fermentations: a case study of gamma linolenic acid production by Mucor hiemalis IRL 51. J. Ind. Microbiol.

13: 212-216

Kida K, Morimura S, Sonoda Y, Obe M, Kondo T (1990) Support media for microbial adhesion in an anaerobic

fluidized-bed reactor. J. Ferm. Bioeng. 69: 354-359

Kim CH, Rao KJ, Youn DJ, Rhee SK (2003) Scale-up of recombinant hirudin production from Saccharomyces

cerevisiae. Biotechnol. Biopro. Eng. 8: 303-305

Kim HS, Ryu DDY (1982) Continuous glutamate production using an immobilized whole-cell system.

Biotechnol. Bioeng. 24: 2167-2174

Kim T-H, Park J-S, Kim H-J, Kim Y, Kim P, Lee H-S (2005) The whcE gene of Corynebacterium glutamicum is

important for survival following heat and oxidative stress. Biochem. Biophys. Res. Commun. 337: 757-764

Kinoshita S, Udaka S, Shimono M (1957) Amino acid fermentation. I. Production of L-glutamic acid by various

microorganisms. J. Gen. Appl. Microbiol. 3: 193-205

Kircher M, Pfefferle W (2001) The fermentative production of L-lysine as an animal feed additive. Chemosphere

43: 27-31

Page 185: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Bibliography 167

Knoll A, Maier B, Tscherrig H, Büchs J (2005) The oxygen mass transfer, carbon dioxide inhibition, heat

removal, and the energy and cost efficiencies of high pressure fermentation. Adv. Biochem. Eng./Biotechnol.

92: 77-99

Kočan M, Schaffer S, Ishige T, Sorger-Hermann U, Wendisch VF, Bott M (2006) Two-component systems of

Corynebacterium glutamicum: deletion analysis and involvement of the PhoS-PhoR system in the phosphate

starvation response. J. Bacteriol. 188: 724-732

Koch V, Rüffer H-M, Schügerl K, Innertsberger E, Menzel H, Weis J (1995) Effect of antifoam agent on the

medium and microbial cell properties and process performance in small and large reactors. Process Biochem.

30: 435-446

Koffas MAG, Jung GY, Stephanopoulos G (2003) Engineering metabolism and product formation in

Corynebacterium glutamicum by coordinated gene overexpression. Metab. Eng. 5: 32-41

Kohls O, Scheper TH (2000) Setup of a fiber optical oxygen multisensor-system and its application in

biotechnology. Sens. Actuators 70: 121-130

Koide K, Yamazoe S, Harada S (1985) Effects of surface-active substances on gas holdup and gas-liquid mass

transfer in bubble column. J. Chem. Eng. Jap. 18: 287-292

Kovárová-Kovar K, Egli T (1998) Growth kinetics of suspended microbial cells: from single-substrate-

controlled growth to mixed-substrate kinetics. Microbiol. Mol. Biol. Rev. 62: 646-666

Lee Y, Tsao GT (1979) Dissolved oxygen electrodes, Adv. Biochem. Eng. Biotechnol. 13: 35-86

Léon Ohl A, Horn H, Hempel DC (2004) Behaviour of biofilm systems under varying hydrodynamic conditions.

Wat. Sci. Technol. 49: 345-351

Lessard RR, Zieminski SA (1971) Bubble coalescence and gas transfer in aqueous electrolytic solutions. Ind.

Eng. Chem. Fundam. 10: 260-269

Li B, Logan BE (2004) Bacterial adhesion to glass and metal-oxide surfaces. Colloids Surfaces B: Biointerfaces

36: 81–90

Lide (1999) Handbook of Chemistry and Physics. CRC Press, Boca Raton

Liepe F, Meusel W, Möckel HO, Platzer B, Weißgräber H (1988) Verfahrenstechnische Berechnungsmethoden

Teil 4, Stoffvereinigung in fluiden Phasen. VCH Publishers, Weinheim

Linek V, Vacek V (1981) Chemical engineering use of catalyzed sulfite oxidation kinetics for determination of

mass-transfer characteristics of gas-liquid contactors. Chem. Eng. Sci. 36: 1147-1768

Losen M, Lingen B, Pohl M, Büchs J (2004) Effect of oxygen-limitation and medium composition on

Escherichia coli in small-scale cultures. Biotechnol. Prog. 20: 1062-1068

Page 186: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Bibliography 168

Lotter S (2003) Leistungseintrag und hydrodynamische Phänomene in geschüttelten Bioreaktoren. PhD thesis

RWTH Aachen. Shaker, Aachen

Maier B, Dietrich C, Büchs J (2001) Correct application of the sulphite oxidation methodology of measuring the

volumetric mass transfer coefficient kLa under non-pressurized and pressurized conditions. Trans. IChemE.

