application of novel metabolic engineering tools for

198
KU LEUVEN FACULTY OF SCIENCE DEPARTEMENT OF BIOLOGY Laboratory of Molecular Cell Biology Dissertation presented in partial fulfilment of the requirements for the degree of Doctor in Biology Application of novel metabolic engineering tools for engineering of the complex trait of glycerol yield in Saccharomyces cerevisiae Georg Hubmann Supervisors: Prof. Dr. Johan Thevelein Prof. Dr. Elke Nevoigt Members of the Examination Committee: Prof. Dr. P. Van Dijck Prof. Dr. B. Teusink Prof. Dr. K. Verstrepen Prof. Dr. S. Guillouet Dr. M. Foulquié-Moreno Leuven 2013

Upload: others

Post on 12-Feb-2022

0 views

Category:

Documents


0 download

TRANSCRIPT

KU LEUVEN

FACULTY OF SCIENCE

DEPARTEMENT OF BIOLOGY

Laboratory of Molecular Cell Biology

Dissertation presented in partial fulfilment

of the requirements for the degree of Doctor in Biology

Application of novel metabolic engineering tools

for engineering of the complex trait of glycerol yield in

Saccharomyces cerevisiae

Georg Hubmann

Supervisors: Prof. Dr. Johan Thevelein

Prof. Dr. Elke Nevoigt

Members of the Examination Committee:

Prof. Dr. P. Van Dijck

Prof. Dr. B. Teusink

Prof. Dr. K. Verstrepen

Prof. Dr. S. Guillouet

Dr. M. Foulquié-Moreno

– Leuven 2013 –

© 2013 KU Leuven, Science, Engineering & Technology

v.u. Leen Cuypers, Arenberg Doctoraatsschool, W. de Croylaan 6, 3001 Heverlee

Alle rechten voorbehouden. Niets uit deze uitgave mag worden vermenigvuldigd en/of openbaar gemaakt

worden door middel van druk, fotokopie, microfilm, elektronisch of op welke andere wijze ook zonder

voorafgaandelijke schriftelijke toestemming van de uitgever.

All rights reserved. No part of the publication may be reproduced in any form by print, photoprint, microfilm,

electronic or any other means without written permission from the publisher.

ISBN 978-90-8649-618-1

Wettelijk depot D/2013/10.705/33

“We have merely scratched the surface of the store of knowledge

which will come to us. I believe that we are now, a-tremble on the

verge of vast discoveries - discoveries so wondrously important

they will upset the present trend of human thought and start it

along completely new lines.”

Acknowledgements

This research work was supported by the DFG (Deutsche Forschungsgemeinschaft), the

Agency for Innovation by Science and Technology (IWT-Flanders) and the EC 7th

Framework program, NEMO.

I want to thank my supervisors, Prof. Dr. Johan Thevelein and Prof. Dr. Elke Nevoigt for

giving me the opportunity to work together for more than four years. Their great supervision,

unbroken support and motivation was enormous and of great help to bring this research work

to completion. I would also like to express my gratitude to the members of my examination

committee, Prof. Dr. Patrick Van Dijck, Prof. Dr. Bas Teusink, Prof. Dr. Kevin Verstrepen,

Prof. Dr. Stephane Guillouet, Dr. M. Foulquié-Moreno and my supervisors, Prof. Dr. Johan

Thevelein and Prof. Dr. Elke Nevoigt for critical reading and commenting of the manuscript.

I would like to thank my colleagues Dr. Maria Foulquié-Moreno, Dr. Yudi Yang, Dr. Thiago

Pais, almost Dr. Mekonnen Demeke, Dr. Jean Paul Meijnen, Dr. Ben Souffriau and of course

all other members of the Genetic Analysis Group for stimulating discussions and good

cooperation during my PhD. Thanks also to the technicians Catherina Coun, Martine De

Jonge, Eef Allegaert, Evy Vanderheyden, Nico Vangoethem, Jan Wouters and Paul

Vandecruys, who organize the daily work in the laboratory and therefore, greatly support

one's own work in every respect. Of course I would like to express my greatest gratitude to the

two most indispensible persons @MCB, Leni Vandoren and Hilde Florquin, for the Dutch

lessons and the introduction to the Flemish culture and for having a open ear for all

problems. Finally, I would like to thank my students Lotte Mathe, Hind Hashweh and Ilke

Suder for their work and good cooperation during the internship or master thesis. Best wishes

for the future and thanks for all. It was a pleasure for me to teach you about the world of

yeast.

Also, I want to thank the team of the Cargill R&D center in Vilvoorde, Dr. Jean-Claude de

Troostembergh, Dr. Luigi Concilio and Nicolas Meurens for good team work and support of

this work.

Many thanks go to all my colleagues and friends, Harish, Dries, Joep, Tom & Tom, Joana,

Huyen and all other @MCB for a pleasant working atmosphere and adventurous time that we

speed together in the lab and outside. Thanks to my friends Frido, Vinay, Vipul and Uli for

the time together in Leuven, for sharing the good and bad times.

Special thanks go to all workers of the NMBS, who decided to go on strike before Christmas

2011. Thanks to their help, I met the loveliest person, who ever since I love and adore at

Leuven's train station on the 23rd

December at 6:27 am in the morning. Nadja, I want to thank

you for sharing so many moments with me in the past year, for your help and support

especially during the final phase of my PhD. I gratefully appreciate this and hope that there

are many more years for us to come.

Georg Hubmann

May 20, 2013

Table of Contents

TABLE OF CONTENTS .................................................................................................. VI

LIST OF FIGURES ........................................................................................................... IX

LIST OF TABLES ............................................................................................................ XI

LIST OF ABBREVIATIONS .......................................................................................... XII

LIST OF GENES AND GENE PRODUCTS ................................................................ XIV

SUMMARY ................................................................................................................ XVI

SAMENVATTING ....................................................................................................... XVII

Introduction ................................................................................................................ 1

CHAPTER I

Literature Overview

1. Yeast of the Saccharomyces sensu stricto family ................................................. 6

2. Alcoholic fermentation ......................................................................................... 7

2.1 Alcoholic beverages ................................................................................................................. 7

2.2 Bioethanol production .............................................................................................................. 9

3. Glycerol formation in Saccharomyces cerevisiae ............................................... 14

3.1 Glycerol Metabolism .............................................................................................................. 15

3.2 Osmoadaption in Saccharomyces cerevisiae .......................................................................... 18

3.3 The role of glycerol in Saccharomyces cerevisiae redox metabolism .................................... 29

3.4 Glycerol formation as complex trait in yeast .......................................................................... 33

Table of Contents VII

4. Engineering of Saccharomyces cerevisiae for Ethanol Fermentation ................ 34

4.1 Metabolic engineering of Saccharomyces cerevisiae – importance and strategies ................ 34

4.2 Previous approaches to optimize yields in ethanol production ............................................... 38

5. Conclusions and scope of the present thesis ....................................................... 49

CHAPTER II

Gpd1 and Gpd2 fine tuning for sustainable reduction of glycerol formation

in Saccharomyces cerevisiae

1. Abstract ............................................................................................................... 52

2. Bibliographic references ..................................................................................... 53

3. Scientific contribution ......................................................................................... 53

4. Manuscript I ........................................................................................................ 54

4.1 Introduction ............................................................................................................................ 54

4.2 Materials and Methods ........................................................................................................... 56

4.3 Results .................................................................................................................................... 62

4.4 Discussion .............................................................................................................................. 72

CHAPTER III

Quantitative trait analysis of yeast biodiversity yields novel gene tools

for metabolic engineering

1. Abstract ............................................................................................................... 78

2. Bibliographic references ..................................................................................... 79

3. Scientific contribution ......................................................................................... 79

4. Manuscript II ....................................................................................................... 80

4.1 Introduction ............................................................................................................................ 80

4.2 Materials and Methods ........................................................................................................... 83

4.3 Results .................................................................................................................................... 93

4.4 Discussion ............................................................................................................................ 108

4.5 Supplementary material ........................................................................................................ 112

VIII Table of Contents

CHAPTER IV

Identification of multiple alleles conferring low glycerol and high ethanol yields

in Saccharomyces cerevisiae ethanolic fermentation.

1. Abstract ............................................................................................................. 120

2. Bibliographic references ................................................................................... 121

3. Scientific contribution ....................................................................................... 121

4. Manuscript III ................................................................................................... 122

4.1 Introduction .......................................................................................................................... 122

4.2 Materials and Methods ......................................................................................................... 125

4.3 Results .................................................................................................................................. 132

4.4 Discussion ............................................................................................................................ 144

4.5 Supplementary material ........................................................................................................ 148

CHAPTER V

General discussion and future perspectives

1. Application and valorisation of the project ....................................................... 156

2. Genetic configurations of strains with reduced glycerol synthesis ................... 157

3. Extending the toolbox for yeast metabolic engineering .................................... 161

4. Designed on the drawing board: Future cell factories ...................................... 162

REFERENCES ................................................................................................................ 165

List of Figures IX

List of Figures

CHAPTER I

Figure 1 Trends of corn and ethanol production in the United States of America since 1981 ........ 11

Figure 2 Annually real price indices of food, sugar and cereals from 1990 to 2011 ....................... 12

Figure 3 Glycerol Metabolism in Saccharomyces cerevisiae. ......................................................... 16

Figure 4 High osmolarity glycerol pathway in Saccharomyces cersvisiae. .................................... 20

Figure 5 Activation of the Sln1 branch in high osmolarity ............................................................. 21

Figure 6 Hog1 dependent cytosolic changes after a hyperosmotic shock ....................................... 25

Figure7 Overview of the differential expression of functional gene families upon

osmotic stress .................................................................................................................... 27

Figure 8 Mechanisms of Hog1 dependent transcriptional regulation upon osmostress .................. 28

Figure 9 Mechanism of NADH oxidation and transport in Saccharomyces cerevisiae .................. 31

Figure 10 Quantitative trait locus (QTL) mapping in Saccharomyces cerevisiae ............................. 37

Figure 11 Major metabolites produced in an ethanol fermentation process

of Saccharomyces cerevisiae ............................................................................................ 39

Figure 12 Metabolic engineering strategies directly targeting glycerol synthesis or transport ......... 42

Figure 13 Changing the NADH/ NADPH cofactor imbalance ......................................................... 46

Figure 14 Co-fermentation of substrates or alternative reduced product formation .......................... 48

CHAPTER II

Figure 1 Schematic overview showing the generation of 11 Saccharomyces cerevisiae

strains with reduced levels of Gpd1 and Gpd2 ................................................................. 63

Figure 2 Specific GPDH activity and glycerol yields of Saccharomyces cerevisiae

wild type and engineered strains ....................................................................................... 64

Figure 3 Growth of Saccharomyces cerevisiae wild type and engineered strains with

reduced levels of Gpd1 and Gpd2 ..................................................................................... 65

Figure 4 Ethanol yields (A), maximal volumetric ethanol production rates (B), biomass yields (C)

and acetate yields (D) of Saccharomyces cerevisiae wild type and engineered strains .... 69

X List of Figures

CHAPTER III

Figure 1 Variation in glycerol and ethanol yield ............................................................................. 94

Figure 2 Glycerol and ethanol yield in the segregants of the diploid parent strains and

in segregants from the cross between the selected haploid parent strains ......................... 95

Figure 3 Plots of SNP variant frequency versus chromosomal position and

corresponding P-values ..................................................................................................... 98

Figure 4 SNP variant frequency and P-values determined in individual segregants

for downscaling of the QTLs .......................................................................................... 100

Figure 5 Identification of SSK1 as the causative gene in the QTL on chromosome XII ............... 101

Figure 6 Glycerol and ethanol yield after deletion or reciprocal exchange of the SSK1 alleles .... 104

Figure 7 Glycerol and ethanol yields and osmostress tolerance in fermentations with the

industrial bioethanol production strain Ethanol Red in which one or two copies

of the ssk1E330N…K356N allele had been introduced ............................................................ 107

CHAPTER IV

Figure 1 Phenotypes of the parental strains ER7A and CBS4C and the segregant 26B ................ 133

Figure 2 Glycerol and ethanol yield in the parental strains, segregant 26B and ER7A ................. 134

Figure 3 Plots of SNP variant frequency versus chromosomal position and corresponding

probability of linkage to the superior or inferior parent .................................................. 136

Figure 4 Linkage analysis of QTLs on chr. II, IV and XIII with different groups

of segregants. .................................................................................................................. 138

Figure 5 Reciprocal hemizygosity analysis ................................................................................... 140

Figure 6 Epistatic analysis of gpd1L164P in segregant 26B, ER7A, the diploid 26B x ER7A

and BY4742 .................................................................................................................... 142

Figure 7 Presence of 26B alleles, smp1R110Q,P269Q, gpd1L164P and hot1P107S,H274Y,

in the selected segregant population ............................................................................... 143

CHAPTER V

Figure 1 Multi - level control of high osmolarity glycerol pathway

in Saccharomyces cersvisiae ........................................................................................... 160

List of Tables XI

List of Tables

CHAPTER II

Table 1 Saccharomyces cerevisiae strains used ............................................................................. 57

Table 2 PCR primers and plasmids used ........................................................................................ 58

Table 3 Fermentation time, product yields and carbon balances of engineered

strains of Saccharomyces cerevisiae ................................................................................. 67

Table 4 Performance of engineered strains of Saccharomyces cerevisiae with reduced

levels of Gpd1 and Gpd2 .................................................................................................. 71

CHAPTER III

Table 1 Saccharomyces cerevisiae strains used ............................................................................. 84

Supplementary Table 1: Primers used ............................................................................................. 112

Supplementary Table 2: Plasmids used in this study ....................................................................... 115

CHAPTER IV

Table 1 Saccharomyces cerevisiae strains used ........................................................................... 125

Supplementary Table 1: Primers used ............................................................................................. 148

Supplementary Table 2: Plasmids used in this study ....................................................................... 153

XII List of Abbreviations

List of Abbreviations

AMP Adenosine monophosphate

ATP Adenosine triphosphate

BC Before christ

bp Base pair(s)

Cmol Moles carbon of a substance

CO2 Carbon dioxide

Co –A Coenzyme A

DHA Dihydroxy acetone

DHAP Dihydroxy acetone phosphate

DNA Deoxyribonulcleic acid

DOI Digital object identifier

DTT Dithiothreitol

dNTP Nucleotide triphosphate

EDTA Ethylendiaminetetraacetic acid

FAD Flavin adenine dinucleotide

FAO Food and Agriculture Organization of the United Nation

GAP Glyceraldehyde 3-phosphate

GAPDH NAD+ dependent glyceraldehyde 3-phosphate dehydrogenase

GPDH NAD+ dependent glycerol 3-phosphate dehydrogenase

HDAC Histone deacetylase complex

HOG High osmolarity glycerol pathway

HPLC High performance liquid chromatrography

LB Luria-Bertani medium

L-G3P Glycerol 3-phosphate

M Molar, mol/l

MAPK Mitogen-activated protein kinase

MAPKK Mitogen-activated protein kinase kinase

MAPKKK Mitogen-activated protein kinase kinase kinase

NADH Nicotinamide adenine dinucleotide (reduced)

NAD+

Nicotinamide adenine dinucleotide (oxidised)

List of Abbreviations XIII

NADPH Nicotinamide adenine dinucleotide phosphate (reduced)

NADP+ Nicotinamide adenine dinucleotide phosphate (oxidised)

O2 Oxygen

OD Optical density

PCR Polymerase chain reaction

PEG Polyethylene glycol

PMSF Phenylmethylsulfonyl flourid

QT Quantitative trait

QTL Quantitative trait locus

RHA Reciprocal hemizygosity analysis

RI Refraction index

RNA Ribonulcleic acid

rpm Rounds per minute

RQ Respiratory quotient

SAGA Spt-Ada-Gcn5-acetyltransferase histone acetylase complex

SDS Sodium dodecyl sulfate

SHF Separate hydrolysis and fermentation

SNP Single nucleotide polymorphism

SNV Single nucleotide variant

SSF Simultaneous saccharification and fermentation

STRE Stress response element

SWI/SNF Multi protein complex involved in nucleosome remodelling

TCA Tricarboxylic acid cylce

TEA Triethanolamine

TF Transcription factor

TPI Triosephosphate isomerase

Tris Tris (hydroxymethyl) aminomethane

UV Ultra violett

v/v Volume per volume [%]

w/v Weight per volume [%]

YD Complex medium (1 % yeast extract, 2 % D-glucose)

YPD Complex medium (1 % yeast extract, 2 % peptone 2 % D-glucose)

XIV List of Genes and Gene Products

List of Genes and Gene Products

ACS1, ACS2 Acetyl-coA synthetase

ADH1, ADH2,

ADH3 Alcohol dehydrogenase

AMD1 Amidase of Zygosaccharomyces rouxii

bleR Phleomycin resistance marker

CDC42 Small rho-like GTPase

CYC8 General transcriptional co-repressor

DAK1, DAK2 Dihydroxyacetone kinase

FPS1 Plasma membrane channel involved in efflux of glycerol

frdA NAD+

dependent fumarate reductase of Escherichia coli

GAL1 Galactose inducable promoter

gapN NADP+

dependent glyceraldehyde 3-phosphate dehydrogenase of

Streptococcus mutans

non-phosphorylating NADP+

dependent glyceraldehyde 3-phosphate

dehydrogenase from Bacillus cereus, or Kluyveromyces lactis

GCY1 NADP+ dependent glycerol dehydrogenase

GDH1 NADP+ dependent glutamate dehydrogenase

GIN11 X - element of the subtelomeric regions

gldA NADH dependent glycerol dehydrogenase

GLN1 Glutamine synthetase

GLT1 NAD+

dependent glutamate synthase

GPD1, GPD2 Glycerol 3-phosphate dehydrogenase

GPP1, GPP2

GRE3 Aldose reductase

GUT1, GUT2 Glycerol kinase

GUP1, GUP2 Plasma membrane protein involved in remodeling GPI anchors

HOG1 Mitogen-activated protein kinase involved in osmoregulation

HOT1 Transcription factor inducing glycerol biosynthetic genes

KanMX6 Kanamycin resistence gene

mgsA NADPH dependent methylglyoxcal synthase

List of Genes and Gene Products XV

mhpF NAD+

dependent acetaldehyde dehydrogenase of Escherichia coli

MSB2 Mucin family member involved in signaling;

MSN2, MSN4 Transcriptional activator; activated in stress conditions

NAT1 Nourseothricin N-acetyl-transferase of Streptomyces noursei

NDI1 NADH:ubiquinone oxidoreductase

NDE1, NDE2 Mitochondrial external NADH dehydrogenase

NMD5 Karyopherin carrier protein involved in nuclear import of proteins

PBS2 MAP kinase kinase of the HOG signaling pathway

PFK26, PFK27 6-phosphofructo-2-kinase

PGK1 3-phosphoglycerate kinase

PTC1 Type 2C protein phosphatase (PP2C); dephosphorylates Hog1

PTP2, PTP3 Phosphotyrosine-specific protein phosphatase

RPD3 Histone deacetylase

SHO1 Transmembrane protein involve in osmosensing

SKO1 Basic leucine zipper transcription factor of the ATF/CREB family

SLN1 Histidine kinase osmosensor

SMP1 Transcription factor of the MADS-box family

SOR1 Sorbitol dehydrogenase

srlD NADH/ NADPH dependent sorbitol 6-phosphate dehydrogenase of

Escherichia coli

SSK1 Cytoplasmic response regulator

SSK2, SSK22 MAP kinase kinase kinase of the HOG signaling pathway

STE11 Signal transducing MEK kinase

STE20 Cdc42p-activated signal transducing kinase

STE50 Adaptor that links G protein-associated Cdc42p-Ste20p complex

STL1 Glycerol proton symporter of the plasma membrane

TEF1 Translation Elongation Factor 1

TDH1, TDH2

TDH3 Glyceraldehyde-3-phosphate dehydrogenase

TUP1 General repressor of transcription

URA3 Orotidine-5'-phosphate decarboxylase

YPD1 Phosphorelay intermediate protein

XVI Summary

Summary

The yeast, Saccharomyces cerevisiae is still the prime species used in ethanol production.

Traditionally, ethanol has been produced from raw materials such as sugar cane or sugar beet

and starch from corn or other grains. Future expansion of ethanol production are planed its

use as a biofuel requiring the utilization new non- food resources, particularly lignocelluloses.

In the present and future ethanol production process, the expenditure for the raw material is a

significant cost factor. Although industrial yeast strains are very efficient in producing

ethanol, the current ethanol yield (90 - 93% of the theoretical maximum) could in principle be

further improved. Particularly, the synthesis of glycerol, a major by-product of alcoholic

fermentation, has been regarded as a wasteful process. The overall goal of this project was to

develop novel industrial yeast strains, which are able to produce more ethanol from the

traditional and future raw materials by reducing the yield of the undesirable by-product

glycerol. The challenge herein is that glycerol serves different important physiological

functions such as redox balancing and osmotic stress tolerance. In this project, two alternative

avenues were used to identify a genetic configuration for a ‘low glycerol’ producer strain. The

first approach was a rational engineering strategy aiming to finely tune the activity of the rate-

limiting enzyme of glycerol synthesis, the glycerol 3-phosphate dehydrogenase. The basic

idea here was to gradually reduce the expression of the corresponding genes, GPD1 and

GPD2, by replacing their native promoters through mutated low strength TEF1 promoter

versions. After the promoter exchange, the mutants were characterized for their residual

glycerol formation, fermentation and growth. In the second approach, the genetic

configuration of wild-type yeast with low glycerol yields were analyzed. Its genetic

determinants, which cause the ‘low glycerol’ phenotype were identified, using an advanced

genetic mapping methodology base on next-generation pooled segregants whole-genome

sequence analysis. The four genes, GPD1 SSK1 SMP1 and HOT1 were identified, which

harbored mutations linked to the 'low glycerol' phenotype. The SSK1 allele of the superior

low producing glycerol strain, ssk1E330N...K356N

was subsequently used as novel gene tool for

targeted genetic improvement of the ethanol yield of a frequently used ethanol production

strain with minimal risk of affecting its other industrially important traits.

Samenvatting XVII

Samenvatting

De gist, Saccharomyces cerevisiae is nog steeds het belangrijkste micro-organisme dat

gebruikt wordt voor de productie van ethanol. Traditioneel wordt ethanol geproduceerd uit

grondstoffen zoals suikerriet of suikerbieten en zetmeel uit maïs of andere granen.

Toekomstige uitbreiding van de ethanolproductie vereist het gebruik van nieuwe suiker- en

zetmeelbronnen die niet als voedsel dienen, voornamelijk lignocellulose, voor de productie

van biobrandstof In het huidige en toekomstige ethanol productieproces, is het verbruik van

grondstoffen een belangrijke kostenfactor. Hoewel industriële giststammen zeer efficiënt zijn

in het produceren van ethanol, kan het huidige ethanol rendement (90 - 93% van het

theoretische maximum) in principe verder verbeterd worden. De synthese van glycerol, een

belangrijk bijproduct van alcoholische gisting, wordt in het bijzonder beschouwd als een

verspilling van grondstoffen. Het doel van dit project was om nieuwe industriële giststammen

te ontwikkelen, die in staat zijn om meer ethanol te produceren uit de traditionele en

toekomstige grondstoffen door de vorming van het ongewenste bijproduct glycerol te

verminderen. De uitdaging ligt in het feit dat glycerol verschillende belangrijke fysiologische

functies in de cel vervult, zoals behouden van het redox evenwicht en beschermen van de cel

bij osmotische stress. In dit project werden twee alternatieven gebruikt om een genetische

configuratie voor een stam met lage glycerol productie te identificeren. De eerste benadering

is een rationele strategie gebaseerd op de beschikbare kennis over het snelheidsbepalend

enzym in glycerol synthese door gist. Het basis idee is de expressie van dit enzym gelijkmatig

te verminderen en zo te onderzoeken hoeveel glycerol moet gevormd worden opdat de cel

goed kan fermenteren en groeien. In de tweede benadering wordt de genetische configuratie

van bestaande wild-type gist stammen met lage glycerol productie tijdens fermentatie

geanalyseerd. De genetische factoren, die dit 'lage glycerol' fenotype veroorzaken worden

geïdentificeerd met behulp van een geavanceerde genetische karteringsmethode gebaseerd op

‘next-generation pooled segregants whole genome sequence analysis’. De geïdentificeerde

causatieve factoren worden gebruikt als nieuw 'gene tools' voor doelgerichte genetische

verbetering van de ethanol opbrengst van een gist stam vaak gebruikt voor ethanolproductie

op een manier waarop zijn andere industrieel belangrijke eigenschappen minimaal worden

beïnvloedt.

Introduction

The production of ethanol by fermentation of grape juices or hydrolyzed cereals originates

from time immemorial. Earliest evidences for preparation of alcoholic beverages and bread

using yeast are dating back to 7000BC in China. Since that time fermentation technology

spread outwards from Asia, Mesopotamia and Egypt (Dequin and Casaregola, 2011) to

become the most important and the most widely used biotechnological process in the world.

Ethyl alcohol is produced in the form of alcoholic beverages such as beer, wine, cider and

other flavoured alcoholic beverages, which contain a low amount of ethanol or in form of

distilled spirits, like whisky, rum and vodka, where the ethanol concentration exceeds more

than a third of the total volume. In 2011, the global alcoholic beverage industry produced a

volume of 206.7 billion litres, reaching a market value of $1 trillion1. Besides the

consumption in alcoholic beverages, pure ethanol is also used as alternative fuel to substitute

mineral fuels in motor vehicles. Over the past decade, demand and production of fuel ethanol

grew exponentially, resulting in a production of 85 billion litres ethanol in 20122 mainly in the

United States and Brazil, which are the two largest producers in the world.

Despite the centuries-old economical importance, the alcoholic fermentation process itself

remained mysterious for a long time and the earliest investigations only started in the 19th

century by the two chemists, Antoine Lavoisier (1789) and Joseph Gay-Lussac (1810). Both

scientists described the “alcoholic fermentation” phenomenon as a chemical conversion of

sugar to carbonic acid and alcohol without being aware of the central role of yeast in this

process. It is astonishing how precisely Gay-Lussac predicted the weight parts of the formed

products, laying the foundation of the equation describing the chemical conversion during

alcoholic fermentation:

C6H12O6 ➝ 2 C2H5OH + 2 CO2

1Alcoholic Drinks: Global Industry Guide, MarketLine March 2012

2F.O. Licht’s World Ethanol & Biofuel Report

2 Introduction

In the early nineteenth century, light microscopy improved considerably, enabling people to

visualize small microbes. This opened up the possibility to study the microbiology of

fermentation and putrefaction. These processes caused the spoilage of vast quantities of wine

in those days. The three scientists, Charles Cagniard-Latour, Friedrich Kützing and Theodor

Schwann discovered and described the microbes responsible for the alcoholic fermentation

for the first time in history. In 1837, Schwann published his observation on alcoholic

fermentation (Schwann, 1837), which describes yeast as a living organism resembling fungi.

Furthermore, he concluded that alcoholic fermentation is started by the development of the

fungus decomposing the available sugar. Schwann’s idea of yeast as a living organism was

opposed by the famous chemists of these days, Jöns Berzelius, Justus von Liebig and

Friedrich Wöhler, which defined fermentation as being a purely physico-chemical

phenomenon. According to their theory, alcoholic fermentation was a chemical

decomposition of sugary plant juice after exposure to air, in which yeast was a body

composed from degenerated molecules (Barnett, 2003). The rejection of this theory and

acceptance of yeast as living organism came nearly two decades later by the work of Louis

Pasteur (Pasteur, 1857), who affirmed yeast as a microorganism causing the alcoholic

fermentation. In addition, he found other fermentation products, for instance glycerol and

succinic acid, besides ethanol and carbon dioxide formed in this physiological process.

In the late nineteenth and early twentieth century, yeast sugar metabolism was further studied

using cell free yeast extracts by Emil Fischer (Fischer, 1894; Fischer and Thierfelder, 1894)

and Eduard Buchner (1897). Their discoveries were the first comprehensive studies of

enzymes and laid the foundation for modern biochemistry. Ever since the discovery of

Lavoisier and Gay-Lussac, one can truly say that studying alcoholic fermentation in yeast has

been the driving force for many new scientific findings in microbiology and biochemistry.

This also explains why Saccharomyces cerevisiae has become a model organism in biology,

which is extraordinary well characterized as a biological system. In fact, the Saccharomyces

cerevisiae genome was the first to be fully sequenced (Goffeau et al, 1996) and many genes

and their function are already known. This paved the way to study systematically the

genotype - phenotype relations. Moreover, yeast is used as model organism for higher

eukaryotes due to the conservation of many pivotal structural and regulatory pathways.

Besides the basic research, our comprehensive knowledge of yeast can be applied to optimize

yeasts for biotechnological applications. In particular, the production of ethanol used as

Introduction 3

biofuel has become a prominent subject in the last decade. Key challenges in engineering

yeasts are the extension of substrate utilization, the optimization of the product formation and

the improvement of robustness and stress tolerance (Stephanopoulos, 2007) in order to create

tailor made Saccharomyces cerevisiae strains, which are perfectly suited for their specific

industrial applications.

During its millennia-long domestication, yeast adapted almost perfectly to the fermentation

environment and became very efficient in producing ethanol, reaching currently 90 - 93% of

the theoretical maximum ethanol yield. However, there is a clear interest of the ethanol

producers in using yeast strains with higher conversion efficiency than that of naturally

appearing strains. Particularly, the synthesis of glycerol, a major by-product of alcoholic

fermentation, has been regarded as a wasteful process, because this metabolite is a highly

reduced product not further converted to ethanol. The overall goal of this project is to develop

novel industrial yeast strains, which are able to produce more ethanol from the traditional and

future raw materials by reducing the yield of the undesirable by-product glycerol. The

challenge herein is that glycerol serves different important physiological functions such as

redox balancing and osmotic stress tolerance.

The goal of the present work is to find genetic configurations, which determine low glycerol

formation along with cellular integrity. Rational engineering was applied to finely tune

glycerol formation and thereafter analyze its impact on the fermentation process. In addition

to rational engineering, a reverse engineering approach was applied. In the latter approach, the

genetic configuration of a natural occurring low glycerol producing Saccharomyces cerevisiae

was elucidated determining the phenotype - genotype relationship of glycerol as a complex

trait. Gene variants, identified as being causative for reduced glycerol formation, were used as

novel gene tools to reduce glycerol yield in industrial yeast. The methods and strategies

presented in this work can in principle be applied on any other phenotype in Saccharomyces

cerevisiae and clearly add to our present toolbox of yeast metabolic engineering.

Chapter I

Literature Overview

6 Chapter I

1. Yeast of the Saccharomyces sensu stricto family

Yeasts are defined as unicellular eukaryotes and are classified in the kingdom of fungi. The

word yeast, descending from the Old English gist or gyst, is a collective term grouping

diverse unrelated species of different taxonomic and phylogenetic origin. This diversity

comprises the commonly known ‘baker’s yeast’, Saccharomyces cerevisiae, used for baking

and preparation of alcoholic beverages and also pathogenic species, like Candida albicans,

causing infections in humans. Baker’s yeast is classified in the Saccharomyces sensu stricto

family, which summarizes all yeast species relevant for the fermentation industry. The yeasts

of the Saccharomyces family, meaning sugar moulds, are able to convert sugar into ethanol

and CO2 via fermentation (Sicard and Legras, 2011) and have been domesticated thousands of

years ago to make beer, wine or bread. Their unique ability is to survive and grow without

oxygen by using the fermentation process (Sicard et al, 2011). This process occurs even in the

presence of oxygen, as long as there is an excess of glucose present. This phenomenon,

named after its discoverer Herbert Grace Crabtree (1928), allows yeasts to produce ethanol

and to repress its respiration. Yeasts exhibiting such a peculiarity are therefore referred as

Crabtree-positive. Once glucose is depleted and oxygen present, the fermentation product,

ethanol is taken up again and consumed via respiration. The yeast must shift their enzymes

and metabolism from fermentation to respiration, causing a small lag phase between two

growth phases, which is referred to as the diauxic shift (Crabtree, 1928).

Beer yeasts were the first microbes to be scientifically studied due to their relatively large cell

size compared to bacteria and due to the financial support from the alcoholic fermentation

industries, which incurred each year losses by putrefaction and false fermentation (Barnett,

2003). The best-known member of the Saccharomyces genus is the species cerevisiae, whose

name is the Latin word for beer referring directly to its main application. Besides the food and

beverage applications, Saccharomyces cerevisiae is one of the most thoroughly researched

eukaryotic microorganisms. Researchers have used it to gather information about the biology

of the eukaryotic cell and ultimately human biology. It was the first eukaryotic genome to be

sequenced (Dujon, 1996; Goffeau et al, 1996) and annotated. Furthermore, it has become an

indispensible model organism in microbiology and biochemistry, exemplified by the work of

Buchner (1897) on cell-free extracts, by Harden & Young’s discoveries of D-fructose 1,6-

bisphoshate and coenzymes (1905), and by Nurse’s findings on the cell cycle (1991).

Chapter I 7

2. Yeast alcoholic fermentation

2.1 Alcoholic beverages

Wine/Cider is made from fermented grapes or other fruits. Wine production has a long

tradition in human history. First evidences of wine preparation are dating back to

Mesopotamia and Egypt (Dequin et al, 2011). During the Roman Empire, wine making

spread around the Mediterranean area and Western Europe. Millennia later, the Spanish

conqueror and the era of imperialism brought grapes and winemaking technology to almost

every place in the world, where the climate is appropriate for growing grapes or other fruits,

which generally contain sucrose as main sugar. After juice extraction from the fruit, the wine

must is ready for fermentation without further pre-treatment or addition of nutrients of any

kind. There is a huge variety of wine and ciders worldwide with diverse aroma bouquet and

flavours. This diversity in wine products originates from the different yeasts, grapes and

vinification conditions, resulting in a variety of chemical compounds produced during this

fermentation process.

Beer is aside from wine one of the most popular alcoholic beverages in the world. It is

produced from grains, like barley, wheat, rice or corn. The grain starch is saccharified prior to

fermentation in the malting process. In the malting process, grains are germinated to produce

the enzymes (α-amylase, β-amylase, proteases and β-glucanase) necessary for starch

hydrolysis. The germination is arrested via a controlled drying process of the grains, which

also preserves the necessary enzymes for the wort preparation. Saccharification itself occurs

in the next step, the mashing process. In this step, the milled grains, referred as malt, are

heated in water up to 72˚C to allow for optimal condition for enzymatic starch degradation.

The solid residuals of the grains are separated from the liquid, resulting in a sweet syrup,

called brewer’s wort, which is subsequently boiled. During boiling, addition of hop to the

wort imparts the characteristic bitter flavour to the beer. After cooling, the wort is ready for

fermentation to either lager beers or ales. These two types of beers are distinguished by the

utilized yeasts and the temperature at which fermentation occurs. For lager beers bottom

fermenting yeast and a temperature of 5-14˚C are used, whereas for ale beers top fermenting

yeast and a temperature of 15-20˚C are used. Historically, authorities have strictly regulated

8 Chapter I

brewing by laws, like the Bavarian purity law of 1516. Today, brewing is a global industrial

business, dominated by multinational companies and a production capacity of over 100 billion

litres.

Sake, sometimes also referred as rice wine, has its origin in the Japanese culture. The

specification as a wine might be misleading, because sake production resembles more

brewing than vinification. In sake preparation, rice starch has to be hydrolysed prior to the

fermentation, comparable to the malting and mashing process in brewing. The mould

Aspergillus oryzae, is responsible for the conversion of starch to fermentable sugar by

producing the enzymes necessary for saccharification. In the initial phase of the fermentation,

only the mould is present, starting the saccharification process. Later yeast and more rice are

added to the mould culture resulting in a simultaneous saccharification and fermentation,

which continues for 2 - 3 weeks at 15 - 20˚C and reaches a final alcohol concentration of 18 -

20% [v/v]. Afterwards, sake is stored for a period of nine to twelve months, in which the sake

develops its characteristic smooth taste.

Distilled beverages/Spirits are alcoholic beverages, which are produced by distillation of

fermented grain, fruit or vegetables. There is a huge variety amongst spirits, all sharing one

common attribute, which is a high alcohol content of at least 20% [v/v]. Whiskies are made

from cereal grains. Its characteristic and unique flavours originates from the grain and its

processing. Apart from whiskey, neutral grain sprits are produced, which are low in flavour

and therefore devoid of any taste. This mostly pure alcohol is used for vodka, gin and other

flavoured distilled beverages. Rum or Cachaça are spirits base on fermented sugar cane juice

or sugar molasses, which are by-products of the sugar refining process. The majority of these

spirits are produced in Latin America and the Caribbean. Fermentation usually occurs rapidly

at higher temperatures of up to 40˚C. Brandy, Schnapps or Eau de vie are produced by

distillation of fermented grapes or other fruits and have their main origin in Europe.

Chapter I 9

2.2 Bioethanol production

Ethanol can substitute gasoline as fuel used for Otto engines. In fact, Henry Ford used ethanol

to fuel one of his first automobiles in the 1880s. Despite Ford’s choice, crude oil became the

most favoured primary source for fuels due to its easy exploitation and low production costs.

Till date, mineral fuels dominate world fuel markets; however views of policy makers started

to change after the first global oil crisis in 1973, during which extensive ethanol production

programs, likewise PROALCOOL in Brazil (Basso et al, 2008; Koizumi, 2003), were

launched. Despite increasing prices and insecure crude oil supply for mineral fuels, most

countries mainly kept on relying on fossil resources. This resulted in steadily increasing prices

for mineral fuels, which led to a renaissance of ethanol as alternative biofuel for substitution.

Today, policy makers regard ethanol production as one feasible alternative to fossil fuels in a

short and medium term perspective for partial substitution of petroleum-based fuels.

Moreover, the stimulation of ethanol production is also for the purpose of lowering carbon

dioxide emissions – the major contributor to human-induced climate change. For instance,

combustion of renewable fuel by motor vehicles releases only the amount of carbon dioxide

previously consumed by plants; thereby reducing the total CO2 emissions, which contribute to

global warming. Therefore, interest in alternative fuels has intensified during the last decade.

In 2012, the annual world production of ethanol exceeded 80 billion liters (F.O.Licht's, 2013)

indicating the importance of ethanol as a partial substitute for petroleum-based liquid fuels.

Ethanol can be mixed with gasoline in various ratios. Currently used blends in Europe are low

ethanol mixtures up to 10% ethanol. The fraction of alternative fuels in blends will increase in

the next years due to the bio-fuel initiatives prompted by governments of countries all over

the world3,4

(FAO, 2008). Currently, most ethanol is produced in Brazil and the United States

of America.

2.2.1 1st Generation bioethanol production

Traditionally, fuel ethanol has been produced from sugar or starch-containing feedstock such

as sugar cane, sugar beet, corn or small grains via microbial fermentation predominantly

3 EU (2003) DIRECTIVE 2003/30/EC OF THE EUROPEAN PARLIAMENT AND OF THE COUNCIL of 8

May 2003 on the promotion of the use of biofuels or other renewable fuels for transport.

4 USA (2007) ENERGY INDEPENDENCE AND SECURITY ACT, 110th Congress Public Law 110-140.

10 Chapter I

using the yeast Saccharomyces cerevisiae. Today, the majority of the produced ethanol still

originates from traditional feedstock - referred to as first-generation biofuel. Feedstock crops

and process types used for ethanol production vary globally, mainly depending on the climate.

After its extraction from the plant, the type of carbohydrate present distinguishes the two

main types of traditional feedstock, which contain either fermentable sugar or starch.

Sugary feedstock, composed of fermentable carbohydrates, such as sucrose, glucose and

fructose, originates from sugar cane, sugar beet or sweet sorghum. In these plants, sugars are

accumulated in a high quantity and are easily extractable. The sugar juice extracted from the

plants is either used for sugar refining or ethanol production, depending on the market price

for sugar or ethanol. Unavoidable waste streams during the sugar refining process are

molasses, which still contain residues of sugar. These molasses are still an ideal and cheap

feedstock for ethanol production. In general, molasses are mixed with the green juice to avoid

fermentation inhibition, caused by the high concentration of inhibitors in molasses. One

advantage of this high inhibitor concentration is that molasses can be stored whereas the juice

is rapidly deteriorated. The utilisation of sugar cane for ethanol production is restricted to

tropical areas around the equator. In Brazilian tropical climate, sugar cane is the most

abundant feedstock used for bioethanol production. After sugar extraction, the plant residual,

so-called bagasse, is further processed for heat (milling and distillation) or electricity

generation. According to Dos Santos (1997), ethanol from sugar cane has the highest energy

ratio, meaning that energy outcome is 6.45 - 9.53 times higher than the energy necessary for

its production. Brazilian bioethanol industry has perfected the production process during the

last three decades. Ethanol plants are mainly operated in fed-batch, referred as Melle-Boinot

process (Basso et al, 2011), or as continuous fermentation. After fermentation, yeast is

recycled and used for the next fermentation, enabling rapid fermentation and a reduced

contamination risk. In temperate climates, like Europe, Canada or Russia, sugar beet is grown

for sugar or ethanol production. Despite a higher sugar content in sugar beets, ethanol

production costs (per litre ethanol) from sugar beet (Basso et al, 2011) are higher and its

energy ratio (Dos Santos, 1997) is lower when compared to sugar cane. Sugar extraction from

beets is more complex than the process necessary for sugar cane and usually plant residuals

are not combusted to generate process energy. These residuals, namely the pulp and tops, are

marketed as high-value cattle feed.

Chapter I 11

Starch based feedstock, originating from corn or small grains, are mainly used for ethanol

production in temperate climates. In the United States of America (USA), currently the

worldwide biggest producer of ethanol, corn is preferably used for feedstock preparation. In

2012, over 300 million tons of corn were harvested in the USA and approximately 40% of the

corn harvest was used for ethanol production5. Since 2002, ethanol production and in lockstep

corn production increased exponentially driven by the American policy, enacting renewable

energy directives, like the Energy Independence and Security Act 6 (2007) (Figure1).

Figure 1 Trends of corn and ethanol production in the United States of America since 1981. Development

of the annual corn production in the USA (white bars) and the percentage of corn consumption for fuel ethanol

production (black bars) from 1981 to 2012 (adapted from Feed Grains: Yearbook tables, United States

Departement of Agriculture, Economic Research Service (USDA, 2012).