79: Part C

Maier U (2002) Gas-Flüssigkeits-Stofftransfer im Schüttelkolben. PhD thesis RWTH Aachen University,

Shaker, Aachen

Maier U, Büchs J (2001) Characterization of the gas-liquid mass transfer in shaking bioreactors. Biochem. Eng.

J. 7: 99-106

Maier U, Losen M, Büchs J (2004) Advances in understanding and modeling the gas-liquid mass transfer in

shake flasks. Biochem. Eng. J. 17: 155-167

Mandel MJ, Silhavy TJ (2005) Starvation for different nutrients in Escherichia coli results in differential

modulation of RpoS levels and stability. J. Bacteriol. 187: 434–442

Maynes D, Webb AR (2002) Velocity profile characterization in sub-millimeter diameter tubes using molecular

tagging velocimetry. Exp. Fluids 32: 3-15

Melo JS, D'Souza SF (1999) Simultaneous filtration and immobilization of cells from a flowing suspension

using a bioreactor containing polyethylenimin coated cotton threads: Application in the continuous inversion of

concentrated sucrose syrups. W. J. Microbiol. Biotechnol. 15: 17-21

Messing RA, Oppermann RA, Kolot FB (1979) Pore dimensions for accumulating biomass. I. Microbes that

reproduce by fission or by budding. Biotechnol. Bioeng. 21: 49-58

Montagnolli W, Zamboni A, Luvizotto-Santos R, Yunes JS (2004) Acute effects of Microcystis aeruginosa from

patos lagoon estuary, southern Brazil, on the microcrustacean Kalliapseudes schubartii (Crustacea: Tanaidacea).

Arch. Environ. Contam. Toxicol. 46: 463-469

Morgenroth E, Eberl HJ, van Loosdrecht MCM, Noguera DR, Pizarro GE, Picioreanu C, Rittmann BE, Schwarz

AO, Wanner O (2004) Comparing biofilm models for a single species biofilm system. Wat. Sci. Technol.

49: 145-154

Mori A (1985) Production of vinegar by immobilized cells. Process Biochem. 20: 67-74

Mrotzek C, Anderlei T, Henzler H-J, Büchs J (2001) Mass transfer resistance of sterile plugs in shaking

bioreactors. Biochem. Eng. J. 7: 107-112

Nicolas-Simonnot MO, Sardin M, Grévillot G (1995) Consequences of ion-exchange and sorption effects on pH

waves propagating in a strong-base anion-exchange column. J. Liquid Chromatography 18: 463-487

Page 187: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Bibliography 169

Noguera DR, Okabe S, Picioreanu C (1999) Biofilm modeling: present status and future directions. Wat. Sci.

Technol. 39: 273-278

Noguera DR, Pizarro G, Stahl DA, Rittmann BE (1999) Simulation of multispecies biofilm development in three

dimensions. Wat. Sci. Tech. 39: 123-130

Nolden L, Ngouto-Nkili C-E, Bendt AK, Krämer R, Burkovski A (2001) Sensing nitrogen limitation in

Corynebacterium glutamicum: the role of glnK and glnD. Mol. Microbiol. 42: 1281-1295

O’Driscoll B, Grahan CG, Hill C (1996) Adaptive acid tolerance response in Listeria monocytogenes: isolation

of an acid-tolerant mutant which demonstrates increased virulence. Appl. Environ. Microbiol. 62: 1693-1698

Owens JD, Legan JD (1987) Determination of the Monod substrate saturation constant for microbial growth.

FEMS Microbiol. Rev. 46: 419-432

Park JK, Jeong GS (2001) Effect of capsule circulation velocity on production of L-Lysine by encapsulated

Corynebacterium glutamicum in an airlift bioreactor. J. Biosci. Bioeng. 91: 81-84

Perry RH, Green DW, Maloney JO (1997) Perry's Chemical Engineers' Handbook. McGraw-Hill Professional

Publishing, New York

Peter CP, Lotter S, Maier U, Büchs J (2004) Impact of out-of-phase conditions on screening results in shaking

flasks experiments. Biochem. Eng. J. 17: 205-215

Peter CP, Suzuki Y, Büchs J (2006) Hydromechanical stress in shake flasks: correlation for the maximum local

energy dissipation rate. Biotechnol. Bioeng. 93: 1164-1176

Peter H, Weil B, Burovski A, Krämer R, Morbach S (1998) Corynebacterium glutamicum is equipped with four

secondary carriers for compatible solutes: identification, sequencing, and characterization of the

proline/ectoine/glycine betaine carrier, EctP. J. Bacteriol. 180: 6005-6012

Picioreanu C, van Loosdrecht MCM, Heijnen JJ (2000) Modelling and predicting biofilm structure. In: Allison

DG, Gilpert P, Lappin-Scott HM, Wilson M (eds). Community structure and co-operation in biofilms.

Cambridge University Press, Cambridge, pp 129-166

Presser KA, Ratkowsky DA, Ross T (1997) Modeling the growth rate of Escherichia coli as a function of pH

and lactic acid concentration. Appl. Environ. Microbiol. 63: 2355–2360

Presser KA, Ross T, Ratkowsky DA (1998) Modeling the growth limits (growth/no growth interface) of

Escherichia coli as a function of temperature, pH, lactic acid concentration, and water activity. Appl. Environ.

Microbiol. 64: 1773–1779

Primm TP, Andersen SJ, Mizrahi V, Avarbock DA, Rubin H, Barry CE (2000) The stringent response of

Mycobacterium tuberculosis is required for long-term survival. J. Bacteriol. 182: 4889–4898

Page 188: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Bibliography 170

Raval KN, Hellwig S, Prakash G, Ramos-Plasencia A, Srivastava A, Büchs J (2003) Necessity of a two-stage

process for the production of azadirachtin-related limonoids in suspension cultures of Azadirachta indica. J.

Biosci. Bioeng. 96: 16-22

Ridgway HF (1977) Source of energy for gliding motility in Flexibacter polymorphus: effects of metabolic and

respiratory inhibitors on gliding movement. J. Bacteriol. 131: 544-556

Rijnaarts HHB, Norde W, Bouwer EJ, Lyklema J, Zehnder AJB (1993) Bacterial adhesion under static and

dynamic conditions. Appl. Environ. Microbiol. 59: 3255-3265

Rischbieter E, Schumpe A (1996) Gas solubilities in aqueous solutions of organic substances. J. Chem. Eng.

Data 41: 809-812

Rittmann BE, McCarty PL (1980) Model of steady-state-biofilm kinetics. Biotechnol. Bioeng. 22: 2343-2357

Roe AJ, O’Byrne C, McLaggan D, Booth IR (2002) Inhibition of Escherichia coli growth by acetic acid: a

problem with methionine biosynthesis and homocysteine toxicity. Microbiol. 148: 2215–2222

Rönsch H (2000) Untersuchung zum Einfluss der Osmoregulation auf die Aminosäureproduktion mit

Corynebacterium glutamicum. PhD thesis university Cologne Germany. Published online university library Köln

Ross T, Ratkowsky DA, Mellefont LA, McMeekin TA (2003) Modelling the effects of temperature, water

activity, pH and lactic acid concentration on the growth rate of Escherichia coli. Int. J. Food Microbiol.

82: 33-43

Roy S, Bauer T, Al-Dahhan M, Lehner P, Turek T (2004) Monoliths as multiphase reactors: a review. AIChE J.

50: 2918-2938

Rutter PR (1980) The physical chemistry of the adhesion of bacteria and other cells. In: Curtis ASG, Pitts JD

(eds) Cell Adhesion and Motility. Cambridge University Press, Cambridge, pp 103–135

Sahoo S, Verma RK, Suresh AK, Rao KK, Bellare J, Suraishkumar GK (2003) Macro-level and genetic-level

responses of Bacillus subtilis to shear stress. Biotechnol. Prog. 19: 1689-1696

Saklani-Jusforgues H, Fontan E, Goossens PL (2000) Effect of acid-adaptation on Listeria monocytogenes

survival and translocation in a murine intragastric infection model. FEMS Microbiol. L. 193: 155–159

Salmond CV, Kroll RG, Booth IR (1984) The effect of food preservatives on pH homeostasis in Escherichia

coli. J. Gen. Microbiol. 130: 2845-2850

Samonin VV, Elikova EE (2004) A study of the adsorption of bacterial cells on porous materials. Microbiol.