The ethanol production process from starch differs substantially from the sugar based process.

The grains are milled to reduce the size and to make the starch accessible for the next step. In

the liquefaction step, water is added to the meal to adjust a dry-matter content of 30 - 35%.

This starch slurry is passed through a heat exchanger (jet cooker), in which the starch granules

gelatinize at 70 - 90˚C. The viscosity of the gelatinized starch increases drastically, which

would complicate the further downstream processing steps. To counteract this, α-amylase is

added to partially hydrolyze starch to soluble maltodextrins (Douglas Crabb and Mitchinson,

5United States Departement of Agriculture (USDA) (2012) Feed Grains: Yearbook tables available at:

http://wwwersusdagov/data-products/feed-grains-database/feed-grains-yearbook-tablesaspx.

6USA (2007) ENERGY INDEPENDENCE AND SECURITY ACT, 110th Congress Public Law 110-140.

12 Chapter I

1997). Complete saccharification of the maltodextrin results after addition of glucoamylase,

which sequentially removes glucose units from the non-reducing end. This step can either

occur ahead (separate hydrolysis and fermentation, SHF) or simultaneous with the

fermentation (simultaneous saccharification and fermentation, SSF). Today, most grain-based

ethanol plants are operated in the continuous SSF process (Madson and Monceaux, 1999),

due to higher ethanol yields obtained in this process when compared to separate hydrolysis. In

contrast to the cell recycle in the Brazilian process, fermentation is started with active dry

yeast, which is simultaneously propagated. The increase of the ethanol yield in SSF compared

to SHF results from improved enzymatic hydrolysis and a controlled release of sugar avoiding

elevated osmotic pressure, which induces glycerol formation. Moreover, the lack of free sugar

also reduces bacterial growth. After fermentation and ethanol distillation, water is evaporated

from the slurry and the fermentation residuals are dried, producing distillers grains used as

cattle feed.

2.2.2 2nd

Generation bioethanol production

In 2007, the US congress enacted the Energy Independence and Security Act (2007), which

regulates the production of renewable fuels until the year 2022. It is stated in the law that

production of renewable fuels should annually increase until 2022 to a final total volume of

36 billion gallons. Utilization of only corn-starch derived bioethanol, referred in the law as

‘conventional biofuel’, will be insufficient to fulfil this ambiguous aim. Moreover, 1st

generation biofuels originate from feedstock and crops, used for feed and food production. Its

expansion already raised prices for food commodities (Figure 2), creating a global food crisis

in 2007 - 2008 and causing political and economical instability.

Figure 2 Annual real price indices of food, sugar and cereals from 1990 to 2011 (published by the Food and

Agriculture Organization of the United Nation (FAO)).

Chapter I 13

To avoid competition with food commodities, future renewable biofuels, termed in the US

law as ‘advanced biofuels’, should derive from non-starch renewable biomass, such as non-

food crops, crop residues, animal waste or algae. Especially, lignocellulosic biomass,

composed of cellulose, hemi-cellulose and lignin, is regarded as future solution in a long-term

perspective ensuring both food and energy security. Ethanol produced from such biomass is

termed 2nd

generation bioethanol or ‘cellulosic ethanol’. In the last decade, researchers of all

disciplines have made huge efforts in developing and implementing new technologies to

produce ‘cellulosic ethanol’ and ‘advanced biofuels’. In order to obtain a mature

economically viable production process, two major challenges need to be solved, which are

the efficient and cost-effective hydrolysis of the lignocellulosic biomass into its sugar

monomers and the construction of microorganisms for the conversion of the sugars into the

desired biofuel.

Advances in pre-treatment methods and hydrolysis have been reviewed by Kumar et al.

(2009) and Van Dyk & Pletschke (2012). Host microorganisms, have been extensively

engineered to adapt them to the new process conditions and utilize them for fuel production

(Darku and Richard, 2001; Nevoigt, 2008; Peralta-Yahya and Keasling, 2010; Peralta-Yahya

et al, 2012). The yeast, Saccharomyces cerevisiae was amongst the primary species

engineered for advanced biofuel production. Particularly, two heterologous pathways for the

utilization of the pentoses, i.e. xylose and arabinose, were established to enable pentose

fermentation of hemi-cellulose hydrolyzates (Nevoigt, 2008). Overall, advanced biofuel

technologies have made an impressive progress over the last ten years and at present several

demonstration cellulosic ethanol plants are worldwide operational or under construction.

14 Chapter I

3. Glycerol formation in Saccharomyces cerevisiae

Firstly mentioned in 1858, glycerol as a product of the alcoholic fermentation was discovered

by Louis Pasteur (Pasteur, 1858). He wrote in a letter to the French Academy of Science:

“Je vous prie de vouloir bien annoncer à l’Académie un résultat curieux et très-

inattendu. C’est la presence constante de la glycerine parmi les produits de la

fermentation alcoolique […] à 3 pour 100 environ du poids du sucre […]”

In general, glycerol is formed at a quantity of 2.0-3.6% [g g-1

] from the sugar during alcoholic

fermentation (Pasteur, 1858). More than a century after Pasteurs’ initial findings, the glycerol

metabolism of yeasts, and particularly that of S. cerevisiae, has been well understood and has

proven to be an indispensible part of its metabolism. In S. cerevisiae, glycerol has two major

physiological functions, which are: i) the regulation of cell turgor especially under high

osmolarity and ii) the replenishment of the cofactor NAD+ to enable the cytosolic redox

balance in the absence of oxygen.

The increase or reduction of the glycerol formation in the yeast S. cerevisiae has been

extensively studied. Reduced glycerol formation is desirable for fuel ethanol production, in

order to increase ethanol yields and to prevent the carbon use for other fermentation products.

Elevated glycerol formation is desirable in wine and beer production (Cambon et al, 2006;

Geertman et al, 2006; Heux et al, 2006; Nevoigt and Stahl, 1996; Remize et al, 1999;

Schuller and Casal, 2005) to improve the mouthfeel of the beverages or to reduce their

alcohol content (Cambon et al, 2006; Eglinton et al, 2002; Ehsani et al, 2009). Moreover,

glycerol can also be the main product in the fermentation of carbohydrates by S. cerevisiae.

During World War I, demands for glycerol increased due to the production of nitro-glycerine.

Steering agents, such as bi-sulfite ions, were added in the fermentation retarding ethanol

production by trapping the intermediate, acetaldehyde (Neuberg and Reinfurth, 1918).

Neuberg’s second form of fermentation increased glycerol synthesis, obtaining conversion

efficiencies from 23% to 28% depending on the process conditions (Wang et al, 2001).

Today, the polyalcohol, glycerol, is an important starting substance for chemical synthesis. Its

demand has increased ever since it was discovered, in the 19th

century.

Chapter I 15

3.1 Glycerol Metabolism

3.1.1 Synthesis

Glycerol synthesis is part of S. cerevisiae central catabolism for the degradation of carbohyd-

rates. It starts from dihydroxy-acetone-phosphate (DHAP), which originates from the split of

fructose 1,6-bisphosphate into DHAP and glyceraldehyde 3-phosphate (GAP). The two

products are the starting point of two metabolic pathways, which lead to the final products

ethanol from GAP and glycerol from DHAP (Figure 3). In S. cerevisiae wild type cells, the

metabolic pathway for glycerol formation is the reduction of DHAP by the NAD+ dependent

glycerol 3-phosphate dehydrogenase (GPDH) and the successive dephosphorylation by the

glycerol 3-phosphate phosphatase.

The key enzyme of the glycerol synthesis, GPDH, is encoded by the two isogenes GPD1 and

GPD2. First discovered by Larson et al. (1993) and Albertyn et al. (1994), the gene GPD1

has been identified as essential for growth under osmotic stress. Moreover, its expression is

induced under reduced water activity. The gpd1∆ mutant was clearly inhibited in growth

under such conditions. In order to increase osmotic stress tolerance, the GPD1 transcription

level is induced by the high osmolarity glycerol (HOG) pathway and leads to intracellular

accumulation of glycerol. A second isogene (GPD2) that encodes the GPDH in S. cerevisiae

was discovered later by Ansell et al. (1997). The transcription of the GPD2 gene was not

affected by changes in external osmolarity, however, gpd2∆ mutants showed poor growth in

the absence of oxygen. The oxygen availability did not directly increase the GPD2

expression, but seemed to be rather linked to the cell’s redox state which is significantly

affected by oxygen availability (Ansell et al, 1997). However, the importance of GPD2 is not

reflect in the specific enzyme activity, since the loss of GPD2 merely reduced the specific

GPDH activity; whereas strains deficient of GPD1, showed a decreased enzymatic activity

(Nissen et al, 2000a). The gpd1∆ gpd2∆ mutant was highly osmo-sensitive and unable to

grow under anaerobic conditions (Bjorkqvist et al, 1997; Nissen et al, 2000a).

The second step in the synthesis of glycerol is the dephosphorylation of L-G3P by the

glycerol 3-phosphatase, which is encoded by the two paralogs, GPP1 (standard name RHR2)

and GPP2 (standard name HOR2) (Norbeck et al, 1996). Mutants lacking GPP1 showed poor

growth under anaerobic conditions and GPP1 expression seemed to be transiently induced

16 Chapter I

under these conditions (Pahlman et al, 2001). The expression of GPP2 is strongly activated

through the HOG pathway in the presence of hyperosmotic or oxidative stress (Norbeck et al,

1996; Pahlman et al, 2001). Single mutants of either GPP1 or GPP2 remain unaffected during

osmotic stress, indicating that the two paralogs can substitute well for each other (Pahlman et

al, 2001). However, the deletion of both genes in the gpp1∆ gpp2∆ mutant caused sensitivity

to osmotic and oxidative stress and a growth inhibition under anaerobic conditions. Devoid of

glycerol 3-phosphatase activity, this mutant showed increased levels of L-G3P and produced

only minor amounts of glycerol (Pahlman et al, 2001).

Figure 3 Glycerol Metabolism in Saccharomyces cerevisiae. Glycerol is synthesized from dihydroxyacetone

phosphate (DHAP) in two enzymatic steps. First, DHAP is reduced to glycerol 3-phosphate (L-G3P) by the

glycerol 3-phosphate dehydrogenase (GPDH, encoded by GPD1 and GPD2). Consecutively, L-G3P is

dephosphorylated by the glycerol 3-phosphatase (GPP, encoded by GPP1 and GPP2). Besides its passive

diffusion, retention and efflux of glycerol are regulated by the Fps1 plasma membrane channel (Luyten et al,

1995). Glycerol uptake occurs via the transporter Stl1 and probably also via Gup1 and Gup2 (Neves et al, 2004).

In S. cerevisiae, glycerol can be dissimilated via L-G3P or dihydroxyacetone (DHA). The L-G3P pathway

dissimilates glycerol by the glycerol kinase, encoded by GUT1 and the mitochondrion-located FAD dependent

glycerol 3-phosphate dehydrogenase, encoded by GUT2. In the DHA pathway, glycerol is first oxidized to DHA

by the glycerol dehydrogenase, encoded by GCY1. DHA is then phosphorylated by the dihydroxyacetone kinase,

encoded by the two isoenzymes DAK1 and DAK2 (Molin et al, 2003). Although the DHA pathway is present and

functional, it might be insignificant for glycerol dissimilation (adapted from Nguyen and Nevoigt (2009)).

Chapter I 17

3.1.2 Glycerol Dissimilation

Under aerobic conditions, S. cerevisiae can use glycerol as a sole carbon and energy source.

In fungi, glycerol dissimilation might occur through two possible pathways (Figure 3), named

after their main intermediate, dihydroxyacetone (DHA) or glycerol 3-phosphate (L-G3P).

Glycerol 3-phosphate pathway. Two mutants, which were deficient in glycerol utilization,

were isolated by Sprague and Cronan (1977). They attributed the growth defect to the lack of

either the glycerol kinase or glycerol 3-phosphate dehydrogenase. This showed that glycerol

is utilized in a two-step process. The gene GUT1 encoded glycerol kinase and its disruption

resulted in a glycerol growth defect (Pavlik et al, 1993). Glycerol 3-phosphate is subsequently

oxidized to DHAP by the mitochondrion-located flavin adenine dinucleotide (FAD)-

dependent glycerol 3-phosphate dehyrdogenase, encoded by GUT2 (Ronnow and Kielland-

Brandt, 1993). DHAP can either enter into the central catabolism via the transformation to

glyceraldehyde 3-phosphate by a triose phosphate isomerase or can serve as a substrate for

lipid synthesis.

Dihydroxyacetone pathway. A second pathway for glycerol dissimilation via DHA, was

proposed by Norbeck and Blomberg (1997) (Figure 3). The first step in the pathway is the

oxidation of glycerol to DHA by the NADP+ dependent glycerol dehydrogenase followed by

the phosphorylation of DHA by the dihydroxyacetone kinase. GCY1, DAK1 and DAK2

encode this two-step pathway. Jung and coworkers recently confirmed that Gcy1 is the

NADP+ dependent glycerol dehydrogenase (Jung et al, 2012). Although its function is

unclear, the pathway might play a role in the regulation of the glycerol concentration during

hyperosmotic stress (Blomberg, 2000) or redox regulation (Costenoble et al, 2000).

3.1.3 Glycerol transport across the membrane

In S. cerevisiae, retention and efflux of glycerol across the plasma membrane are regulated by

the plasma membrane channel, Fps1 (Luyten et al, 1995; Sutherland et al, 1997; Van Aelst et

al, 1991). Mutants lacking FPS1 are sensitive to hypo-osmotic shock, indicating that Fps1 is

facilitating the efflux of glycerol for adaptation to sudden drops in osmolarity (Tamas et al,

18 Chapter I

1999). During growth on glycerol, yeast cells actively take up glycerol by the glycerol proton

symporter Stl1 (Ferreira et al, 2005; Zhao et al, 1994). Other glycerol uptake mechanisms

probably involve the two plasma membrane proteins, Gup1 and Gup2 (Holst et al, 2000).

3.2 Osmoadaptation in Saccharomyces cerevisiae

3.2.1 Accumulation of compatible solutes conferring osmotic stress tolerance

The availability of water is crucial for microbial growth and limitation of water becomes

detrimental or lethal for the organism. In solutions, the term aw (water activity) defines the

thermodynamically available water, taking into account the solutes present in the solution.

Microorganisms developed mechanisms to avoid water efflux and to survive in low aw

environments. These mechanisms can be divided into two basic strategies, which are either to

balance inorganic ions (usually KCl) or the production and accumulation of small organic

molecules with osmotic potential (Grant, 2004). Organic osmolytes, also named as compatible

solutes, are protecting the cells against protein or enzyme inactivation and prevent the

denaturation of macromolecules in cells facing low water activity. As implied by their name,

such small molecules are compatible with cellular functionality even at a high intracellular

concentration (Brown, 1978). In general, the most important compatible solutes in

microorganisms are uncharged or zwitterionic molecules and can be divided into the

following groups: i) polyols (glycerol, arabitol, trehalose and sucrose), ii) amino acids

(proline, glutamate and glutamine), and iii) ectoines, (ectoine and β-hydroxyectoine) (Grant,

2004).

In yeast and fungi, the polyol glycerol is the most prominent compatible solute to regulate the

cell turgor at high extracellular osmolarity and upon changes in the external water potential

(Brown, 1978; Grant, 2004; Nevoigt and Stahl, 1997). In S. cerevisiae, glycerol production

increased in medium containing high concentrations of NaCl (Blomberg and Adler, 1989).

Moreover, the measured specific GPDH activity was six fold higher than under basal

conditions. The yeast cells also favoured the production of glycerol by decreasing the

metabolic flux toward ethanol, in that the activity of the alcohol dehydrogenase was reduced.

Chapter I 19

A high level of GPDH expression, however, is not the only prerequisite for an enhanced

osmotolerance of S. cerevisiae, because ethanol-grown cells display also a high specific

GPDH activity level but less tolerance towards a reduced water potential (Andre et al, 1991).

Glycerol was produced, not only in cultures containing glucose as a carbon source, but also

when raffinose or ethanol served as the sole carbon source in high salinity media (Andre et al,

1991). Under hypotonic conditions, S. cerevisiae cells release glycerol into the surrounding

medium; whereas at high salinity it retains glycerol. Osmotolerance of S. cerevisiae can be

acquired or enhanced, by conditioning the cells with non-lethal NaCl concentrations. This

indicates that protein synthesis is required to establish a state of osmotolerance.

In contrast to S. cerevisiae, xerotolerant yeasts such as Zygosaccharomyces rouxii use a

different adaptation mechanism to respond to a hyperosmotic shock. Unlike S. cerevisiae,

which seems to enhance intracellular glycerol by increasing its synthesis through higher

activity of GPDH or phosphofructokinase, Z. rouxii retains and accumulates a higher

intracellular level of glycerol against the concentration gradient (Edgley and Brown, 1983;

Lages et al, 1999). Accumulation and synthesis of compatible solutes are highly regulated in

cells due to the rapid reaction necessary for adaptation to sudden changes in osmolarity in the

environment (Hohmann, 2002).

3.2.2 Osmotic sensing and signalling

Generally, S. cerevisiae accumulates glycerol as compatible solute to re-establish cellular

turgor under hyperosmotic conditions (Blomberg et al, 1989). This adaptation in the yeast is

exerted by a complex mechanism, enabling the cells to respond and adapt their physiology.

Changes in osmolarity are sensed and stimulate a rapid cellular response, thereby maintaining

cellular activity in the new growth environment. The cellular response mechanism is

ubiquitous in eukaryotes and follows basic principles commonly found in signalling

pathways, e.g. those controlling mating pheromone response, filamentous growth and other

stress responses. These signalling pathways are only responsive to a specific stimulus, and

execute a distinct cellular response for appropriate adaptation of cellular physiology. The

cellular responses involved are specific to their stimuli, avoiding erroneous cross-talk and

unwanted responses (Schwartz and Madhani, 2004). In yeast, osmo-adaptation is primarily

based on one signalling pathway, referred to as high-osmolarity glycerol (HOG) pathway. The

20 Chapter I

HOG pathway is a typical example of a mitogen-activated protein kinase (MAPK) pathway,

regulated by a three-tiered cascade of kinases. Central in the HOG pathway is the MAPK,

Hog1 (Brewster et al, 1993), which is activated by the MAP kinase kinase (MAPKK), Pbs2

(Brewster et al, 1993). Pbs2 itself is phosphorylated and activated by a third type of kinase, a

MAP kinase kinase kinase (MAPKKK). As shown in Figure 4, there are two different

enzymes of this type involved in Pbs2 activation. This cascade of kinases allows for multistep

regulation, for integration of several upstream control and sensing systems (Figure 4).

Figure 4 High osmolarity glycerol pathway in Saccharomyces cerevisiae. The yeast uses a mitogen-activated

protein kinase (MAPK) signalling pathway to respond to high osmolarity. The MAPK cascade is ubiquitous in

eukaryotes and usually composed of three sequentially acting kinases: the MAP kinase kinase kinase (MPKKK)

phosphorylates the MAP kinase kinase (MAPKK), which finally phosphorylates and activates the MAPK (e.g.

Hog1). The two independent upsteam osmo-sensing pathways, the Sln1-branch and Sho1-branch, activate the

MAPKK, Pbs2. The Sln1-branch is composed of the transmembrane protein Sln1, which forms together with

Ypd1 and Ssk1 a phosphorelay system responsible for inactivation of the MAPKKK, Ssk2 and Ssk22, under

basal conditions. The Sho1-branch is composed of the Sho1 transmembrane protein, with its C-terminal

elongation, the SH3 domain. This domain anchors Pbs2 to the membrane, where it interacts with the MAPKKK,

Ste11. This branch includes Ste20 and Ste50, which are part of the pheromone-response and filamentous MAPK

pathways. A third osmo-sensing branch is constituted by Msb2, which probably converges with the Sho1 branch

before activation of Pbs2. Active phosphorylated Hog1 is imported into the nucleus, where it interacts with

transcription factors (e.g. Hot1) to increase the expression of osmotic stress relevant genes, amongst them GPD1

and GPP2. After rapid signal amplification in the initial minutes, Hog1 is thereafter inactivated and

dephosphorylated by the phosphatases Ptp2, Ptp3 and Ptc1 to prepare the cell for the consecutive signal.

Chapter I 21

The two independent upstream osmo-sensing pathways, the Sln1-branch and Sho1-branch

activate the MAPKK Pbs2 (Hohmann, 2002; O'Rourke et al, 2002). Both branches may not

directly sense osmotic stress but rather act in response to mechanical stimuli in the plasma

membrane upon changes in osmolarity. Sln1 negatively regulates the HOG pathway and is

composed of two transmembrane domains and an intracellular histidine kinase (Hohmann,

2002; O'Rourke et al, 2002). Its intracellular kinase domain forms together with Ypd1 and

Ssk1 a phosphorelay system, a signaling concept more commonly present in prokaryotes

(Maeda et al, 1994). In this phosphorelay system, phosphate is transferred from the sensor

histidine-kinase to the response regulator Ssk1, via the intermediate protein Ypd1. Ssk1

subsequently interacts with the MAPKKK, Ssk2 and Ssk22, to start the signaling process.

Under low osmolarity, Sln1 is constitutively active and inactivates its downstream target Ssk1

by phosphorylation (Figure 5). A change to high osmolarity inactivates Sln1, leading to rapid

Ssk1 dephosphorylation, which then enables binding of Ssk1 with Ssk2 and Ssk22. This

triggers the auto-phosphorylation of the MAPKKK, and subsequent phosphorylation of Pbs2

and Hog1 (Hohmann, 2002; O'Rourke et al, 2002). During a hyperosmotic shock, the level of

dephosphorylated Ssk1 increases rapidly, suggesting that an unkown phosphatase might be

involved (Hohmann, 2002; O'Rourke et al, 2002). Avoiding hyperactivation of the HOG

pathway, the dephosphorylated Ssk1 is rapidly degraded by the ubiquitin-proteasome system

(Sato et al, 2003) to down-regulate the signal and prepare the cell for the next signal.

Figure 5 Activation of the Sln1 branch in high osmolarity conditions (modified from Posas et al. (1996)).

Dashed boxes indicate inactive elements. Arched arrows indicate phospho-transfer reactions, whereas straight

arrows simply indicate signal flow. At low osmolarity, Sln1 is activated and phosphorylates an aspartate residue

of Ssk1 via the intermediate Ypd1. Phosphorylated Ssk1 is inactive and inhibits the signal transduction. At high

osmolarity, Sln1 is inactivated, resulting in accumulation of dephosphorylated Ssk1. The dephosphorylated form

of Ssk1 interacts with the MAPKKK Ssk2/Ssk22, which then activates the MAPKK, Pbs2 and the MAPK,

Hog1.

22 Chapter I

The Sho1-branch is composed of the Sho1 transmembrane protein, with its C-terminal

elongation, the SH3 domain. Although Sho1 has been proposed as the osmo-sensor in this

branch, no activation mechanism was found for the protein. Recent investigations by

Zarrinpar et al. (2004) gave evidence that Sho1 together with Pbs2 acts as co-scaffold

proteins, linking all components of the Hog pathway to the membrane. The C-terminal, SH3

domain of Sho1 anchors the two scaffolds, Sho1 and Pbs2 (Raitt et al, 2000), thereby

enabling the interaction of Pbs2 with the MAPKKK Ste11 and the MAPK Hog1. The Sho1-

dependent activation of Hog1 involves two other membrane-localized proteins, Cdc42 and

Ste20. Upon osmotic stress, both genes are required for phosphorylation of Ste11, which in

turn activates Pbs2 involving Ste50 as a cofactor for phosphorylation; however, the exact

induction mechanism of Ste20 is still unclear. Other studies revealed Msb2, a mucin-like

transmembrane protein, as a potential osmo-sensor in the Sho1-branch, which monitors

movements between the cell wall and the plasma membrane (Hohmann, 2009; O'Rourke et al,

2002).

After rapid signal amplification in the initial minutes, the HOG pathway is brought back to a

basal level to prepare the cell for a consecutive signal. The constitutive activation of the HOG

pathway, by for example deletion of SLN1, is lethal for the yeast cells (Hohmann, 2002). The

inactivation of the HOG pathway occurs through dephosphorylation of Hog1 by the negative

regulators Ptp2, Ptp3 and Ptc1 (Warmka et al, 2001; Wurgler-Murphy et al, 1997). Ptp2 and

Ptp3 are phosphotyrosine specific phosphatases and remove the phosphate at Tyr176. Ptc1 is

a serine/threonine-specific phosphatase, which dephosphorylates Thr174. The activity of both

phosphatases is required to limit Hog1 acitivity, even under basal conditions (O'Rourke et al,

2002).

Once activated through dual phosphorylation, Hog1 regulates cellular adaptation processes

for short-term (<2 min) and long-term (>30 min) protection against osmotic stress (Westfall et

al, 2004).

Chapter I 23

3.2.3 Hog1 dependent short-term osmoadaptation

Increases in external osmolarity causes a yeast cell response within a few milliseconds, which

are visible through changes in the cell volume immediately after the shock. The cell shrinkage

is caused by the efflux of water due to a very high hydraulic permeability for yeast membrane

(Gervais and Beney, 2001). A recent study by Petelenz-Kurdziel et al. (2011) quantified the

dynamics of cell volume changes during hyperosmotic stress. In agreement with the earlier

studies of Gervais and Beney (2001), they found that the shrinkage is proportional to the

stress intensity until the cell volume reaches approximately 55% of its unstressed volume .

This volume is maintained even in the case of further increases in osmolarity (Petelenz-

Kurdziel et al, 2011). The severity of the cell shrinkage during osmotic stress forces the yeast

cell to adapt quickly to the new conditions within a time range of 5 to 15 min (Mettetal et al,

2008). These adaptations allow yeast cells to regain their initial, unstressed volume

approximately 45 minutes after the osmotic shock. This time span is too short for

transcriptional changes, which lead to cellular protection. Thus Hog1 triggers a rapid

cytoplasmic response that does not require protein synthesis. During osmo-adaptation, these

cytoplasmic processes may play a more critical role for an instant protection of the cells than

the longer lasting process of de-novo protein synthesis. Moreover, Westfall et al. (2008)

showed that cells deficient in the nuclear import of Hog1 or in which Hog1 was anchored at

the plasma membrane were still able to withstand a long-term hyperosmotic shock by ionic

and non-ionic solutes. In these cells, the usual Hog1 dependent changes in the transcriptional

program were not found, suggesting that Hog1 dependent gene expression is not critical for

survival of a hyperosmotic shock; however fast-acting Hog1 dependent processes, which do

not require de-novo protein synthesis, control the intracellular glycerol formation to re-

establish the osmotic balance (Mettetal et al, 2008). In fact, changes in yeast metabolism may

account for 80% of the increase in the glycerol flux, whereas the induction of gene expression

results in only a 20% increase. Contrary to the general assumption that glycerol flux increases

due to Hog1 dependent transcriptional activation of GPD1 and GPP2, these findings support

that the osmotic adaptations are mainly effected through rapidly induced metabolic shifts

towards glycerol production (Bouwman et al, 2011; Mettetal et al, 2008; Westfall et al,

2008).

24 Chapter I

As depicted in Figure 6, Hog1 controls glycerol formation and accumulation in order to

establish the metabolic conditions and to maintain a high intracellular glycerol level. These

mechanisms are briefly highlighted in the following paragraphs.

Control of glycerol flux across the plasma membrane. Glycerol transport has a crucial

function in short-term osmoadaptation to hyperosmotic and hypoosmotic shock. The plasma

membrane channel Fps1 reduces efflux of glycerol under hyperosmotic conditions and

facilitates its efflux in the absence of osmotic stress and particularly during a hypoosmotic

shock. In fact, a high intracellular glycerol concentration was present in fps1∆ mutants, which

caused a growth defect under anaerobic conditions and sensitivity to hypoosmotic shocks

(Tamas et al, 1999). There is clear evidence that Hog1 targets Fps1 to acquire tolerance

against arsenite (Thorsen et al, 2006) or acetic acid (Mollapour and Piper, 2006), yet the

Hog1 dependent control mechanism of Fps1 during a hyperosmotic shock is still unclear. A

recent study by Geijer et al. (2012) has shed some light on this mechanism. The Fps1 channel

might be regulated by the interplay of its transmembrane helices with its extended cytosolic

termini; both are required to fine-tune glycerol flux through the channel, allowing the

adaptation to a wide range of extracellular osmolarities. In addition to preventing glycerol

efflux by closing the Fps1 channel, intracellular glycerol accumulation during osmotic shock

seems to be facilitated by a transient induction of the glycerol/H+ symporter Stl1which

transports glycerol from the medium into the cell (Ferreira et al, 2005).

Adjustments of glycolytic flux. Kühn et al. (2008) investigated the changes in glycolysis

during osmoadaptation, suggesting an activation of glycolysis and a redirection of central

metabolic fluxes to glycerol rather than to pyruvate. Their model is supported by experimental

findings of Dihazi et al. (2004) and Westfall et al. (2008). First, the model suggested a Hog1

dependent activation of the upstream part of glycolysis, possibly via stimulation of the

activity of phosphofructo-kinase. This enzyme undergoes allosterical control by AMP and

fructose 2,6-bisphosphate, which both increase phosphofructo-kinase activity. The latter,

fructose 2,6-bisphosphate is the product of the phosphorylation of fructose 6-phosphate by 6-

phosphofructo-2-kinase, which is encoded by PFK26 and PFK27. Dihazi et al. (2004)

showed a Hog1 dependent increase in Pfk26 activity resulting in higher activity of the upper

part of glycolysis under hyperosmotic conditions. The second finding of Westfall et al. (2008)

Chapter I 25

postulates a shift of the glycolytic flux towards glycerol production at the costs of ethanol

formation caused by a Hog1-mediated activity loss of glyceraldehyde 3-phosphate-

dehydrogenase, Tdh1 Tdh2 and Tdh3. The decreased activity leads to lower conversion of

GAP into 1,3-bisphosphoglycerate and partially redirects the glycolytic flux to DHAP, the

first intermediate of glycerol synthesis. Westfall et al. (2008) further observed a Hog1

dependent decrease in activity of Tdh1 and Tdh3. Similar rerouting of the carbon-flux by

altering NAD+ dependent glyceraldehyde 3-phosphate dehydrogenase (GAPDH) and triose-

phosphate isomerase (TPI) levels has been associated with oxidative stress (Ralser et al,

2007).

Independent of Hog1, changes in metabolism, e.g. in reaction equilibria, may simply result

from the rapid shrinking of the cells. In particular, cell-shrinking releases NADH bound to

proteins, stimulating the production of glycerol (Bouwman et al, 2011). Furthermore, the

changes in redox status can cause shifts in the cellular distribution of Gpd1, as a recent study

of the subcellular distribution of Gpd1 during osmotic stress has shown (Jung et al, 2010).

The redistribution of Gpd1 may be a spatial metabolic regulation mechanism, which further

shifts the glycolytic flux towards the production of glycerol.

Figure 6 Hog1 dependent cytosolic changes after a hyperosmotic shock. After activation, Hog1 regulates

several steps to increase glycerol formation and accumulation, which are: i) the transcriptional activation of

GPD1 and GPP2 involving the transcription factor HOT1; ii) expression of the glycerol symporter Stl1 for

glycerol uptake from the medium; iii) activation of phosphofructo-2 kinase, encoded by PFK26; iv) control of

the plasma membrane channel Fps1 and v) down-regulation of glyceraldehyde 3-phosphate dehydrogenase

(GAPDH), encoded by TDH1, TDH2 and TDH3.

26 Chapter I

Additional mechanisms. Besides the shift in glycerol accumulation and production, Hog1

regulates several other cellular processes upon hyperosmotic stress, like regulation of protein

biosynthesis (Bilsland-Marchesan et al, 2000; Teige et al, 2001), cell-cycle progression

(Alexander et al, 2001) and ion transport (Proft and Struhl, 2004). Lastly, phosphorylated

Hog1 is imported into the nucleus by the karyopherin Nmd5 (Ferrigno et al, 1998). Once

inside the nucleus, Hog1 participates directly or indirectly in the regulation of gene expression

of approximately 600 genes (O'Rourke and Herskowitz, 2004).

3.2.4 Hog1 dependent long-term osmoadaptation

Besides the rapid Hog1 dependent cytosolic changes, there is also evidence for a Hog1

dependent transcriptional program to survive and grow under the new environmental

conditions. Transcriptional changes are most pronounced 20 min after the shock, suggesting

that transcriptional regulation becomes relevant only 30 min after the osmotic shock. Thus,

transcriptional regulation plays a role in the long-term osmo-adaptation and in faster response

to consecutive stimuli (Mettetal et al, 2008). Aside from its minor contribution to instant

osmo-adaptation, the numerous changes in gene transcription occur within the adaptation

process. Interestingly, the induction patterns of osmo-responsive genes, are often similar and

transient in nature (Gasch et al, 2000; O'Rourke et al, 2004; Posas et al, 2000; Yale and

Bohnert, 2001). The maximum change is reached within 20 min; after that expression is

rapidly reverted to the basal level. O’Rourke and Herskowitz (2004) analyzed the yeast

transcriptome in a time span of 180 min after a hyperosmotic shock and uncovered a set of

2277 genes, differentially expressed during this period. Approximately 600 genes showed a

strong Hog1 dependent change in expression. The most significant changes in expression

were found in genes, which were either involved in carbohydrate metabolism, e.g. glycerol,

trehalose and glycogen metabolism, or in general stress response (Martinez-Montanes et al,

2010; O'Rourke et al, 2004; Posas et al, 2000). Figure 7 depicts an overview of the functional

gene families, which show changes in expression after a hyperosmotic shock.

In general, activated nuclear Hog1 associates with the osmo-responsive genes by specific

chromatin-bound proteins. The Hog1 kinase then recruits RNA polymerase II (Pol II) and

other transcription factors (Alepuz et al, 2003; Alepuz et al, 2001; Martinez-Montanes et al,

Chapter I 27

2010), required for transcription initiation. Additionally, recent findings showed that Hog1

remodels the chromatin structure at stress-responsive loci, resulting in higher levels of

transcription (Nadal-Ribelles et al, 2012). Furthermore, Hog1 behaves as a transcriptional

elongation factor and remains associated with Pol II to ensure a fast and proper mRNA

production during the transcriptional process (de Nadal and Posas, 2011; Proft et al, 2006).

The transcriptional re-programming regulated by Hog1 has been classified in three main

mechanisms, through which Hog1 controls different stages in the transcription cycle: A)

direct regulation of transcription factor or DNA bound proteins; B) stimulation of the

recruitment of Pol II at osmo-responsive promoters and C) chromatin modifying activities at

osmo-stress responsive genes (Figure 8).

Figure 7 Overview of the differential expression of functional gene families upon osmotic stress (adapted

from Martinez-Montanes et al. (2010)).

Once in the nucleus, active Hog1 targets directly several transcription factors (TF) and

regulates their activity by phosphorylation. A direct interaction with Hog1 has been shown for

several TFs, including the activators, Hot1 (Alepuz et al, 2003; Rep et al, 2000) and Smp1 (de

Nadal et al, 2003), or the repressor Sko1 (Proft et al, 2001; Proft and Struhl, 2002). The Hog1

dependent activation mechanism is well described for the repressor Sko1, which switches

from repression to activation upon osmo-stress (Figure 8A). Under basal conditions, this

28 Chapter I

repressor shuts down gene expression by the recruitment of the Tup1 - Cyc8 co-repressor

complex (Proft et al, 2002; Rep et al, 2001).

Figure 8 Mechanisms of Hog1 dependent transcriptional regulation upon osmostress. Hog1 modulates the

initiation of transcription by A) direct phosphorylation of promoter-specific transcription factors (TF), by B)

stimulating the recruitment of RNA Pol II and other required cofactors and by C) recruitment of the Rpd3

histone deacetylase complex (HDAC) (adapted from deNadal and Posas (2011)).

After phosphorylation by Hog1, Sko1 releases the Tup1 - Cyc8 complex and switches into an

activator, recruiting the ‘Spt - Ada - Gcn5 - acetyltransferase’ (SAGA) histone acetylase

complex and SWI/SNF chromatin modifying complexes (Proft et al, 2002). Apart from direct

targeting, Hog1 also associates with general stress-responsive transcription factors (Figure

8B), for example Msn2 and Msn4, which induce gene expression by the stress response

elements (STRE) (de Nadal and Posas, 2010; Martinez-Pastor et al, 1996). Although this

Hog1 dependent regulation mechanism is still unclear, both activators, Msn2 and Msn4, are

stimulated by the MAPK upon osmo-stress. Similar interactions with Hog1 were reported for

osmo-induced TF Hot1 (Rep et al, 1999). Lastly, Hog1 is involved in chromatin re-structuring

by the Rpd3 histone deacetylation complex (Figure 8C). The Rpd3 deacetylase complex is

recruited to the promoters by Hog1 and alters DNA accessibility to allow for the binding and

recruitment of other transcriptional activators thereby inducing osmo-responsive genes (de

Nadal et al, 2010).

Chapter I 29

3.3 The role of glycerol in Saccharomyces cerevisiae redox metabolism

3.3.1 Redox metabolism in Saccharomyces cerevisiae

Dissimilation of nutrients, e.g. carbohydrates generates energy and cell constituents required

for growth and cell proliferation. In general, the dissimilatory pathways (catabolism) involve

several steps of oxidative reactions, which transfer electrons from the nutrients itself or its

metabolized intermediates to an electron acceptor. In cells, final electron donors and acceptors

are often not involved in the same reaction and electrons must be transferred from one redox-

reaction to the next one. In many organisms, two main important electron transport carriers

exist, which are the coenzymes, nicotinamide adenine dinucleotide (NAD+) and nicotinamide

adenine dinucleotide phosphate (NADP+). First discovered by Harden and Young (1905) and

Warburg (1932), both coenzymes are limited in the cells and must be continuously cycled

between their oxidative NAD(P)+ and reduced NAD(P)H + H

+ state. In yeast, both redox

couples are functionally separated by their predominant role in either catabolism involving

mainly NAD(H) or anabolism involving mainly NADP(H) (van Dijken and Scheffers, 1986).

This separation is even more pronounced due to the lack of transhydrogenase activity

resulting in a chemical compartmentation of the two coenzymes (van Dijken et al, 1986). As a

consequence of this chemical compartmentation, a small part of the sugar is metabolised via

the hexose mono-phosphate pathway, also referred as the pentose phosphate pathway (PPP),

to generate the reducing power NADPH for the anabolic processes, such as synthesis of

amino acids or fatty acids. Besides the role in anabolism, NADP(H) plays an essential role in

yeast’s resistance to oxidative stress. The NADPH dependent glutathione reductase requires

this coenzyme to regenerate the oxidized tri-peptide glutathione (γ - L - glutamyl - L -

cystinyl - glycine) as the main reducing agent, which protects yeast cells against oxidants, free

radicals, and alkylating agents (Boles et al, 1993). NADP(H) metabolism is not further

described in this literature review due to its minor role in glycerol metabolism. The next

section is focused solely on the coenzyme NAD(H), with regards to its cellular metabolism

and compartmentation.

30 Chapter I

3.3.2 Cellular metabolism and compartmentation of NAD(H)

NAD(H) is predominantly used in redox reactions of the central catabolism, comprising

glycolysis and the tricarboxylic acid (TCA/ Krebs) cycle, (Bakker et al, 2001; van Dijken et

al, 1986). Aside from its metabolic function, NAD(H) has also an important role in

transcriptional regulation or cellular aging processes, which are not further addressed (Lin and

Guarente, 2003).

Continuous circulation of the coenzyme steadily maintains the concentrations of both, the

oxidized NAD+ and the reduced NADH, in order to preserve a proper redox state within the

cell. Moreover, the coenzyme is cycled in each cellular compartment where it was generated

due to the impermeability of membranes for pyridine nucleotides. The total ratio of

NAD:NADH is approximately 3 in yeast (Anderson et al, 2002), including the free and bound

forms of the coenzyme. The ratio of the free form, also referred as ‘cytosolic pool’ is

classically inferred from the cell’s redox potential resulting in a NAD:NADH ratio of 500

(Lin et al, 2003; Williamson et al, 1967). However, the reported numbers for the cytosolic

pool diverge considerably (Lin et al, 2003). In the cytosol, the reduction of NAD+ is mainly

catalyzed by the glyceraldehyde 3-phosphate dehydrogenase. The mitochondrial NADH

originates from the reactions catalyzed by pyruvate dehydrogenase, isocitrate dehydrogenase

and α-ketoglutarate dehydrogenase. The generated NADH is re-oxidized via five main

mechanisms existing in S. cerevisiae, which are: i) ethanol production by the alcohol

dehydrogenase, Adh; ii) glycerol production by cytosolic glycerol 3-phosphate dehydrogenase

Gpd1/2; iii) respiration of cytosolic NADH via the external mitochondrial NADH

dehydrogenases, Nde1/2; iv) respiration of cytosolic NADH via the glycerol 3-phosphate

shuttle; and v) oxidation of intramitochondrial NADH via a mitochondrial ‘internal’ NADH

dehydrogenase, Ndi1 (Bakker et al, 2001). Depending on the carbon source, NADH turnover

in the cytosol and mitochondrion is different. In order to establish the redox balance between

these two compartments, two main electron transport mechanisms are known, which

circumvent the impermeability of the inner mitochondrial membrane for the NAD(H)

cofactors. NADH transfer is mediated via the glycerol 3-phosphate or ethanol - acetaldehyde

redox-shuttle (Figure 9). Electrons are shuttled through the mitochondrial membrane in the

form of reduced metabolites, L-G3P or ethanol, which are able to pass the membrane. Both

shuttle mechanisms are involved in electron transfer from cytosolic NADH to the respiratory

chain.

Chapter I 31

Figure 9 Mechanism of NADH oxidation and transport in Saccharomyces cerevisiae. There are at least five

mechanisms for NADH oxidation: i) ethanol production by the Adh, alcohol dehydrogenase; ii) glycerol

production by cytosolic glycerol 3-phosphate dehydrogenase Gpd1/2; iii) respiration of cytosolic NADH via the

external mitochondrial NADH dehydrogenases, Nde1/2; iv) respiration of cytosolic NADH via the glycerol 3-

phosphate shuttle; and v) oxidation of intramitochondrial NADH via a mitochondrial ‘internal’ NADH

dehyrogenase, Ndi1; Cytosolic NADH can be transferred to the mitochondrial matrix via the glycerol 3-

phosphate or ethanol-acetaldehyde shuttle. The respiratory chain of S. cerevisiae is schematically indicated by

bc1 (bc1 complex); cox (cytochrome c oxidase); Q (ubiquinone) (adapted from Bakker et al. (2001)).

During growth, the formation of biomass results in net NADH production, which is re-

oxidized by generating reduced metabolites (Bakker et al, 2001). The ethanol production is

itself a redox-neutral process, and thus ethanol cannot serve as a redox sink for excess

NADH. A closed redox balance is realized by S. cerevisiae through i) the production of

glycerol in fermentative cultures and ii) the mitochondrial reduction of oxygen in respiring

cells. During the respiratory growth of S. cerevisiae, cytosolic NADH is mainly re-oxidized

by the respiratory chain. The external NADH dehydrogenase (Nde) and internal NADH

dehydrogenase (Ndi), both localized in the mitochondrial inner membrane, directly transfer

the reducing equivalents of their compartment to the respiratory chain. The glycerol 3-

phosphate shuttle is an alternative way for oxidizing excess cytosolic NADH. The shuttle

consists of the two metabolites DHAP and G3P. Both substances are transported through the

outer mitochondrial membrane. The cytosolic generation of G3P by GPDH consumes NADH.

32 Chapter I

After diffusion through the outer mitochondrial membrane, G3P is oxidized by the

membrane-bound glycerol 3-phosphate ubiquinone oxidoreductase, Gut2. The generated

DHAP returns to the cytosol, where it can be again reduced by GPDH activity.

An export of NADH from the mitochondria to the cytosol is necessary for growth in the

absence of oxygen, when the mechanisms of NADH re-oxidation by the respiratory chain do

not operate. Generally the excess NADH of biomass formation originates from amino acid

synthesis, which occurs mainly in the mitochondrial matrix. According to Valadi et al.

(2004), the addition of lysine and glutamine/glutamic acid to the culture medium restored

growth of respiratory-deficient yeasts lacking GPD2, which codes for Gpd2, under aerobic

conditions. Moreover, the amino acid synthesis occurs mainly in the mitochondria, and half of

the NADH generated during amino acid synthesis is produced in the mitochondrial matrix

(Valadi et al, 2004). Bakker et al. (2001) proposed a possible involvement of the second

shuttle, the ethanol-acetaldehyde shuttle, in order to enable re-oxidation of mitochondrial

NADH in anaerobically grown cultures. Supposedly, this shuttle plays an important role in

the export of intra-mitochondrial NADH. The shuttle involves three alcohol dehydrogenases,

Adh. One of them, located in the mitochondrial matrix, reduces acetaldehyde to ethanol by

the consumption of NADH. The inverse reaction appears in the cytosol and is linked to

NADH generation. The re-oxidation of NADH takes place through the Gpd2 catalysed

reduction of DHAP to glycerol 3-phosphate. The shuttle is driven by the NADH/NAD+ ratio

gradient over the inner mitochondrial membrane (Valadi et al, 2004). Gpd1 is also able to

drive this shuttle instead of Gpd2, although the NADH/NAD+ gradient across the

mitochondrial membrane might be increased - due to the peroxisomal localization of Gpd1

(Valadi et al, 2004). In respiro-fermentative cultures, all mechanisms of NADH re-oxidation

described above are possible; whereas in the absence of oxygen the formation of glycerol is

the only possibility left to re-oxidize excess NADH as already described by Neuberg and

Reinfurth (1918). In S. cerevisiae, glycerol production is closely linked to the maintenance of

the cell’s redox-state. Neuberg’s approach enhanced glycerol formation of S. cerevisiae by

adding sulfite to the medium, which trapped acetaldehyde. Ethanol formation, which

consumes glycolytic NADH, was therefore impaired, lacking the intermediate acetaldehyde;

consequently the re-oxidation of the cofactor was managed through glycerol production.

Chapter I 33

3.4 Glycerol yield as complex trait in yeast

Glycerol formation or dissimilation is required for many cellular processes, as exemplified for

osmo-adaption and redox-balancing in the previous chapters. These cellular processes involve

several hundred cellular reactions, enzymes and their encoding genes; hence glycerol

formation shows by itself a particular variation caused by its polygenic nature. Generally, the

amount of glycerol formed in a population of S. cerevisiae subspecies varies continuously a

low yield of 0.03 gram per gram consumed glucose to high yield of 0.09 gram per gram

consumed glucose (Hubmann et al, 2013) and can take any measurable value between these

two extremes. Frequently found glycerol phenotypes in natural yeast populations show

glycerol formation between those two extremes of either very high or very low, resulting in a

Gaussian distributed trait (Hubmann et al, 2013). Such a phenotypic appearance, i.e.

continuous and normally distributed, distinctly typifies glycerol formation as a quantitative

trait (QT) (Swinnen et al, 2012b). In yeast strains, genetic as well as environmental variations

determine the amount of formed glycerol (Albers et al, 1996; Bideaux et al, 2006). Apart

from the environmental variation; genetic variation results from the sum of many small

effects of genetic differences (Brookfield, 2012). This polygenic nature caused a complex

pattern of inheritance, usually divergent from Mendelian inherited traits, which makes its

characterization difficult. Apart from the characterization, this polygenic nature is

advantageous for the modification of glycerol formation, which is possible and has been done

in several ways to either increase or decrease the amount of glycerol produced by yeast during

growth or alcoholic fermentation.

34 Chapter I

4. Engineering of Saccharomyces cerevisiae for ethanol fermentation

4.1 Metabolic engineering of Saccharomyces cerevisiae – importance and strategies

4.1.1 Saccharomyces cerevisiae, a key ‘cell factory of the future’

The yeast S. cerevisiae has become a key microorganism in the biotech industry. Beside its

traditional role for ethanol production, this yeast has developed in recent years into an

important ‘cell factory’ for the production of other bulk or fine chemicals, food ingredients

and pharmaceuticals (Hong and Nielsen, 2012). Obviously, there are good reasons why this

yeast has been selected multiple times as the ‘microbial host of choice’ (Fischer et al, 2008).

First, its physiology, genetics and biochemistry have been comprehensively investigated and

described by generations of researchers for well over a century. Today, this great amount of

accumulated knowledge of yeast as a cellular system allows us to engineer this

microorganism in manifold ways, adapting and improving it more and more for its industrial

purpose. Secondly, its outstanding genetic competence to take up and accept foreign DNA

facilitates targeted manipulations within genes or chromosomes. Besides that, the availability

of an almost fully annotated reference-genome (Goffeau et al, 1996) and the capability of

homologous recombination allows for fast and efficient strain development. Finally,

manipulation of the genotype is accomplished in manifold ways due to the availability of a

large genetic tool set for engineering, comprising several types of expression vectors, reporter

genes and selectable or counter-selectable markers for highly efficient transformation

(Nevoigt, 2008). The cellular responses to genetic modification can now be studied

comprehensively due to the availability of ‘omics’ platforms and other system biology

approaches (Fischer et al, 2008; Nevoigt, 2008). In the last decade, key challenges in

engineering yeasts for biofuel production were the extension of substrate utilization (Fischer

et al, 2008; Hong et al, 2012), the optimization of product formation and the improvement of

robustness and stress tolerance (Fischer et al, 2008; Stephanopoulos, 2007) in order to create

tailor-made Saccharomyces cerevisiae strains perfectly suited for various industrial

applications.

Chapter I 35

4.1.2 Novel strategies and tools for engineering of Saccharomyces cerevisae

In today’s bioethanol industry, many production strains originate from simple non-targeted

engineering strategies. However, such strategies often apply a mechanism to select a strain

with a desired phenotypic trait. Traits such as, for example reduced glycerol formation, are

not selectable. This excludes methodologies, such as evolutionary engineering or random

mutagenesis, which are based on a selection mechanism to sort out improved strains in a huge

population. However, an advantage of those selection-based strategies is the independency of

a priori knowledge on the genetics or on the physiology of the microorganism; hence, such

non-targeted approaches are until today a key driver of microbial strain improvements (Oud et

al, 2012). Moreover, many industrially relevant traits of S. cerevisiae are far from being

completely elucidated, particularly with regards to complex traits such as glycerol formation.

Consequently, it is difficult to predict, how genetic modification affect the whole cellular

system. This is often the principal reason for the failure of many rationally deduced

approaches. Avoiding failure, successful engineering strategies are based on a fundamental

knowledge of the molecular mechanisms underlying the traits of interest, allowing the precise

prediction of cellular behaviour after the genetic modification. In case detailed knowledge on

the trait of interest in the microorganism is not available, one has to apply engineering

strategies, which aims to elucidate complexity, shedding light into the ‘black box’ of a

complex trait. Such an engineering concept, as Bailey et al. (1996) proposed, encompasses

two steps: the ‘black box’ system is first analyzed in order to elucidate the molecular

principles and subsequently it is reconstructed (Oud et al, 2012). Bailey and co-workers

(1996) specified their concept as ‘inverse metabolic engineering’. However, its principles are

based on the more commonly known discipline of reverse engineering. In the framework of

metabolic engineering, reverse approaches start with an existing microbial strain with superior

features relative to a reference strain. In this regard, the huge natural phenotypic diversity

allows finding a superior yeast strain or a superior species for almost every phenotype.

Biodiversity therefore constitutes a remarkable “treasure chest” for improving industrially

used microorganisms.

Phenotypic diversity originates in the first instance from genetic diversity. Thus, one has to

find the genetic basis of the superior feature and establish the genotype-phenotype

relationship; this analysis turned out to be the bottleneck of reverse engineering due to the

36 Chapter I

lack of genome-wide methods for detection of such genetic elements. However, its

attractiveness lies within the possibility of discovering novel genes and mutations involved in

the phenotype, which would never have been found by any rational method. Once this

unknown genetic basis has been identified, targeted genetic modifications can be used to

improve the industrial strain. Since only beneficial, naturally-occurring mutations are

transferred, reverse engineering approaches suffer less from side-effects, commonly present in

rational engineering or non-targeted strain improvement, such as mutagenesis or evolutionary

engineering (Oud et al, 2012).

For many years, resolving the complexity of the phenotype-genotype relationship restricted

reverse engineering approaches. However, recent advances in analytical technologies,

generally referred as ‘omics’ technologies, partially resolved this problem, allowing now for a

better and faster identification of genetic determinants over the whole genome. For instance,

the development of highly efficient DNA sequencing methodologies, referred as 2nd

generation sequencing technology, simplified genomic localization and identification of the

multiple mutant genes involved in complex traits (Ehrenreich et al, 2010; Parts et al, 2011;

Swinnen et al, 2012a). Locating or mapping such causative determinants is possible by

analyzing the inheritance of genomic regions in offspring of parents, where one of the two

exhibits the superior phenotype. As depicted in Figure 10, genetic mapping of causative

mutations in yeast requires two strains of opposite extreme phenotypes, i.e. one exhibiting the

superior trait (Trait+) and the other showing inferior behaviour (Trait

-). After crossing the two

parental strains, meiotic segregants are isolated, phenotyped and individually tested and

selected for the phenotype Trait+. In such a selected offspring population, the mutations,

which are causative for Trait+, are preferentially inherited from the superior parent and

therefore overrepresented in the selected population. The localization of these mutations is

possible by tracking molecular markers, such as single nucleotide polymorphisms, which

distinguish the Trait+ from the Trait

- parent. The availability of efficient sequencing methods

now allows for the rapid tracking of SNPs or other molecular markers over the whole genome

in individuals or pooled-segregant populations (Swinnen et al, 2012a; Swinnen et al, 2012b).

Scoring the parental markers in the selected offspring population will result in a higher

frequency of the markers in close proximity to the genetic elements important for the Trait+

phenotype. On the other hand, the frequency of a randomly selected population remains at

Chapter I 37

about 50% for the whole genome (Swinnen et al, 2012a; Swinnen et al, 2012b).

Chromosomal regions, which show a statistically significant deviation, are termed

quantitative trait loci. The locus itself is then further dissected to identify the causative genes

using for instance the method of reciprocal hemizygosity analysis (Steinmetz et al, 2002).

Once the gene or mutation has been localized and identified, it can be transferred to its

destination industrial strain, thereby establishing to a certain extent the desired trait. In this

way, genetic mapping will become very powerful for identification of genetic determinants in

reverse engineering approaches.

Figure 10 Quantitative trait locus (QTL) mapping in Saccharomyces cerevisiae. A quantitative trait (QT)

varies continuously in yeast populations between two extremes of either Trait+ or Trait-, and can take any

measurable value between these two extremes. QTL mapping aims to locate the genetic determinants in the

genome, using two yeast strains of opposite extreme phenotypes, i.e. one exhibiting Trait+ and the other Trait-. In

the offspring population of segregants, selected for the phenotype Trait+, molecular markers, such as single

nucleotide polymorphisms, which distinguish the Trait+ from the Trait- parent, are tracked, and a higher

frequency of Trait+ associated molecular markers will be observed. The SNP variant frequency of the selected

population significantly deviates in the chromosomal loci, which are linked to the Trait+ phenotype, whereas the

SNP variant frequency of a randomly selected population remains constant at about 50% for the whole genome

(adapted from Swinnen et al. (2012b)).

38 Chapter I

4.2 Previous approaches to optimized yields in ethanol production

In industrial ethanol production, yeast strains must show distinct key features to allow for a

competitive production process. These key features comprise the ability to produce ethanol in

a high titer, yield and at the fastest rate possible; additionally the yeast should tolerate a

number of stresses and, ideally, be able to ferment other sugars than hexoses (Fischer et al,

2008). The final ethanol titer is an important cost factor in the downstream purification

process, a low ethanol titer caused by growth inhibiting compounds or a low ethanol tolerance

of the yeast strain, will increase purification costs. Manufacturing as well as capital costs are

lowered by increasing ethanol productivity of the yeast. Conversion efficiency and yield are

important determinants for competitiveness, since the major portion of the total expenditure in

ethanol production is allotted to the cost of the fermentation feedstock. Thus, the ethanol

industry is always seeking for new yeast strains that could meet their standards, because even

small improvements in the ethanol titer, yield or productivity would have a large impact on

earnings and cost competitiveness of the whole production process (Basso et al, 2008;

Stephanopoulos, 2007).

Although S. cerevisiae is an efficient ethanol producer, it can still be improved in several

ways in order to increase the economic viability of the industrial ethanol production process.

The classical way, mutagenesis or breeding, often affects not only glycerol metabolism but

also other good traits of the production strains. Furthermore, shifting to a new production

strain is often difficult and is connected with changes in the entire process in order to find the

optimal conditions for the new production strains. Therefore, the aim of previous engineering

approaches as well as the current work is to develop novel strains based on existing industrial

yeast strains. Microbiological and metabolic engineering have leveraged the finding of

optimized fermentation conditions and suitable yeast strains for ethanol production. Rate

limiting steps have to be identified and abolished in order to improve the manufacturing

process and its economic viability. Thus, there is a strong interest of the ethanol producers in

using and in further developing novel strategies to engineer tailored strains (Stephanopoulos,

2007).

In ethanol production, a prominent engineered trait in yeast is the conversion efficiency,

which is the ratio of the product to the consumed substrate, which is usually referred as

ethanol yield. As Figure 11 reveals, the major products in ethanol fermentation are ethanol

Chapter I 39

and carbon dioxide (CO2). By-products are biomass, glycerol, and acetate. If the substrate

glucose is only converted into CO2 and ethanol, S. cerevisiae produces one mole of CO2 per

mole of ethanol; as a result the theoretical maximum ethanol yield is 51% (gram ethanol per

gram substrate). However, ethanol yields in practice are lower since some carbon is used for

the synthesis of cell components during yeast growth and the generation of fermentation by-

products such as glycerol and acetate. In fact, ethanol yield in industrial processes reaches

roughly 90-93% of the maximal theoretical value (Bai et al, 2008). One of the major by-

products of the alcoholic fermentation by S. cerevisiae is glycerol. Its synthesis has been

regarded as a wasteful process, because this metabolite is a fairly highly reduced fermentation

product not further converted to ethanol. The challenge herein is that glycerol has important

biological functions such as redox balancing and osmotic stress tolerance. Glycerol as by-

product in alcoholic fermentation is necessary for the growth of S. cerevisiae; however the

requirements might be lower than the normally produced level. Engineering glycerol

metabolism, one should consider the necessity of glycerol, as a redox sink for excess NADH,

which is generated during biomass synthesis, and as a compatible solute to enable osmotic

stress tolerance. Therefore, glycerol formation is important for the yeast cell and various

metabolic engineering strategies attempted to find compromises between higher ethanol

yields and the yeast’s minimal glycerol requirements.

Figure 11 Major metabolites produced in an ethanol fermentation process by Saccharomyces cerevisiae.

Anaerobic batch fermentation of CEN.PK 113-7D; 5% glucose was used as single carbon source. The amounts

of ethanol, CO2, glycerol, biomass and acetate are presented as part of the molar carbon fraction used from the

consumed glucose during anaerobic fermentation.

40 Chapter I

4.2.1 Optimization of Process Conditions

Various attempts were made to optimize the alcoholic fermentation process. Not only could

genetic modification of the yeast improve the overall yield of ethanol, but also the variation of

fermentation conditions was able to minimize the formation of by-product. In general,

glycerol formation can be influenced by the availability of oxygen and by the nitrogen source

present in the medium.

Aeration strategy. The effects of different aeration strategies in a fed-batch process were

studied extensively by Alfenore et al. (2004). Non-limited oxygen conditions enhanced

average productivity, viability, and the final ethanol titer. Glycerol formation was also

strongly reduced in comparison to limited oxygen conditions, and the limitation of oxygen

resulted in higher ethanol yields. In contrast to glycerol, the yield of other fermentation

products, in general biomass and acetate, was higher under non-limiting oxygen conditions.

Another approach aimed to minimize the glycerol production in a respiratory quotient (RQ)

controlled fed-batch process (Bideaux et al, 2006). The RQ value was maintained between

four and five times the ratio of CO2 production to O2 consumption, during the fermentation

process. The overall glycerol yield obtained by this aeration strategy was minimal in

comparison to micro-aerated or fully aerated processes (RQ = 1). The RQ-controlled process

could maintain growth and biomass formation on the one hand, but on the other hand, resulted

in a decreased ethanol yield and lower productivity of S. cerevisiae. Cell viability was also

impaired by oxygen limitation under micro-aerated or RQ-controlled processes, which led to

a decreased final ethanol titer.

Influence of the nitrogen source. A sufficient supply of nitrogen and other nutrient

supplements is essential for a rapid and complete fermentation process (Arshad et al, 2008;

Jones and Ingledew, 1994). Albers et al. (1996) studied the influence of the nitrogen source

on S. cerevisiae growth and product formation under anaerobic conditions. Three different

nitrogen sources - ammonium, glutamic acid, and a mixture of amino acids - were used to

examine the effect on glycerol formation and growth behaviour. Earlier studies noted that

high amino acid concentrations in the wort increased the fermentation rate of yeasts in

brewing processes. With the addition of the amino acid, in the form of a mixture or a single

amino acid (glutamic acid), Albers et al. (1996) could stimulate growth and biomass

Chapter I 41

formation. In comparison to ammonia-grown cultures; the glycerol yield decreased by 19%

when glutamic acid was used as the nitrogen source and by 50% with the use of a mixture of

amino acids. The ethanol yields, however, increased by 8% and by 14% respectively. The

different nitrogen sources did not influence biomass composition in terms of carbohydrate

storage, protein content, or protein composition. Interestingly, a preference for certain amino

acids, in particular glutamic acid, aspartic acid, and asparagine, was observed. These amino

acids were involved in the up-take of other amino acids and also preferred by the cells as

additional carbon sources (Albers et al, 1996).

4.2.2 Genetic optimization of metabolic pathways in Saccharomyces cerevisiae

Restructuring of metabolic networks by altering pathway or flux rates can improve the

production of a desirable metabolite or abolish the production of unwanted by-products

(Bailey, 1991). A young engineering discipline, which deals with such scientific issues, is

metabolic engineering. This engineering discipline aims to iteratively optimize product

formation using targeted genetic perturbation combined with cellular system analysis until the

final goal, i.e. the improved strain, is accomplished (Nevoigt, 2008). Traditionally, the places

of the perturbations in the metabolic network are ‘rationally’ deduced from available

information about the pathways, enzymes, and their regulation. In terms of optimized ethanol

production, different strategies were pursued to direct the glycolytic flux of S. cerevisiae

toward the production of ethanol. These strategies can be classified in three groups, which

are: i) the direct targeting of glycerol synthesis, ii) the change of cofactor utilization to

improve redox balancing, and iii) the co-fermentation of an oxidized substrate.

Directly targeting glycerol synthesis. The first approaches directly influenced the synthesis

or the transport of glycerol (Figure 12). Mutants, either deleted in the genes, GPD1 or GPD2,

and the gpd1∆ gpd2∆ mutant were described in several studies (Albertyn et al, 1994; Ansell

et al, 1997; Bjorkqvist et al, 1997; Nissen et al, 2000a; Valadi et al, 1998). The earlier studies

of Bjorkqvist et al. (1997) and Nissen et al. (2000a) examined the physiology of these S.

cerevisiae strains under aerobic and anaerobic conditions. Under aerobic conditions the

gpd1∆ and gpd2∆ single deletion strains behaved almost identical to their parental strain,

which grew at rates up to 0.5 h-1

. The growth behaviour of the gpd1∆ gpd2∆ mutant was

42 Chapter I

obviously impaired. The physiology of the single or double mutants diverged from the wild

type much more in the absence of oxygen. The wild type and the gpd1∆ mutant responded in

a similar way to the shift from aerobic to anaerobic conditions. Both strains adapted quickly

to the anoxic conditions and their fermentative capacity was only temporarily decreased. The

ethanol yield increased in both single mutants, gpd1∆ and gpd2∆, whereas the glycerol yield

correspondingly decreased. A more drastic consequence was observed for the gpd1∆ gpd2∆

mutant. Growth and ethanol formation were stopped under anaerobic conditions and could

only be restored by the addition of acetoin, which was used as an alternative redox sink for

excess NADH via production of butanediol. The results suggested that anaerobic conversion

of glucose to ethanol by S. cerevisiae without glycerol formation was not possible (Bjorkqvist

et al, 1997; Nissen et al, 2000a).

Figure 12 Metabolic engineering strategies directly targeting glycerol synthesis or transport. Glycerol

formation in S. cerevisiae can be decreased or even inhibited by modification of the structural pathway enzymes,

glycerol 3-phosphate dehyrogenase, encoded by GPD1 and GPD2, and glycerol 3-phosphatase, encoded by

GPP1 and GPP2, or the plasma membrane channel, encoded by FPS1, which facilitates glycerol efflux.

Chapter I 43

In anaerobic fermentations, GPDH activity was examined in addition to product formation

and growth behaviour of the gpd1∆ and gpd2∆ mutants. The specific GPDH activity, in cell

extracts of the gpd1∆ mutant, was lower than that measured in the gpd2∆ strain, which was

higher compared to the wild type. However the anaerobic growth behaviour of these strains

was not related to their specific GPDH activity.

Since the GPDH is the rate-limiting step in glycerol synthesis (Cronwright et al, 2002;

Remize et al, 2001), the consecutive second step catalyzed by the glycerol 3-phosphatase, was

less attractive for modification in the glycerol pathway. Only the gpp1∆ gpp2∆ mutant, which

is devoid of glycerol 3-phosphatase activity, produced a small amount of glycerol; whereas

the mutants lacking either GPP1 or GPP2 did not show any changes in glycerol formation

(Pahlman et al, 2001). Under anaerobic conditions, all gpp mutants showed a significant

increase in intracellular levels of glycerol 3-phosphate, which has been of interest for the

production of glycerol 3-phosphate (L-G3P) as a starting material for the enzymatic synthesis

of monosaccharides and glycerophospholipids (Nguyen et al, 2004). Interestingly, the

mutants engineered for high L-G3P production, produced glycerol at a later phase in the

fermentation, despite the complete elimination of glycerol 3-phosphatase activity. These

findings suggest that dephosphorylation of L-G3P also occurs by unspecific phosphatases

besides the main phosphatases Gpp1 and Gpp2 (Popp et al, 2008).

Zhang et al. (2007) investigated whether a change in the transport of glycerol across the

membrane, has an influence on glycerol formation. In fact, they observed a decrease in

glycerol and a concomitant increase in ethanol yield in fps1∆ mutants. The plasma membrane

channel Fps1 facilitates the efflux of glycerol across the plasma membrane during osmo-

adaptation (Luyten et al, 1995; Sutherland et al, 1997; Van Aelst et al, 1991). In the fps1∆

mutants, glycerol export is severely reduced and may only be possible via passive diffusion.

Probably, this enhanced the intracellular glycerol level, resulting in an inhibitory effect on its

own production.

Changing the cofactor in enzymatic reactions. Another way to reduce glycerol formation of

S. cerevisiae in alcoholic fermentation is to change the cofactor usage in redox reactions of

biomass synthetic pathways. Bro et al. (2006) tested different strategies of cofactor exchange

in S. cerevisiae’s redox metabolism for improved ethanol production. Their metabolic model

tested three main scenarios, which theoretically allowed for reduced glycerol and increased

44 Chapter I

ethanol due to a change in cofactor utilization. These scenarios are: i) substitution of

NADP(H) dependent reactions in biomass formation with NAD(H) dependent reactions, ii)

introduction of a reaction which either directly or via a cycle converts NADH into NADPH,

iii) substitution of glycerol production with production of ethanol, which occurs with a net

oxidation of NADH (Figure 13).

Nissen et al. (2000b) successfully confirmed the first scenario, e.g. the exchange of cofactors

in biomass synthesis, by changing the cofactor requirement in ammonium assimilation. The

approach pursued the goal of reducing surplus formation of NADH during biomass

production by using NADH for ammonium assimilation. Consequently, NADH accumulation

was less severe during anaerobic growth of S. cerevisiae. Since less NADH had to be

regenerated via glycerol synthesis in the engineered S. cerevisiae strains, these strains

produced less glycerol - despite their unaffected opportunity to synthesize the metabolite. As

depicted in Figure 13, cofactor utilization in ammonium assimilation can easily be changed

from NADPH to NADH by the over-expression of the pathway, using NADH as cofactor, and

deleting the enzymatic reaction requiring NADPH as cofactor. Ammonium assimilation is

mainly catalyzed by the NADPH dependent glutamate dehydrogenase (encoded by GDH1).

Another pathway, existing in S. cerevisiae, synthesizes glutamate in two coupled reactions,

which leads to a net consumption of NADH (glutamate synthase, encoded by GLT1) and ATP

(glutamine synthase, encoded by GLN1). Finally, minor accumulation of cytocolic NADH

resulted in a reduction of the glycerol yield by 38% and an increase of the ethanol yield by

10%. This was obtained in anaerobic batch fermentation in a minimal medium. In addition,

the maximum specific growth rate of the engineered strains was slightly lower than that of the

wild type (Nissen et al, 2000b). A disadvantage of Nissen’s strategy was that biomass growth

was necessary in order to provide ATP, which was required in addition to NADH for

ammonium assimilation. Furthermore industrial media often contain other nitrogen sources

than ammonia, and as long as there are amino acids present in the medium, amino acid

biosynthesis is largely downregulated.

In addition, Nissen and co-workers (Nissen et al, 2001; Nissen et al, 2000b) investigated the

feasibility of the second scenario by introducing a transhydrogenase system in a gpd1∆ gpd2∆

mutant, thereby converting excess NADH to NADPH. However, the expression of the

Chapter I 45

cytoplasmic transhydrogenase from Azotobacter vinelandii in the double mutant affected a

further decrease in µmax, because the transhydrogenase converted NADPH into NADH.

Therefore it was not possible to introduce an alternative pathway for re-oxidation of excess

NADH. In fact, yeast is devoid of any transhydrogenase activity. However, a cyclic

transhydrogenase system has been proposed to suppress the growth defect on glucose of

mutants lacking phosphoglucose isomerase (Boles et al, 1993). This artificial cyclic

transhydrogenase system was created by a substrate cycling between 2-oxoglutarate and

glutamate. Over-expression of NADH dependent glutamate dehydrogenase caused an

elevated oxidative deamination of glutamate to 2-oxoglutarate - accompanied by the reduction

of NAD+. 2-Oxoglutarate was aminated to glutamate by the NADPH dependent glutamate

dehydrogenase, whereby NADPH was consumed. This cyclic transhydrogenase system

oxidized NADPH to NADP+ under the generation of NADH, which can be used by other

NADH consuming systems (Boles et al, 1993).

The third scenario predicted the reduction of glycerol production by introducing reactions,

which change the redox neutral state of NADH in ethanol production to a net oxidation of

NADH. Bro et al. (2006) investigated the success of this scenario by over-expressing the

NADP+ dependent glyceraldehyde 3-phosphate dehydrogenase of Streptococcus mutans

(gapN). This enzyme partially substituted the NAD+ reducing two-step reaction from

glyceraldehyde 3-phosphate to 3-phosphogylcerate, which is normally catalyzed by the

NAD+-dependent glyceraldehyde 3-phosphate dehydrogenase, GAPDH, and the phospho-

glycerate kinase, PGK. The heterologous gapN uses NADP+ instead of NAD

+ as cofactor for

the redox reaction. Thus, the reaction contributes to reduce the cytosolic NADH formation,

consequently, less glycerol has to be produced to re-oxidize NADH. The in vivo evaluation of

this strategy resulted in a 40% decreased glycerol yield and a 3% increased ethanol yield. The

growth of the strain expressing gapN was unaffected under anaerobic conditions. Recently, a

similar approach described by Guo Zp and co-workers (2011) used the non-phosphorylating

NADP+-dependent glyceraldehyde 3-phosphate dehydrogenase from Bacillus cereus or

Kluyveromyces lactis. The expression of both gapN in a gpd2∆ mutant reduced glycerol yield

by half and concomitantly increased ethanol yield by 7% (Guo et al, 2011). Interestingly, the

maximum specific growth rate of the mutants expressing gapN increased compared to the

gpd2∆ mutant, and were indistinguishable from the wild type strain in anaerobic batch

fermentations.

46 Chapter I

Figure 13 Changing the NADH/NADPH cofactor imbalance. Glycerol formation decreases, if cytosolic

NADH excess is recycled via alternative routes. Three metabolic engineering strategies have been proposed: i)

substitution of NADP(H) dependent reactions in biomass formation with NAD(H) dependent reactions, ii)

introduction of a reaction which either directly or via a cycle converts NADH into NADPH, iii) substitution of

glycerol production with production of ethanol, which occurs with a net oxidation of NADH. The first strategy

was tested by Nissen et al. (2000b) with a changed cofactor utilization in ammonium assimilation. The following

enzymes were involved in the study: NADPH dependent glutamate dehydrogenase (GDH1); NADH dependent

glutamate dehydrogenase (GDH2); glutamate synthase (GLT1) and glutamine synthetase (GLN1). Additionally,

Nissen and co-workers (Nissen et al, 2001; Nissen et al, 2000a) tested the second strategy, over-expressing the

transhydrogenase of Azotobacter vinelandii in the gpd1∆ gpd2∆ mutant. Bro et al. (2006) expressed the NADP -

dependent glyceraldehyde 3-P dehyrogenase, gapN of Streptococcus mutans to replace the NADH dependent

glyceraldehyde 3-P dehyrogenase (Tdh 1/2 and 3) and phosphoglycerate kinase (Pgk), resulting in a net

oxidation of NADH in ethanol production.

Chapter I 47

Co-fermentation of oxidized substrates or production of alternative products. The third

scenario solves the NADH redox imbalance by co-fermentation of sugar together with

oxidized substrates. Alternatively, glycerol synthesis is replaced by formation of another

reduced product.

As already known from earlier studies on the glycerol pathway, the growth of the gpd1∆

gpd2∆ mutant was restored by the addition of acetoin, an intermediate of butanoate

metabolism, which is further reduced to butanediol. Acetoin served as an acceptor and redox

sink for excess NADH present in the gpd1∆ gpd2∆ mutant through its further reduction to

butanediol (Bjorkqvist et al, 1997). Several recent publications used this strategy to

completely eliminate glycerol formation and to improve ethanol yield towards the possible

theoretical maximum (Figure 14). Guadalupe Medina et al. (2010) introduced a heterologous

pathway to enable co-fermentation of glucose together with acetate, which is available in

significant amounts in starch or lignocellulosic hydrolysates. Generally, the acetate

dissimilation is subject to glucose repression and its consumption starts only after glucose

exhaustion; hence, acetate is usually not consumed during fermentation (Vilela-Moura et al,

2011). When cells are grown on acetate as a sole carbon source, acetate is mainly metabolized

to acetyl coenzyme-A (Co-A) by the acetyl-Co A synthetase, encoded by ACS1 and ACS2 and

then further oxidized in the tricarboxylic acid cycle. Besides CoA, the synthesis in S.

cerevisiae requires ATP. In contrast to S. cerevisiae, Escherichia coli catalyzes the same

reaction by the NAD+-dependent acetaldehyde dehydrogenase, mhpF. Therefore, expression

of mhpF in S. cerevisiae enables the re-oxidation of NADH through the reduction of acetic

acid during fermentation (Guadalupe Medina et al, 2010). This strategy allowed the complete

elimination of glycerol production via co-fermentation of acetate; however the osmo-

sensitivity of the gpd1∆ gpd2∆ mutant remains a problem. Guo et al. (2011) compared the

mhpF, as functional replacement for glycerol as a redox sink, with a reduction of fumarate,

catalyzed by the NAD+

dependent fumarate reductase, frdA of Escherichia coli. The

production of reduced metabolites, such as sorbitol or propane-1,2-diol, has been tested as a

functional glycerol alternative, restoring the redox balance in mutants deficient in glycerol

formation (Jain et al, 2011). As depicted in Figure 14, heterologous pathways were introduced

for sorbitol and 1,2 propanediol formation. The ethanol yields in this approach did not

48 Chapter I

improve, despite the absence of glycerol and the restored growth ability of the double mutant

under anaerobic conditions.

Figure 14 Co-fermentation of substrates or alternative reduced product formation. Glycerol formation

can be functionally replaced as a redox sink by the NADH dependent reduction of acetate or fumarate or by

producing other reduced metabolites, such as sorbitol or 1,2 propanediol.

Chapter I 49

5. Conclusions and scope of the present thesis

Various attempts using metabolic engineering have focused on reducing glycerol formation.

Many of these mostly rationally deduced approaches successfully reduced glycerol formation

and concomitantly increased ethanol yields. However, the resulting strains suffered from

unwanted side-effects, such as osmo-sensitivity or reduced ethanol productivity. Other

approaches, requiring the addition of co-fermented substrates, are not attractive due to their

difficult or costly implementation in a large-scale industrial process. For instance, the

production of alternative metabolites, such as sorbitol or 1,2 propanediol, reduced glycerol

formation but did not result in the desired increase in ethanol (Jain et al, 2011). Hence, such

approaches are not redirecting the carbon flux efficiently toward ethanol.

Glycerol metabolism and its regulation in yeast have been extensively studied for more than a

century. This showed that glycerol formation has essential functions in the yeast cell and

cannot be completely abolished. Improving yields in ethanol production, the major challenge

is to provide the yeast’s minimal requirements in glycerol synthesis. For optimal ethanol

yields, industrial strains should be modified and cultivated in a way that allows cells to

accumulate only the necessary amount of glycerol dependent on the environmental

conditions, like osmolarity or oxygen availability. An appropriately controlled glycerol

formation enables both, cellular integrity and the prevention of product losses through

excessive by-product formation.

The goal of the present work was to find genetic configurations, which allow for low glycerol

formation while maintaining proper cellular integrity. Rational engineering was applied in this

work in order to fine-tune glycerol formation by adjusting the activity of the rate-controlling

enzyme GPDH. The strains were afterwards analyzed in several fermentation setups. In

parallel, a reverse engineering approach was carried out in order to elucidate the genetic

configuration of a natural occurring low-glycerol producing S. cerevisiae strain. Mutations,

which were causative for the low glycerol yield, were identified throughout the whole yeast

genome. Four causative mutations were identified and can be used as novel gene tools to

reduce glycerol yield in an industrially important yeast strain. The methods and strategies

presented in this work clearly expand our present toolbox of yeast metabolic engineering and

the engineering strategies applied in this work can be extended to any other interesting

phenotype in S. cerevisiae.

Chapter II

Gpd1 and Gpd2 fine-tuning for sustainable

reduction of glycerol formation in Saccharomyces

cerevisiae

52 Chapter II

1. Abstract

Gpd1 and Gpd2 are the two isoforms of glycerol 3-phosphate dehydrogenase (GPDH), i.e. the

rate-controlling enzyme of glycerol formation in S. cerevisiae. The two isoenzymes play

crucial roles in osmoregulation and redox balancing. Past approaches to increase ethanol yield

at the cost of reduced glycerol yield have most often been based on deletion of either one or

two isogenes (GPD1 and GPD2). While single deletions of GPD1 or GPD2 reduced glycerol

formation only slightly, the gpd1 gpd2 double deletion strain produced zero glycerol but

showed an osmo-sensitive phenotype and abolished anaerobic growth. Our current approach

has aimed at generating "intermediate" phenotypes by reducing both isoenzyme activities

without abolishing them. To this end, the GPD1 promoter was replaced in a gpd2

background by two lower-strength TEF1 promoter mutants. In the same manner, the activity

of the GPD2 promoter was reduced in a gpd1∆ background. The resulting strains were

crossed to obtain different combinations of residual GPD1 and GPD2 expression levels.

Among our engineered strains we identified four candidates showing improved ethanol yields

compared to the wild type. In contrast to a gpd1 gpd2 double deletion strain these strains

were able to completely ferment the sugars under quasi-anaerobic conditions in both minimal

medium and during Simultaneous Saccharification and Fermentation (SSF) of liquefied wheat

mash (wheat liquefact). This result implies that our strains can tolerate the ethanol

concentration at the end of the wheat liquefact SSF (up to 90 g l-1

). Moreover, a few of these

strains showed no significant reduction in osmotic stress tolerance compared to the wild type.

Chapter II 53

2. Bibliographic reference

Pagliardini, J., Hubmann, G., Bideaux, C., Alfenore, S., Nevoigt, E. and Guillouet, S. E.

Quantitative evaluation of yeast's requirement for glycerol formation in very

high ethanol performance fed-batch process.

Microb. Cell Fact. May 2010 vol. 9 no. 36.

DOI: http://dx.doi.org/10.1186/1475-2859-9-36

Hubmann, G., Guillouet, S. and Nevoigt, E.

Gpd1 and Gpd2 fine-tuning for sustainable reduction of glycerol formation in

Saccharomyces cerevisiae.

Appl. Environ. Microbiol. September 2011 vol. 77 no. 17 5857-5867

DOI: http://dx.doi.org/10.1128/AEM.05338-11

European Patent

METHOD OF MODIFYING A YEAST CELL FOR THE PRODUCTION OF

ETHANOL

Publication number: EP2222860

Publication date: 2010-09-01

Inventor(s): NEVOIGT, E., GUILLOUET, S., BIDEAUX, C., ALFENORE,

S., HUBMANN, G.

Pagliardini, J ., Hubmann, G., Alfenore, S., Nevoigt, E., Bideaux, C., Guillouet, S.E.

The metabolic costs of improving ethanol yield by reducing glycerol formation

capacity under anaerobic conditions in Saccharomyces cerevisiae

Microb. Cell Fact. March 2013 vol. 12 no. 29.

DOI: http://dx.doi.org/10.1186/1475-2859-12-29

3. Scientific contribution

The author participated in the project conception, experimental work, data analysis and article

writing.

54 Chapter II

4. Manuscript I: Gpd1 and Gpd2 fine-tuning for sustainable reduction of

glycerol formation in Saccharomyces cerevisiae

4.1 Introduction

The major portion of total expenditure in today’s bioethanol industry is allotted to feedstock

costs (Galbe et al, 2007). Therefore, it has to be ensured that all polysaccharides present in the

raw material are efficiently converted into fermentable sugars plus that all sugars are

completely converted into ethanol with a high yield (ethanol per sugar consumed). This work

focuses on the question of how to improve the conversion yield (sugar to ethanol) in the yeast

S. cerevisiae.

The maximum theoretical ethanol yield of glucose fermentation by S. cerevisiae is 0.51 g (g

glucose)-1

. However, the ethanol yields in current industrial processes only reach 90 - 93% of

this maximal theoretical value (Bai et al, 2008). In fact, some carbon is used for the formation

of biomass and certain by-products particularly glycerol. As even small improvements in

ethanol yield would have significant impacts on profits in large-scale ethanol production,

there is a great industrial interest to reduce by-product formation and increase ethanol yield.

A promising route towards an increase in ethanol yield has been the reduction of glycerol

formation (Bro et al, 2006). Indeed, glycerol production in S. cerevisiae can be quite

substantial. For example, 2.0 - 3.6 g glycerol per 100 g consumed glucose have been already

reported by Pasteur (1858). Glycerol formation depends on yeast strain, fermentation

conditions (Alfenore et al, 2004; Bideaux et al, 2006; Gardner et al, 1993) and medium

composition, especially the type of nitrogen source (Albers et al, 1996). Therefore, glycerol

yields (per glucose consumed) reported for different strains and conditions vary and can even

significantly exceed the above mentioned value reported by Pasteur (1858).

Glycerol formation results primarily from balancing the net NADH surplus generated during

yeast growth, particularly under anaerobic conditions (Ansell et al, 1997; Bakker et al, 2001;

Rigoulet et al, 2004; Valadi et al, 2004). In addition, intracellular glycerol is involved in

osmo-adaptation (Hohmann, 2002), oxidative stress protection (Pahlman et al, 2001), and

response to heat shock (Aldiguier et al, 2004; Siderius et al, 2000).