73: 696–701. Translated from Mikrobiologiya (2004) 73: 810–816

Satterfield CN, Özel F (1977) Some characteristics of two-phase flow in monolithic catalyst structures. Ind. Eng.

Chem. Fundam. 16: 61-67

Page 189: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Bibliography 171

Schumpe A, Deckwer WD (1979) Estimation of O2 and CO2 solubilities in fermentation media. Biotechnol.

Bioeng. 21: 1075-1078

Schumpe A, Quicker G, Deckwer WD (1982) Gas solubilities in microbial culture media. Adv. Biochem. Eng.

48: 1-38

Seletzky JM, Noack U, Fricke J, Hahn S, Büchs J (2006a) Metabolic activity of Corynebacterium glutamicum

grown on L-lactic acid under stress. Appl. Biotechnol. Bioeng. 72: 1297-1307

Seletzky JM, Otten K, Lotter S, Fricke J, Peter CP, Maier HR, Büchs J (2006b) A simple and inexpensive

method to investigate heterogeneous catalysis (microbiological, enzymic, or inorganic) using standard histology

and microbiology laboratory equipment: assembly, mass transfer properties, hydrodynamic conditions, and

evaluation. Biotec. Histochem. 81: 133-138

Seletzky JM, Noack U, Hahn S, Knoll A, Amoabediny G, Büchs J (2006c) An experimental comparison of

respiration measuring techniques in fermenters and shake flasks: exhaust gas analyzer vs. RAMOS device vs.

respirometer. J. Ind. Microbiol. Biotechnol. 34: 23-130

Seletzky JM, Noak U, Fricke J, Welk E, Eberhard W, Knocke C, Büchs J (2007) Scale-up from shake flasks to

fermenters in batch and continuous mode with Corynebacterium glutamicum on lactic acid based on oxygen

transfer and pH. Biotechnol. Bioeng. 98: 800-811

Shah AH, Hameed A, Ahmad S, Khan GM (2002) Optimization of culture conditions for L-lysine fermentation

by Corynebacterium glutamicum. OnLine J. Bio. Sci. 2: 151-156

Shelef LA (1994) Antimicrobial effects of lactate: a review. J. Food Prot. 57: 445-450

Silberbach M, Maier B, Zimmermann M, Büchs J (2003) Glucose oxidation by Gluconobacter oxydans:

characterization in shaking-flasks, scale-up and optimization of the pH profile. Appl. Microbiol. Biotechnol.

62: 92-98

Silberbach M, Schäfer M, Hüser AT, Kalinowski J, Pühler A, Krämer R, Burkovski A (2005) Adaptation of

Corynebacterium glutamicum to ammonium limitation: a global analysis using transcriptome and proteome

techniques. Appl. Environ. Microbiol. 71: 2391-2402

Sipkema EM, De Koning W, Ganzeveld KJ, Janssen DB, Beenackers A (1998) Experimental pulse technique for

microbial kinetics in continuous culture. J. Biotechnol. 64: 159-176

Skjerdal OT, Sletta H, Flenstad SG, Josefsen KD, Levine DW, Ellingsen TE (1996) Changes in intracellular

composition in response to hyperosmotic stress of NaCl, sucrose, or glutamic acid in Brevibacterium

lactofermentum and Corynebacterium glutamicum. Appl. Microbiol. Biotechnol. 44: 635-642

Sobotka M, Prokop A, Dunn IJ, Einsele E (1982) Review of methods for the measurement of oxygen transfer in

microbial systems. Ann. Rep. Ferm. Proc. 5: 127-210

Page 190: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Bibliography 172

Stansen C, Uy D, Delaunay S, Eggeling L, Goergen J-L, Wendisch VF (2005) Characterization of a

Corynebacterium glutamicum lactate utilization operon induced during temperature-triggered glutamate

production. Appl. Environ. Microbiol. 71: 5920-5928

Stein A (1987) Mischzeiten beim rühren von begasten viskosen Flüssigkeiten. Chem. Ing. Tech. 9: 750-751