A first group of past metabolic engineering approaches towards reduction of glycerol

formation in yeast targeted either enzymes (Bjorkqvist et al, 1997; Nissen et al, 2000a; Valadi

Chapter II 55

et al, 1998) and/or a plasma membrane transporter (Cao et al, 2007; Zhang et al, 2007) which

are directly involved in formation and intracellular accumulation of glycerol. A second type

of approaches attempted reducing net NADH production during yeast growth (Bro et al,

2006; Nissen et al, 2000b). Moreover, a number of studies have combined strategies of the

first and the second type (Cao et al, 2007; Kong et al, 2007a; Kong et al, 2006; Kong et al,

2007b). Recently, Guadalupe Medina et al. (2010) have opened up a third type of approaches,

namely by providing an alternative route of NADH re-oxidation. The additional value of this

latter approach is that acetic acid, an unwanted compound in plant biomass hydrolyzates is

converted into ethanol thereby serving as a redox sink and replacing glycerol formation.

However, the approach of Guadalupe Medina et al. (2010) also included a double deletion of

GPD1 and GPD2 and was thus accompanied by osmostress sensitivity and a strongly reduced

specific growth rate. A recently published route in order to overcome the osmosensitivity of

mutants defective in GPD1, has been the overexpression of trehalose synthesis genes TPS1

and TPS2 (Guo et al, 2011).

The glycerol 3-phosphate dehydrogenase (GPDH) is the rate-controlling enzyme in the

glycerol formation pathway of S. cerevisiae (Cronwright et al, 2002). Therefore, many

previous attempts of group one and several combined strategies included deleting either one

or both isogenes encoding GPDH. The phenotypes after deleting GPD1 and/or GPD2 have

already been tested in different yeast strain backgrounds, media and oxygenation conditions.

Generally speaking, the single deletion of GPD1 resulted in strains sensitive to osmotic stress

(Albertyn et al, 1994), while the deletion of GPD2 reduced growth under anaerobiosis

(Bjorkqvist et al, 1997; Nissen et al, 2000a). However, neither the deletion of GPD1 nor

GPD2 resulted in a noticeable change in glycerol yield at least when prototrophic strains were

studied (Nissen et al, 2000a). In contrast, the deletion of both isogenes led to complete loss of

glycerol formation under all tested conditions. However, the gpd1∆ gpd2∆ mutant is of no

practical relevance since growth and ethanol productivity was abolished under anaerobic

conditions and even strongly reduced under aerobic conditions (Ansell et al, 1997; Bjorkqvist

et al, 1997; Nissen et al, 2000a).

In order to study S. cerevisiae strains which have a glycerol formation capacity ranging

between that of the gpd2∆ single mutant (100%) and the gpd1∆ gpd2∆ double mutant (0%),

we recently replaced the native GPD1 promoter in a gpd2∆ background by two well-

characterized TEF1 promoter mutant versions (Nevoigt et al, 2006; Pagliardini et al, 2010).

56 Chapter II

The genetic modifications were accompanied by 61% and 88% reduction in glycerol yield on

glucose and by 20 and 30% reduction in maximal aerobic growth rate compared to the wild

type. Interestingly, our engineered (“intermediate”) strains referred to as TEFmut2 and

TEFmut7 showed a 2 and 5% increase in ethanol yield and could well cope with process

stress which is in remarkable contrast to a gpd1Δ gpd2Δ double deletion strain. These results

were obtained in a Very High Ethanol Performance (VHEP) fed-batch process with aeration

(Pagliardini et al, 2010). It remained to be tested, how these strains will perform under

anaerobic conditions. This was the first motivation for our current work. A second motivation

resulted from the fact that Gpd2 activity was completely abolished in the “intermediate”

strains described by Pagliardini et al. (2010) and the residual GPDH activity was solely based

on the Gpd1 isoenzyme. As mentioned above, Gpd2 plays a crucial role in redox balancing

under anaerobiosis. In fact, industrial bioethanol production usually occurs under anaerobic

conditions. In the current work, two isogenic strains with abolished Gpd1 activity and two

residual levels of Gpd2 activity were constructed. These novel strains have the opposite

mating type compared to the strains described by Pagliardini et al. (2010) and allowed to

eventually obtain all possible combinations of Gpd1 and Gpd2 residual activities by mating.

Our set of 12 different strains formed a sound basis in order to study the impact of partial

reduction of Gpd1 and Gpd2 activity on physiology under laboratory conditions as well as

conditions relevant in bioethanol industry.

4.2 Materials and Methods

Microbial strains and cultivation conditions

The Escherichia coli strain DH5αTM

(Invitrogen Corp., Carlsbad) was used for amplification

of plasmids. The strain was grown in Luria-Bertani (LB) medium (0.5 % yeast extract, 1 %

peptone, 1% NaCl, pH 7) at 37˚C. E. coli transformation and isolation of plasmid DNA was

carried out using standard techniques (Sambrook et al, 1989). All engineered Saccharomyces

cerevisiae strains (Table 1) generated in this study have been derived from the prototrophic

haploid wild-type strains CEN.PK 113-7D (Mat a) and CEN.PK 113-1A (Mat α). The

medium used for yeast strain maintenance was YD, which contained 1% (v/w) yeast extract

and 1% (v/w) glucose.

Chapter II 57

Table 1 Saccharomyces cerevisiae strains used.

Group Strain Code Genotype Source

reference

I

CEN.PK 113-7D A B MATa van Dijken

et al. (2000) CEN.PK 113-1A A B MATα provided by

P. Kötter IA 1

☐ a B gpd1∆::loxP-bleR-loxP This study

CE 1* A b gpd2∆::loxP-bleR-loxP This study

CE 1.01* a b gpd1∆::loxP-KanMX4-loxP

gpd2∆::loxP-bleR-loxP

This study

II

CE 1.71* a7 b gpd2∆::loxP-bleR-loxP

GPD1p∆::loxP-KanMX4-loxP-TEF1p mutant 7

Pagliardini

et al. (2010)

CE 1.21* a2b gpd2∆::loxP-bleR-loxP

GPD1p∆::loxP-KanMX4-loxP-TEF1p mutant 2

Pagliardini

et al. (2010)

IA 1.743☐ a b7 gpd1∆::loxP-bleR-loxP

GPD2p∆::loxP-KanMX4-loxP-TEF1p mutant 7

This study

IA 1.232☐ a b2 gpd1∆::loxP-bleR-loxP

GPD2p∆::loxP-KanMX4-loxP-TEF1p mutant 2

This study

III

HGW 12-P4▲ a7 b7 GPD1p∆::loxP-KanMX4-loxP-TEF1p mutant 7

GPD2p∆::loxP-KanMX4-loxP-TEF1p mutant 7

This study

HGW 15-F6▲ a7 b2 GPD1p∆::loxP-KanMX4-loxP-TEF1p mutant 7

GPD2p∆::loxP-KanMX4-loxP-TEF1p mutant 2

This study

HGW 14-J6▲ a2 b7 GPD1p∆::loxP-KanMX4-loxP-TEF1p mutant 2

GPD2p∆::loxP-KanMX4-loxP-TEF1p mutant 7

This study

HGW 17-H5▲ a2 b2 GPD1p∆::loxP-KanMX4-loxP-TEF1p mutant 2

GPD2p∆::loxP-KanMX4-loxP-TEF1p mutant 2

This study

* Strains derived from CEN.PK 113-7D

☐ Strains derived from CEN.PK 113-1A

▲ Strains isolated after mating of * strains with ☐ strains of group II and sporulation

Abbreviations for strain codes:

A GPD1 controlled by the natural promoter B GPD2 controlled by the natural promoter

a gpd1∆ b gpd2∆

a7 GPD1 controlled by TEFmut7 promoter b7 GPD2 controlled by TEFmut7 promoter

a2 GPD1 controlled by TEFmut2 promoter b2 GPD2 controlled by TEFmut2 promoter

58 Chapter II

Yeast strain construction

Genetic modifications of S. cerevisiae carried out within this study comprised the deletion of

GPD1 and the replacement of the native GPD2 promoter by the two low-activity promoters,

TEFmut2 and TEFmut7. Both promoters are mutated version of the constitutive TEF1

promtor and were characterized and published by Nevoigt and co-workers (2006). Deletion of

GPD1 was carried out using the method described by Güldener et al. (Gueldener et al, 2002).

The primers for the amplification of the GPD1 disruption and promoter replacement cassettes

as well as for the verification of the correct integration are listed in Table 2.

Table 2 PCR primers and plasmids used as templates for amplification of GPD1 disruption and GPD2

promoter replacement cassettes and for verification of their correct integration in the S. cerevisiae genome.

PCR application Primer

code Sequence*

Amplification of GPD1 deletion

cassette; templates: pUG6 or pUG66

(Gueldener et al, 2002)

P56 CATCAAATCTATCCAACCTAATTCGCACGTAGACTGGCTTGGTAT

cagctgaagcttcgtacgc

P57 CGACGTCCTTGCCCTCGCCTCTGAAATCCTTTGGAATGTGGTAAG

gcataggccactagtggatctg

Verification of GPD1 deletion

P58 CCGCACAACAAGTATCAGA

P59 AAGTAAGGTCTGTGGAACAA

GPD2 promoter replaced byTEFmut7;

template:p416-loxP-KmR-TEFmut7-

yECitrine (Nevoigt et al, 2006)

A2631 TAGCTTACGGACCTATTGCCATTGTTATTCCGATTAATCTATTGT

cagctgaagcttcgtacgc

A2633 TGCGTTCGCTTAAGGAATGTGTATCTTGTTAATCTTCTGACAGCAAGCA

T ttttctagaaaactag

GPD2 promoter replaced byTEFmut2;

template: p416-loxP-KmR-TEFmut2-

yECitrine (Nevoigt et al, 2006)

A2631 See above

A2632 TGCGTTCGCTTAAGGAATGTGTATCTTGTTAATCTTCTGACAGCAAGCA

T ttttctagaaaacttgg

Verification of GPD2 promoter

replacement

A2766 ACGACGATGGCTCTGCCATTG

A2767 AGGATCGGCCACTAGATTATGG

* Nucleotides in capital letters were derived from S. cerevisiae genomic sequence for homologous recombinations at GPD1 and GPD2 locus.

Sequences in low chase letters served as primers for amplification of the deletion or promoter cassettes from respective plasmids (Gueldener

et al, 2002; Nevoigt et al, 2006).

Gene-disruption and promoter-replacement cassettes were amplified by DNA polymerases

with proofreading activity. The PCR reaction conditions were adapted to the guidelines of the

manufacturer. Transformation of S. cerevisiae was carried out according to Gietz and Schiestl

(1991) using treatment with lithium acetate and polyethylene glycol. After the transformation,

cells were incubated in YD for at least 4 h at 30 ˚C to allow expression of the antibiotic

Chapter II 59

resistance genes. In order to select positive yeast transformants, YD agar plates were

supplemented with 7.5 µg/ml phleomycin or 200 µg/ml geneticin G418.

The deletion of GPD1 in the gpd2∆ background (Ab) was carried out using the loxP-

KanMX4-loxP disruption cassette amplified from pUG6 plasmid (Gueldener et al, 2002). The

resulting strain is referred to as the gpd1∆ gpd2∆ double deletion strain (ab). The loxP-bleR-

loxP disruption cassette located on the pUG66 plasmid (Gueldener et al, 2002) was used to

delete GPD1 in Mat α background in order to construct strain aB. The correct integration of

the disruption cassettes into the GPD1 gene locus was verified by PCR (Table 2).

In order to replace the native GPD2 promoter by promoters of much lower activities, the

TEF1 promoter mutants 2 and 7 of our previously published promoter collection for fine-

tuning gene expression in yeast (Nevoigt et al, 2006) were used. Both promoters were located

on CEN/ARS plasmids, which contained the loxP-KanMX-loxP cassette upstream of the

TEF1 promoter mutant. The replacement of the native GPD2 promoter by the low-strength

promoters in gpd1∆ background (strain aB) was confirmed by PCR diagnosis using primers

listed in Table 2.

Mating, sporulation and tetrad analysis

In order to obtain all combinations of promoter engineered GPD1 and GPD2 expression

levels, the strains a7b and a2b (both MATa) described by Pagliardini et al. (2010) were mated

with the newly created strains ab7 and ab2 (both MATα). The diploids were grown over night

in YD medium, harvested and transferred to sporulation plates containing 1,5% agar and 1%

potassium acetate at pH 6. The plates were incubated at 25˚C for 5 days. Asci were pretreated

with 1000 U/ml lyticase (Sigma) to degrade the ascus walls before spore dissection. Dissected

spores were grown on YD medium and segregants afterwards replica plated to YD medium

containing either the antibiotic phleomycin or Geneticin (G418) with the above mentioned

concentrations in order to check the presence of the resistance markers KanMX and bleR. The

segregants which grew in the presence of both antibiotics were parental types (a7b, a2b, ab7

and ab2). Recombination types which had lost the geneticin resistance but were resistant to

phleomycin corresponded to the double deletion strain gpd1∆ gpd2∆ (ab). In recombination

types with resistance to geneticin and sensitivity against phleomycin both GPD1 and GPD2

promoters were replaced by the TEF1 promoter mutant versions TEFmut2 or TEFmut7 (a7b7,

60 Chapter II

a7b2, a2b7 and a2b2). The pre-selected segregants were checked by PCR for the presence of the

promoter cassettes or gene deletions. PCR primers for GPD1 promoter replacement and

GPD2 deletion have been described in Pagliardini et al. (2010), while those for GPD2

promoter replacement and GPD1 deletion are given in Table 2.

Measurement of specific GPDH activity

To determine the specific GPDH activity, yeast strains were grown in minimal medium

(Verduyn et al, 1992) containing 2% (w/v) glucose in shake flasks. The GPDH activity was

measured in exponentially growing cells (OD600 was roughly 1) according to a previously

described method (Gancedo et al, 1968; Nevoigt and Stahl, 1996). One unit of enzyme

activity (U) corresponds to the conversion of 1 µmol NADH to NAD+ per minute (Verduyn et

al, 1992).

Growth tests on solid medium

Cells were grown over night in YD medium. The OD600 was determined and adjusted with

0.85% NaCl solution to OD600 of 1. A 1:10 serial dilution of these cultures was prepared in

triplicate using 0.85% NaCl solution, and 5 l of each dilution was spotted on different solid

media. The minimal medium used for the spot test contained 1.7g l-1

Yeast Nitrogen Base

(Fisher Scientific), 5 g l-1

NH4SO4 and 2% glucose (w/v). For testing osmotolerance, the

glucose concentration was increased to 25% (w/v). Growth was monitored for 2 to 3 days at

30˚C. Anaerobic incubation of agar plates was carried out in an anaerobic jar. The anaerobic

environment was generated by Anaerocult®

A (MERCK, Darmstadt) and controlled with

indicator stripes.

Batch-fermentations under quasi-anaerobic conditions in both minimal and industrial

medium (wheat liquefact)

The fermentations were carried out in 250 ml Erlenmeyer flasks containing 150 ml of medium

inoculated with yeast cells adjusting an initial OD600 of 1. The inoculum was prepared in

minimal medium containing 2% (w/v) glucose. Flasks were equipped with air locks, ensuring

Chapter II 61

the exclusion of oxygen but allowing the release of CO2. This setup was chosen to imitate

industrial bioethanol production and can be considered to provide quasi-anaerobic conditions

for the major part of the fermentation, i.e. after a short initial phase during which cells

consumed residual oxygen. The fermentations were performed at 30˚C and cultures were

continuously mixed at 200 rpm using a magnetic stirrer.

The synthetic minimal medium used for the fermentation experiments was composed as

follows: (NH4)2SO4 3g l-1

, KH2PO4 3.51g l-1

, K2HPO4 3H2O 2.1 g l-1

, MgSO4 7H2O 0.74 g l-1

,

EDTA (disodium salt) 1 mg l-1

, CaCl2 2H2O 6 mg l-1

, ZnSO4 7H2O 9 mg l-1

, FeSO4 7H2O

6mg l-1

, H3BO3 2 mg l-1

, MnCl2 4H2O 1.23 mg l-1

, Na2MoO4 2H2O 0.6 mg l-1

, CoCl2 6H2O

0.8 mg l-1

, CuSO4 5H2O 0.5 mg l-1

, KI 0.2 mg l-1

, D-Biotin 0.05 mg l-1

, p-Aminobenzoic acid

1 mg l-1

, Nicotinic acid 1 mg l-1

, Ca pantothenate 8 mg l-1

, Pyridoxine HCl 5 mg l-1

, Thiamine

HCl 5 mg l-1

, m-Inositol 25 mg l-1

, and D-glucose 50 g l-1

.

The wheat liquefact (24.5% dry mass content) used in the SSF fermentation was kindly

provided by a local ethanol producer. The pH was adjusted to 4.5 with sulfuric acid. In order

to obtain wheat hydrolyzate from wheat liquefact, enzymatic starch hydrolysis was performed

using Spirizyme®

(Novozyme, Denmark). In order to perform the SSF of the wheat liquefact,

Spirizyme®

was added up to 0.05% (w/w) of the wheat liquefact dry mass. At the same time,

wheat liquefact was inoculated with yeast OD600 of 1. We added 5 mg l-1

of the antibiotic

chloramphenicol to inhibit bacterial growth in the first hours of the fermentation. All

Erlenmeyer flasks were weighed during fermentation to monitor CO2 production as readout

for the volumetric productivity.

In order to obtain the glucose concentration after the complete hydrolysis of the liquefact (for

product yield calculations), an aliquot was incubated at 60˚C for 24h with the Spirizyme®

(without yeast).

Determination of biomass in minimal medium fermentations

OD600 was determined at the beginning and at the end of the fermentation. In addition, the

yeast dry mass was determined at the end of fermentation by filtering 50 ml of culture through

pre-weighed nitrocellulose filters with a pore size of 0.45 µm. Filters were washed once with

distilled water and kept at 80˚C for two days. Afterwards, they were weighed again. The

62 Chapter II

biomass at the start of the minimal medium fermentation was estimated from OD600 using the

dry mass OD600/ratio obtained from samples taken at the end of fermentation.

Determination of ethanol, glycerol and glucose

Glucose, glycerol and ethanol concentrations in the fermentation supernatants were

determined by HPLC (Waters®

isocratic BreezeTM

HPLC, ion exchange column

WAT010290) and refractive index detection (Waters, 2414 RI detector) using the following

conditions: 75˚C column temperature, 5 mM H2SO4 used as eluent and 1 ml min-1

flow rate.

Yield calculations

The product yields in the minimal medium fermentations were calculated from the final

product concentration (g l-1

) and the difference in glucose concentrations at start and end of a

fermentation (consumed glucose in g l-1

). The product yields in the SSF were based on the

final product concentrations and the equivalent initial glucose concentration (the latter was

measured in a completely hydrolyzed sample of wheat liquefact as described above).

Maximal volumetric production rate

The fermentation tubes were weighed throughout the fermentation process. The loss of the

medium mass was recorded and related to the initial medium mass. The maximal volumetric

ethanol production rate rmax (g l-1

h-1

) was calculated from the monitoring of the mass loss

during anerobic fermentation as described by Zwietering et al. (1990).

4.3 Results

Genetic modifications of GPD1 and GPD2

The goal of the current work was to study the effect of partly reduced Gpd1 and Gpd2 cellular

levels (separately and in combinations) on physiology of S. cerevisiae. The CEN.PK

background was used as a model genetic background, since strains of the CEN.PK family

Chapter II 63

have been previously proposed as "acceptable references for quantitative yeast research"

(van Dijken et al, 2000). We exclusively used prototrophic strains in order to avoid any

potential artefacts based on auxotrophic markers (Pronk, 2002). The strains with several

residual activities of Gpd1p and Gpd2p in different combinations were obtained as described

in Material and Methods and depicted in Figure 1.

Figure 1 Schematic overview showing the generation of 11 S. cerevisiae strains with reduced levels of

Gpd1 and Gpd2 by genetic engineering and mating. The isogenic haploid and prototrophic wild-type strains

CEN.PK 113-7D and CEN.PK 113-1A were used as starting strains.

The resulting strains listed in Table 1 were classified into three groups. The group I contains

the wild type (AB), the gpd1∆ (aB) and gpd2∆ (Ab) single deletion as well as the gpd1∆

gpd2∆ double deletion strain (ab). These strains served as references in our study since these

genetic modifications correspond to those which have already been characterized by other

authors even though in a different genetic background (Albertyn et al, 1994; Ansell et al,

1997; Bjorkqvist et al, 1997; Guo et al, 2011; Nissen et al, 2000a; Remize et al, 2003; Valadi

et al, 1998). Strains of group II each have one deleted isogene and the other isogene under

control of either the TEFmut2 or the TEFmut7 promoter. These promoters are mutated

versions of the S. cerevisiae TEF1 promoter (Nevoigt et al, 2006), i.e. not regulated by

osmostress or anarobiosis and weaker than the native promoters of GPD1 and GPD2. Group

III comprises strains having residual levels of both isogenes, i.e. the expression of both GPD1

and GPD2 gene is de- and downregulated by the TEFmut2 or the TEFmut7 promoter.

64 Chapter II

Figure 2 Specific GPDH activity and glycerol yields of S. cerevisiae wild type and engineered strains with

reduced levels of Gpd1 and Gpd2. (A) GPDH activity measured in yeast cell extracts of aerobically grown cells

harvested during the exponential growth phase. (B) Glycerol yield obtained with the strains in synthetic minimal

medium under quasi-anaerobic conditions are plotted against the specific GPDH activity shown in Figure 3A

(explanation in the text). Strains are indicated as follows: ▽ for wild type CEN.PK 113-7D (AB), ◆ for gpd1∆

(aB), ● for gpd2∆ (Ab), ▲ for gpd1∆ gpd2∆ (ab), and for all intermediate strains (ab2, ab7, a2b, a7b, a2b7,

a7b2, a2b2, a7b7). All results shown are mean values including standard deviations from three biological

replicates.

Specific GPDH activities of all engineered strains and the wild type (CEN.PK 113-7D) were

measured in exponentially growing cells from shake flask experiments (Figure 2). The

deletion of the GPD1 (strain aB) reduced GPDH activity to 26%; whereas the deletion of

GPD2 (strain Ab) led to an increase of GPDH activity by 12% when compared to the wild

type. The GPDH activity measured in the double deletion gpd1∆ gpd2∆ strain (ab) was close

to zero. These relative GPDH activities (% of wild-type activities) measured for group I

reference strains more or less matched the findings obtained by other authors when studying

the GPDH gene deletions in different genetic backgrounds (Nissen et al, 2000a; Valadi et al,

2004). The increased GPDH activity in the gpd2∆ deletion strain has been traced back to a

kind of autoregulation, i.e. an increased transcription level of GPD1 (Valadi et al, 2004). This

auto-regulation should be abolished in all engineered strains which do not have the native

GPD1 promoter. The main objective of the current study was to generate “intermediate”

strains, i.e. specific GPDH activities between the level of the wild type (AB) and the gpd1∆

Chapter II 65

single deletion strain (aB). When looking at the GPDH activities of the group II strains

(Figure 2A), it becomes obvious that only strain a7b represents such an “intermediate” strain

while the levels of the other three strains are equal or lower compared to strain aB. In

contrast, strains of group III show finely graded residual activities between those of AB and

aB.

Figure 3 Growth of S. cerevisiae wild type and engineered strains with reduced levels of Gpd1 and Gpd2

under conditions, which challenge glycerol formation, i.e. osmotic stress (synthetic minimal medium plus 25%

glucose) and NADH accumulation (synthetic minimal medium with 2% glucose under anaerobiosis). Minimal

medium under aerobiosis was used as reference conditions. Growth tests were performed on solid media. Cells

were pre-grown in YD. The OD600 of all cultures was adjusted to 1, serial 1:10 dilutions were prepared and 5 l

of each dilution was spotted.

Growth during osmotic stress or anaerobic conditions

As mentioned in the introduction, the formation of glycerol has several important biological

functions in S. cerevisiae. It enables growth under osmotic stress and its formation maintains

cytosolic NADH/NAD+ balance. Surplus cytosolic NADH is generated during biomass

formation, particularly when yeast grows in minimal medium and has to synthesize all amino

66 Chapter II

acids de novo (Albers et al, 1996). Moreover, maintaining cytosolic redox balance is

particularly important under anaerobiosis when cells cannot reoxidize the surplus NADH by

the external NADH dehydrogenases coupled to the respiratory chain (Bakker et al, 2001).

Therefore, we tested our set of strains under conditions, which particularly challenge glycerol

formation such as minimal medium in presence and absence of oxygen as well as minimal

medium with 25% (w/v) glucose (to test osmotolerance). In the presence of 25% glucose,

growth on solid medium was unaffected as long as GPD1 expression was driven by the native

GPD1 promoter. Interestingly, 25% glucose did also not negatively influence the growth if

GPD1 expression was controlled by the TEFmut7 promoter which has been shown to confer

16% of native TEF1 promoter activity (Nevoigt et al, 2006). The TEFmut7 promoter is

assumed to be much less active than the native GPD1 promoter particularly during osmostress

where the native GPD1 promoter is strongly induced. Growth in 25% glucose was however

severely decreased if GPD1 expression was either under control of the weak TEFmut2

promoter (7% of native TEF1 promoter) or completely abolished by gene deletion.

Without osmotic stress, virtually all strains grew fairly well on minimal medium in the

presence of oxygen, however the combination of minimal medium with anaerobiosis

negatively influenced the growth of the strains with reduced Gpd1 and/or Gpd2 activities

(Figure 3). It seems that the impact of a reduced GPDH activity on growth under anaerobic

conditions in minimal medium is independent of whether Gpd1 or Gpd2 activity was reduced.

For example, there was no significant difference in growth between the strains a7b and ab7 or

between the strains a7b2 and a2b7 under anaerobic conditions (Figure 3). Remarkably, these

pairs of strains showed clear differences in growth under osmotic stress, respectively.

In contrast to several previous studies, which showed that mutants lacking GPD2 showed

poor growth under anaerobic conditions (Bjorkqvist et al, 1997; Nissen et al, 2000a), our

gpd2∆ deletion strain (Ab) did not show any remarkable growth problems under anoxic

conditions in minimal medium (Figure 3; Table 3). Moreover, expression of GPD2 has been

shown to be stimulated by cytosolic NADH accumulation as it occurs e.g. during anoxic

conditions in minimal medium (Ansell et al, 1997). In fact, this upregulation was impossible

in strains where GPD2 was under control of the TEFmut promoters. Our results put an

isoform-specific role of Gpd2 under anaerobic conditions somewhat into question. It should,

however, be emphasized that a growth test on solid medium does not allow a deduction of

Chapter II 67

specific growth rates and is less sensitive than liquid medium when it comes to the detection

of only slight differences in specific growth rates.

Fermentations in defined minimal medium

We first chose defined minimal medium in order to investigate the fermentation performance

of our strains with different GPD1/2 expression levels, particularly their glycerol and ethanol

yields as well as their maximal volumetric ethanol production rates. One advantage of using

synthetic minimal medium is a generally high demand in NAD+ regeneration particularly

under anaerobic conditions as explained above. In fact, slight differences in glycerol

production capacity between the 12 strains were expected to be more pronounced in minimal

medium than in industrially used media which are complex and contain amino acids (see

below). Another major advantage of using defined minimal medium is the possibility to

calculate precise carbon balances.

Table 3 Fermentation time, product yields and carbon balances of engineered strains of S. cerevisiae with

reduced levels of Gpd1 and Gpd2 as well as the isogenic wild type calculated at the end of fermentations in

synthetic minimal medium under quasi-anaerobic conditions.

Gen

oty

pe

Tim

e (h

) Yields of fermentation products

(C-mol ratio product per glucose consumed)

Ca

rbo

n

reco

ver

y

Ethanol Carbon

dioxide1 Glycerol Acetate Biomass

Gro

up

I

A B 25.8 0.586 0.275 0.058 0.008 0.064 99.1%

a B 28.5 0.595 0.276 0.050 0.006 0.075 100.2%

A b 27.1 0.595 0.279 0.046 0.005 0.070 99.4%

a b n.d.2 0.6302 0.3092 0.0052 0.0002 0.0442 98.8%2

Gro

up

II

a7 b 68.2 0.623 0.295 0.021 0.001 0.065 100.6%

a2 b 152.3 0.632 0.324 0.009 0.000 0.050 101.6%

a b7 75.3 0.615 0.288 0.022 0.001 0.064 99.0%

a b2 140.5 0.627 0.311 0.017 0.001 0.054 101.0%

Gro

up

III

a7 b7 51.0 0.613 0.288 0.032 0.002 0.060 99.3%

a7 b2 48.0 0.610 0.295 0.025 0.001 0.070 100.1%

a2 b7 54.2 0.621 0.287 0.023 0.001 0.067 99.9%

a2 b2 110.0 0.622 0.301 0.019 0.000 0.062 100.5%

1Carbon dioxide production was calculated from the total weight loss of the medium during fermentation.

68 Chapter II

2 The gpd1∆ gpd2∆ double deletion was not able to completely consume the sugar.

The small-scale batch fermentations in synthetic minimal medium were carried out under

quasi-anaerobic conditions (see Material and Methods). Glucose concentration used for this

experiment was 50 g l-1

which was not expected to cause osmotic stress (Wang et al, 1979).

Apart from the double deletion gpd1∆ gpd2∆ strain, all strains were able to completely

consume the sugar even though the required time was different (Table 3). Product yields are

also shown in Table 3. Although obtained under different conditions, we plotted the glycerol

yield (obtained under quasi-anaerobic conditions) against the corresponding GPDH activity

measured in shake flasks. The decision to measure all GPDH activities under aerobic

conditions was due to comparable growth characteristics, which was not achievable under

anaerobic conditions. A fairly good correlation can be seen in Figure 3B, when disregarding

the wild type and gpd1∆ deletion strain. In these two strains, GPD2 expression is under

control of the native GPD2 promoter. This promoter has been shown to be induced by NADH

accumulation under quasi-anaerobic conditions (Ansell et al, 1997; Valadi et al, 2004) and

the GPDH activity measured under aerobic conditions clearly underestimated the actual value

under quasi-anaerobic conditions. As all other strains have a deregulated or no GPD2 gene,

we do not assume major differences between anaerobic and aerobic GPDH activity.

To evaluate the effect of reduced glycerol formation capacity on other fermentation

parameters, we decided to plot the relevant fermentation product yields against the glycerol

yield. Figure 4A shows that any reduction in glycerol yield leads to a corresponding increase

in the ethanol yield. This inverse correlation can be seen over the entire range of glycerol

yields. However, the metabolic shift from glycerol to ethanol is accompanied by a reduction

in the maximal volumetric ethanol production rate (g h-1

l-1

; calculation see Material and

Methods) as visible in Figure 4B. We can conclude that even a slight reduction in glycerol

formation capacity resulted in a decrease in growth rate and/or specific ethanol production

rate. This is in contrast to the biomass yield, where we can only see a reduction at glycerol

yields lower than 50% of the wild-type level (Figure 4C). Interestingly, acetate yields

exhibited a nice direct correlation with glycerol yield (Figure 4D), particularly in the range

between 50 and 100% of wild-type glycerol yield. Carbon balances were nicely closed for

virtually all strains (Table 3).

Chapter II 69

Figure 4 Ethanol yields (A), maximal volumetric ethanol production rates (B), biomass yields (C) and

acetate yields (D) of S. cerevisiae wild type and engineered strains with reduced levels of Gpd1 and Gpd2

obtained in synthetic minimal medium under quasi-anaerobic conditions were plotted against the corresponding

glycerol yield, respectively: () strains of group I, () group II and () group III . All results shown are mean

values including standard deviations from two biological replicates. Yields were calculated from the final

product concentration divided by the consumed glucose. For calculating the maximal volumetric ethanol

production rate, the flasks were weighed throughout the fermentation process. The loss of the medium mass was

recorded. It was assumed that the mass loss of the medium is only due to production CO2 plus ethanol is

produced in equimolar amounts. The maximal volumetric ethanol production rate rmax [g l-1 h-1] resulted by

applying the mass balance on the fermentation system (see Material and Methods).

SSF in wheat liquefact

It was of interest to investigate our set of strains in media/conditions relevant to industrial

practice. Hydrolyzed starch from corn and other small grains is a common carbon source in

fuel ethanol production. For our study, wheat liquefact was provided by a local bioethanol

70 Chapter II

producer. Wheat liquefact contains mainly dextrins and results from milling, cooking and α-

amylase treatment of wheat kernels (Kelsall and Lyons, 2003). In order to further convert the

dextrins into fermentable sugars, a treatment with glucoamylase is required. Two different

scenarios are common in industry (Kelsall et al, 2003): Separate Hydrolysis and Fermentation

(SHF) and Simultaneous Saccharification and Fermentation (SSF). In SHF, the glucoamylase

treatment of wheat liquefact is performed before the hydrolyzate is inoculated with yeast. The

sugar concentration at the end of the mashing process was 200 g l-1

and thus represents an

osmotic stress for the yeast cells. In the SSF scenario, the hydrolytic enzymes plus the yeast

are added to the liquefact at the same time, i.e. saccharification and fermentation occur

simultaneously. This SSF process which is nowadays more frequently used in industrial

practice does not allow an accumulation of sugars released from starch and thus avoids

osmotic stress. Based on our plate growth assays (Figure 3), we already knew that osmotic

stress was deleterious for the strains with very low GPD1 expression levels. Therefore, we

decided to investigate our strains in the SSF process. The SSF experiments in wheat liquefact

revealed that not all strains were able to completely ferment the available sugars in an

acceptable time. In fact, the wild type required 53 hours to finish fermentation and, apart from

the single deletion strains (aB and Ab), only four further strains (a7b, a7b7, a7b2 and a2b7)

completed the fermentation within a period no longer than 150 hours (Table 4).

The glycerol yield obtained with the wild type in the SSF was only half of the glycerol yield

in minimal medium (Figure 4 and Table 3) because of two reasons. Firstly, wheat mash

(hydrolyzate) is rich in nutrients and thus also contains amino acids even though assimilable

nitrogen in wheat hydrolyzate is not sufficient to support fermentation at the fastest rate

(Thomas and Ingledew, 1990). Secondly, it is a well known fact that the major part of

glycerol formation is coupled to growth (Boender et al, 2009). Glycerol yield per sugar

consumed becomes lower if there is a longer production phase after finishing cell growth.

This was the case in wheat liquefact since the total amount of glucose (180 g l-1

) was much

higher compared to the experiments in minimal medium (50 g l-1

). Due to the generally lower

glycerol yields in wheat liquefact, the differences between strains were, as expected, less

pronounced in the SSF compared to minimal medium. For example, the reduction of the

glycerol yield in the a7b strain was only 33% (Table 4) when compared to the wild type, while

the same strain showed a reduction by 63% (Table 3) in the fermentations in minimal

medium.

Chapter II 71

Table 4 Performance of engineered strains of S. cerevisiae with reduced levels of Gpd1 and Gpd2 as well

as the isogenic wild type simultaneous saccharification and fermentation (SSF) of wheat liquefact under quasi-

anaerobic conditions. Results shown are mean values of two independent experiments including standard

deviation. Ethanol yield, glycerol yield and maximal volumetric production rate was only calculated if

fermentation was completed within 150 hours or less. Fermentation was considered to be completed when the

mass of the medium did not change anymore.

Gen

oty

pe

Fermentation

time Yields on total substrate Max. volumetric

ethanol

production rate Glycerol Ethanol

[h] [g g-1] [g g-1] [g l-1h-1]

Gro

up

I

A B 53 0.027 ± 0.002 0.487 ± 0.004 4.8 ± 0.1

a B 59 0.028 ± 0.001 0.486 ± 0.001 4.5 ± 0.7

A b 64 0.024 ± 0.002 0.489 ± 0.001 3.9 ± 0.2

a b >150 n.d. n.d. n.d.

Gro

up

II

a7 b 107 0.018 ± 0.000 0.492 ± 0.003 2.1 ± 0.1

a2 b >150 n.d. n.d. n.d.

a b7 >150 n.d. n.d. n.d.

a b2 >150 n.d. n.d. n.d.

Gro

up

III

a7 b7 87 0.020 ± 0.001 0.490 ± 0.006 2.7 ± 0.2

a7 b2 99 0.019 ± 0.001 0.492 ± 0.003 2.5 ± 0.1

a2 b7 150 0.019 ± 0.001 0.491 ± 0.004 1.6 ± 0.1

a2 b2 >150 n.d. n.d. n.d.

Regarding the four “intermediate” strains which were able to completely ferment the available

sugars, the reduction in the glycerol yield compared to the wild type was 33.3% in strain a7b,

25.9% in strain a7b7, and 29.6% in strains a7b2 and a2b7. Consistent with this, these strains

also showed an increase in ethanol yield which was 1.0% in strain a7b, 0.6% in strain a7b7,

1.0% in strain a7b2 and 0.8% in strain a2b7. Although the increases in ethanol yield do not

seem to be significant (when looking at the standard deviations), we still believe that the

slight increases are true due to the following considerations. Firstly, we measured a slight

increase in the ethanol yield in all four strains considered here (i.e. a7 b, a7b7, a7b2 and a2b7).

72 Chapter II

Secondly, a reduction in glycerol yield by 30% compared to the wild type (from 0.027 g g-1

to

0.019 g g-1

) should be, in theory, accompanied by an increase in the ethanol yield by 0.84%

(from 0.487 g g-1

to 0.491 g g-1

) provided that all carbon is redirected from glycerol to

ethanol. However, we previously showed that the increase in the ethanol yield with the a7b

and a2b strains observed in aerobic ethanolic fermentation was due to redirection of the

carbon from not only glycerol but also biomass to ethanol (Pagliardini et al, 2010).

It has to be emphasized once more, that even slight increases in the ethanol yield are not

irrelevant and could have major impacts on profits of a bioethanol company. However, the

increases in ethanol yield in SSF also come along with a decrease in maximal volumetric

ethanol production rates accompanied by prolonged fermentation times (Table 4). For

example, the two strains which show a 1% increase in ethanol yield (strains a7b and a7b2),

show a maximal volumetric production rate which is decreased by about 56 or 44% compared

to the wild type, which means that fermentation time was roughly doubled for these strains

(Table 4).

4.4 Discussion

Even slight reductions of glycerol production (which result in ethanol yield increases without

impacting volumetric productivity) would be of great interest for the bioethanol industry due

to the fact that ethanol is a bulk product and feedstock is a major cost factor. The glycerol 3-

phosphate dehydrogenase (GPDH) is known to be the rate-controlling enzyme of glycerol

formation and the isogenes GPD1 and GPD2 encode for it. On one hand, many approaches to

reduce glycerol formation have included the deletion of one or both isogenes. On the other

hand, the two gene products have been shown to fulfil different important biological functions

in the cell. Thus, the complete deletion of one gene might have negative impacts in particular

when it comes to industrial bioethanol production subjecting the yeast to several stress

conditions (Kelsall et al, 2003). We wondered whether it is more sustainable for the cell if

isoenzyme activities are not completely shut down to zero but rather reduced to a certain

residual level. Here, several prototrophic strains were constructed which have different

expression levels of both isogenes (i.e. gene expression is driven by either the native

promoter, the TEFmut7 promoter, the TEFmut2 promoter or completely abolished by gene

Chapter II 73

deletion). Different combinations of GPD1 and GPD2 expression levels were generated. The

resulting set of strains represents a comprehensive tool for in-depth quantitative studies about

the impact of reduced Gpd1 and/or Gpd2 levels on physiology of S. cerevisiae.

The use of prototrophic strains is particularly important in studying glycerol formation since

the flux through this pathway is strongly dependent on whether amino acids have to be de

novo synthesized or taken up from the medium (see below). In addition, medium composition

strongly influences glycerol formation. We therefore believe that it is crucial to test

engineered strains in real industrial media, which - in terms of free amino acid availability -

strongly differ from media used in fundamental yeast research such as synthetic minimal

medium or YPD complex medium. To our knowledge, this is the first study of strains with

reduced GPDH activity in a medium relevant in industrial bioethanol production.

S. cerevisiae strains deleted in GPD1 have been proven to be osmosensitive (Albertyn et al,

1994). Although osmotic stress can be virtually avoided during bioethanol production by

using the SSF process, we studied the behaviour of our strains when subjected to osmostress.

This was particularly interesting because the replacement of the native promoters of GPD1

and GPD2 by TEF1 promoter versions must be accompanied by a loss of native gene

regulation by osmostress and NADH accumulation, respectively. Interestingly, the three

“intermediate” strains a7b, a7b7 and a7b2 did not show any significant loss of osmotolerance

(Figure 3) despite the replacement of the native GPD1 promoter by a constitutive and weaker

promoter. The residual GPDH activity in these strains was between 60 and 78% of the wild

type (Figure 2A) under non-osmostress conditions and the activity of the wild type is known

to even increase significantly by osmotic stress (Albertyn et al, 1994). Our result showing that

GPD1 expression driven by the weaker and non-osmoresponsive TEFmut7 promoter is

sufficient to cope with osmotic stress such as the wild type is particularly interesting in the

context of recent findings published by other authors suggesting that GPD1 upregulation only

plays a minor (if any) role for survival of osmotic shock (Bouwman et al, 2011; Mettetal et al,

2008; Westfall et al, 2008). Instead metabolic changes seem to be far more important for

increased glycerol formation and counteracting osmotic stress.

Another interesting result of our study was the striking difference regarding how one GPD

isogene was able to replace the other one for its major biological function such as osmostress

tolerance and NADH balancing under anaerobiosis, respectively. Obviously, the major

function of GPD2, i.e. redox balancing, can be easier complemented by GPD1, while GPD2

74 Chapter II

is less efficient in taking over the role of GPD1 in osmostress response. In fact, no significant

differences in growth between the strains a7b and ab7 or between the strains a7b2 and a2b7 in

minimal medium under anaerobic conditions could be detected while these pairs of strains

showed clear differences under osmotic stress, respectively (Figure 3). It rather seems that the

role of GPDH in NADH balancing can be virtually equally fulfilled by both isoenzymes. This

even seems to be confirmed by the nice correlation of GPDH activities and glycerol yields in

Figure 2B (provided that the GPDH activities measured in this study are the sum of residual

Gpd1 and Gpd2 activities).

In contrast to redox balancing, osmostress tolerance was rather dependent on the residual

level of only one isoenzyme which is GPD1. Probably, the action of the metabolic regulation

on Gpd1 for glycerol formation during osmostress recently proposed by Bouwman et al.

(Bouwman et al, 2011) is stronger than on Gpd2. The detailed mechanism of this regulation is

not yet known. However, it has been assumed by Jung et al. (2010) that Gpd1 re-localizes to

the nucleus during osmostress. If this re-localization solely occurs on Gpd1 but not on Gpd2,

it could explain the inability of Gpd2 to act in the same way as Gpd1 after osmostress. We

have to emphasize that the expression levels of GPD1 and GPD2 are driven by opposite

promoter strengths in the strains a7b2 and a2b7. Thus, we can expect that the levels of

transcription for GPD1 in a7b2 should be comparable to the level for GPD2 in a2b7 and vice

versa. Nevertheless, we cannot completely exclude that a different efficiency of translation

was the reason for the difference in phenotype.

The metabolic shift from glycerol to ethanol was accompanied by a reduction in metabolic

rates as shown here by the maximal volumetric ethanol production rates (see discussion

below). However, the biomass yield (per glucose consumed) was not reduced as long as

glycerol yield was above 50% of wild type level. The question arise which route of NADH

reoxidation was used then instead of glycerol formation. The reduction of NADH production

during acetate production in strains with reduced GPDH activity observed by us and other

authors (Guo et al, 2010; Valadi et al, 2004) can also not serve as a quantitative explanation

since acetate production was extremely low. We rather believe that cells with reduced

glycerol formation capacity were to a certain part (i.e. up to 50% reduction of glycerol yield)

able to adjust their metabolism for maximizing biomass yield. In this regard, the reader might

be referred to the discussion of Valadi et al. (2004) who proposed that NADP+ could probably

substitute for NAD+ as a cofactor in biosynthetic pathways to a larger extent than expected.

Chapter II 75

A major industrially relevant question of our study was to which level Gpd1 and Gpd2

expression can be reduced without a negative impact on maximal volumetric ethanol

production rate. Based on our results, we can conclude that even a small reduction in glycerol

yield leads to a corresponding negative impact on maximal volumetric ethanol production rate

in CEN.PK, which means that this strain does not produce more glycerol than absolutely

necessary for fulfilling the crucial biological functions. It has to be however mentioned that

three industrial S. cerevisiae strains have been demonstrated to show a 60 to 80% higher

glycerol yield compared to the model strain CEN.PK used in the current study (Devantier et

al, 2005). Therefore, it remains to be studied whether these commercial strains differ from

CEN.PK in that they produce more glycerol than absolutely necessary for maximal growth.

The work presented here shows that higher yields in ethanol seem to be feasible if longer

fermentation times were acceptable. However, a higher risk of contamination impairs such an

approach. In fact, industrial ethanol production is carried out under non-sterile conditions and

quick ethanol production by yeast is required in order to efficiently inhibit the growth of lactic

acid bacteria and compete for the available sugars.

In summary, strains with intermediate activities of Gpd1 and Gpd2 showed an improved

ethanol yield combined with the ability to completely ferment the sugars in both minimal

medium and wheat liquefact (SSF). Our result implies that these strains were able to tolerate

the high ethanol concentration at the end of the wheat liquefact SSF (up to 90 g l-1

), i.e.

fermentation did not stick due to an elevated sensitivity against high ethanol caused by

reduced GPDH activity. In fact, the gpd1∆ gpd2∆ double deletion strain has been shown to be

severely affected in ethanol tolerance (Boulahya, 2005). Moreover, three of our strains were

comparable to the wild type in terms of osmotolerance. Thus, compared to the gpd1∆ gpd2∆

double deletion strain, which virtually cannot grow and ferment at all under anaerobic

conditions, our “intermediate” strains represent a clear improvement, even though maximal

volumetric ethanol production rate was only 50% in these strains when compared to wild

type. Although strains solely based on partly reduced levels of both Gpd1 and Gpd2 might not

be directly applicable in bio-ethanol production, they seem to be a good starting point for

further metabolic engineering approaches which provide alternative pathways for NADH

reoxidation such as the strategy recently presented by Guandalupe Medina et al. (2010).

Chapter III

Quantitative trait analysis of yeast biodiversity

yields novel gene tools for metabolic engineering

78 Chapter III

1. Abstract

Engineering of metabolic pathways by genetic modification has been restricted largely to

enzyme-encoding structural genes. The product yield of such pathways is a quantitative

genetic trait. Out of 52 Saccharomyces cerevisiae strains phenotyped in small-scale

fermentations, we identified strain CBS6412 as having unusually low glycerol production and

higher ethanol yield as compared to an industrial reference strain. We mapped the QTLs

underlying this quantitative trait with pooled-segregant whole-genome sequencing using 20

superior segregants selected from a total of 257. Plots of SNP variant frequency against SNP

chromosomal position revealed one major and one minor locus. Downscaling of the major

locus and reciprocal hemizygosity analysis identified an allele of SSK1, ssk1E330N…K356N

,

expressing a truncated and partially mistranslated protein, as causative gene. The diploid

CBS6412 parent was homozygous for ssk1E330N…K356N

. This allele affected growth and

volumetric productivity less than the gene deletion. Introduction of the ssk1E330N…K356N

allele

in the industrial reference strain resulted in stronger reduction of the glycerol/ethanol ratio

compared to SSK1 deletion and also compromised volumetric productivity and osmotolerance

less. Our results show that polygenic analysis of yeast biodiversity can provide superior novel

gene tools for metabolic engineering.

Chapter III 79

2. Bibliographic reference

Hubmann, G., Foulquié-Moreno, MR., Nevoigt, E., Duitama, J., Meurens, N., Pais, TM.,

Mathé, L., Saerens, S., Nguyen, HTT., Swinnen, S., Verstrepen, KJ., Concilio, L., de

Troostembergh, JC., Thevelein, JM.

Quantitative trait analysis of yeast biodiversity yields novel gene tools for

metabolic engineering

Manuscript accepted in Metabolic Engineering (available Online 18th

March 2013)

DOI: http://dx.doi.org/10.1016/j.ymben.2013.02.006

Patent Application Intellectual Property Office

Mutant yeast strain with decreased glycerol production

Application number: GB1217028.8

Date Loged: 25 Sept. 2012

Inventor(s): Thevelein J.M., Hubmann G., Foulquié-Moreno M.R.

Swinnen, S., Schaerlaekens, K., Pais, T., Claesen, J., Hubmann, G., Yang, Y., Demeke, M.,

Foulquie-Moreno, MR., Goovaerts, A., Souvereyns, K., Clement, L., Dumortier, F.,

Thevelein JM.

Identification of novel causative genes determining the complex trait of high

ethanol tolerance in yeast using pooled-segregant whole-genome sequence

analysis.

Genome Res. 2012. 22: 975-984

DOI: http://dx.doi.org/10.1101/gr.131698.111

3. Scientific contribution

The author participated in the project conception, experimental work, data analysis and article

writing.

80 Chapter III

4. Manuscript II: Quantitative trait analysis of yeast biodiversity yields

novel gene tools for metabolic engineering

4.1 Introduction

Up to now, the targeted genetic engineering of microorganisms has concentrated largely on

the modification of structural genes encoding enzymes in metabolic pathways. This has been

done either by up- or downregulation of gene expression or by modification of kinetic

characteristics, substrate specificity or regulatory properties of the constituent enzymes

(Nevoigt, 2008). However, targeted engineering has shown only limited success when it

comes to complex traits determined by multiple genes and largely unknown regulatory

networks. Empirical approaches, on the other hand, such as mutagenesis and genome

shuffling, have often been used more successfully as a strategy to alter such complex

phenotypes (Nevoigt, 2008). These approaches have mainly been applied to improve stress

tolerance, which allows direct selection of improved mutants. Alternatively, improvement of

stress tolerance in various species has been achieved using genomic, metagenomic or cDNA

libraries, expressed in a sensitive host microorganism (Nicolaou et al, 2010). Rational

attempts have been carried out to engineer regulatory factors in order to simultaneously and

randomly alter the regulation of many genes at a time. In bacterial systems, transcription of

multiple genes can be modified easily through mutagenesis of σ factors, which has been

referred to as global cellular transcription machinery engineering (Alper and Stephanopoulos,

2007). This strategy has also been used to improve ethanol tolerance and yield in

Saccharomyces cerevisiae (Alper et al, 2006). However, eukaryotic systems regulating gene

expression are by far more complex than those of bacteria. In comparison to bacterial

systems, the changes in the phenotype caused by mutations in transcription factors are still

difficult to predict and may also cause unwanted side-effects on other important properties.

Genetic engineering of metabolic pathways in industrial eukaryotic microorganisms, such as

yeast, is clearly limited by a lack of knowledge on regulatory factors and their mechanisms of

action. This is particularly true under the conditions that occur in industrial applications.

Moreover, the structural genes and regulatory factors involved in metabolic pathways contain

Chapter III 81

many mutations in natural and industrial yeast strains, which create large phenotypic

diversity, further complicating the understanding of the interplay between the functioning of

the structural pathway and its regulatory systems. On the other hand, reverse engineering

makes use of this phenotypic diversity by targeting genes identified by genetic analysis of

natural and industrial strains with interesting traits (Bailey et al, 1996; Nevoigt, 2008). Most

of these traits, however, are complex and only recently methodologies have become available

for efficient mapping and identification of the multiple mutant genes responsible for such

complex traits (Swinnen et al, 2012b).

The exceptional capacity of the yeast S. cerevisiae for anaerobic production of ethanol is the

basis of nearly all industrial production of alcoholic beverages and fuel ethanol. Apart from

carbon dioxide, glycerol is the most important byproduct in yeast ethanolic fermentation.

Under anaerobic conditions, glycerol production is closely connected to the growth rate of the

cells. The withdrawal of intermediates from glycolysis for biosynthetic purposes necessitates

regeneration of NAD+ to sustain the redox balance and in the absence of oxygen this is

accomplished by formation of glycerol (Bakker et al, 2001). A second function for glycerol

production in yeast is its use as a compatible osmolyte under conditions of hyperosmotic

stress (Blomberg and Adler, 1989; Hohmann, 2002).

Glycerol is synthesized in two steps from dihydroxyacetone phosphate by NAD+ dependent

glycerol 3-phosphate dehydrogenase (GPDH) and glycerol 3-phosphate phosphatase, encoded

by GPD1 and GPD2, and GPP1 and GPP2, respectively (Albertyn et al, 1994; Ansell et al,

1997). Enhanced expression of GPD1 is a major factor responsible for stimulation of glycerol

production under osmostress (Albertyn et al, 1994; Larsson et al, 1993; Nevoigt and Stahl,

1997). The high osmolarity glycerol (HOG) pathway, responsible for osmostress-induced

glycerol production and other cellular adaptations, has been characterized in great detail

(Brewster et al, 1993; Hohmann, 2002). Changes in extracellular osmolarity are sensed via

two independent transmembrane proteins, Sho1 and Sln1, that both activate the HOG Map

kinase pathway. The Sln1-branch plays the most prominent role and acts through a

phosphotransfer system, composed of Sln1, Ypd1 and Ssk1. The two pathways converge on

the phosphorylation of Pbs2, which activates the Map kinase Hog1. This causes translocation

of Hog1 into the nucleus, where it associates with several transcriptional regulators, i.a. Sko1,

82 Chapter III

Msn2, Smp1 and Hot1. These Hog1-regulator complexes induce GPD1 expression to enhance

the formation of glycerol under osmostress (Hohmann, 2002). Retention of glycerol within

the cells and its efflux upon relief of osmostress are controlled by the Fps1 plasma membrane

channel (Luyten et al, 1995).

Engineering of glycerol production in yeast has attracted considerable attention. Higher

glycerol levels are desirable in wine and beer production as well as industrial glycerol

production (Cambon et al, 2006; Geertman et al, 2006; Heux et al, 2006; Nevoigt and Stahl,

1996; Remize et al, 1999; Schuller and Casal, 2005). Multiple genetic modifications have

been used to raise glycerol production and counteract the side-effect of higher acetate

production (Cambon et al, 2006; Eglinton et al, 2002; Ehsani et al, 2009). Lower glycerol

levels are highly desirable in ethanol fuel production because they are usually associated with

increased ethanol yields (Basso et al, 2008; Bro et al, 2006; Nissen et al, 2000a; Nissen et al,

2000b). High ethanol yield is a key characteristic of bioethanol production strains, reaching

approximately 90-93% of the theoretical maximum of 0.51 g ethanol per g glucose in current

industrial processes (Bai et al, 2008). Despite the high ethanol yield, part of the sugar is still

used for yeast growth and glycerol production. Glycerol yield can reach up to 2.0 - 3.6 g per

100 g consumed glucose as already reported by Pasteur (Pasteur, 1858). Glycerol yields

strongly depend on fermentation conditions (Alfenore et al, 2004; Bideaux et al, 2006;

Gardner et al, 1993) and medium composition, especially the type of nitrogen source used

(Albers et al, 1996).

A key challenge in industrial ethanol production is lowering glycerol yield without

compromising osmostress tolerance and growth rate under anaerobic conditions.

Osmotolerance is an important trait for industrial production, storage and utilization of yeast

and growth rate is closely correlated with ethanol production rate under anaerobic conditions.

Hence, diminution of GPD1 and/or GPD2 expression is not an option since it likely

compromises osmostress tolerance and growth under anaerobic conditions (Ansell et al, 1997;

Bjorkqvist et al, 1997; Nissen et al, 2000a). Even strains with fine-tuned reduction in GPDH

activity obtained with promotor engineering still showed a significant drop in osmotolerance

and/or growth rate resulting in lower ethanol productivity (Hubmann et al, 2011; Pagliardini

et al, 2010). Hence, despite considerable efforts, metabolic engineering of the structural genes

Chapter III 83

for glycerol synthesis, GPD1 and GPD2, has met with little final success because of negative

side-effects on growth, fermentation and osmotolerance (Bjorkqvist et al, 1997; Bro et al,

2006; Cao et al, 2007; Guo et al, 2011; Guo et al, 2009, 2010; Hubmann et al, 2011; Kong et

al, 2007a; Kong et al, 2006; Kong et al, 2007b; Nissen et al, 2000a; Nissen et al, 2000b).

Reverse metabolic engineering is an attractive alternative (Bailey et al, 1996), but the

identification of the genetic basis of complex traits, such as glycerol yield in fermentation, has

remained for many years an important bottleneck. The availability of genome-wide methods

for scoring SNPs as genetic markers has facilitated simultaneous mapping of multiple linked

loci referred to as quantitative trait loci (QTLs) (Brem et al, 2002; Deutschbauer and Davis,

2005; Steinmetz et al, 2002; Winzeler et al, 1998). Next generation sequencing methods now

allow very efficient QTL mapping using whole-genome sequence analysis of pooled

segregants displaying the trait of interest (Ehrenreich et al, 2010; Parts et al, 2011; Swinnen et

al, 2012a).

We have now applied pooled-segregant whole-genome sequence analysis for identification of

genetic elements determining glycerol yield in yeast fermentation. We were able to identify a

yeast strain, CBS6412, with unusually low glycerol production and concomitantly higher

ethanol production. Using 20 - 44 segregants, we succeeded in identifying major and minor

QTLs and successfully identified SSK1 as the causative allele in the locus with the strongest

linkage. Introduction of the mutant SSK1 allele in the industrial target strain lowered the

glycerol/ethanol ratio more than deletion of SSK1, without compromising osmotolerance or

ethanol productivity more. Hence, our results show that QTL analysis of selected yeast strains

can provide interesting new gene tools for reverse metabolic engineering.

4.2 Materials and Methods

Microbial strains and cultivation conditions

All S. cerevisiae strains used are listed in Table 1. Strain CBS6412 was originally indicated as

sake yeast Kyokai No.7 in the CBS collection, but comparison of the genome sequence

revealed that this indication was erroneous. E. coli strain DH5αTM

(Invitrogen Corp.,

84 Chapter III

Carlsbad) was used for amplification of plasmids. The strain was grown in Luria-Bertani (LB)

medium containing 0.5% (w/v) yeast extract, 1% (w/v) Bacto tryptone, 1% (w/v) NaCl, (pH

7.5) at 37˚C. E. coli transformation and isolation of plasmid DNA was carried out using

standard techniques (Sambrook et al, 1989). Transformants were selected on LB medium

containing 100µg/ml ampicillin.

Table 1. Saccharomyces cerevisiae strains used.

Strain Genotype Source

CBS6412 Diploid, ssk1E330N…K356N/ssk1E330N…K356N CBS-KNAW

Ethanol Red Diploid, SSK1/SSK1 Fermentis, S. I. Lesaffre

ER7A Segregant 7A of Ethanol Red, Matα This study

CBS4C Segregant 4C of CBS6412, Mata This study

ER7A ssk1∆ ER7A, ssk1∆ (marker removed) This study

CBS4C ssk1∆ CBS4C, ssk1∆ (marker removed) This study

ER7A/ CBS4C Hybrid diploid ER7A x CBS4C This study

ER7A/ CBS4C ssk1∆ Hybrid diploid ER7A x CBS4C ssk1∆

This study

ER7A ssk1∆/ CBS4C Hybrid diploid ER7A ssk1∆ x CBS4C This study

ER7A ssk1∆/ CBS4C ssk1∆ Hybrid diploid ER7A ssk1∆ x CBS4C ssk1∆ This study

ER7A ssk1E330N…K356N ER7A ssk1E330N…K356N (SSK1 allele exchanged) This study

CBS4C SSK1 CBS4C SSK1 (SSK1 allele exchanged) This study

HG5 Ethanol Red ssk1∆/ssk1∆ This study

HG7 Ethanol Red ssk1E330N…K356N/ssk1∆ This study

HG8 Ethanol Red ssk1E330N…K356N/ ssk1E330N…K356N This study

SSK1 alleles: SSK1 = SSK1 allele of Ethanol Red, ssk1E330N…K356N = SSK1 allele of CBS6412, ssk1∆ = deletion of SSK1

Chapter III 85

Mating, sporulation and tetrad analysis

Mating, sporulation and dissection of asci were carried out according to standard procedures

(Sherman and Hicks, 1991). Mating type of segregants was determined by diagnostic PCR for

the MAT locus (Huxley et al, 1990).

Fermentation conditions

A selection of 52 S. cerevisiae wild type strains (Figure 1) was screened in 250 ml oxygen-

limited and stirred fermentations containing 1% (w/v) yeast extract, 2% (w/v) peptone and

12% (w/v) glucose. Screening of the selected parent strains and the segregants was performed

in 15 ml falcon tubes containing 5 ml minimal medium containing 1.9 g l-1

yeast nitrogen base

(Difco), 5 g l-1

ammonium sulphate, 250 mg l-1

leucine, 50 mg l-1

uracil, 100 mg l-1

histidine,

30 mg l-1

lysine, 20 mg l-1

methionine and 50 g l-1

glucose. Fermentations were inoculated

with an initial OD of 1 and their progress followed by weight loss. Selected segregants were

also tested in 100 ml oxygen-limited stirred fermentations. All fermentations were carried out

at 30˚C.

High gravity fermentations were carried out in fermentation tubes containing 250 ml of YP

and 33% (w/v) glucose. Precultures used as inoculum were first grown on YP 2% (w/v)

glucose for 24 hours and then on YP 10% (w/v) glucose up to an OD600 of 1. The

fermentations were inoculated with 5.10

7 cells/ml and kept at 25°C. Stirring was applied for

the first 4h (120 rpm). When the weight loss was stable for 2 consecutive days, the

fermentation was considered to be finished.

SHF (Separate Hydrolysis and Fermentation) fermentations were carried out with wheat

liquefact (24.5% dry mass content) acquired from a local ethanol plant. After adjustment of

the pH to 4.5 with sulfuric acid, it was treated with Dextrozyme®

(Novozyme, Denmark) for

24h at 60˚C to obtain hydrolysate. The latter was boiled at 100˚C for 20 min and then cooled.

86 Chapter III

Oxygen-limited fermentations were carried out with 100 ml of this medium inoculated with 5

ml of yeast suspension. The fermentations were performed at 30˚C and were continuously

stirred at 200 rpm.

Assessment of osmotolerance was performed in fermentations containing minimal medium

with or without 0.7M or 1.4M NaCl, or 1M or 2M sorbitol. The fermentations were

continuously stirred at 200 rpm.

Determination of fermentation parameters

In all fermentations weight loss was used to follow the progress of the fermentation. Glucose,

glycerol and ethanol in the medium were determined by HPLC (Waters® isocratic BreezeTM

HPLC, ion exchange column WAT010290). Column temperature was 75˚C, 5 mM H2SO4

was used as eluent with a flow rate of 1 ml min-1

and refractive index detection was used

(Waters, 2414 RI detector). Biomass was determined by OD600 at the beginning and the end

of fermentation and yeast dry mass also at the end. The product yield was calculated from the

final product concentration (g l-1

) and the difference in glucose concentration at the start and

end of fermentation (consumed glucose in g l-1

). The product yields in the SSF were based on

the final product concentrations and the equivalent initial glucose concentration (the latter was

measured in a completely hydrolyzed sample of wheat liquefact).

Determination of maximum aerobic growth rate

The aerobic cultivations were carried out in Erlenmeyer flasks containing 50 ml medium,

which was composed of 1% (w/v) yeast extract and 2% (w/v) glucose. The strains were pre-

cultured overnight at 30˚C and 200 rpm in small tubes containing 3 ml of the same medium.

Chapter III 87

The pre-culture were then used to inoculate the main culture up to an initial OD600 of 0.2. The

Erlenmeyer flasks were incubated at 30˚C in a rotary shaker at 200 rpm. Samples were taken

every second hour in order to determine the OD600. Non-linear regression was used to fit a

polynomial curve to the experimental progression of the OD600 values. The growth rate was

calculated by differentiating the continuous curve. Specific growth rates were calculated by

dividing the growth rate by the biomass estimated from the continuous curve at the same time.

The maximum of the curve was used as the maxium growth rate.

Aerobic growth spot assay

The aerobic growth behaviour was tested on solid medium, which contained 1% (w/v) yeast

extract, 2% (w/v) glucose and 1.5% (w/v) agar. The strains were pre-cultured overnight at

30˚C and 200 rpm in small tubes containing 3 ml of the same liquid medium. The OD600 was

determined for all pre-cultures and the amount of cells was harvested to obtain an OD600 of 1

in 1 ml distilled water. After re-suspension of the cells in 1 ml water, these samples were

subsequently diluted six times with a dilution factor of 1:5. All dilutions were transferred to a

microtiter plate, and 5 µl of each dilution was transferred to the solid medium. The plates

were incubated for two days at 30˚C. Growth was monitored after 24 h and at the end.

DNA methods

Yeast genomic DNA was extracted with Phenol/Chloroform/Isoamyl-alcohol (25:24:1)

(Hoffman and Winston, 1987) and further purified with diethyl-ether extraction or ethanol

precipitation if required. PCR was performed with high-fidelity polymerases PhusionTM

(Finnzymes) or ExTaqTM

(TaKaRa) for cloning, amplification of deletion or insertion

88 Chapter III

cassettes, and sequencing purposes. Sequencing was carried out using the dideoxy chain-

termination method (Sanger and Coulson, 1975) at the VIB Genetic Service Facility

(Antwerp). The sequences were analyzed with geneious (Geneious Basic 5.3.4), SeqMan

(Lasergene Coresuite 8) or CLC DNA workbench (CLC bio) software. Lists of primers and

plasmids are provided as Supplementary Tables 1 and 2, respectively.

Pooled-segregant whole-genome sequence analysis

After crossing the two parent strains CBS4C and ER7A, the 20 most superior segregants

(lowest glycerol production) were assembled in the 'selected pool' while 20 random

segregants were used to assemble the 'unselected pool'. The two pools were made by

combining equal amounts of cells based on OD600. High molecular weight DNA (3 µg, ~

20kb fragments) was isolated from the pools and parent strains according to Johnston and

Aust (1994). The purity of the DNA sample was estimated from UV measurement (260/280 =

1.7-2.0). The DNA samples were provided to GATC Biotech AG (Konstanz, Germany) and

BGI (Hong Kong, China) for whole-genome sequence analysis by Illumina technology.

QTL analysis based on the distribution of SNP variant frequency over the length of the

chromosomes was carried out as described by Swinnen et al. (2012a). The short read

sequences obtained from the parental strains and the pools were mapped against the known

S288c reference sequence using the mapping software Bfast (Homer et al, 2009). After

pairing, unique alignments for the CBS4C strain were selected and homozygous variants, i.e.

SNPs and small indels, were called using SNVQ (Duitama et al, 2012). In addition, regions

with coverage below 0.5 or above 1.5 of the average coverage were identified and SNPs of

those regions were filtered out. For each polymorphic position the variant calls in the aligned

Chapter III 89

reads for the ER7A strain were then extracted and variants were filtered out for which the

coverage of the reference variant was too small (<20x) or too large (>150x) or SNPs of both

parents coincided but were different from the reference. Finally, the number of calls to the

reference and the alternative variant of each selected polymorphic position was determined

from the set of aligned reads corresponding with the segregant pools. The SNP variant

frequencies were calculated by dividing the number of the alternative variant by the total

number of aligned reads. A very high or a very low frequency was a sign of a one-sided SNP

segregation preferentially coming from one parent, indicating a genetic linkage to the trait of

interest. Genetic linkage was statistically confirmed using the methods described earlier

(Swinnen et al, 2012a).

Detection of SNP markers by allele specific PCR

Individual SNPs were scored by allele specific PCR. The forward and reverse primer

contained the nucleotide of ER7A or CBS4C as the 3’ terminal nucleotide. The annealing

temperature was optimized using DNA extracts of ER7A and CBS4C so as to allow only

hybridization with primers containing a complete match. A list of primers is provided in

Supplementary Table 1.

Reciprocal hemizygosity analysis (RHA)

For RHA analysis (Steinmetz et al, 2002), two diploid strains were constructed by crossing

CBS4C and ER7A wild type or ssk1∆ strains, so that the resulting diploids only contained a

single SSK1 allele, either CBS4C derived ssk1E330N…K356N

or ER7A derived SSK1. Deletion

cassettes were constructed essentially as described by Gueldener et al. (2002) with the

90 Chapter III

phleomycin resistance marker bleR and SSK1 gene deletion was confirmed by PCR. The

selection marker was removed using the Cre-loxP system. The removal of the selection

marker was verified by phleomycin sensitivity as well as by PCR. RHA was performed with

three independent isolates of all tested diploids.

Construction of SSK1 insertion cassettes

The repeat region H1 was PCR amplified with the primers A-6101 and A-6103 using genomic

DNA of CBS4C and ER7A as template. The resulting PCR fragment was digested with KpnI

and SalI and purified from an agarose gel. SSK1 was PCR amplified from genomic DNA of

CBS4C and ER7A and the primers A-6100 and A-6102. The obtained product of around

2800bp was digested with SalI and XmaI. The cloning vector pBluescriptII SK(+)

(Fermentas) was digested with KpnI and XmaI and ligated with the repeat region H1 and the

SSK1 allele of the respective strain. The construct was verified using Sanger sequencing. The

two selectable and counter-selectable systems, AMD1 and NAT1-GIN11, were used to

introduce the insertion cassette. During counter-selection, the marker genes spontaneously

looped out via the H1 repeat region, leaving no scars of non-S. cerevisiae DNA in the

genome. The AMD1 marker of Zygosaccharomyces rouxii was cut out of the plasmid pF6a-

AMD1-MX6 using SacI and BglII (Shepherd and Piper, 2010). The fragment was gel purified

and ligated with pUG66, which was also digested with the same two enzymes. The resulting

plasmid pUG-AMD was used for PCR amplification of the AMD1 marker using the primers

A-5166 and A-6770. The PCR product as well as the H1-SSK1 plasmids were digested with

SalI and ligated, resulting in plasmids pBluescriptII_AMD1_ssk1E330N…K356N

and

pBluescriptII_AMD1_SSK1. The selection marker NAT1 was amplified from pAG25 using

primers A-7116 and A-7117. The GIN11 counter-selection marker (Akada et al, 2002) was

Chapter III 91

amplified from pG119 using primers A-7118 and A-7119. Both fragments were sequentially

digested with DraIII and SalI and ligated with the H1-SSK1 plasmid, which was previously

digested with SalI resulting in the plasmids pBluescriptII_NAT1_GIN11_ssk1E330N…K356N

and

pBluescriptII_NAT1_GIN11_SSK1. The insertion cassette H1-loxP-AMD1-loxP-SSK1 and

H1-loxP-AMD1-loxP- ssk1E330N…K356N

were amplified from pBluescriptII_AMD1_SSK1 and

pBluescriptII_AMD1_ssk1E330N…K356N

using the outside flanking primers matching the M13

primer binding sites of the plasmid pBluescriptII SK(+). The PCR product was purified and

used for transformation. Cassettes with H1-NAT1-GIN11-SSK or H1-NAT1-GIN11-

ssk1E330N…K356N

were digested with BspHI and digestion products were used for

transformation. Yeast was transformed with the LiAc/PEG method (Gietz et al, 1992).

Reciprocal SSK1 allele replacement in CBS4C and ER7A

Site-directed modification of the CBS4C and ER7A SSK1 locus was carried out using a two

step-method. In the first step, the SSK1 insertion cassettes (see above) were transferred to the

SSK1 deletion strains, CBS4C ssk1∆ and ER7A ssk1∆. After transformation, positive clones

were selected on YD agar plates containing 200µg/ml ClonNat. The presence of the insertion

cassette was verified by PCR using the primers A-5168 and A-7301. In the second step, the

marker genes were removed by selection of spontaneous loop-outs on galactose-containing

medium after induction of the counter-selectable marker GIN11 (Akada et al, 2002; Akada et

al, 1999; Olesen et al, 2000). Positive looped-out clones were identified by ClonNat

sensitivity and verified by PCR using the forward primer A-5168 and the SSK1 allele specific

reverse primers A-5126 and A-5127. The inserted SSK1 alleles were verified by Sanger

sequencing.

92 Chapter III

SSK1 allele replacement in the industrial strain Ethanol Red

The ssk1E330N…K356N

allele of CBS4C was inserted twice in the Ethanol Red derivative, HG5

(see Table 1), which had both SSK1 alleles deleted. The latter strain was constructed by

introducing a disruption cassette flanked with loxP sites using homologous recombination

(Gueldener et al, 2002; Kotaka et al, 2009). The disruption cassette was constructed with

homologous sequences (H1/H2) corresponding to the 5’ and 3’ end of the SSK1 ORF

surrounding the phleomycin resistance-gene bleR used as selectable marker. The selectable

marker, bleR, was removed by Cre recombinase. A second disruption cassette was constructed

with the recombination sites H1* and H2

*, which were located inside the first homologous

integration sites, H1 and H2, enabling specific recombination into the 2nd

SSK1 allele of the

diploid strain. Gene disruption was verified by PCR. The bleR marker gene was again

removed using the Cre/loxP system. The double deletion was confirmed by PCR using

primers located outside the integration site.

The two ssk1E330N…K356N

insertion cassettes, H1-loxP-AMD1-loxP-ssk1E330N…K356N

and H1-

NAT1-GIN11-ssk1E330N…K356N

, were successively transformed. After the 1st transformation of

the H1-loxP-AMD1-loxP-ssk1E330N…K356N

cassette, transformants were selected based on

hydrolysis by Amd1 of acetamide used as sole nitrogen source. The correct integration of the

insertion cassette was verified by PCR using the primers A-5168 and A-5894 inside the AMD

gene. In the 2nd

transformation, the H1-NAT1-GIN11-ssk1E330N…K356N

was transferred. Positive

transformants were selected on acetamide medium, containing 200µg/ml ClonNat. Correct

integration in the 2nd

chromosome was verified by PCR using the primers A-5168 and A-7301

to test the presence of the insertion cassette as well as the primers A-5168 and A-5169 to

verify the disappearance of the two ORF deletions of the Ethanol Red ssk1∆/∆. Counter-

Chapter III 93

selection was simultaneously applied for both marker systems using medium with 100mM

fluoroacetamide and 0.04% galactose (induction of GIN11).

Data deposition

Sequencing data is deposited at the SRA database (NCBI), http://www.ncbi.nlm.nih.gov/sra,

with the account number SRA054394.

4.3 Results

Selection of parent strains for genetic mapping of low glycerol yield

We have evaluated 52 diploid S. cerevisiae strains from diverse origins for the ratio between

the amount of glycerol and ethanol produced in small-scale (250 ml) fermentations with

complex medium containing 12% glucose. A continuous and normal distribution of the trait

was observed (Figure 1). The CBS6412 strainshowed the lowest glycerol yield (0.043 g g-1

)

of all strains tested, which was about 63% of that of the reference industrial strain Ethanol

Red (0.068 g g-1

) (Figure 2A), an industrial strain commonly used for bioethanol production

with corn and wheat starch hydrolysate. As it had both a low glycerol/ethanol ratio and a low

glycerol yield, CBS6412 was chosen as the superior strain and Ethanol Red was used as the

inferior strain. In order to obtain haploid strains for genetic mapping analysis, the two diploid

strains were sporulated and segregants were tested in small-scale fermentations. Glycerol

yields of the segregants were normally distributed around those of the diploid parents (Figure

2B), indicating a highly heritable phenotype. The CBS6412 segregant, CBS4C, had an even

lower glycerol yield than its parental diploid (Figure 2A), indicating acquirement of one or

more beneficial, recessive alleles present in heterozygous form in the diploid strain. CBS4C

94 Chapter III

was selected as the superior parent strain for the genetic mapping. The Ethanol Red segregant

ER7A had a glycerol yield closest to its parental diploid and served as inferior parent strain.

Figure 1. Variation in glycerol and ethanol yield. (A) Glycerol/ethanol ratio in 52 natural and industrial S.

cerevisiae strains. The selected diploids used for quantitative trait analyis, Ethanol Red (inferior parent, target

industrial bioethanol production strain) and CBS6412 (superior parent) are marked in black. (B) Normal

distribution of the glycerol/ethanol ratio around a mean value of 5.7% for the 52 strains. Fermentations were

carried out in 250 ml oxygen-limited and stirred medium containing 1% (w/v) yeast extract, 2% (w/v) peptone

and 12% (w/v) glucose. The glucose was essentially fermented to completion by all strains.

Chapter III 95

Figure 2. Glycerol and ethanol yield in the segregants of the diploid parent strains and in segregants

from the cross between the selected haploid parent strains. (A) Glycerol and ethanol yield of the diploids,

Ethanol Red and CBS6412, and the CBS6412 segregant, CBS4C, showing the lowest glycerol yield of all tested

segregants of CBS6412. Fermentations were carried out in 100 ml minimal medium with 10% glucose. (B)

Distribution of the glycerol yield in the haploid segregants of CBS6412 (black bars) and Ethanol Red (white

bars). The distribution was normal around the value of the diploid parents, CBS6412 (left small black square on

top) and Ethanol Red (right small black square on top). (C) Distribution of the glycerol yield in the segregants

from the cross CBS4C x ER7A (white bars). All segregants were screened in 5 ml fermentations. After

evaluation of the 48 segregants with the lowest glycerol yield in 100 ml minimal medium with 5% glucose, the

44 segregants with the lowest glycerol production were selected for pooled-segregant whole-genome sequence

analysis (black bars). The glycerol yield of the haploid parents CBS4C and ER7A is indicated with small black

squares on top.

Construction of the CBS4C/ER7A hybrid and selection of superior segregants with low

glycerol yield

The CBS4C and ER7A haploid strains were crossed with each other and 257 segregants were

isolated and first characterized for glycerol and ethanol yield in 5 ml fermentations with 5%

glucose in minimal medium. Figure 2C shows a histogram of the glycerol yield in the

96 Chapter III

segregant population in comparison with that of the CBS4C and ER7A haploid parents. The

glycerol yield showed a normal distribution and most segregants had a glycerol yield close to

the average (0.063 g g-1

, ±142% of the CBS4C glycerol yield). We re-tested the 48 segregants

with a glycerol yield below 120% of the CBS4C parent in 100 ml small-scale fermentations;

44 segregants showed the same low glycerol yield also under these conditions. Among these,

the 20 segregants showing the lowest glycerol yield (≤ 0.054 g g-1

) were selected for QTL

mapping with pooled-segregant whole-genome sequence analysis. The 24 remaining

segregants were used for subsequent validation of the results as described below. A second

pool with 20 randomly selected segregants was also subjected to pooled-segregant whole-

genome sequence analysis and used as control.

QTL mapping using pooled-segregant whole-genome sequence analysis

The genomic DNA of the selected and random pools, as well as the two parent strains, was

extracted and submitted to custom sequence analysis using Illumina HiSeq 2000 technology

(GATC Biotech AG, Konstanz, Germany; BGI, Hong Kong, China). The sequence reads of

the CBS4C and ER7A parent strains were aligned with the S288c standard sequence, which

allowed to identify 21,818 SNPs between CBS4C and ER7A. The SNPs were filtered as

described previously (Duitama et al, 2012). The variant frequency of the quality-selected

SNPs in the DNA of the two pools was then plotted against the SNP position on the

chromosome. The scattered raw data were smoothened by fitting smoothing splines in the

generalized linear mixed model framework as previously described (Swinnen et al, 2012a).

The results are shown in Figure 3. A prominent QTL with strong linkage was present on

chromosome XII (between 135,000 and 200,000 bp) and is shown in more detail in Figure

4B. Individual SNPs from that region, as well as from the QTL with lower linkage on

Chapter III 97

chromosome II (Figure 4A), were scored by allele specific PCR detection in the 20 individual

segregants of the selected pool (Figure 4A, B). The precise SNP variant frequency determined

in this way was used to verify the linkage of the two regions on chromosome II and XII,

respectively. This revealed a very strong linkage with low glycerol yield for the QTL on

chromosome XII with the minimal P-value being 1.45.10

-4, while the P-values for the QTL on

chromosome II only dropped just below the 0.05 threshold for significance (0.009). The same

SNPs were also scored by allele specific PCR in the 24 remaining segregants with a glycerol

yield below 120% of the CBS4C parent. Calculation of the P-values for the whole group of 44

segregants no longer revealed significant linkage for the QTL on chromosome II. On the other

hand, the P-values for the QTL on chromosome XII dropped to 9.10

-11, strongly increasing

significance of the linkage. Hence, we concentrated the further analysis on the QTL of

chromosome XII.

98 Chapter III

Figure 3. Plots of SNP variant frequency versus chromosomal position and corresponding P-values. The

variation in SNP variant frequency is shown for all 16 yeast chromosomes (raw data: small grey circles;

smoothened data: black line; statistical confidence interval: grey lines). Significant upward deviations from the

average of 0.5 indicate linkage to the superior parent CBS4C, while significant downward deviations indicate

linkage to the inferior parent ER7A. The smoothened line was determined as described previously (Swinnen et

al, 2012a). Strong candidate QTLs were found on chromosome II (at position 500,000 – 700,000 bp) and

chromosome XII (at position 135,000 – 200,000 bp), but only for the latter the P-value dropped below the

significance limit of 0.05.

Chapter III 99

Identification of SSK1 as a causative gene in the QTL on chromosome XII

The 20,000 bp region with the strongest linkage in the QTL on chromosome XII contained 13

genes, of which four genes contained non-synonymous mutations in the ORF (Figure 4B).

One of those four genes, SSK1, was located in the centre of the QTL, which had a slightly

stronger linkage. Ssk1 has a known function in the HOG pathway. Sequence comparison of

the SSK1 alleles of the parental strains CBS4C and ER7A with the allele of the reference

strain S288c revealed ten polymorphisms between the sequence of the SSK1 ORF in CBS4C

and ER7A. A single base pair deletion at position 162,907 bp of Chr. XII was the most

prominent mutation in the CBS4C SSK1 ORF, since it caused a reading frame shift and a new

stop codon at position 357 in the protein. This resulted in a new primary amino acid sequence

from position 330 until 356, while the wild type Ssk1 protein had a total length of 712 amino

acids. Hence, we named the new allele ssk1E330N…K356N

. The dramatic change in amino acid

sequence and the truncation would normally be expected to result in a completely inactive

protein and therefore in a phenotype similar to that of the ssk1∆ strain. However, this was not

the case. The ssk1E330N…K356N

allele caused a different phenotype compared to deletion of

SSK1 (see below).

Next, we evaluated SSK1 as possible causative gene using reciprocal hemizygosity analysis

(RHA) (Steinmetz et al, 2002). For that purpose, two CBS4C/ER7A hybrid and hemizygous

diploid strains were constructed differing only in a single SSK1 allele, the other allele being

deleted. The diploid strain with the single ssk1E330N…K356N

allele derived from CBS4C showed

a significantly reduced glycerol yield and a significantly higher ethanol yield than the diploid

strain with the SSK1 allele from the ER7A strain (Figure 5). This showed that ssk1E330N…K356N

was a causative gene in the QTL on chromosome XII.

100 Chapter III

Figure 4. SNP variant frequency and P-values determined in individual segregants for downscaling of

the QTLs. Top: SNP variant frequency map of chromosome II (A) and chromosome XII (B) determined with

the pool of 20 selected segregants (raw data: small circles; smoothened data: black line; statistical confidence

interval: black stippled lines) and the pool of 20 unselected segregants (raw data: small triangles; smoothened

data: grey line; statistical confidence interval: grey stippled lines). Middle: SNP variant frequency of seven

selected SNPs in the candidate regions on chromosome II (at position 500,000 – 700,000 bp) (A) and

chromosome XII (at position 135,000 – 200,000 bp) (B), determined in the individual 20 most superior

segregants (▢) and the individual 44 most superior segregants (●). Smoothened lines: SNP variant frequency

determined with the pool of 20 selected segregants (black line) and the pool of 20 unselected segregants (grey

line). Bottom: P-values for the same seven SNPs in the regions on chromosome II (A) and XII (B). The

statistical confidence line (P-value ≤ 0.05) is also indicated. The region on chromosome XII was significantly

linked. Lowest panel: overview of all genes present and all SNPs identified in the region with the highest linkage

of the QTL on chromosome XII (154,000 bp – 175,000 bp). Genes marked with a star contained a non-

synonymous mutation in the ORF.