Sternström TA (1989) Bacterial hydrophobicity, an overall parameter for the measurement of adhesion potential

to soil particles. Appl. Environ. Microbiol. 55: 142-147

Stewart PS (1998) A review of experimental measurements of effective diffusive permeabilities and effective

diffusion coefficients in biofilms. Biotechnol. Bioeng. 59: 261-272

Stewart PS (2003) Diffusion in biofilms. J. Bacteriol. 185: 1485-1491

Stöckmann C, Losen M, Dahlems U, Knocke C, Gellissen G, Büchs J (2003a) Effect of oxygen supply on the

passaging, stabilizing and screening of recombinant Hansenula polymorpha production strains in test tube

cultures. FEMS Yeast Res. 4: 195-205

Stöckmann C, Maier U, Anderlei T, Knocke C, Gellissen G, Büchs J (2003b) The oxygen transfer rate as key

parameter for the characterization of Hansenula polymorpha screening cultures. J. Ind. Microbiol. Biotechnol.

30: 613-622

Støle-Hansen K, Foss BA (1997) Controlling pH in a precipitation process. Modelling Identification Control

18: 261-272

Sumino Y, Sonoi K, Doi M (1993) Scale-up of purine nucleoside fermentation from shaking flask to a stirred-

tank fermentor. Appl. Microbiol. Biotechnol. 38: 581-585

Takashi H, Yoshida F (1979) Oxygen transfer in a bubble column for fermentation with Aspergillus niger.

Ferment. Technol. 57: 349-356

Tata M, Bower P, Bromberg S, Duncombe D, Fehring J, Lau V, Ryder D, Stassi P (1999) Immobilized yeast

bioreactor systems for continuous beer fermentation. Biotechnol. Prog. 15: 105-113

van Loosdrecht MCM, Heijnen JJ, Eberl H, Kreft J, Picioreanu C (2002) Mathematical modeling of biofilm

structures. A. v. Leeuwenhoek, 81: 245-256

van Loosdrecht MCM, Lyklema J, Norde W, Schraa G, Zehnder AJB (1987) The role of bacterial cell wall

hydrophobicity in adhesion. Appl. Environ. Microbiol. 53: 1893-1897

van Loosdrecht MCM, Norde W, Lyklema J, Zehnder AJB (1990) Hydrophobic and electrostatic parameters in

bacterial adhesion. Aquat. Sci. 52: 103-114

Van't Riet K (1979) Review of Measuring methods and results in nonviscous gas-liquid mass transfer in stirred

vessels. Ind. Eng. Chem. Process Des. Dev. 18: 357-364

Page 191: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Bibliography 173

Varela C, Agosin E, Baez M, Klapa M, Stephanopoulos G (2003) Metabolic flux redistribution in

Corynebaterium glutamicum in response to osmotic stress. Appl. Microbiol. Biotechnol. 60: 547-555

Vasconcelos JM, Orvalho SP, Alves SS (2002) Gas-liquid mass transfer to single bubbles: Effect of surface

contamination. AIChE J. 48: 1145-1154

Voit H, Zeppenfeld R, Mersmann A (1987) Calculation of primary bubble volume in gravitational and

centrifugal fields. Chem. Eng. Technol. 10: 99-103

Walker SL, Hill JE, Redman JA, Elimelech M (2005) Influence of growth phase on adhesion kinetics of

Escherichia coli D21g. Appl. Environ. Microbiol. 71: 3093-3099

Wang S, Wang DIC (1989) Pore dimension effects in the cell loading of a porous carrier. Biotechnol. Bioeng.

33: 915-917

Watson SP, Clements MO, Foster SJ (1998) Characterization of the starvation-survival response of

Staphylococcus aureus. J. Bacteriol. 180: 1750–1758

Wehmeier L, Brockmann-Gretza O, Pisabarro A, Tauch A, Pühler A, Martin JF, Kalinowski J (2001) A

Corynebacterium glutamicum mutant with a defined deletion within the rplK gene is impaired in (p)ppGpp

accumulation upon amino acid starvation. Microbiol. 147: 691–700

Wehmeier L, Schäfer A, Burkovski A, Krämer R, Mechold U, Malke H, Pühler A, Kalinowski J (1998) The role

of the Corynebacterium glutamicum rel gene in (p)ppGpp metabolism. Microbiol. 144: 1853–1862

Westerlund T, Skrifavars B, Karilla S (1995) On the uniqueness in pH calculations. Chem. Eng. Sci. 40: 973-976