Chapter III 101

Figure 5. Identification of SSK1 as the causative gene in the QTL on chromosome XII. Diploid strains

constructed for reciprocal hemizygosity analysis (RHA) with either the deletion of the ssk1E330N…K356N allele of

CBS4C or the deletion of the SSK1 allele of ER7A. Glycerol and ethanol yield of the two hemizygous diploid

strains. The difference in glycerol and ethanol yield for the two diploids was significant by the Student t-test.

Fermentations were carried out in 100 ml minimal medium with 5% glucose.

To evaluate whether the ssk1E330N…K356N

allele of CBS4C behaved as a recessive allele and

whether it caused the same phenotype as the deletion of SSK1, we also constructed a

CBS4C/ER7A hybrid diploid strain with both SSK1 alleles deleted and compared its

phenotype with that of CBS4C/ER7A with its original SSK1 alleles. The glycerol and ethanol

yields of these strains were similar to that of the hemizygous diploid strain with the SSK1

allele from ER7A or the ssk1E330N…K356N

allele from CBS4C, respectively (Figure 6A). This

indicates that the ssk1E330N…K356N

allele from CBS4C is a recessive allele and that

ssk1E330N…K356N

behaves as a loss of function allele, at least in the hybrid background and the

fermentation conditions used (100 ml anaerobic fermentations in minimal medium containing

102 Chapter III

5% glucose). When the glycerol yield (0.043 g g-1

) of the CBS4C parent strain was

normalized to 100%, the glycerol yield of ER7A (147%) and that of the diploids ER7A/

CBS4C (145%) and ER7A/ CBS4C ssk1∆ (148%) was very similar (Figure 6A). In contrast,

the strains ER7A ssk1∆/ CBS4C ssk1E330N…K356N

and ER7A ssk1∆/ CBS4C ssk1∆ had a

glycerol yield of 119% and 122%, respectively, (Figure 6A) suggesting that ssk1E330N…K356N

was responsible for the majority of the reduction in glycerol yield in CBS4C compared to

ER7A. This agrees with the result of the pooled-segregant whole-genome mapping, which

revealed the SSK1 locus as the only QTL with significant linkage.

To confirm the importance of SSK1 in an alternative way, we reciprocally exchanged the

SSK1 alleles of CBS4C and ER7A by homologous recombination. The ssk1E330N…K356N

allele

was introduced as new allele in ER7A ssk1∆. In the same manner, SSK1 of ER7A was

introduced in CBS4C ssk1∆. The ER7A ssk1E330N…K356N

strain showed reduced glycerol yield

and enhanced ethanol yield comparable to the ER7A ssk1∆ strain (Figure 6B). Deletion of

ssk1E330N…K356N

in CBS4C resulted in a further reduction in glycerol yield when compared to

the original CBS4C. Concomitantly, ethanol yield increased in the CBS4C ssk1∆ strain when

compared to the original CBS4C strain, which may be due to a growth deficiency caused by

ssk1∆ in that strain background. This growth defect was suggested also by the maximal

volumetric ethanol production rate, which was strongly reduced in the CBS4C ssk1∆ strain

compared to the original CBS4C strain with the ssk1E330N…K356N

allele (Figure 6B). In the

ER7A background there was no difference. This again shows that at least in the CBS4C

background the effect of ssk1∆ is different from that of ssk1E330N…K356N

.

Chapter III 103

As shown by the aerobic growth assay in Figure 6C, we observed that the CBS4C ssk1∆

strain grew more poorly than the original CBS4C strain, which contains the ssk1E330N…K356N

allele. A clear difference in growth of these two strains was also observed in aerobic liquid

cultures, where the two strains, CBS4C and CBS4C ssk1∆, grew at a maximum rate of 0.47

and 0.35 h-1

, respectively (Figure 6C). This difference in growth was again not observed in

the ER7A strain background. The introduction of wild type SSK1 in CBS4C ssk1∆ enhanced

its glycerol yield and reduced its ethanol yield, and it increased the maximal volumetric

ethanol production rate (Figure 6B). These results confirmed SSK1 as a causative allele for

reduced glycerol and enhanced ethanol production in CBS4C. They also show that the

ssk1E330N…K356N

allele causes a different effect compared to the deletion of SSK1 and that it

also differs in causing no apparent growth defect in the CBS4C background as opposed to the

deletion of SSK1.

Given the recessive character of the ssk1E330N…K356N

allele, we tested its presence in the

original diploid strain CBS6412 and found it to be present in two copies (data not shown).

This suggests that the unusual allele may provide a selective advantage in specific

environmental niches.

104 Chapter III

Figure 6. Glycerol and ethanol yield after deletion or reciprocal exchange of the SSK1 alleles. (A)

Comparison of the glycerol and ethanol yield of the two hemizygous strains with that of the ER7A and CBS4C

parental strains and the SSK1/ssk1E330N…K356N and ssk1∆/ssk1∆ diploids. (B) Glycerol and ethanol yield and

maximal volumetric ethanol production rate in the ER7A and CBS4C parental strains, SSK1 deletion versions

and the ER7A and CBS4C strains with reciprocal exchange of the SSK1 and ssk1E330N…K356N alleles.

Fermentations were carried out in 100 ml minimal medium with 5% glucose. (C) Growth comparison of the

strains, as detailed in (B), under aerobic conditions. Fivefold dilutions of overnight pre-grown cells were spotted

on YPD medium and incubated at 30˚C for two days. The maximum growth rate in liquid cultures was

determined from OD600 measurements in 50 ml YPD shake flasks incubated at 30˚C.

Chapter III 105

Reduction of the glycerol/ethanol ratio in an industrial bio-ethanol strain using

ssk1E330N…K356N

as a novel gene tool

To test the functionality of ssk1E330N…K356N

as a novel gene tool for reduction of glycerol yield

under industrially relevant conditions, both SSK1 alleles of the industrial bio-ethanol

production strain, Ethanol Red, were replaced by the ssk1E330N…K356N

variant using

homologous recombination. In addition, an Ethanol Red ssk1∆/ssk1∆ strain and an Ethanol

Red ssk1E330N…K356N

/ssk1∆ strain were constructed. These strains were tested in fermentations

with minimal medium (5% (w/v) glucose), high gravity medium (YP with 33% (w/v) glucose)

and wheat hydrolysate (SHF: Separate Hydrolysis and Fermentation). The results are shown

in Figure 7A. The double deletion of SSK1 reduced the glycerol yield. Interestingly, further

reduction of glycerol yield was observed after introduction of one copy of ssk1E330N…K356N

,

while introduction of the second copy of ssk1E330N…K356

lowered glycerol yield even more.

Ethanol yields clearly increased in all Ethanol Red mutants compared to the wild type strain

in the minimal medium. The reduction of glycerol yield under high gravity or SHF conditions

was generally less pronounced compared to minimal medium. Thus, the concomitant increase

in ethanol yield in the Ethanol Red mutants was less obvious. Nevertheless, particularly the

result obtained in minimal medium indicated that in the Ethanol Red diploid background the

ssk1E330N…K356N

allele did not simply behave as a loss-of-function allele but had a stronger

reducing effect on the glycerol/ethanol ratio than deletion of the SSK1 gene. These results

confirm the usefulness of the ssk1E330N…K356N

allele as a novel gene tool for lowering glycerol

production in industrial yeast strains. Identification of other mutant alleles in the CBS4C

strain and introduction of these alleles in the Ethanol Red strain with two ssk1E330N…K356N

alleles, may allow to reduce glycerol yield even more, especially under the industrially-

relevant conditions.

106 Chapter III

The novel gene tool ssk1E330N…K356N

retains its positive effect under high osmolarity

conditions

Several previous studies successfully reduced glycerol yield in S. cerevisiae with a

concomitant increase in ethanol yield. However, many of the resulting strains showed a

significantly reduced maximal volumetric ethanol production rate and increased sensitivity

against osmotic stress (Bjorkqvist et al, 1997; Guadalupe Medina et al, 2010; Hubmann et al,

2011; Nissen et al, 2000a). In order to address this issue, we determined both the

glycerol/ethanol ratio and the maximal volumetric ethanol production rate in the Ethanol Red

strains containing one or two ssk1E330N…K356N

alleles under conditions of high osmolarity. In

general, the cells produced higher levels of glycerol under hyperosmotic stress, i.e. in the

presence of 1.4 M and 2 M sorbitol or 0.7 M and 1 M NaCl (Figure 7B). In spite of this, a

similar improvement in the glycerol/ethanol ratio was observed in the Ethanol Red strains

containing one or two ssk1E330N…K356N

alleles. The maximal volumetric ethanol production rate

dropped with increasing osmolarity but this drop was not correlated with the presence or the

number of ssk1E330N…K356N

alleles. Hence, the ssk1E330N…K356N

allele does not appear to cause

an increase in osmosensitivity and retains its positive effect under conditions of high

osmolarity. Close examination of the effect of ssk1E330N…K356N

on glycerol production in the

Ethanol Red background also allows to make a quantitative assessment of the contribution of

this allele to the phenotype. The initial glycerol yield was 167% of the CBS4C yield while the

double insertion of ssk1E330N…K356N

caused a drop to 128% of the CBS4C yield. Hence, the

ssk1 mutation appears to determine 50-60% of the trait.

Chapter III 107

Figure 7. Glycerol and ethanol yields and osmostress tolerance in fermentations with the industrial

bioethanol production strain Ethanol Red in which one or two copies of the ssk1E330N…K356N

allele had been

introduced. (A) Glycerol and ethanol yields in fermentations with minimal medium (5% (w/v) glucose), high

gravity medium (YP with 33% (w/v) glucose) and wheat hydrolyzate (SHF: Separate Hydrolysis and

Fermentation). (B) Glycerol/ethanol ratio and maximal volumetric ethanol production rate (rmax in g l-1 h-1) in

fermentations with minimal medium (5% glucose) in the presence of NaCl (0, 0.7 and 1M) or sorbitol (0, 1.4 and

2M).

108 Chapter III

4.4 Discussion

Up to now, QTL analysis in S. cerevisiae has been performed mainly with selectable traits

easily scorable in high throughput assays, so that many hundreds, thousands or, with

appropriate selection, even millions (Parts et al, 2011), of segregants can be used. Stress

tolerance for instance is a trait that can be screened easily with large numbers of strains

simultaneously. For many industrially-important traits, such as fermentation kinetics and

product yields, on the other hand, segregants have to be phenotyped individually, for instance

in small-scale fermentations. This requires much more work and is usually limited to a few

hundred strains. Glycerol yield is a non-selectable metabolic trait and up to now no such

quantitative traits have been submitted to polygenic analysis in yeast. In this work we have

shown that genetic mapping of QTLs underlying this metabolic trait can be performed

successfully using pooled-segregant whole-genome sequence analysis with only 20

segregants out of a total of 257 phenotyped. This shows that genetic analysis of polygenic

metabolic traits can now be used as a novel tool for identifying interesting alleles for reverse

metabolic engineering. This approach should in principle be useful for reverse metabolic

engineering of any phenotype and any target industrial yeast strain of which a mating-

competent haploid segregant can be obtained.

The conspicuous deficiency in our understanding of the interplay between metabolic

pathways and cellular regulation has made successful metabolic engineering and synthetic

biology a more daunting task than originally anticipated (Yadav et al, 2012). An important

issue in this respect is also the inability to mimic industrial-scale conditions precisely in the

laboratory, and therefore the inability to test for all possible side-effects which may occur at

industrial scale. This creates high risk for taking multi-engineered strains into industrial scale.

Chapter III 109

As shown in the present paper, appropriate screening of yeast biodiversity can identify strains

with superior performance for a specific metabolic trait, like reduced glycerol production. We

have shown that the responsible mutant alleles can be used successfully as new gene tools to

reverse engineer an industrial yeast strain, minimizing the risk of side-effects on other

industrially-important properties.

Application of pooled-segregant whole-genome sequence analysis resulted in the

identification of one major QTL located on chromosome XII, and a minor QTL located on

chromosome II. Besides this major QTL, two to three other QTLs were expected in CBS4C,

which may have been revealed by testing more segregants or defining more stringent

conditions for selection. For instance, the weakly linked region on chromosome II was

initially also confirmed by scoring the SNPs in the 20 individual segregants, the inclusion of

the 24 segregants with a somewhat weaker superior phenotype abolished the significance.

This seems to indicate that the causative gene in this QTL is additive to the major mutation in

SSK1 in further lowering glycerol production.

We identified a mutant SSK1 allele as a causative gene in the QTL on chromosome XII. Ssk1

is part of a “two component” signal transduction system, which signals via a multistep

phosphorelay mechanism from the plasma membrane osmo-sensor Sln1 to the redundant

MAP kinase kinase kinases (MAPKKK) Ssk2 and Ssk22 (Hohmann, 2002). Upon osmostress,

Ssk1 is rapidly dephosphorylated and the resulting protein binds with and activates Ssk2 and

Ssk22. Upon completion of osmoadaptation, dephosphorylated Ssk1 is degraded by the Ubc7-

dependent ubiquitin-proteasome system, causing downregulation of the HOG pathway (Sato

et al, 2003). The ssk1E330N…K356N

protein has lost part of its carboxyterminal end including the

110 Chapter III

response regulatory domain (amino acid position 505 - 647). However, at high osmolarity

Ssk1 is unphosphorylated, which enables the protein to interact with the non-catalytic

inhibitory domain of Ssk2 and Ssk22 MAPKKK (Maeda et al, 1995). Hence, one would

normally expect that absence of the functional regulatory domain causes overactivation of the

HOG pathway. In fact, hyperactivation of Hog, by deletion of negative regulators like PTC1

or PTP2 (Warmka et al, 2001; Wurgler-Murphy et al, 1997), is lethal in S. cerevisiae.

Therefore, the ssk1E330N…K356N

is probably not just a simple loss-of-function allele. This is

underscored by the observation that it does not cause the same effect as deletion of SSK1 for

glycerol/ethanol yield, ethanol productivity and growth in the CBS4C background.

A major goal of our work was to identify specific alleles, which would allow to reduce

glycerol yield without or with less-pronounced side-effects on osmotolerance or volumetric

ethanol productivity. The ssk1E330N…K356N

allele apparently fulfills these criteria. Introduction

of ssk1E330N…K356N

in the target strain Ethanol Red caused an even stronger reduction of

glycerol production than the deletion of SSK1, without causing a stronger effect on volumetric

productivity. Under anaerobic conditions glycerol production is required for reoxidation of

the NADH that is produced in glycolysis for the synthesis of intermediates that are withdrawn

for biosynthetic purposes. Under osmostress conditions, glycerol is produced as a compatible

osmolyte to prevent cellular dehydration. A possible explanation is that under both conditions,

the Ethanol Red strain produces an excess of glycerol and that therefore a partial reduction

can be achieved without seriously compromising ethanol productivity or osmotolerance.

The original diploid parent strain CBS6412 had two copies of the ssk1E330N…K356N

allele,

supporting that this mutation was also causative for the unusually low glycerol production in

Chapter III 111

that strain. This suggests that the mutant protein may provide a specific advantage in certain

environmental niches, for instance by enhancing ethanol production without compromising

osmostress tolerance. In this genetic background, the ssk1E330N…K356N

allele also did not

compromise growth as opposed to the deletion of SSK1. The precise origin of strain

CBS6412, however, is not known. The industrial strain Ethanol Red, with two copies of the

ssk1E330N…K356N

allele, showed a reduction in the glycerol/ethanol ratio, but not to the same

extent as in CBS6412. Identification of causative mutations in the other QTLs may allow

further reduction of the glycerol/ethanol ratio. The successful engineering of the industrial

target strain makes us conclude that the huge diversity of natural and industrial S. cerevisiae

strains available in culture collections may constitute a rich source of novel gene tools for

reverse metabolic engineering. In contrast to classical direct metabolic engineering, these

gene tools are not limited by our knowledge of the underlying molecular mechanisms.

112 Chapter III

4.5 Supplementary material

Supplementary Table 1: Primers used

No. Function Sequence

A-4770 SNP check Pair 1, CBS4C Chr II w 535197 GAAACTGAAACCAGGAGGAG

A-4735 SNP check Pair 1, ER7A Chr II w 535197 GAAACTGAAACCAGGAGGAA

A-4736 SNP check Pair 1, CBS4C Chr II c 535985 CTTTATGTAGTCTGGATTTTAG

A-4737 SNP check Pair 1, ER7A Chr II c 535985 CTTTATGTAGTCTGGATTTTAA

A-4738 SNP check Pair 2, CBS4C Chr II w 598260 TTCAAGTTAAATCGAATTGTAT

A-4739 SNP check Pair 2, ER7A Chr II w 598260 TTCAAGTTAAATCGAATTGTAC

A-4740 SNP check Pair 2, CBS4C Chr II c 599590 ATATCAATGTAAACACGTCA

A-4741 SNP check Pair 2, ER7A Chr II c 599590 ATATCAATGTAAACACGTCG

A-4742 SNP check Pair 3, CBS4C Chr II w 680091 GATATTAGTGTACATACGTTGC

A-4743 SNP check Pair 3, ER7A Chr II w 680091 GATATTAGTGTACATACGTTGA

A-4744 SNP check Pair 3, CBS4C Chr II c 681213 TTCCTTTTGAAGTGTCCTCG

A-4745 SNP check Pair 3, ER7A Chr II c 681213 TTCCTTTTGAAGTGTCCTCT

A-4746 SNP check Pair 4, CBS4C Chr XII w 169083 TCTCCATTACCAGCTGAA

A-4747 SNP check Pair 4, ER7A Chr XII w 169083 TCTCCATTACCAGCTGAG

A-4748 SNP check Pair 4, CBS4C Chr XII c 169876 GCATATATATATTTTAAGAAAATT

A-4749 SNP check Pair 4, ER7A Chr XII c 169876 GCATATATATATTTTAAGAAAATC

A-4750 SNP check Pair 5, CBS4C Chr XII w 198118 CAGAGTGGCAGACATTATCG

A-4751 SNP check Pair 5, ER7A Chr XII w 198118 CAGAGTGGCAGACATTATCA

A-4752 SNP check Pair 5, CBS4C Chr XII c 198478 GTAGCTGCCACAAAGCAC

A-4753 SNP check Pair 5, ER7A Chr XII c 198478 GTAGCTGCCACAAAGCAT

A-5009 SNP check Pair 6, CBS4C Chr II w 650007 TCTGCGTCTCTACGTTCTTG

A-5010 SNP check Pair 6, ER7A Chr II w 650007 TCTGCGTCTCTACGTTCTTA

A-5011 SNP check Pair 6, CBS4C Chr II c 650708 CACGCGTCGTTCTCGTC

A-5012 SNP check Pair 6, ER7A Chr II c 650708 CACGCGTCGTTCTCGTT

A-5013 SNP check Pair 7, CBS4C Chr II w 707209 ACATGTACACAAATCTTGAT

Chapter III 113

A-5014 SNP check Pair 7, ER7A Chr II w 707209 ACATGTACACAAATCTTGAC

A-5015 SNP check Pair 7, CBS4C Chr II c 708732 AGAAGAGAAGATCAAGCGTA

A-5016 SNP check Pair 7, ER7A Chr II c 708732 AGAAGAGAAGATCAAGCGTG

A-5017 SNP check Pair 8, CBS4C Chr II w 137510 CCCATTTTCTCGAATTGCAGA

A-5018 SNP check Pair 8, ER7A Chr II w 137510 CCCATTTTCTCGAATTGCAGG

A-5019 SNP check Pair 8, CBS4C Chr II c 138796 TGGCTTATGCAGGCGGTAAT

A-5020 SNP check Pair 8, ER7A Chr II c 138796 TGGCTTATGCAGGCGGTAAC

A-5049 SNP check Pair 9, CBS4C Chr II w 664541 GAATATGGATATGTAGCCAC

A-5050 SNP check Pair 9, ER7A Chr II w 664541 GAATATGGATATGTAGCCAT

A-5051 SNP check Pair 9, CBS4C Chr II c 665836 ATGGTCTTCAGAGGTCCC

A-5052 SNP check Pair 9, ER7A Chr II c 665836 ATGGTCTTCAGAGGTCCT

A-5053 SNP check Pair 10, CBS4C Chr II w 694207 CATGACAGTGAGTCTGAGTC

A-5054 SNP check Pair 10, ER7A Chr II w 694207 CATGACAGTGAGTCTGAGTT

A-5055 SNP check Pair 10, CBS4C Chr II c 695455 TTCAACAATCCTCAAAATCC

A-5056 SNP check Pair 10, ER7A Chr II c 695455 TTCAACAATCCTCAAAATCT

A-5057 SNP check Pair 11, CBS4C Chr XII w 155264 TCTTTTTGAGCTTAGGAGCG

A-5058 SNP check Pair 11, ER7A Chr XII w 155264 TCTTTTTGAGCTTAGGAGCA

A-5059 SNP check Pair 11, CBS4C Chr XII c 155773 GGTCCCCGTGTTTAAGAGTG

A-5060 SNP check Pair 11, ER7A Chr XII c 155773 GGTCCCCGTGTTTAAGAGTA

A-5061 SNP check Pair 12, CBS4C Chr XII w 179598 GGCACGTCATTATCGTCCAA

A-5062 SNP check Pair 12, ER7A Chr XII w 179598 GGCACGTCATTATCGTCCAG

A-5063 SNP check Pair 12, CBS4C Chr XII c 180560 GTTTTCCATTGCCGCTATTCT

A-5064 SNP check Pair 12, ER7A Chr XII c 180560 GTTTTCCATTGCCGCTATTCC

A-5118 SNP check Pair 13, CBS4C Chr XII w 162202 AAGTCCCTAATCTGCTTGGA

A-5119 SNP check Pair 13, ER7A Chr XII w 162202 AAGTCCCTAATCTGCTTGGC

A-5126 SNP check Pair 13, CBS4C Chr XII c 162909 GATCATATTTCTCCGGGGA

A-5127 SNP check Pair 13, ER7A Chr XII c 162909 GATCATATTTCTCCGGGCGA

A-5122 SNP check Pair 14, CBS4C Chr XII w 173098 GAGAAAGTCCATAAAGAATG

A-5123 SNP check Pair 14, ER7A Chr XII w 173098 GAGAAAGTCCATAAAGAATA

A-5124 SNP check Pair 14, CBS4C Chr XII c 174280 TCTTCACGGCAACTAATTTC

A-5125 SNP check Pair 14, ER7A Chr XII c 174280 TCTTCACGGCAACTAATTTT

A-5128 Sequencing SSK1 AAAAATTCTTTGTTTTAC

A-5129 Sequencing SSK1 AAGTTCGGCCGTTTTGTAT

114 Chapter III

A-5130 Sequencing SSK1 CCCACTCTGTAATTTTCT

A-5131 Sequencing SSK1 CGCCTTCCTTCCAAATAT

A-5132 Sequencing SSK1 ACATGGGTTGCGTTATGC

A-5133 Sequencing SSK1 CTTTACCTTTAAGATCTG

A-5134 Sequencing SSK1 AAGTGCAATCATTAACTG

A-5135 Sequencing SSK1 CGGCAAAGGCGTCGTATA

A-5136 Sequencing SSK1 TCAGTTTCTGAAAAAACC

A-5164 1st SSK1 deletion cassette, long flanking H1 site CACGATCACTCTTCCATCATA

A-5165 1st SSK1 deletion cassette, long flanking H1 site GGTTGTACCGTGTAGAGGACATT

A-5166* 1st SSK1 deletion cassette TCGTTACATTCTATCATAATGTCCTCTACACGGTA

CAACCCAGCTGAAGCTTCGTACGC

A-5167* 1st SSK1 deletion cassette

TTCGGCCGTTTTGTATAAGAAATATTGGAAAGGCT

GCTGTAAATCAAAAACGAATCGCATAGGCCACTA

GTGGATCTG

A-6113 2nd SSK1 deletion cassette CCCACTCTGTAATTTTCTTACTAAGCCAGTGTAAA

TTCACTGGTTTAGTCCAGCTGAAGCTTCGTACGC

A-6114 2nd SSK1 deletion cassette

TATACGACGCCTTTGCCGCCGTGCCGACAGTGGCC

GCGACTACCAATGTGGCATAGGCCACTAGTGGAT

CTG

A-5168 Verification of deletion, upstream of SSK1 TGCCAGTCAAGATTTCCCATA

A-5169 Verification of deletion, downstream of SSK1 TCCATGCCTATAATTATCGCGTTT

A-5894 Check primer AMD1 TCTCTAGCCTTCTGATAGGC

A-5895 Check primer AMD1 AGCCTTGGAATAGGTAGACG

A-6101 SSK1 insertion cassette, H1 TTTTTTGGTACCCACGATCACTCTTCCATCATA

KpnI

A-6103 SSK1 insertion cassette, H1 ATATATGTCGACGGTTGTACCGTGTAGAGGACATT

SalI

A-6100 SSK1 insertion cassette, SSK1 allele ATATATGTCGACCACGATCACTCTTCCATCATA

SalI

A-6102 SSK1 insertion cassette, SSK1 allele TTTTTTCCCGGGTCCATGCCTATAATTATCGCGTTT

XmaI

A-6770 SSK1 insertion cassette, pUG-AMD ATATATGTCGACGCATAGGCCACTAGTGGATCTG

SalI

A-7116 SSK1 insertion cassette, pAG25 (NAT1) ATGCGGCATCAGAGCAGATTGTA

A-7117 SSK1 insertion cassette, pAG25 (NAT1) TATATATACACGTAGTGGATCTG

DraIII

A-7118 SSK1 insertion cassette, pG119 (GIN11) CGAAATCGGCAAAATCCCTTAT

Chapter III 115

A-7119 SSK1 insertion cassette, pG119 (GIN11) TTCGCTCCTCTTTTAATGCCTTT

A-7301 Check primer NAT1 ACCCATCCAGTGCCTCGATG

Nucleotides in capital letters are derived from S. cerevisiae genome for homologous recombination at SSK1 gene locus. respectively.

Sequences with underlined letters functioned as primers for amplification of the deletion or promoter cassettes from respective plasmids

(Gueldener et al, 2002).

Supplementary Table 2: Plasmids used in this study

Plasmid Description Reference

pUG66 E. coli/ vector containing, Amp+, loxP-bleR-loxP

disruption cassette

Gueldener

et al., 2002

pFA6a-AMD1-MX6 vector containing AMD1 gene of Z.rouxii

pUG-AMD E. coli vector containing, Amp+, loxP-AMD1-

loxP disruption cassette This study

pNAT-Cre E. coli/ S. cerevisiae shuttle vector containing,

Amp+, NAT1, Cre recombinase This study

pBluescriptII SK(+) E.coli cloning vector Stratagene

pBluescriptII_AMD1_ssk1E330N…K356N

E. coli vector containing, Amp+, H1(repeat

region), AMD1 (selection/counterselction),

ssk1E330N…K356N of CBS4C

This study

pBluescriptII_NAT1_GIN11_ssk1E330N…K356N

E. coli vector containing, Amp+, H1(repeat

region), NAT1/GIN11 (selection/counterselction),

ssk1E330N…K356N of CBS4C

This study

pBluescriptII_AMD1_SSK1

E. coli vector containing, Amp+, H1(repeat

region), AMD1 (selection/counterselction), SSK1

of ER7A

This study

pBluescriptII_NAT1_GIN11_SSK1

E. coli vector containing, Amp+, H1(repeat

region), NAT1/GIN11 (selection/counterselction),

SSK1 of ER7A

This study

116 Chapter III

Calculations

Yields

Absolute yield

Relative yield

Carbon balance

The yields, based on product mass, were transferred into carbon based yields, using the

following coefficients:

Biomass 25.356 g/Cmol1 Ethanol 23 g/Cmol

Glycerol 30.66 g/Cmol Acetat 30 g/Cmol

CO2 44 g/Cmol Glucose 30 g/Cmol

The amount of CO2 produced during the fermentation resulted from the difference of the

initial weight and the final weight.

Calculation of the volumentric production rate

The fermentation tubes were weighed throughout the fermentation process to calculate the

relative weight loss m[%].

m % m0 mt

m0

100mCO2

m0

100

A curve was fit to the weight loss over time using a polynomial function of 8th

degree. The

derivate of this function resulted the normalized maximum fermentation rate [h-1

] during the

time of the fermentation.

Massbalance batch - fermentation:

dmCO2

dt = rCO2

V rCO2

1

V

dmCO2

dt

rCO2 =

1

V

m

t

Assumption => CO2 only produced together with EtOH (maximum yield)

Molar production rates of CO2 and EtOH are equal

1 Biomass formula used to convert dry weights into molar carbon concentrations was C1H1.79O0.64N0.16 (ENSC,

Toulouse).

YS / i cifinal

(csini cS

final)

Y% YS / i

segr

YS / i

CBS4C

Chapter III 117

rCO2

1

V

dnCO 2

dt rEtOH

1

V

dnEtOH

dt

nCO2 =

mCO2

MWCO2

1

nEtOH = mEtOH

MWEtOH

2

2

1 1 nEtOH

nCO2

mEtOH

MW CO 2

MWEtOH

mCO2

mEtOH mCO2

MWEtOH

MWCO 2

Use the weightloss [%]:

m % m0 m t

m0

100 mCO2

m0

100

mCO2

m % m0

100

rEtOH 1

V

dmCO2

dt

1

V

d mCO2

MW EtOH

MWCO 2

dt

rEtOH 1

V

MW EtOH

MW CO 2

dmCO2

dt

rEtOH 1

V

MW EtOH

MW CO2

m0

100

dm[%]

dt

maximum volumetric productivity rmax

Chapter IV

Identification of multiple alleles

conferring low glycerol and high ethanol yield

in Saccharomyces cerevisiae

ethanolic fermentation.

120 Chapter IV

1. Abstract

Genetic engineering of complex traits can make use of multiple strategies to create a desired

phenotype; however many engineering approaches suffer from undesirable side effects on

essential functions. A precedent-setting example in Saccharomyces cerevisiae is the reduction

of glycerol synthesis, which has been engineered in manifold ways for the purpose of

increasing ethanol yield. Many approaches relied on the manipulation of one or two genes

with marginal success because of unexpected side-effects; however complex traits must be

addressed by combinatorial gene modifications allowing for compatibility with normal

cellular functionality. Studying Saccharomyces cerevisiae biodiversity for specific alleles

causing lower glycerol and higher ethanol yield can provide new gene tools to design such

combinatorial engineering approaches. We previously identified ssk1E330N…K356N

as causative

allele in strain CBS6412, which displays a low glycerol and high ethanol yield. We have now

identified a single segregant, 26B, that lacks ssk1E330N…K356N

, and still shows similar low

glycerol/high ethanol production as the superior parent. Using segregants from the backcross

of 26B with the inferior parent strain, we applied pooled-segregant whole-genome sequencing

and identified three minor QTLs linked to low glycerol and high ethanol yield. Three alleles

known to be involved in regulation or metabolism of glycerol, smp1R110Q,P269Q

, hot1P107S,H274Y

and gpd1L164P

were identified as causative genes. All three genes separately caused a

significant drop in the glycerol yield in yeast fermentation, while gpd1L164P

was epistatically

suppressed by the other two alleles. Our results show that natural yeast strains harbor multiple

specific alleles controlling glycerol yield. These alleles can be used as gene tools for

engineering industrial yeast strains, minimizing the risk of negatively affecting other essential

functions. These gene tools can act at the transcriptional, regulatory and/or structural gene

level, distributing the impact over multiple targets and thus minimizing possible side-effects.

In addition, the results suggest complex trait analysis as a promising new avenue to identify

new components involved in cellular functions.

Chapter IV 121

2. Bibliographic reference

Hubmann, G., Mathé, L., Foulquié-Moreno, MR, Duitama, J, Nevoigt, E., and Thevelein, JM.

Identification of multiple alleles conferring low glycerol and high ethanol yield in

Saccharomyces cerevisiae ethanolic fermentation

(Manuscript submitted)

Duitama, J., Sanchez-Rodriguez, A., Goovaerts, A., Pulido-Tamayo, S., Hubmann, G.,

Foulquie-Moreno, MR., Thevelein, JM., Verstrepen, K. and Marchal, K.

Improved linkage analysis of Quantitative Trait Loci using bulk segregants

unveils a novel determinant of high ethanol tolerance in yeast

(Manuscript submitted)

3. Scientific contribution

The author participated in the project conception, experimental work, data analysis and article

writing.

122 Chapter IV

4. Manuscript III: Identification of multiple alleles conferring low

glycerol and high ethanol yield in Saccharomyces cerevisiae ethanolic

fermentation

4.1 Introduction

Glycerol formation is of great importance in the yeast S. cerevisiae and serves two major

physiological functions: i) the regulation of cell turgor especially under high osmolarity and

ii) recycling of the cofactor NAD+ to maintain the redox balance, particularly in the absence

of oxygen. In the first case, glycerol is produced as a compatible osmolyte during osmostress.

The HOG pathway mediates the stimulation of glycerol production during osmostress and has

been elucidated and characterized in great detail (Hohmann, 2002). It involves osmosensing

proteins at the level of the plasma membrane, a MAP kinase signalling pathway and

transcription factors that regulate expression of stress related target genes important for proper

adaptation to osmotic stress. In the second case, glycerol is produced to maintain the redox

balance during anaerobic growth. This is due to the fact that biosynthesis results in an excess

of NADH, which mainly is regenerated through glycerol formation from dihydroxy aceton 3-

phosphate (DHAP). Besides CO2, glycerol is the main by-product in the alcoholic

fermentation. It is synthesized by the consecutive action of glycerol 3-phosphate

dehydrogenase, encoded by GPD1 and GPD2, and glycerol 3-phosphate phosphatase,

encoded by GPP1 and GPP2 (Albertyn et al, 1994; Ansell et al, 1997).

Glycerol formation is a complex quantitative trait and is highly variable among isolates of

Saccharomyces cerevisiae (Hubmann et al, 2013). Low glycerol formation is essential for

maximal yield in bioethanol production (Bro et al, 2006; Nissen et al, 2000) while high

glycerol formation is important for production of wines with a good mouth-feel and a reduced

alcohol level (Schmidtke et al, 2012; Schuller and Casal, 2005). Rational genetic engineering

of glycerol production by modification of the main structural gene, GPD1, encoding glycerol-

3-phosphate dehydrogenase, the rate limiting enzyme of glycerol synthesis; has not been

successful in obtaining industrial yeast strains with low glycerol and high ethanol yield

because of the strong side-effects caused on other phenotypic traits. Deletion and even

Chapter IV 123

reduced expression of GPD1 lowers growth and fermentation rates (Bjorkqvist et al, 1997;

Guadalupe Medina et al, 2010; Hubmann et al, 2011; Nissen et al, 2000; Pagliardini et al,

2010) while overexpression causes redox imbalance and overproduction of acetate (Nevoigt

and Stahl, 1996; Remize et al, 1999; Schmidtke et al, 2012; Schuller et al, 2005). Hence,

studying glycerol formation as complex trait may provide new insights, gene targets and lastly

gene tools for more directed sustainable engineering of glycerol formation without causing

negative side-effects on other essential traits.

Like many other industrially relevant traits in yeasts, glycerol formation has a complex

phenotype/genotype relationship, meaning polygenic inheritance and environmental

variability. So far, such traits are poorly understood with regards to their regulatory network

and gene - gene interactions. For complex traits, knowledge based engineering has been

difficult and mainly addressed through targeted deletion and/or overexpression of one

structural or regulatory gene, which very often results in undesirable side-effects on other

essential functions (Albertyn et al, 1994; Guadalupe Medina et al, 2010; Nissen et al, 2000).

In awareness of these difficulties, strain improvement on complex traits has mainly been

achieved through non-targeted approaches, which do not require a priori knowledge about

gene targets. Such approaches, also referred as random engineering strategies, were the key

driver in microbial engineering (Nevoigt, 2008; Oud et al, 2012) and resulted in most of

today’s commercially used yeasts. Despite these easy to use ‘black box’ approaches, non-

targeted improvements are laborious with a high risk of affecting other traits and the acquired

improvements can also not be transferred to other strains. Hence, identification of causative

genes and gene - gene interactions is essential to overcome these drawbacks of non-targeted

approaches (Oud et al, 2012). Before the ‘-omics’ era, genome-scale methodologies for

identification of causative genetic determinants were limited and therefore this was a major

bottleneck in non-targeted approaches and reverse engineering (Bailey et al, 1996). The

developments of genome scale ‘-omics’ technologies facilitate such identification. For

instance, locating or mapping causative determinants is possible by analyzing the inheritance

of genomic regions in offspring of parents, where one of the two exhibits the superior

phenotype. An offspring population selected for the superior phenotype shows a deviation

from the usual 50% inheritance in the linked genomic regions. A genomic region with such

deviation is called a quantitative trait locus (QTL). Several genetic markers, the most

124 Chapter IV

prominent being the single nucleotide polymorphisms (SNPs), have been used to score

inheritance of both parental types. Recent advances in sequencing technologies facilitated

SNP scoring in genome-wide mapping studies allowing for simultaneous detection of

multiple linked QTLs over the whole genome (Brem et al, 2002; Deutschbauer and Davis,

2005; Steinmetz et al, 2002; Swinnen et al, 2012a; Winzeler et al, 1998). Moreover, methods

for rapid dissection of the mapped QTLs, such as reciprocal hemizygosity analysis, were

developed to identify the causative genes in the locus (Steinmetz et al, 2002). Certainly, QTL

analysis became very powerful for identification of genetic determinants using non-targeted

approaches (Hubmann et al, 2013). So far, major causative genes determining a trait were

easily identified and were even mapped to the gene level, using several random inbreeding

steps (Parts et al, 2011). Reliable identification and analysis of minor QTLs and their

causative genes has remained challenging because of weak linkage and the small contribution

to the phenotype, which are easily masked by major causative genes and/or can be replaced

by other minor causative genes. This often results in the insignificance of the minor QTLs. It

remains highly important in genetic analysis of complex traits to develop alternative

methodologies to analyze minor QTLs in an efficient and reliable way. One strategy for

identification of minor QTLs is the sequential elimination of the causative gene identified in

the major QTL using a backcross with a segregants that lacks this causative gene but shows

the phenotype of interest (Birkeland et al, 2010; Demogines et al, 2008; Sinha et al, 2008). A

disadvantage of this strategy is that the phenotypic difference between parent and segregants

often becomes smaller and that therefore larger numbers of segregants are required for

reliable phenotyping and QTL mapping. Another strategy to identify minor QTLs is to

increase the stringency of phenotypic screening. Swinnen et al. (2012a) showed that selection

of yeast segregants tolerant to 17% ethanol versus 16% ethanol, strengthened the linkage of

several minor QTLs, facilitating their further analysis. However, this methodology also results

in higher numbers of segregants being required in the phenotypic screening.

We conducted QTL mapping of the trait of ‘low glycerol yield’, for which we previously

identified the sake yeast CBS6412, as a ‘low glycerol’ producer (Hubmann et al, 2013). The

glycerol yield of this strain was about 63% of that of Ethanol Red, a commercial strain

commonly used for bioethanol production. CBS6412 and Ethanol Red were chosen as the

parent strains to conduct QTL mapping of ‘low glycerol’ as trait of interest. The two haploid

Chapter IV 125

segregants, CBS4C and ER7A, were closest in terms of glycerol yield to their parental diploid

progenitors CBS6412 and Ethanol Red. Downscaling of the major locus on chromosome XII

and reciprocal hemizygosity analysis identified and validated ssk1E330N…K356N

, a SSK1 mutant

allele, encoding a truncated, partially mistranslated protein, as causative for the low-glycerol

phenotype. Although the ssk1E330N…K356N

allele was the major effector, it did not completely

confer the ‘low glycerol’ phenotype observed in CBS6412.

In the present paper, we present a targeted backcross (Demogines et al, 2008; Sinha et al,

2008) of a segregant of the cross CBS4C/ER7A, which was selected for the phenotype of

interest, ‘low glycerol’ yield, but lacked the ssk1E330N…K356N

allele. Hence, in the backcross

with the inferior parent ER7A the ssk1E330N…K356N

allele was entirely absent. Mapping by

pooled-segregant whole-genome sequence analysis and reciprocal hemizygosity analysis led

to the identification in minor QTLs of three new causative alleles of known genes in glycerol

metabolism and its regulation, each by itself causing a reduction in glycerol yield.

4.2 Materials and Methods

Microbial strains and cultivation conditions

All S. cerevisiae strains used are listed in Table 1. E. coli strain DH5αTM

(Invitrogen Corp.,

Carlsbad) was used for amplification of plasmids. The strain was grown in Luria-Bertani (LB)

medium containing 0.5% (w/v) yeast extract, 1% (w/v) Bacto tryptone, 1% (w/v) NaCl, (pH

7.5) at 37˚C. E. coli transformation and isolation of plasmid DNA was carried out using

standard techniques (Sambrook et al, 1989). Transformants were selected on LB medium

containing 100µg/ml ampicillin.

Table 1 Saccharomyces cerevisiae strains used.

Strain Genotype Source

CBS6412 Diploid, ssk1E330N…K356N/ssk1E330N…K356N CBS-KNAW

Ethanol Red Diploid, SSK1/SSK1

Fermentis, S. I.