Weuster-Botz D, Altenbach-Rehm J, Hawrylenko A (2001) Process-engineering characterization of small-scale

bubble columns for microbial process development. Biop. Biosyst. Eng. 24: 3-11

Wittmann C, Kim HM, John G, Heinzle E (2003) Characterization and application of an optical sensor for

quantification of dissolved O2 in shake-flasks. Biotechnol. L. 25: 377-380

Wittmann C, Yang T, Kochems I, Heinzle E (2001) Dynamic respiratory measurements of Corynebacterium

glutamicum using membrane mass spectrometry. J. Microbiol. Biotechnol. 11: 40-49

Wojnowska-Baryla I, Stolarczyk E, Babuchowski A (1999) The efficiency of the COD and nitrogen removal

from municipal waste by biomass immobilized on the ceramic carrier. Med. Fac. Landbouww. Univ. Gent

64: 233-236

Wolf A, Krämer R, Morbach S (2003) Three pathways for trehalose metabolism in Corynebacterium

glutamicum ATCC 13032 and their significance in response to osmotic stress. Mol. Microbiol. 49: 1119-1134

Wolf A, Krämer R, Morbach S (2003) Three pathways for trehalose metabolism in Corynebacterium

glutamicum ATCC 13032 and their significance in response to osmotic stress. Molecular Microbiol.

49: 1119-1134

Page 192: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Bibliography 174

Yagi H, Yoshida F (1974) Oxygen absorption in fermenters - effects of surfactants, antifoaming agents, and

sterilized cells. Ferment. Technol. 52: 905-916

Yongming Z, Liping H, Jianlong W, Jutang Y, Hanchang S, Yi Q (2002) An internal airlift loop bioreactor with

Burkholderia pickttii immobilized onto ceramic honeycomb support for degradation of quinoline. Biochem. Eng.

J. 11: 149-157

Zhang TC, Bishop PL (1994) Experimental determination of the dissolved oxygen boundary layer and mass

transfer resistance near the fluid-biofilm interface. Wat. Sci. Tech. 30: 47-58

Page 193: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

Curriculum vitae 175

Curriculum vitae

Name, Vorname Seletzky, Juri Martin

Geburtsdatum 05.11.1973

Geburtsort Bonn

Staatsangehörigkeit deutsch

Schulausbildung

1984 - 1993 Gymnasium, Otto-Kühne-Schule, Bonn

1990 - 1991 High-School, Hudson, Ohio/USA

Zivildienst

1993 - 1994 Altenheim, Haus am Redoutenpark, Bonn

Studium

1994 - 2000 Universität Siegen, Studium Wirtschaftsingenieurwesen

1997 - 1998 Technische Universität Turin Italien, Auslandsstudium

SAA Management Schule Turin/Italien, Auslandsstudium

Promotion

2001 - 2006 RWTH Aachen, Lehrstuhl für Bioverfahrenstechnik

Page 194: Process Development and Scale-up from Shake Flask to ... · Process Development and Scale-up from Shake Flask to Fermenter of Suspended and Immobilized Aerobic Microorganisms Juri

The aim of this study was to develop and advance techniques required to carry out process

developments with suspended and immobilized aerobic microorganisms cultivated in batch or

continuous mode. The applicability is demonstrated with the biological model system

Corynebacterium glutamicum grown on lactic acid. Respiration measurement in shake flasks

is introduced as a screening method to characterize the behavior of microorganisms exposed

to stress factors such as carbon starvation, anaerobic conditions, organic acids, osmolarity and

pH. It is evaluated and compared with the standard measuring methods, exhaust gas analysis

and respirometry. A simple and inexpensive shake flask screening method to perform

immobilization studies with solid supports is introduced. This method is applied to study the

adhesion of dormant and growing cells to ceramic and glass support materials with different

porosities and surface modifications. A novel everyday scale-up strategy from shake flasks to

fermenters based on empirical correlations of the volumetric mass transfer coefficient and the

pH change of the medium is developed and applied for batch and continuous cultures. Mixing

time, gas-liquid mass transfer, the influence of antifoam agent, power input, gas hold-up,

productivity, and long term behavior of a novel external loop biofilm reactor with a

honeycomb ceramic monolith as immobilization carrier are determined. With modeling the

productivities of biofilm and standard reactors are compared for aerobic processes with

growth associated product formation. To identify optimal operating conditions, the influences

of dilution rate, mass transfer, carbon source concentration, and surface area on the

productivity is modeled.

PhD thesis Biochemical Engineering, RWTH Aachen University, Juri Martin Seletzky, 2007


Recommended