Lesaffre

ER7A Segregant 7A of Ethanol Red, Matα This study

126 Chapter IV

CBS4C Segregant 4C of CBS6412, Mata, ssk1E330N…K356N This study

26B Segregant of the cross ER7A x CBS4C, Matα, SSK1 This study

26B MatA Mating type switch of 26B to MatA This study

26B x ER7A Hybrid diploid 26B x ER7A This study

26B smp1∆ x ER7A Hybrid diploid 26B smp1∆ x ER7A

This study

26B x ER7A smp1∆ Hybrid diploid 26B x ER7A smp1∆ This study

26B gpd1∆ x ER7A Hybrid diploid 26B gpd1∆ x ER7A This study

26B x ER7A gpd1∆ Hybrid diploid 26B x ER7A gpd1∆ This study

26B hot1∆ x ER7A Hybrid diploid 26B hot1∆ x ER7A This study

26B x ER7A hot1∆ Hybrid diploid 26B x ER7A hot1∆ This study

BY 4742 gpd1∆ Ycplac33 Haploid, gpd1∆, Ycplac33 This study

BY 4742 gpd1∆ Ycplac33

GPD1-ER7A

Haploid, gpd1∆, Ycplac33 GPD1-ER7A This study

BY 4742 gpd1∆ Ycplac33

gpd1L164P-CBS4C

Haploid, gpd1∆, Ycplac33 gpd1L164P-CBS4C This study

26B gpd1∆ Ycplac33 Haploid, ura3∆, gpd1∆, Ycplac33 This study

26B gpd1∆ Ycplac33 GPD1-

ER7A

Haploid, ura3∆, gpd1∆, Ycplac33 GPD1-ER7A This study

26B gpd1∆ Ycplac33 gpd1L164P-

CBS4C

Haploid, ura3∆, gpd1∆, Ycplac33 gpd1L164P -

CBS4C

This study

ER7A gpd1∆ Ycplac33 Haploid, ura3∆, gpd1∆, Ycplac33 This study

ER7A gpd1∆ Ycplac33 GPD1-

ER7A

Haploid, ura3∆, gpd1∆, Ycplac33 GPD1-ER7A This study

ER7A gpd1∆ Ycplac33

gpd1L164P-CBS4C

Haploid, ura3∆, gpd1∆, Ycplac33 gpd1L164P -

CBS4C

This study

26B x ER7A gpd1∆/∆ Ycplac33 Haploid, ura3∆/∆, gpd1∆/∆, Ycplac33 This study

26B x ER7A gpd1∆/∆ Ycplac33

GPD1-ER7A

Haploid, ura3∆/∆, gpd1∆/∆, Ycplac33 GPD1-ER7A This study

Chapter IV 127

26B x ER7A gpd1∆/∆ Ycplac33

gpd1L164P-CBS4C

Haploid, ura3∆/∆, gpd1∆/∆, Ycplac33 gpd1L164P -

CBS4C

This study

Mating, sporulation and internal crosses

Mating and sporulation were carried out according to standard procedures (Sherman and

Hicks, 1991). Mating type of segregants was determined by diagnostic PCR for the MAT

locus (Huxley et al, 1990). Meiotic spores of the cross 26B/ER7A were isolated by random

spore analysis as described by Treco and Winston (2001). The hybrid 26B/ER7A strain was

plated on a sporulation plate to initiate sporulation for a period of about two weeks up to 1

month until asci were observed. The yeast cells were washed off the sporulation plate and

suspended in 25 ml of MilliQ water, in a 300 ml Erlenmeyer flask together with sterile 0.45

mm glass beads. 500 µl Zymolase (10mg/ml) and 10 µl of β-mercaptoethanol were added to

the cell suspension in order to degrade the asci. This cell suspension was incubated overnight

at 30˚C by shaking at 200 rpm. The cell suspension was transferred to a 50 ml tube together

with the glass beads, and shaken vigorously. The cell debris was centrifuged at 20000 rpm for

20 min. The supernatant was discarded and the pellet was suspended in 5 ml of Nonidet P-40

and placed on ice for 15 min. This was followed by 4 rounds of sonication (30 s, 75%). The

cell suspension rested 2 min on ice between two rounds. After that, the cell suspension was

centrifuged for 10 min at 3000 rpm, the supernatant was discarded, and the cells were re-

suspended in 1.5% Nonidet P-40. This procedure was repeated once, followed by 4 more

rounds of sonication and incubation on ice. Lastly, the cell suspension was centrifuged for 10

min at 3000 rpm. The supernatant was discarded, and cells were re-suspended in 300 ml of

MilliQ water. The cell suspension was diluted to obtain single colonies on plates. Plates were

incubated at 30˚C until single colonies were visible. These single colonies were re-plated and

checked for mating type to confirm haploidy. Usually, this procedure yielded 90% haploids.

Internal crossing of 26B and ER7A was carried out as follows. First, diploids were isolated

from the cross between ER7A and 26B. For this purpose, the mating type of single colonies

resulting from the cross between ER7A and 26B was checked. In the first step of internal

crossing, the diploids were streaked out on a sporulation plate to initiate sporulation until

sufficient asci were visible. In the second step, spores were isolated using random spore

analysis (see further). The isolated spores were all plated on YD and incubated for 2 days at

128 Chapter IV

30˚C to ensure that enough diploids had formed. In the last step, newly formed diploids were

transferred to a new sporulation plate to start the next cycle of internal crossing.

Fermentation conditions

The 26B, ER7A and CBS4C strains were tested in two fermentations: minimal medium and

YP 10% glucose. The minimal medium was composed of 1.9 g l-1

yeast nitrogen base (Difco),

5 g l-1

ammonium sulphate, 250 mg l-1

leucine, 50 mg l-1

uracil, 100 mg l-1

histidine, 30 mg l-1

lysine, 20 mg l-1

methionine and 50 g l-1

glucose. The inoculum was prepared in minimal

medium containing 2% (w/v) glucose to inoculate the fermentation with an initial OD of 1.

Fermentations were carried out in Erlenmeyer flasks, which were equipped with air locks,

ensuring the exclusion of oxygen but allowing the release of CO2. The fermentations were

performed at 30˚C and cultures were continuously stirred at 200 rpm.

The parental strains, ER7A and 26B, as well as CBS4C, were additionally tested in YPD,

which contained 0.2% (w/v) yeast extract, 0.6% (w/v) peptone, and 10% (w/v) glucose. The

free nitrogen content present in YPD was adjusted to that present in wheat liquefact. The

fermentations were carried out in cylindrical glass tubes, which were closed with a rubber

stopper containing a glass pipe, sealed off with a cotton plug to release CO2. 100 ml

fermentation medium was added to each tube. Inoculum cultures were grown statically

overnight at 30˚C in 5 ml of YD medium, one day ahead of the fermentation and this culture

was used completely to inoculate the 100ml fermentation tube. The empty weight, starting

weight and weight after 72h fermentation of the tubes was measured to determine the net

weight loss, due to conversion of sugar to carbon dioxide, during the fermentation.

Fermentations ran for 72 hours at 30˚C and were stirred at 200 rpm. After the 72 hours, the

final weight was determined. The fermentation broth was cooled for 1 night at 4˚C prior to the

analysis of glycerol and ethanol concentration, in order to minimize evaporation of ethanol

during the sample taking. The fermentations of the initial screen of 26B/ER7A segregants,

those of the reciprocal hemizygosity analysis, and those of the gpd1∆ strains were carried out

in 100 ml YP 10% glucose.

Screening of additional selected segregants was downscaled to a 5ml fermentation. The pre-

culture was started one day ahead of the fermentation in 3 ml of YD medium. Cultures were

Chapter IV 129

grown statically overnight at 30˚C. The next day, the pre-cultures were used to inoculate the

5ml fermentations at a rate of 1/20. Fermentations were kept for 96h at 30˚C and afterwards

placed at 4˚C overnight prior to the analysis of the fermentation broth in order to prevent

ethanol evaporation.

Determination of fermentation parameters

In all fermentations weight loss was used to follow the progress of the fermentation. Glucose,

glycerol and ethanol in the medium were determined by HPLC (Waters® isocratic BreezeTM

HPLC, ion exchange column WAT010290). Column temperature was 75˚C, 5 mM H2SO4

was used as eluent with a flow rate of 1 ml min-1

and refractive index detection was used

(Waters, 2414 RI detector). The product yield was calculated from the final product

concentration (g l-1

) and the difference in glucose concentration at the start and end of the

fermentation (consumed glucose in g l-1

). Yields of screenings, RHA and the gpd1∆ strain

constructs were related to the yield of 26B in order to decrease variance between different

experiments.

DNA methods

Yeast genomic DNA was extracted with Phenol/Chloroform/Isoamyl-alcohol (25:24:1)

(Hoffman and Winston, 1987) and further purified with diethyl-ether extraction or ethanol

precipitation if required. PCR was performed with high-fidelity polymerases PhusionTM

(Finnzymes) or ExTaqTM

(TaKaRa) for cloning and amplification of deletion or insertion

cassettes, and sequencing purposes. Sequencing was carried out using the dideoxy chain-

termination method (Sanger and Coulson, 1975) at the VIB Genetic Service Facility

(Antwerp). The sequences were analyzed with geneious (Geneious Basic 5.3.4), SeqMan

(Lasergene Coresuite 8) or CLC DNA workbench (CLC bio) software.

Pooled-segregant whole-genome sequence analysis

The segregant 26 was isolated from the cross of CBS4C and ER7A. This segregant was

backcrossed with its own parent ER7A. From this backcross, the 22 most superior segregants

130 Chapter IV

(lowest glycerol production) were assembled in the 'selected pool' while 22 random

segregants were used to assemble the 'unselected pool'. The two pools were made by

combining equal amounts of cells based on OD600. High molecular weight DNA (3 µg, ~

20kb fragments) was isolated from the pools and parent strains according to Johnston and

Aust (1994). The purity of the DNA sample was estimated from UV measurement (260/280 =

1.7 - 2.0). The DNA samples were provided to BGI (Hong Kong, China) for whole-genome

sequence analysis by Illumina technology.

Mapping of short read sequences, variant calling and QTL analysis were carried out as

described earlier by Swinnen et al. (2012a) and by Hubmann et al. (2013). The SNP variant

frequencies were calculated by dividing the number of the alternative variant by the total

number of aligned reads. A very high or a very low frequency was a sign of a one-sided SNP

segregation preferentially coming from one parent, indicating a genetic linkage to the trait of

interest. Genetic linkage was statistically confirmed using EXPloRA (Duitama et al.

submitted for publication) or the methods described earlier (Swinnen et al, 2012a).

Detection of SNP markers by allele specific PCR

Individual SNPs were scored by PCR. The forward and reverse primer contained the

nucleotide of ER7A or CBS4C as the 3’ terminal nucleotide, respectively. The annealing

temperature was optimized using DNA extracts of ER7A and CBS4C so as to allow only

hybridization with primers containing an exact match.

Reciprocal hemizygosity analysis (RHA)

For RHA analysis (Steinmetz et al, 2002), two diploid strains were constructed by crossing

26B and ER7A wild type or deletion strains for the candidate gene, so that the resulting

diploids only contained a single allele from either 26B or ER7A for the candidate gene being

evaluated. The SMP1, HOT1 and GPD1 deletions were carried out as described by Gueldner

et al. (2002) using the bleR marker conferring phleomycine resistance. After the

transformation, the gene deletions were verified by PCR using primers outside the deleted

gene. In case of SMP1 and GPD1 the deletion cassette enlarged the PCR fragment indicating

Chapter IV 131

the integration of the bleR cassette in the locus. In case of HOT1 the deletion cassette was

smaller than the deleted fragment of HOT1, thus a decrease in the PCR fragment showed

correct deletion. Each deletion strain was crossed with the wild type strain of the other parent,

resulting in two hybrid 26B/ER7A diploids. The correct genotype of these diploids was

reconfirmed by the presence of two PCR fragments, one resulting from the wild type allele

the other from the deleted allele. RHA was performed with three independent isolates of all

tested diploids.

Construction of YCplac GPD1 plasmids

The primers, A-3709 and A-3743, were used to PCR amplify genomic DNA of CBS4C and

ER7A chr. IV (410523 - 413479), containing the promoter, GPD1 ORF, and terminator. The

resulting PCR fragment was digested with KpnI, purified, and ligated to the plasmid

YCplac33, which was digested with KpnI prior to ligation. The constructs YCplac33 GPD1-

ER7A and YCplac33 gpd1L164P

-CBS4C were verified using Sanger sequencing. URA3 was

deleted in the strain 26B gpd1∆ and ER7A gpd1∆. Both strains were mated to obtain the

ER7A x 26B gpd1∆/∆ ura3∆/∆. The deletion cassette was constructed as described by

Gueldner et al. (2002) with the geneticin resistance marker KanMX6. Transformants were

selected for the two selectable markers, phleomycin and geneticin, to avoid a cassette switch.

URA3 gene deletion was confirmed by PCR and absence of growth on SD-ura plates. The

YCplac33 GPD1-ER7A, YCplac33 gpd1L164P

-CBS4C and the empty plasmid were transferred

to the strain BY4742 gpd1∆, 26B gpd1∆ ura3∆, ER7A gpd1∆ ura3∆, and ER7A/26B

gpd1∆/∆ ura3∆/∆, using the LiAc/PEG transformation method (Gietz et al, 1992).

Data deposition

Sequencing data have been deposited at the SRA database (NCBI),

http://www.ncbi.nlm.nih.gov/sra, with the account number SRA059109.

132 Chapter IV

4.3 Results

Selection of a segregant without the ssk1E330N…K356N

allele

Previous work has identified the Saccharomyces cerevisiae strain CBS6412 as a strain with

an unusually low glycerol yield and a concomitant higher ethanol yield. Genetic analysis

identified the ssk1E330N…K356N

mutation as a major causative mutation (Hubmann et al 2013)

(Figure 1A). Among the 44 segregants of the cross CBS4C/ER7A, which were selected for

‘low glycerol’, only a single segregant, 26B, was present, in which ssk1E330N…K356N

was absent

(Figure 1A). Despite the absence of ssk1E330N…K356N

in this segregant, its glycerol yield was

equally low and its ethanol yield equally high as the superior parent strain CBS4C, both in

minimal medium with 5% glucose and rich yeast extract-peptone medium with 10% glucose

(Figure 1B). Hence, 26B showed the same phenotypic difference with the inferior parent

strain ER7A as CBS4C (Figure 1B). In this work, the fermentation medium, referred to as YP

10% glucose, was modified to adapt the conditions to more industrial-like conditions. In the

previous study, fermentations were carried out in minimal medium, which results in higher

glycerol yields when compared to those obtained in industrial scale. Moreover, the elevated

sugar concentration caused a higher osmotic pressure leading to the induction of osmostress

genes and induced glycerol formation. In this way, we increased the selection strength on the

low glycerol phenotype allowing for the detection of weakly linked genes, which confer both

a low glycerol production phenotype as well as osmo-tolerance.

Targeted backcross with the inferior parent ER7A

The mating type of segregant 26B was switched from Matα to MatA by induction of the HO-

gene located on the plasmid pFL39 GAL1 HO KanMX to allow mating with the inferior

parent ER7A (Matα). Glycerol yield of the hybrid diploid, 26B/ER7A showed an intermediate

phenotype between ER7A and 26B (Figure 1B). The ethanol yield obtained in YP 10%

glucose distinguished more clearly between the four strains when compared to minimal

medium. Strain ER7A showed the lowest ethanol yield (0.475 g g-1

) among all tested strains.

The ethanol yield was higher in the diploid 26B/ER7A (0.481 g g-1

) and the segregant 26B

(0.484 g g-1

) and reached the maximum for the CBS4C strain (0.488 g g-1

). 260 meiotic

segregants were characterized for glycerol and ethanol yield in fermentations with 100 ml YP

Chapter IV 133

10% glucose. 26B, ER7A and the hybrid diploid were used as controls in each batch of

fermentations.

Figure 1 Phenotypes of the parental strains ER7A and CBS4C and the segregant 26B. (A) Scheme of the

crossings to map mutations linked to the low glycerol yield phenotype. The initial parental cross of ER7A and

CBS4C resulted in the segregant 26B with a low glycerol phenotype but without the ssk1E330N…K356N allele. The

26B segregant was crossed back with the inferior parent ER7A to find other linked mutations. (B) Glycerol and

ethanol yield obtained in minimal medium with 5% glucose and in YP 10% glucose for the parental strains,

ER7A and CBS4C, the segregant 26B, and the zygote 26B x ER7A.

Glycerol and ethanol yield of the segregants in each batch were normalized to those of 26B,

which were set to 100%. ER7A and the diploid 26B/ER7A showed an average glycerol yield

of 146% and 124% and a concomitantly decreased ethanol yield of 98.1% and 99.4% (Figure

2A). The glycerol and ethanol yield showed a Gaussian distribution amongst the segregants,

which extended over the range of the two parental stains. Both population means were located

closely to the hybrid diploid 26B/ER7A. In general, glycerol and ethanol yield of the

134 Chapter IV

segregant population correlated inversely (as determined with a Pearson test), meaning that

low glycerol yield usually resulted in high ethanol yield. Exceptions to this correlation were

strains with an unusually low ethanol yield that failed to show a correspondingly higher

glycerol yield. Application of two criteria, a glycerol yield lower than 120% of 26B and an

ethanol yield higher than 99% of 26B, resulted in the selection of a set of 34 superior

segregants. We re-checked the 34 segregants showing a glycerol yield below 120% of 26B in

100 ml fermentations; 22 segregants showed a stable relative glycerol yield combined with a

high ethanol yield (Figure 2B). These 22 segregants were selected for QTL mapping with

pooled-segregant whole-genome sequence analysis. A second pool with 22 randomly selected

segregants was also subjected to pooled-segregant whole-genome sequence analysis and used

as the unselected control pool (Figure 2B).

Figure 2 Glycerol and ethanol yield in parental strains, hybrid diploid and segregants. (A) Glycerol and

ethanol yield in the parental strains, 26B (■) and ER7A (▲),the hybrid diploid strain 26B/ER7A (●) and in

segregants of 26B/ER7A (◯). Fermentations were carried out in 100 ml YP with 10% glucose. Glycerol and

Ethanol yields of all segregants, ER7A and the diploid 26B/ER7A were related to the yield of 26B, which was

set as 100%. (B,C) Distribution of the glycerol and ethanol yield in the unselected (B) and selected (C) segregant

pool of 26B/ER7A. The criteria for selection of “low glycerol” segregants (<120% glycerol yield, >99% ethanol

yield) are indicated with stippled lines. The 22 selected segregants were fermented twice to confirm low glycerol

production. These segregants were used for pooled-segregant whole-genome sequence analysis. The glycerol

and ethanol yield of the parental strains, 26B and ER7A, and diploid 26B/ER7A are indicated as in (A).

Chapter IV 135

Pooled-segregant whole-genome sequence analysis and QTL mapping

The genomic DNA of the selected and unselected pools, as well as the parent strain 26B, was

extracted and submitted to custom sequence analysis using Illumina HiSeq 2000 technology

(BGI, Hong Kong, China). The parent strain ER7A has been sequenced in our previous study

(data accession number SRA054394). Read mapping and SNP filtering were carried out as

described previously (Hubmann et al 2013). The SNP variant frequency was plotted against

the SNP chromosomal position (Figure 3). Of the total number of 21,818 SNPs between

CBS4C and ER7A, 5,596 SNPs of CBS4C were found back in 26B. These SNPs were used

for mapping minor QTLs in the genomic areas that were not identical between 26B and

ER7A. The other genomic areas were completely devoid of SNPs because they were identical

between the 26B and ER7A parents (Figure 3). The scattered raw SNP variant frequencies

were smoothened and a confidence interval was calculated, as previously described (Swinnen

et al, 2012a). The Hidden-Markov Model, EXPloRA (Duitama et al. submitted for

publication), was used to evaluate whether candidate regions showed significant linkage to

the low glycerol phenotype. EXPloRA reported six candidate QTLs: on chr. I (3859-11045),

chr. II (584232-619637), chr. IV (316389-375978 and 696486-748140), and chr. XIII

(600902-610995 and 634582-640415) for the selected segregant pool. The locus on chr. I was

present in both the selected and unselected pool and was thus likely linked to an inadvertently

selected trait, such as sporulation capacity or spore viability. It was excluded from further

analysis. The locus on chr. II was also present in the previous mapping with the two original

parents, CBS4C and ER7A, but the linkage was not significant (Hubmann et al 2013). The

backcross has now confirmed its usefulness. On chr. IV and XIII, new QTLs were detected,

which were not present in the mapping with the original parent strains CBS4C and ER7A.

EXPloRA also reported two significantly linked loci on chr. VI (169586-170209) and chr. VII

(472620-493523) for the unselected pool. Both loci were linked to the inferior parent, ER7A.

For the region on chr. VII, the linked locus with the inferior parent genome was also present

in the selected pool. Both loci likely represent linkage to inadvertently selected traits, such as

sporulation capacity or spore viability. It is unclear why the locus on chr. VI was only present

in the unselected pool. Since both loci were not linked to the low glycerol phenotype they

were not investigated further.

136 Chapter IV

Figure 3 Plots of SNP variant frequency versus chromosomal position and corresponding probability of

linkage to the superior or inferior parent. Plots of SNP variant frequency versus chromosomal position of all

16 yeast chromosomes for the selected (raw data: light grey triangles; smoothened data: red line) and unselected

pool (raw data: light grey circles; smoothened data: green line). Significant upward deviations from the average

of 0.5 indicate linkage to the superior parent 26B, while significant downward deviations indicate linkage to the

inferior parent ER7A. The smoothened line was determined as described previously (Swinnen et al, 2012a).

Linked regions were detected with EXPLoRA (Duitama et al. submitted for publication).

The QTLs on chr. II, IV and XIII were further investigated in detail. Selected individual SNPs

were scored in the 22 individual superior segregants to determine the SNP variant frequency

precisely and determine the statistical significance of the putative linkage. However, using the

binomial test previously described (Swinnen et al, 2012a) none of the three loci was found to

Chapter IV 137

be significantly linked to the genome of the superior parent strain 26B with the number of

segregants available. Therefore, we screened 400 additional F1 segregants of the diploid

26B/ER7A for low glycerol and high ethanol yield. In addition, we performed four rounds of

random inbreeding to increase the recombination frequency (Parts et al, 2011) and

subsequently evaluated 400 F5 segregants in small-scale fermentations for glycerol/ethanol

yield. The results for all tested segregants, including the 260 tested once for QTL analysis, are

shown in Figure 4A. The glycerol and ethanol yield are expressed as percentage of that of the

superior parent strain 26B. There was again a clear inverse relationship between glycerol and

ethanol yield. From the 800 segregants, we selected in total 48 superior segregants, 22 F1

segregants and 26 F5 segregants (Figure 4B). Subsequently, individual SNPs in the putative

QTLs were scored on chr. II, IV and XIII in all individual segregants, i.e. the 22 segregants of

the sequenced selected pool, the 22 additional selected F1 segregants and the 26 selected F5

segregants to calculate the SNP variant frequency and the corresponding P-value. Both

parameters were also determined for the two combined groups, 44 F1 segregants and the total

of 70 segregants (Figure 4C). Due to the low association of the three regions to the trait (70%-

80%), the P-values for the small segregant population of 22 or 26 individuals were not

significant. On the other hand, the three QTLs mapped with the large combined segregant

populations of 44 and 70 (Figure 4C) were significant. The region on chromosome II was

confirmed to be linked from position 504240 till 680091. The P-values reached a minimum

for the SNP markers, 504240 and 535197. In contrast, the highest association in the

sequenced pool was found at position 598260 predicted by EXPloRA (Duitama et al.

submitted for publication). On chromosome IV, EXPloRA predicted the two regions, 316389-

375978 and 696486-748140, as being significantly linked. The highest SNP variant

frequencies were found in the first region from 310000 to 410000. In all pools, the highest

frequency was found at position 411831, which was followed by a clear drop in the

frequency. The second region on chromosome IV was not significant in all tested segregant

pools. On chromosome XIII, EXPloRA reported two closely located regions, 600902-610995

and 634582-640415, which probably derived from a single QTL. In the total segregant pool,

the SNP at position, 606166 and 620234, showed the highest association, reaching a minimal

P-value at position 606166. The upstream region on chromosome XIII was not significantly

linked.

138 Chapter IV

Figure 4 Linkage analysis of QTLs on chr. II, IV and XIII with different groups of segregants. (A)

Glycerol and ethanol yield of the parental strains, 26B (■) and ER7A (▲), and the hybrid diploid strain

26B/ER7A (●).Glycerol and ethanol yield of the first isolated F1 segregants from 26B/ER7A (◯), of the

additional F1 segregants (□) and of the F5 segregants (◇). Fermentations were carried out in 5 ml YP 10%

glucose. Glycerol and ethanol yield of all segregants, ER7A and the diploid 26B/ER7A were related to the yield

of 26B, which was set as 100% (B) Segregants were selected for low glycerol (<120% glycerol yield, stippled

line) and high ethanol (>99% ethanol yield, stippled line) after each round of screening, resulting in the

following segregant groups: 22 F1 segregants used for pooled-segregant whole-genome sequence analysis (◯),

22 additional selected F1 segregants (□), and 26 F5 segregants (◇). These segregants were reconfirmed in 100

ml YP 10% glucose. (C) SNP variant frequency (top) and respective P-value (bottom) were determined by

allele-specific PCR in individual segregants of the sequenced selected pool (●), additional F1 selected pool (◯),

the total F1 selection of 44 (▲), the selection of F5 segregants (∆), and the total selection of all 70 segregants (■)

to fine-map the QTLs on chr. II, IV and XIII, which were detected with EXPloRA. The statistical confidence

line (for P-value ≤ 0.05) is indicated with a stippled line.

Chapter IV 139

Identification of causative genes in the QTLs on chr. II, IV and XIII

Three candidate genes in the three QTLs were selected based on their known function in

glycerol metabolism. SMP1, which is located in the QTL on chr. II (594,864 to 593,506 bp),

encodes a putative transcription factor involved in regulating glycerol production during the

response to osmostress (de Nadal et al, 2003). The gene is located in the chromosomal region

from 584,232 to 619,637 bp, which was predicted as most significantly linked by the

EXPloRA model. The 26B SMP1 allele has two point mutations, which are changing the

primary protein sequence at position 110 from arginine to glycine and at position 269 from

proline to glycine. Hence, we have named this allele smp1R110Q,P269Q

. On chr. IV, the SNP

with the highest linkage was located at position 411,831 bp (Figure 4C), which is within the

open reading frame of GPD1 (411,825 to 413,000 bp), the structural gene for the NAD+-

dependent cytosolic glycerol 3-phosphate dehydrogenase (Albertyn et al, 1994; Larsson et al,

1993). The GPD1 allele of 26B harbors a point mutation, changing leucine at position 164

into proline. This mutation was found earlier (DDBJ database data, accession number

AY598965). The GPD1 allele of 26B was named gpd1L164P

. On chr. XIII, the SNP with the

highest linkage was located at position 606,166 bp (Figure 4C), which is within the open

reading frame of HOT1 (605,981 to 608,140 bp). HOT1 encodes a transcription factor

required for the response to osmotic stress of glycerol biosynthetic genes, including GPD1,

and other HOG-pathway regulated genes (Alepuz et al, 2003; Rep et al, 1999). The 26B

HOT1 allele contains two non-synonymous point mutations, changing proline at position 107

to serine and histidine at position 274 to tyrosine. We have named the HOT1 allele of 26B,

hot1P107S,H274Y

.

The relevance of smp1R110Q,P269Q

, gpd1L164P

and hot1P107S,H274Y

for the low glycerol/high

ethanol phenotype was further investigated using reciprocal hemizygosity analysis (RHA)

(Steinmetz et al, 2002). For that purpose, two 26B/ER7A hybrid and hemizygous diploid

26B/ER7A were constructed, in which each pair contained a single copy of the superior allele

or the inferior allele of SMP1, GPD1 or HOT1, respectively. The three pairs of hemizygous

diploids were tested in the same 100 ml YP 10% glucose fermentations as used for the

screening. The parent strains 26B and ER7A and the hybrid diploid 26B/ER7A were added as

controls. The glycerol and ethanol yields were expressed as percentage of those of 26B, which

were set at 100%. The significance of any differences between the strains was evaluated using

140 Chapter IV

a two-tailed unpaired t-test with a P-value < 0.05 considered to indicate a significant

difference. The results of the RHA are shown in Figure 5.

Figure 5 Reciprocal hemizygosity analysis (RHA). RHA for the candidate genes, SMP1 (chromosome II),

GPD1 (chromosome IV), and HOT1 (chromosome XIII) to evaluate them as causative genes in the QTLs. For

RHA, diploids were constructed with either the deletion of the ER7A allele or the deletion of the 26B allele.

Glycerol and ethanol yield of the two hemizygous diploids were related to the parental strain 26B. The Student t-

test was used to confirm significant differences in glycerol and ethanol yield for the two diploids and is indicated

with *.

The results indicate that both smp1R110Q,P269Q

and hot1P107S,H274Y

, but not gpd1L164P

, derived

from the superior parent 26B, cause a significant drop in the glycerol yield compared to the

alleles of the inferior parent strain ER7A. Ethanol yield increased in the presence of

smp1R110Q,P269Q

and hot1P107S,H274Y

when compared to the inferior allele; however, significant

differences were only observed for the hot1P107S,H274Y

allele with a P-value < 0.05 but not for

smp1R110Q,P269Q

. The smp1R110Q,P269Q

gene is a causative gene in the QTL on chr. II; however it

cannot be excluded that this QTL may contain a second causative gene, especially since

smp1R110Q,P269Q

is not located in the region with the strongest linkage (lowest P-value). The

RHA with the GPD1 alleles failed to show any difference both for glycerol and ethanol yield

(Figure 5). Hence, the superior character of the gpd1L164P

allele could not be confirmed with

Chapter IV 141

RHA. This is remarkable because the SNP with the strongest linkage (lowest P-value) in the

QTL on chr. IV was located in the open reading frame of GPD1 and showed very strong

linkage to the phenotype. The hot1P107S,H274Y

allele of the superior strain 26B, on the other

hand, caused a reduction in glycerol and an increase in ethanol production, and both changes

were significant (P-value < 0.05) (Figure 5). Hence, these results indicate that hot1P107S,H274Y

is a causative allele in the QTL on chr. XIII and because it contains the SNP with the

strongest linkage (lowest P-value) it is likely the main causative allele in this QTL.

Expression of the gpd1L164P

allele in gpd1∆ strains

Several explanations could account for the failure to confirm the superior character of the

gpd1L164P

allele from 26B in the RHA test. A closely located gene may be the real causative

gene, the gpd1L164P

allele may be effective only in a haploid genetic background or the effect

of the gpd1L164P

allele may be suppressed through epistasis by one or both of the other two

superior alleles, smp1R110Q,P269Q

and hot1P107S,H274Y

. To distinguish between these possibilities,

the gpd1L164P

allele was amplified from CBS4C and the GPD1 allele from strain ER7A by

PCR (410,523 to 413,479 bp, including promotor, ORF and terminator). The PCR fragment

was ligated in the centromeric plasmid YCplac33, resulting in plasmids YCplac33-gpd1L164P

-

CBS4C and YCplac33-GPD1-ER7A. Both plasmids were transformed into gpd1∆ strains of

the two parents 26B and ER7A, the hybrid diploid 26B/ER7A and the lab strain BY4742

(Giaever et al, 2002; Winzeler et al, 1999). All strains were tested in 100ml scale

fermentations with YP 10% glucose. Glycerol and ethanol yield were determined after 120h

fermentation and were expressed as percentage of those of 26B. The results are shown in

Figure 6. The expression of the gpd1L164P

-CBS4C or GPD1-ER7A allele in the gpd1∆ strains

of the superior parent 26B and the hybrid diploid 26B/ER7A enhanced glycerol yield and

reduced ethanol yield to the same extent for the two alleles. However, in the gpd1∆ strains of

the inferior parent ER7A and the lab strain BY4742, the expression of gpd1L164P

-CBS4C allele

significantly reduced glycerol yield and increased ethanol yield compared to the expression of

the GPD1-ER7A allele in the respective strains. The latter confirms that the gpd1L164P

-CBS4C

allele is superior for the low glycerol/ high ethanol phenotype compared to the GPD1-ER7A

allele. The difference between the two alleles is not dependent on the haploid or diploid

background of the strain but seems to be related with the presence of the two other superior

142 Chapter IV

alleles, smp1R110Q,P269Q

and hot1P107S,H274Y

. In 26B and 26B/ER7A, no differential effect on

glycerol and ethanol yield was observed in the presence of both alleles gpd1L164P

-CBS4C and

GPD1-ER7A. In contrast, the glycerol and ethanol yield between the two expressed GPD1

alleles significantly differed in ER7A and BY4742, probably due to the absence of both

superior alleles, smp1R110Q,P269Q

and hot1P107S,H274Y

. Hence, the gpd1L164P

allele is causative for

the ‘low glycerol’ phenotype, however the gpd1L164P

alleles phenotypically interferes with the

smp1R110Q,P269Q

and hot1P107S,H274Y

alleles in 26B.

Figure 6 Expression of gpd1L164P

-CBS4C and GPD1-ER7A in segregant 26B, ER7A, the diploid 26B/ER7A

and BY4742. Glycerol and ethanol yield in the gpd1∆ strains, 26B, ER7A, 26B/ER7A and BY4742, harbouring

the plasmids YCplac33, YCplac33-GPD1-ER7A, and YCplac33-gpd1L164P-CBS4C. Fermentations were carried

out in 100 ml YP 10% glucose. Glycerol and ethanol yield of the strains were related to the yield of 26B, which

was set as 100%, to decrease experimental variations. In the BY 4742 and ER7A backgrounds, which lack the

smp1R110Q,P269Q and hot1P107S,H274Y alleles, the gpd1L164P allele clearly reduced glycerol yield and concomitantly

increased ethanol yield compared to the wild type GPD1 allele. In the strains 26B and 26B/ER7A, which contain

the smp1R110Q,P269Q and hot1P107S,H274Y alleles, the gpd1L164P allele resulted in a similar glycerol yield as the wild

type GPD1 allele.

Chapter IV 143

Figure 7 Presence of the 26B alleles, smp1R110Q,P269Q

, gpd1L164P

and hot1P107S,H274Y

, in the selected segregant

population. Mean values are indicated for each group.

We have scored the final 70 superior segregants with a glycerol yield < 120% and an ethanol

yield > 99% of that of the superior parent 26B, for the presence of the three causative alleles,

smp1R110Q,P269Q

, gpd1L164P

and hot1P107S,H274Y

. The results are shown in Figure 7. The largest

group of superior segregants contained all three mutant alleles, followed by smaller groups

with only two of the three mutant alleles and finally the three smallest groups with only one

mutant allele. Hence, there was a tendency that the glycerol yield decreased and the ethanol

yield increased with the number of mutant alleles present in a selected segregant. Although

the means of the groups showed differences depending on the number of mutant alleles, these

differences were not significant due to the small population size and high phenotypic variance

within the groups. The frequency of the three mutant alleles in the 70 superior segregants was

smp1R110Q,P269Q

: 76%, gpd1L164P

: 83% and hot1P107S,H274Y

: 70%. Therefore, the gpd1L164P

allele

showed a slight preference for the phenotype over the two other mutant alleles, despite its

epistatic suppression by its two transcriptional regulators, smp1R110Q,P269Q

and hot1P107S,H274Y

.

The selected segregants frequently possess at least two out of three mutant alleles, hence at

least two alleles were required to confer the superior phenotype for the majority of the

selected segregants.

144 Chapter IV

4.4 Discussion

Identification of alleles causative for reduced glycerol yield in segregant 26B

The goal of the present work was to investigate whether natural yeast strains may harbour

specific alleles that would allow reducing the glycerol yield and increasing the ethanol yield

in yeast fermentation. Here, we successfully identified three mutant alleles, which separately

and together reduce the glycerol yield in a subtle way without affecting at least not in a

conspicuous way the overall rate and characteristics of the fermentation process. Combined

with the previous discovery of the ssk1E330N…K356N

allele in strain CBS4C (Hubmann et al

2013), the current results indicate that the original diploid parent strain CBS6412 contains at

least four specific alleles of well-known structural and regulatory genes filtered by natural

selection and evolution for compatibility with survival in the natural environment. The chance

that these alleles exert significant negative effects on other essential functions of the yeast

cells is probably not completely absent but at least minimized compared to drastic genetic

modifications like gene deletion or overexpression. Screening of biodiversity for such specific

alleles is therefore a fruitful strategy to identify mutant alleles that can be used as specific

gene tools for strain improvement by targeted genetic modification.

Identification of minor QTLs and causative genes

Non-selectable quantitative traits, like glycerol or ethanol formation in yeast display a

polygenic inheritance. Causative genes contribute to varying degrees to the phenotype. While

identification of major QTLs has become straightforward with pooled-segregant whole-

genome sequence analysis (Deutschbauer et al, 2005; Ehrenreich et al, 2010; Parts et al,

2011; Swinnen et al, 2012b), identification of minor QTLs remains challenging. In the present

work, we successfully implemented a targeted backcross (Sinha et al, 2008) for identification

of minor QTLs. This approach requires an F1 segregant displaying the phenotype-of-interest,

in which one or more superior alleles are absent. The F1 segregant 26B descending from the

parental cross ER7A/CBS4C was identified as such a segregant, in which the major causative

gene, ssk1E330N…K356N

, previously found in QTL mapping (Hubmann et al 2013), was absent.

Moreover, the ratio of selected to the total number of F1 segregants from ER7A/CBS4C

indicated that two or three additional mutations besides the major gene ssk1E330N…K356N

were

present in CBS4C. The absence of the ssk1E330N…K356N

allele in the segregant 26B required

Chapter IV 145

that all or most minor QTLs are present in this segregant to confer the phenotype-of-interest.

Interestingly, these combined mutations resulted in an equally low glycerol yield in the

segregant 26B when compared to the parental strain CBS4C. This indicates that the effect of

the major causative allele, ssk1E330N…K356N

functionally interferes with the other minor linked

alleles, smp1R110Q,P269Q

, gpd1L164P

and hot1P107S,H274Y

, i.e. ssk1E330N…K356N

allele suppresses the

effect of the minor contributing alleles. In genetic mapping, such gene-gene interactions result

in insignificant linkage disequilibrium for the suppressed QTLs, which results in particular

from parental crosses with all causative mutations present. To avoid this problem, backcross

strategies have been successfully applied to detect minor QTLs (Birkeland et al, 2010;

Demogines et al, 2008; Sinha et al, 2008). In the present work, we eliminated the major allele

ssk1E330N…K356N

by backcrossing the segregant 26B with its inferior parent ER7A. In principle

this approach can be repeated with a segregant of the new generation that still displays the

low glycerol phenotype. As shown in Figure 7, several of the new generation segregants

contained only one of the three causative alleles and still displayed a low glycerol yield under

the cut-off of 120%. This suggests that there are still additional alleles present in these strains

able to confer low glycerol yield in the absence of the other alleles and that a new backcross

of such a segregant with the inferior parent may allow identification of additional alleles

conferring low glycerol yield.

Random coincidence versus linkage in small segregant populations

In small populations of segregants, random coincidence can easily cause falsely predicted

QTLs (Beavis, 1998), which are difficult to distinguish from QTLs with significant, but weak

linkage. An unselected pool helps in adjusting the significance threshold for minor QTLs in

order to exclude false positives caused by random coincidence. Higher stringency in QTL

selection can eliminate false QTLs but also weakly linked true QTLs. It is well known that a

higher number of segregants increases the reliability of minor QTL detection. In previous

work, many minor QTLs could be identified by using millions of segregants and a selectable

phenotype (Parts et al, 2011). However, many quantitative traits are not selectable and the

phenotypic screening of large numbers of segregants can be very laborious, as in the case of

low glycerol/high ethanol yield, and is thus often limited to small numbers of segregants.

Furthermore, the sequencing depth considerably determines the accuracy of SNP variant

frequencies in pooled DNA samples. Generally, a low sequencing depth causes high variation

146 Chapter IV

of the SNP variant frequency, even if a large population of segregants is used. In practice, the

sequencing depth should be adjusted to the number of pooled segregants. If the number of

segregants exceeds the sequencing depth, the surplus of segregants no longer enhances the

reliability of mapping and is thus useless. In this case, the selected segregants could be split

into different pools to enable cross-validation of the population subsets. An essential

difference between a false QTL caused by random coincidence and a true minor QTL is that

the latter should be consistently reproducible in different selected segregant populations,

whereas a false QTL caused by random coincidence is not. Therefore, we used three different

pools of segregants resulting in three independent pools of small-size with segregants

displaying low glycerol yield. This allowed us to distinguish the false QTLs on chr IV

(696486-748140) and chr XIII (634582-640415) from the true, weakly linked QTLs on the

same chromosomes, chrIV (316389-375978) and chrXIII (600902-610995). Resampling of

more selected segregants resulted in a clear differentiation of false QTLs from weakly linked

QTLs. Moreover, all three pools combined retrieved equal analysis power as one huge

population. Hence, both strategies enable a reliable identification of minor QTLs.

Novel mutant alleles and their epistatic interactions

The discovery of three novel alleles of known components of glycerol metabolism and its

regulation suggests that complex trait analysis could be a promising avenue for identification

of novel components involved in cellular functions. It seems plausible that further mapping

and analysis of even more minor QTLs, as discussed previously, may easily result in

identification of new genes involved in the regulation of glycerol metabolism.

In this work we have identified three novel alleles of components with a well-established role

in glycerol metabolism and its regulation. Smp1 is a transcription factor, belonging to the

MEF2 family, that regulates the expression of stress-responsive genes, such as GPD1. Its

DNA binding domain is located at the amino acid residues 1-90 (Dodou and Treisman, 1997).

Upon osmotic stress, Smp1 is phosphorylated by Hog1, which physically interacts with its C-

terminal domain. Four different phosphorylation sites were identified, i.e. Ser348, Ser357,

Thr365, and Ser376, all located within a region coincident with the Hog1 binding domain.

Phosphorylation of Smp1 is essential for its functioning, since an allele unable to be

phosphorylated caused an impaired stress response (de Nadal et al, 2003). The point

Chapter IV 147

mutations in the 26B allele, smp1R110Q,P269Q

, are not located in the domain for DNA

recognition or the Hog1 binding domain. However, the change of a proline to a glycine, close

to the phosphorylation sites, might change Smp1 structure, thereby influencing its ability to

be bound and/or phosphorylated by Hog1.

GPD1 encodes NAD+ dependent cytosolic glycerol 3-phosphate dehydrogenase. It catalyzes

the conversion of dihydroxyacetone phosphate (DHAP) to glycerol 3-phosphate through the

oxidation of NADH. The expression of GPD1 is induced by the HOG-pathway and it is

essential for growth under high osmolarity (Albertyn et al, 1994). Possible domains for

binding of NADH, H+ and DHAP have been predicted based on similarity with proteins with

a comparable function (Marchler-Bauer et al, 2009). The single point mutation present in the

26B allele, gpd1L164P

, may be located in the putative NADH-binding domain, but the location

of this domain is not well predicted. This mutation was found earlier and called a ‘natural

variant’ (DDBJ database data, accession number AY598965). No linkage of gpd1L164P

with

low glycerol yield was observed in RHA but its effect was revealed by expression of the two

parental GPD1 alleles in GPD1-deficient mutants. The L164P mutation could reduce the

intrinsic activity of the Gpd1 enzyme and results in reduced glycerol production.

The gpd1L164P

allele was clearly subject to epistatic suppression. In the BY 4742 and ER7A

backgrounds, which lack the smp1R110Q,P269Q

and hot1P107S,H274Y

alleles, the allele clearly

reduced glycerol yield compared to the wild type GPD1 allele (Figure 6). On the other hand,

its expression in the strains 26B and 26B/ER7A, which contain the smp1R110Q,P269Q

and

hot1P107S,H274Y

alleles, resulted in a similar glycerol yield as the wild type GPD1 allele. The

epistatic effect may be explained at the biochemical level by the fact that the reduction in

expression of GPD1, caused by the smp1R110Q,P269Q

and hot1P107S,H274Y

alleles, is so strong that

the mutation in GPD1 itself has no significant effect anymore. Hot1 activates transcription of

GPD1 and other HOG-dependent genes under osmo-stress (Alepuz et al, 2003; Rep et al,

2000; Rep et al, 1999). Alepuz et al. (2003) proposed that Hot1 serves as an anchor for Hog1,

which directly recruits the RNA polymerase II complex. The position of the Hog1 binding

domain in Hot1 is unknown. It is unclear how the two mutations in the 26B allele,

hot1P107S,H274Y

, could affect the functioning of the protein.

148 Chapter IV

4.5 Supplementary material

Supplementary Table 1: Primers used

No. Function Sequence

A-4770 SNP check Pair 1, CBS4C Chr II w 535197 GAAACTGAAACCAGGAGGAG

A-4735 SNP check Pair 1, ER7A Chr II w 535197 GAAACTGAAACCAGGAGGAA

A-4736 SNP check Pair 1, CBS4C Chr II c 535985 CTTTATGTAGTCTGGATTTTAG

A-4737 SNP check Pair 1, ER7A Chr II c 535985 CTTTATGTAGTCTGGATTTTAA

A-4738 SNP check Pair 2, CBS4C Chr II w 598260 TTCAAGTTAAATCGAATTGTAT

A-4739 SNP check Pair 2, ER7A Chr II w 598260 TTCAAGTTAAATCGAATTGTAC

A-4740 SNP check Pair 2, CBS4C Chr II c 599590 ATATCAATGTAAACACGTCA

A-4741 SNP check Pair 2, ER7A Chr II c 599590 ATATCAATGTAAACACGTCG

A-4742 SNP check Pair 3, CBS4C Chr II w 680091 GATATTAGTGTACATACGTTGC

A-4743 SNP check Pair 3, ER7A Chr II w 680091 GATATTAGTGTACATACGTTGA

A-4744 SNP check Pair 3, CBS4C Chr II c 681213 TTCCTTTTGAAGTGTCCTCG

A-4745 SNP check Pair 3, ER7A Chr II c 681213 TTCCTTTTGAAGTGTCCTCT

A-4746 SNP check Pair 4, CBS4C Chr XII w 169083 TCTCCATTACCAGCTGAA

A-4747 SNP check Pair 4, ER7A Chr XII w 169083 TCTCCATTACCAGCTGAG

A-4748 SNP check Pair 4, CBS4C Chr XII c 169876 GCATATATATATTTTAAGAAAATT

A-4749 SNP check Pair 4, ER7A Chr XII c 169876 GCATATATATATTTTAAGAAAATC

A-4750 SNP check Pair 5, CBS4C Chr XII w 198118 CAGAGTGGCAGACATTATCG

A-4751 SNP check Pair 5, ER7A Chr XII w 198118 CAGAGTGGCAGACATTATCA

A-4752 SNP check Pair 5, CBS4C Chr XII c 198478 GTAGCTGCCACAAAGCAC

A-4753 SNP check Pair 5, ER7A Chr XII c 198478 GTAGCTGCCACAAAGCAT

A-5009 SNP check Pair 6, CBS4C Chr II w 650007 TCTGCGTCTCTACGTTCTTG

A-5010 SNP check Pair 6, ER7A Chr II w 650007 TCTGCGTCTCTACGTTCTTA

A-5011 SNP check Pair 6, CBS4C Chr II c 650708 CACGCGTCGTTCTCGTC

A-5012 SNP check Pair 6, ER7A Chr II c 650708 CACGCGTCGTTCTCGTT

A-5013 SNP check Pair 7, CBS4C Chr II w 707209 ACATGTACACAAATCTTGAT

A-5014 SNP check Pair 7, ER7A Chr II w 707209 ACATGTACACAAATCTTGAC

A-5015 SNP check Pair 7, CBS4C Chr II c 708732 AGAAGAGAAGATCAAGCGTA

Chapter IV 149

A-5016 SNP check Pair 7, ER7A Chr II c 708732 AGAAGAGAAGATCAAGCGTG

A-5017 SNP check Pair 8, CBS4C Chr II w 137510 CCCATTTTCTCGAATTGCAGA

A-5018 SNP check Pair 8, ER7A Chr II w 137510 CCCATTTTCTCGAATTGCAGG

A-5019 SNP check Pair 8, CBS4C Chr II c 138796 TGGCTTATGCAGGCGGTAAT

A-5020 SNP check Pair 8, ER7A Chr II c 138796 TGGCTTATGCAGGCGGTAAC

A-5049 SNP check Pair 9, CBS4C Chr II w 664541 GAATATGGATATGTAGCCAC

A-5050 SNP check Pair 9, ER7A Chr II w 664541 GAATATGGATATGTAGCCAT

A-5051 SNP check Pair 9, CBS4C Chr II c 665836 ATGGTCTTCAGAGGTCCC

A-5052 SNP check Pair 9, ER7A Chr II c 665836 ATGGTCTTCAGAGGTCCT

A-5053 SNP check Pair 10, CBS4C Chr II w 694207 CATGACAGTGAGTCTGAGTC

A-5054 SNP check Pair 10, ER7A Chr II w 694207 CATGACAGTGAGTCTGAGTT

A-5055 SNP check Pair 10, CBS4C Chr II c 695455 TTCAACAATCCTCAAAATCC

A-5056 SNP check Pair 10, ER7A Chr II c 695455 TTCAACAATCCTCAAAATCT

A-8483 SNP check Pair 15, CBS4C Chr IV 342149 w ATTTGGTGAGCAGATCATT

A-8482 SNP check Pair 15, ER7A Chr IV 342149 w ATTTGGTGAGCAGATCATC

A-8485 SNP check Pair 15, CBS4C Chr IV c 343682 GAATGGTCTGCATTCAGA

A-8484 SNP check Pair 15, ER7A Chr IV c 343682 GAATGGTCTGCATTCAGG

A-8487 SNP check Pair 16, CBS4C Chr IV w 374477 GGTAGTTCTGCCGGTTTG

A-8486 SNP check Pair 16, ER7A Chr IV w 374477 GGTAGTTCTGCCGGTTTA

A-8489 SNP check Pair 16, CBS4C Chr IV c 375389 TGGAGTAGATGAGTTGAATG

A-8488 SNP check Pair 16, ER7A Chr XII c 375389 TGGAGTAGATGAGTTGAATA

A-8491 SNP check Pair 17, CBS4C Chr IV w 716135 CGTTGCTTTTGGCATCCA

A-8490 SNP check Pair 17, ER7A Chr IV w 716135 CGTTGCTTTTGGCATCCG

A-8493 SNP check Pair 17, CBS4C Chr IV c 716689 CGAGTTACAACGATGCGT

A-8492 SNP check Pair 17, ER7A Chr IV c 716689 CGAGTTACAACGATGCGC

A-8495 SNP check Pair 18, CBS4C Chr XIII w

606166 CTATAGGTATCAATCTTGAT

A-8494 SNP check Pair 18, ER7A Chr XIII w 606166 CTATAGGTATCAATCTTGAC

A-8497 SNP check Pair 18, CBS4C Chr XIII c 606800 GCTGGTTGATACCAAAATA

A-8496 SNP check Pair 18, ER7A Chr XIII c 606800 GCTGGTTGATACCAAAATG

A-8499 SNP check Pair 19, CBS4C Chr XIII w

637066 GAATTCAACTCCCTGAGC

A-8498 SNP check Pair 19, ER7A Chr XIII w 637066 GAATTCAACTCCCTGAGT

A-8501 SNP check Pair 19, CBS4C Chr XIII c 637703 GGACCTAAATGACGTCTG

150 Chapter IV

A-8500 SNP check Pair 19, ER7A Chr XIII c 637703 GGACCTAAATGACGTCTA

A-9111 SNP check Pair 20, CBS4C Chr II w 557132 CCAAAGGAAAGACCATGCTC

A-9112 SNP check Pair 20, ER7A Chr II w 557132 CCAAAGGAAAGACCATGCTT

A-9113 SNP check Pair 20, CBS4C Chr II c 558463 GACTTCCAAATCCGAGACG

A-9114 SNP check Pair 20, ER7A Chr II c 558463 GACTTCCAAATCCGAGACA

A-9115 SNP check Pair 21, CBS4C Chr II w 316389 GAATTGCGACTTTGACCATA

A-9116 SNP check Pair 21, ER7A Chr II w 316389 GAATTGCGACTTTGACCATG

A-9117 SNP check Pair 21, CBS4C Chr II c 316965 AGCTAAATAATAAGTAAGATCCG

A-9118 SNP check Pair 21, ER7A Chr II c 316965 AGCTAAATAATAAGTAAGGTCCC

A-9119 SNP check Pair 22, CBS4C Chr II w 333847 TCTGATTTATTCTTTTATCCTGT

A-9120 SNP check Pair 22, ER7A Chr II w 333847 TCTGATTTATTCTTTTATCCTGC

A-9121 SNP check Pair 22, CBS4C Chr II c 334511 CACGATGCCCGCTTTGTTCC

A-9122 SNP check Pair 22, ER7A Chr II c 334511 CACGATGCCCGCTTTGTTTT

A-9736 SNP check Pair 23, CBS4C Chr II w 459460 ATGATATTCAGCAAATTTGCTT

A-9737 SNP check Pair 23, ER7A Chr II w 459460 ATGATATTCAGCAAATTTGCTG

A-9738 SNP check Pair 23, CBS4C Chr II c 460849 CAGTGGTAGTGCTCCTTA

A-9739 SNP check Pair 23, ER7A Chr II c 460849 CAGTGGTAGTGCTCCTTG

A-9740 SNP check Pair 24, CBS4C Chr II w 504240 GACTTTTTGCTGTGCTGT

A-9741 SNP check Pair 24, ER7A Chr II w 504240 GACTTTTTGCTGTGCTGC

A-9742 SNP check Pair 24, CBS4C Chr II c 505225 TGTTGTTGCTGCTGTTGT

A-9743 SNP check Pair 24, ER7A Chr II c 505225 TGTTGTTGCTGCTGTTGC

A9744 SNP check Pair 25, CBS4C Chr IV w 439631 AGTTAAACATTATTATTCGTT

A-9745 SNP check Pair 25, ER7A Chr IV w 439631 AGTTAAACATTATTATTCGTG

A-9746 SNP check Pair 25, CBS4C Chr IV c 440415 CAAAGTGTCAGATGCT

A-9747 SNP check Pair 25, ER7A Chr IV c 440415 CAAAGTGTCAGATGCC

A-9748 SNP check Pair 26, CBS4C Chr IV w 470659 TCAAAAGTCAAGTCTTGTTGAGC

A-9749 SNP check Pair 26, ER7A Chr IV w 470659 TCAAAAGTCAAGTCTTGTTGAGT

A-9750 SNP check Pair 26, CBS4C Chr IV c 471261 TCGCTGCATGCACTTCTC

A-9751 SNP check Pair 26, ER7A Chr IV c 471261 TCGCTGCATGCACTTCTT

A-9752 SNP check Pair 27, CBS4C Chr IV w 498006 GCACGCTATCAACTTTCTTA

A-9753 SNP check Pair 27, ER7A Chr IV w 498006 GCACGCTATCAACTTTCTTG

A-9754 SNP check Pair 27, CBS4C Chr IV c 499592 ATAATAGGTTCCCCTCCT

Chapter IV 151

A-9755 SNP check Pair 27, ER7A Chr IV c 499592 ATAATAGGTTCCCCTCCC

A-9962 SNP check Pair 28, CBS4C Chr XIII w

566283 AATCGTGCTGATGATTCCAT

A-9963 SNP check Pair 28, ER7A Chr XIII w 566283 AATCGTGCTGATGATTCCAC

A-9964 SNP check Pair 28, CBS4C Chr XIII c 566967 GTGGCAATCGCATTGGAT

A-9965 SNP check Pair 28, ER7A Chr XIII c 566967 GTGGCAATCGCATTGGAC

A-9966 SNP check Pair 29, CBS4C Chr XIII w

580772 TGATAGCGTTTATTCCAG

A-9967 SNP check Pair 29, ER7A Chr XIII w 580772 TGATAGCGTTTATTCCAA

A-9968 SNP check Pair 29, CBS4C Chr XIII c 582017 GAGATCAAAATGTTCTTGAC

A-9969 SNP check Pair 29, ER7A Chr XIII c 582017 GAGATCAAAATGTTCTTGAT

A-9970 SNP check Pair 30, CBS4C Chr XIII w

619387 GCCGGTAGTTGCGCT

A-9971 SNP check Pair 30, ER7A Chr XIII w 619387 GCCGGTAGTTGCGCC

A-9972 SNP check Pair 30, CBS4C Chr XIII c 620234 TCTCTTCACTTTCTTCTTCAC

A-9973 SNP check Pair 30, ER7A Chr XIII c 620234 TCTCTTCACTTTCTTCTTCAT

A-3913 SNP check Pair GPD1, CBS4C Chr IV w

411831

CACAAATATTGATAATATAAAGATGTCTGC

C

A-3914 SNP check Pair GPD1, ER7A Chr IV w

411831

CACAAATATTGATAATATAAAGATGTCTGC

T

A-3915 SNP check Pair GPD1, CBS4C Chr IV c

412313 AGTTCCTCAGTGATGTAAGAGGATG

A-3916 SNP check Pair GPD1, ER7A Chr IV c

412313 AGTTCCTCAGTGATGTAAGAGGATA

A-8396 HOT1 deletion cassette, long flanking site

ATAGGATAACCATCAGCTTCGATTATTCTA

CACT

GATGAGCCCTTTCTTCCAGCTGAAGCTTCGT

ACGC

A-8397 HOT1 deletion cassette, long flanking site

GTGCCGTTTTACCTCTCTCCATACCGTTCAG

AATG

AACTTGTACAATCTTGCATAGGCCACTAGT

GGATCTG

A-8398 Verification of deletion, upstream of HOT1 TGGGTATTGCGATTCTTTGC

A-8399 Verification of deletion, downstream of HOT1 TTCGCCATCCTCATCATCTT

A-8392 SMP1 deletion cassette, long flanking site

AATTGAACCTATCAAAGATGATAGAAATCG

TAC

AGTTACTTTCATAAAGCCAGCTGAAGCTTC

GTACGC

A-8393 SMP1 deletion cassette, long flanking site

AGGCAATGAAGGCTTTGGCAAATTCTCGTA

AGGCG

ACGGGTAAAAATTATGCATAGGCCACTAGT

GGATCTG

A-8394 Verification of deletion, upstream of SMP1 GCTCTAGATAAGCAAACACAA

A-8395 Verification of deletion, downstream of SMP1 TCCATCCTTTCAATCGCAAT

A-8478 GPD1 deletion cassette, long flanking site

CATCAAATCTATCCAACCTAATTCGCACGT

AGA

CTGGCTTGGTATCAGCTGAAGCTTCGTACG

152 Chapter IV

C

A-8479 GPD1 deletion cassette, long flanking site

CGACGTCCTTGCCCTCGCCTCTGAAATCCTT

TGG

AATGTGGTAAGGCATAGGCCACTAGTGGAT

CTG

A-8480 Verification of deletion, upstream of GPD1 CCGCACAACAAGTATCAGA

A-8481 Verification of deletion, downstream of GPD1 AAGTAAGGTCTGTGGAACAA

B-231 URA3 deletion cassette, long flanking site

ATGTCGAAAGCTACATATAAGGAACGTGCT

GCTACTCATCCTAGTCCTGTCAGCTGAAGC

TTCGTACGC

B-232 URA3 deletion cassette, long flanking site

TTAGTTTTGCTGGCCGCATCTTCTCAAATAT

GCTTCCCAGCCTGCTTTTCGCATAGGCCACT

AGTGGATCTG

A-3864 Verification of deletion, inside KanMX6 GTTGTATTGATGTTGGACGA

A-2970 Verification of deletion, downstream of URA3 TGGCGAGGTATTGGATAGTTCC

1648 MAT locus GGTATTTGCTAGAACTACACTGA

5584 MAT-alpha-Sc (399-425) GACTACTTCGCGCAACAGTATAATTTT

5585 MAT-a-Sc (324-347) AAGAAAGCAAAGCCTTAATTCCAA

A-3709 GPD1 promoter ORF terminator cloning

TTTCAAGGTACCAAACATATGCGCGCCACG

T

KpnI

A-3743 GPD1 promoter ORF terminator cloning CTACTAGGTACCATTTTGCGTCGCGCACGT

KpnI

Nucleotides in capital letters are derived from S. cerevisiae genome for homologous recombination at HOT1,

SMP1 or GPD1 locus. Sequences with underlined letters functioned as primers for amplification of the deletion

or promoter cassettes from respective plasmids (Güldner et al, 2002).

Chapter IV 153

Supplementary Table 2: Plasmids used in this study

Plasmid Description Reference

pUG6 E. coli/ vector containing, Amp+, loxP-

KanMX6-loxP disruption cassette

Güldner et al., 2002

pUG66 E. coli/ vector containing, Amp+, loxP-bleR-

loxP disruption cassette

Güldner et al., 2002

pFL39 GAL1 HO KanMX vector containing HO gene

Ycplac33 yeast shuttle vector URA3

Ycplac33 GPD1-ER7A yeast shuttle vector URA3 GPD1-ER7A

Ycplac33 gpd1L164P yeast shuttle vector URA3, gpd1L164P

Yields

Absolute yield

Relative yield

 

YS/ i = cifinal

(csini -cS

final )

 

Y% = YS/ i

segr

YS/ i

26B

Chapter V

General discussion and future perspectives

156 Chapter V

1. Application and valorisation of the project

Bioethanol is increasingly used as biofuel for transportation purposes, replacing petroleum

derived fuels. The biofuel markets boomed in recent years and its industry has grown rapidly

worldwide. Also in Europe, several bioethanol production plants have recently been

constructed and are now actively producing bioethanol. New legislation has been implied that

imposes the mandatory addition of biofuel in existing fossil fuels, and new European goals

have been established for replacement of fossil fuels by biofuels, guaranteeing steady sales

and likely further expansion of the bioethanol market and industry. Currently, wheat and to a

lesser extent corn and sugar beet, are used as substrates in the European plants. As these

substrates are also used as food an feed, there is a strong drive to replace them with

lignocellulosic wastes and dedicated energy crops. However, current lignocellulosic bio-

ethanol production processes require a large increase in efficiency and cost reduction to make

them commercially viable.

In addition to engineering dedicated yeast strains for second generation bioethanol, the

improvement of ethanol yield has been a target for strain improvement. This is valid for both

first and second generation bioethanol production. In tropical regions, sugar cane will likely

remain the substrate of choice for bioethanol production because of the high cost-efficiency of

the production process, which results in a very competitive ethanol price. However, even in

this case yield improvement is highly desirable. In fact, the high osmolarity of the sucrose

solution extracted from sugar cane, glycerol overproduction and concomitant loss in ethanol

yield is a continuous challenge. Glycerol levels in bio-ethanol production can vary between

about 1% and 6 - 7% (w/v) depending on the particular process used. Obviously, production

of 6 - 7% (w/v) glycerol means a loss of about 5% (w/v) ethanol, which is a huge loss in

economic terms. The impact of improved industrial yeast strains with lower glycerol

production and concomitantly higher ethanol yield can thus be very high. The knowledge

obtained in the present project on mutations causing reduced glycerol production and

enhanced ethanol yield is therefore of great value for the development of novel industrial

yeast strains with high conversion efficiency.

Besides that, the findings of the current work might be relevant for the enforcement of

glycerol formation and a concomitant reduction in ethanol titre. Higher glycerol levels are

desirable in wine production because they improve the mouthfeel and reduce the ethanol

Chapter V 157

concentration of the wine. Yeast strains, which exhibit this desirable trait, are currently in

very high demand in the wine industry.

2. Optimization of ethanol production

The goal of the present work was to find genetic configurations of S. cerevisiae, which

confers a ‘low glycerol’ phenotype. Glycerol synthesis is indispensible for cellular and

metabolic integrity of S. cerevisiae. Nevertheless, glycerol formation in this yeast species can

vary considerably, depending on environmental and genetic constraints implying that there is

room for engineering glycerol yield without severe effects on growth and stress tolerance

To define the minimal glycerol requirement in a fermentation process, one has to mainly

consider the nitrogen source, the availability of oxygen and the osmotic pressure (particularly

determined by sugar concentration) since these factors greatly influence redox balancing. In

fact, optimizing fermentation conditions and medium composition are important factors in

reducing by-product formation in an alcoholic fermentation process, as demonstrated in

several studies (Albers et al, 1996; Alfenore et al, 2004; Bideaux et al, 2006). Certainly, some

of the suggested improvements might be indeed readily implemented in the industrial process,

because such optimizations do not require de-novo yeast strain development construction of

novel yeast strains and they impact the ethanol industry much faster. For instance, significant

ethanol yield improvement in starch-based ethanol production was achieved through

integration of starch saccharification in the fermentation process. Running saccharification

and fermentation simultaneously has several advantages, such as the fast and complete

enzymatic hydrolysis of sugars, the prevention of high osmotic pressure and a concomitant

reduction of glycerol formation, and the inhibition of bacterial growth due to the much lower

levels of free sugar substrate. Furthermore, the incorporation of yeast propagation in the SSF

technology has also found rapid implementation in the starch-based processes due to high-

sustained ethanol yield and the lower production costs (Madson and Monceaux, 1999). In

Brazilian ethanol production, most fermentations are operated in fed-batch mode referred to

as Melle-Boinot process. The duration of one fermentation cycle is only approximately 8h,

due to the high cell density, with the yeast taking roughly one tenth of the bioreactor’s

158 Chapter V

volume. After every fermentation cycle, the yeast cells are re-used in the subsequent cycle,

avoiding strong yeast propagation and biomass formation and hence, glycerol formation.

Furthermore, slow sugar feeding keeps the sugar concentration low, thereby lowering the

osmotic pressure imposed by the sugar (Basso et al, 2011).

Apart from process optimization, genetic prerequisites were established in the yeasts for

obtaining a highly productive strain. Yeast strains have undergone a long process of selection

becoming more and more adapted to the specific applications by human activities (Sicard and

Legras, 2011). More recently, dedicated breeding programs have supported this selection

process and have put major strains forth that are used in today’s ethanol industry (Basso et al,

2008). Still, recent advances in metabolic engineering evidenced that even these highly

efficient natural production strains can be improved using targeted genetic modifications.

The present work investigated the feasibility of gradually reducing glycerol formation in S.

cerevisiae by controlling its synthesis. In considering the role of glycerol as redox sink for the

anaerobic excess NADH and as compatible solute to adapt to hyperosmotic conditions, the

cellular requirement in glycerol production clearly depends on the medium composition and

on the availability of an alternative redox sink, such as oxygen.

The key enzyme of the glycerol synthetic pathway is glycerol 3-phosphate dehydrogenase

(GPDH), encoded by GPD1 and GPD2. GPDH activity strongly relates to transcriptional

expression of GPD1 and/or GPD2. Enhanced transcription, resulting in higher GPDH

activity, is observed under osmotic stress conditions and under anaerobic conditions. The

exchange of the natural GPD1 and GPD2 promoters by lower-strength promoters deregulated

gene expression and clearly limited glycerol synthesis. However, the low expression of GPD1

and GPD2 resulted in reduced growth of the promoter replaced strains at high extra-cellular

osmolarity or anaerobic conditions similar to the growth behaviour that was observed for the

gpd1Δ gpd2Δ mutant. The GPD1 and GPD2 mutants were thoroughly characterized for

growth and product formation during both aerobic and anaerobic very high ethanol

performance fermentation (Pagliardini et al, 2013; Pagliardini et al, 2010). Under aerobic

conditions, the yeast’s growth and metabolic activity was not influenced by the low GPDH

activity and was at wild-type level in aerobic fermentation (Pagliardini et al, 2010).

Productivity and viability was maintained in spite of lowered GPDH activity. In contrast, the

Chapter V 159

same strains tested under anaerobic conditions showed strongly impaired growth and

fermentation abilities besides the reduction in glycerol yield (Pagliardini et al, 2013).

Interestingly, cell viability in the presence of high ethanol concentrations was drastically

lower under anoxic conditions, pointing out that even limited changes in minor pathways,

such as glycerol formation, impact the global cellular network. In this context, promoter

engineering provided a unique tool to estimate the glycerol demand of S. cerevisiae under

hyperosmotic or anaerobic conditions. Moreover, it allowed to investigate the metabolic and

physiological changes in yeast at low versus abolished glycerol formation. Surprisingly,

changes were not only found for the usual suspects, i.e. osmo-tolerance and redox balancing,

but also for cell viability in the presence of high ethanol concentrations.

Deletion of the genes, coding for the glycerol pathway enzymes, caused severe side-effects on

several industrially relevant traits, as was shown by inhibiting or gradually reducing glycerol

formation. Hence, future engineering approaches to reduce glycerol should allow for cellular

integrity, improving the strain in a subtle manner. Supposedly, lowering glycerol formation

requires the modification of several target genes, which are linked to this phenotype. Several

attempts focused on solving the redox imbalance, caused by glycerol reduction or

abolishment. However, complete abolishment of glycerol formation makes cells susceptible to

change in osmolarity. One way to re-establish the osmotolerant state is the replacement of

glycerol by other compatible solutes, as recently shown for trehalose (Guo et al, 2011).

Alternatively, it is possible to allow for controlled but not excessive glycerol production to

enable cellular homeostasis upon changes in osmolarity. As depicted in Figure 1, the high

osmolarity glycerol pathway in S. cerevisiae enables a multi-level control of glycerol

formation in response to high osmolarity.

The primary level of manipulation affects or interrupts the osmo-sensing or signal

transduction process. In the present work, the frameshift mutation found in Ssk1 apparently

prevents signal transduction and hence also activation of the MAPK Hog1 (Figure 1A). The

deletion of SSK1 inhibits signal transduction of the Sln1 branch but this does not result in an

osmo-sensitive state. In fact, O’Rourke and Herskowitz (2004) suggested that the Sln1-branch

operates under conditions of moderate osmolarity. Severe osmo-stress disables the sensor and

requires the activity of the Sho1-branch and the general stress response pathway (O'Rourke et

al, 2004). The SSK1 variant, ssk1E330N…K356N

, might permanently disable signal transduction

160 Chapter V

from the Sln1-branch. Still, yeast cells are able to withstand modest osmotic stress without

glycerol accumulation. The ssk1E330N…K356N

variant is therefore an interesting gene tool to

render yeast less sensitive to changes in osmolarity while activation of Hog1 only occurs

during extreme hyperosmolarity. Besides Hog1 activation, downstream signalling targets or

promoters of structural genes can be targeted to modify the induction of transcription of

GPD1 and GPP2, coding for the structural enzymes of glycerol synthesis (Figure 1B).

Figure 1 Multi-level control of high osmolarity glycerol pathway in Saccharomyces cerevisiae. Cellular

glycerol formation in response to high osmolarity can be controlled at several levels: (A) the interruption of the

osmo-sensing process or the signal transduction by Ssk1 prevents the activation of the MAPK Hog1; (B) after

Hog1 activation, GPD1 promoter replacement or the utilization of the low-activity transcriptional activators,

Hot1 or Smp1, reduces the transcription of the GDP1 gene; (C) expressing allelic variants of the glycerol 3-

phosphate dehydrogenase (GPDH) with lower activity compared to the natural GPD1 allele.

Chapter V 161

In fact, the deregulation of GPD1 transcription by promoter replacement probably disabled

the induction of its expression during hyperosmotic conditions. However, this deregulation

resulted in side-effects; hence more subtle modification is necessary to allow for cellular

integrity. Such modifications were found in the modified transcriptional activator, Hot1 or

Smp1, both required for transcriptional induction of osmo-responsive genes. There is evidence

that mutations in HOT1 significantly reduce GPD1 transcription (Rep et al, 1999); however,

it remains to be elucidated whether the mutations we found in HOT1 cause a decrease in

activity of the transcription factor.

Finally, reduction of glycerol formation in yeast is possible using low-activity variants of

glycerol 3-phosphate dehydrogenase (GPDH). The GPDH activity measured in the superior

parental strains, CBS4C and segregant 26B, was consistently lower than that in ER7A. Future

experiments should give an answer to the question whether this is due to a decreased GPD1

expression level, which result from the allelic variants of GPD1, SMP1 and HOT1.

3. Extending the toolbox for yeast metabolic engineering

Metabolic engineering, as a new engineering discipline, emerged from the objective of

developing novel microorganisms, which can better fulfil industrial requirements. Classical

approaches of strain improvement, like mutagenesis or breeding, reached their limitations,

because today’s objectives in microorganism engineering aim far beyond the natural substrate

or product range of the species itself. The implementation of new dissimilation or bio-

synthesis pathways requires an extended genetic tool set, which enables the establishment of

the phenotype relevant genetic configuration. In recent years, the genetic engineering of S.

cerevisiae has made significant progress and genetic modifications are possible in manifold

ways due to the availability of a large genetic tool set, comprising several types of expression

vectors, reporter genes and selectable or counter-selectable markers for a highly efficient

transformation (Nevoigt, 2008). Through the present work, we showed that this toolbox is still

extendable particularly with regards to tools for fine adjustment of cellular activities.

Promoter engineering and the identification of mutations defining the polygenic trait ‘low

glycerol production’ in natural strains, provide two novel avenues for the improvement of the

currently used bioethanol production yeast strains, with minimal risk of affecting other

162 Chapter V

commercially important properties of the strains. Promoter engineering is a modern advanced

tool in rational metabolic engineering and helps finding a compromise between extreme

phenotypes, in our case the desired phenotype of low glycerol production and the undesired

phenotype of growth inhibition under anaerobic conditions. The promoter replacement

cassettes (Nevoigt et al, 2006), which we have used to modify glycerol synthesis in the

CEN.PK strain family, are now available for the modification of glycerol formation in

production strains.

Besides controlling the transcription of the isoenzymes catalysing the rate-controlling step of

glycerol formation, the natural gene variants of GPD1, SMP1 and HOT1, which were found

in CBS6412 and found to be linked to the ‘low glycerol’ phenotype in this work, offer a

further potential for the fine adjustment of glycerol formation and the improvement of the

currently used production strains. The rapid and efficient identification of such mutant alleles

over the whole genome has been facilitated by development of the genetic analysis platform

‘pooled-segregant whole-genome sequence analysis’. This platform is an excellent tool for

reverse engineering of S. cerevisiae, particularly with regards to the characterization of non-

selectable quantitative traits, which undergo polygenic inheritance in yeast.

4. Designed on the drawing board: Future cell factories

Back in the early days of metabolic engineering, this discipline was defined ‘as the purposeful

modification of intermediary metabolism using recombinant DNA techniques’ (Cameron and

Tong, 1993). Such modification comprised the production improvement of host own or of

new chemicals, the extensions of the substrate range of a host organism, and the addition of

new catabolic activities. In those days, the complexity of a metabolic network was the critical

step in the implementation of the new metabolic constraints (Stephanopoulos and Sinskey,

1993). Today, powerful experimental and mathematical methods are available to determine

this network, allowing for the precise quantification of in vivo fluxes and the subsequent

identification of critical nodes in the metabolic networks. As a relatively young scientific

discipline, metabolic engineering currently revolutionizes chemical manufacturing, using

knowledge on cellular metabolism for the production of valuable chemicals. So far, the re-

engineering of existing organisms has been the prime goal of metabolic engineering. Beyond

Chapter V 163

that, what are the future challenges in metabolic engineering and towards which future is this

discipline heading? Generally, the goal of many engineering disciplines is the creation of new

products, which are helpful in our daily life. For instance the invention of the steam engine or

fuel engines was revolutionary in the sense, that it replaced the force of humans or animals by

a physical or a chemical process for the generation of the necessary force for movements.

Accordingly, the creation of new ‘molecular machines’ may allow us to manufacture nano-

molecular products far beyond human visibility. First steps are made by Gibson and

colleagues (2010), which recently synthesized and assembled the genome of Mycoplasma

mycoides and Voigt and co-workers (Moon et al, 2012), which aims to implement genetic

programs to control the metabolic networks, similar to circuits in electronics. Our future

microbial production strains might soon be designed on the drawing board as a ‘cell factory’,

perfectly suited for its purpose.

References

166 References

Akada R, Hirosawa I, Kawahata M, Hoshida H, Nishizawa Y (2002) Sets of integrating plasmids and gene disruption

cassettes containing improved counter-selection markers designed for repeated use in budding yeast. Yeast 19: 393-402.

Akada R, Matsuo K, Aritomi K, Nishizawa Y (1999) Construction of recombinant sake yeast containing a dominant

FAS2 mutation without extraneous sequences by a two-step gene replacement protocol. J Biosci Bioeng 87: 43-48.

Albers E, Larsson C, Liden G, Niklasson C, Gustafsson L (1996) Influence of the nitrogen source on Saccharomyces

cerevisiae anaerobic growth and product formation. Appl Environ Microbiol 62: 3187-3195.

Albertyn J, Hohmann S, Thevelein JM, Prior BA (1994) GPD1, which encodes glycerol-3-phosphate dehydrogenase, is

essential for growth under osmotic stress in Saccharomyces cerevisiae, and its expression is regulated by the high-

osmolarity glycerol response pathway. Mol Cell Biol 14: 4135-4144.

Aldiguier AS, Alfenore S, Cameleyre X, Goma G, Uribelarrea JL, Guillouet SE, Molina-Jouve C (2004) Synergistic

temperature and ethanol effect on Saccharomyces cerevisiae dynamic behaviour in ethanol bio-fuel production.

Bioprocess Biosyst Eng 26: 217-222.

Alepuz PM, de Nadal E, Zapater M, Ammerer G, Posas F (2003) Osmostress-induced transcription by Hot1 depends on a

Hog1-mediated recruitment of the RNA Pol II. EMBO J 22: 2433-2442.

Alepuz PM, Jovanovic A, Reiser V, Ammerer G (2001) Stress-induced map kinase Hog1 is part of transcription

activation complexes. Mol Cell 7: 767-777.

Alexander MR, Tyers M, Perret M, Craig BM, Fang KS, Gustin MC (2001) Regulation of cell cycle progression by

Swe1p and Hog1p following hypertonic stress. Mol Biol Cell 12: 53-62.

Alfenore S, Cameleyre X, Benbadis L, Bideaux C, Uribelarrea JL, Goma G, Molina-Jouve C, Guillouet SE (2004)

Aeration strategy: a need for very high ethanol performance in Saccharomyces cerevisiae fed-batch process. Appl

Microbiol Biotechnol 63: 537-542.

Alfenore S, Molina-Jouve C, Guillouet SE, Uribelarrea JL, Goma G, Benbadis L (2002) Improving ethanol production

and viability of Saccharomyces cerevisiae by a vitamin feeding strategy during fed-batch process. Appl Microbiol

Biotechnol 60: 67-72.

Alper H, Fischer C, Nevoigt E, Stephanopoulos G (2005) Tuning genetic control through promoter engineering. Proc

Natl Acad Sci U S A 102: 12678-12683.

Alper H, Moxley J, Nevoigt E, Fink GR, Stephanopoulos G (2006) Engineering yeast transcription machinery for

improved ethanol tolerance and production. Science 314: 1565-1568.

Alper H, Stephanopoulos G (2007) Global transcription machinery engineering: a new approach for improving cellular

phenotype. Metab Eng 9: 258-267.

Ambroset C, Petit M, Brion C, Sanchez I, Delobel P, Guerin C, Chiapello H, Nicolas P, Bigey F, Dequin S, Blondin B

(2011) Deciphering the molecular basis of wine yeast fermentation traits using a combined genetic and genomic

approach. G3 (Bethesda) 1: 263-281.

Anderlund M, Nissen TL, Nielsen J, Villadsen J, Rydstrom J, Hahn-Hagerdal B, Kielland-Brandt MC (1999) Expression

of the Escherichia coli pntA and pntB genes, encoding nicotinamide nucleotide transhydrogenase, in Saccharomyces

cerevisiae and its effect on product formation during anaerobic glucose fermentation. Appl Environ Microbiol 65: 2333-

2340.

Anderson RM, Bitterman KJ, Wood JG, Medvedik O, Cohen H, Lin SS, Manchester JK, Gordon JI, Sinclair DA (2002)

Manipulation of a nuclear NAD+ salvage pathway delays aging without altering steady-state NAD+ levels. J Biol Chem

277: 18881-18890.

Andre L, Hemming A, Adler L (1991) Osmoregulation in Saccharomyces cerevisiae. Studies on the osmotic induction of

glycerol production and glycerol-3-phosphate dehydrogenase (NAD+). FEBS Lett 286: 13-17.

Ansell R, Granath K, Hohmann S, Thevelein JM, Adler L (1997) The two isoenzymes for yeast NAD+-dependent

glycerol 3-phosphate dehydrogenase encoded by GPD1 and GPD2 have distinct roles in osmoadaptation and redox

regulation. Embo J 16: 2179-2187.

References 167

Arshad M, Khan ZM, Khalil ur R, Shah FA, Rajoka MI (2008) Optimization of process variables for minimization of

byproduct formation during fermentation of blackstrap molasses to ethanol at industrial scale. Lett Appl Microbiol 47:

410-414.

Bai FW, Anderson WA, Moo-Young M (2008) Ethanol fermentation technologies from sugar and starch feedstocks.

Biotechnol Adv 26: 89-105.

Bailey JE (1991) Toward a science of metabolic engineering. Science 252: 1668-1675.

Bailey JE, Sburlati A, Hatzimanikatis V, Lee K, Renner WA, Tsai PS (1996) Inverse metabolic engineering: A strategy

for directed genetic engineering of useful phenotypes. Biotechnol Bioeng 52: 109-121.

Bakker BM, Overkamp KM, van Maris AJ, Kotter P, Luttik MA, van Dijken JP, Pronk JT (2001) Stoichiometry and

compartmentation of NADH metabolism in Saccharomyces cerevisiae. FEMS Microbiol Rev 25: 15-37.

Barnett JA (2003) Beginnings of microbiology and biochemistry: the contribution of yeast research. Microbiology 149:

557-567.

Basso LC, Basso TO, Rocha SN (eds) (2011) Ethanol Production in Brazil: The Industrial Process and Its Impact on

Yeast Fermentation, Biofuel Production-Recent Developments and Prospects: InTech.

Basso LC, de Amorim HV, de Oliveira AJ, Lopes ML (2008) Yeast selection for fuel ethanol production in Brazil. FEMS

Yeast Res 8: 1155-1163.

Beavis W (1998) QTL analyses: Power, precision, and accuracy. In Molecular Dissection of Complex Traits, Paterson

AH (ed), pp 145-161: CRC PressINC.

Bideaux C, Alfenore S, Cameleyre X, Molina-Jouve C, Uribelarrea JL, Guillouet SE (2006) Minimization of glycerol

production during the high-performance fed-batch ethanolic fermentation process in Saccharomyces cerevisiae, using a

metabolic model as a prediction tool. Appl Environ Microbiol 72: 2134-2140.

Bilsland-Marchesan E, Arino J, Saito H, Sunnerhagen P, Posas F (2000) Rck2 kinase is a substrate for the osmotic stress-

activated mitogen-activated protein kinase Hog1. Mol Cell Biol 20: 3887-3895.

Birkeland SR, Jin N, Ozdemir AC, Lyons RH, Jr., Weisman LS, Wilson TE (2010) Discovery of mutations in

Saccharomyces cerevisiae by pooled linkage analysis and whole-genome sequencing. Genetics 186: 1127-1137.

Bjorkqvist S, Ansell R, Adler L, Liden G (1997) Physiological response to anaerobicity of glycerol-3-phosphate

dehydrogenase mutants of Saccharomyces cerevisiae. Appl Environ Microbiol 63: 128-132.

Blomberg A (2000) Metabolic surprises in Saccharomyces cerevisiae during adaptation to saline conditions: questions,

some answers and a model. FEMS Microbiol Lett 182: 1-8.

Blomberg A, Adler L (1989) Roles of glycerol and glycerol-3-phosphate dehydrogenase (NAD+) in acquired

osmotolerance of Saccharomyces cerevisiae. J Bacteriol 171: 1087-1092.

Boender LG, de Hulster EA, van Maris AJ, Daran-Lapujade PA, Pronk JT (2009) Quantitative physiology of

Saccharomyces cerevisiae at near-zero specific growth rates. Appl Environ Microbiol 75: 5607-5614.

Boles E, Lehnert W, Zimmermann FK (1993) The role of the NAD-dependent glutamate dehydrogenase in restoring

growth on glucose of a Saccharomyces cerevisiae phosphoglucose isomerase mutant. Eur J Biochem 217: 469-477.

Boulahya (2005) Evaluation des potentialités fermentaires de souches mutées de S. cerevisiae en vue d'une production

nulle de glycérol dans une fermentation éthanolique. Université de Toulouse, INSA Toulouse, France.

Bouwman J, Kiewiet J, Lindenbergh A, van Eunen K, Siderius M, Bakker BM (2011) Metabolic regulation rather than de

novo enzyme synthesis dominates the osmo-adaptation of yeast. Yeast 28: 43-53.

Brem RB, Yvert G, Clinton R, Kruglyak L (2002) Genetic dissection of transcriptional regulation in budding yeast.

Science 296: 752-755.

Brewster J, de Valoir T, Dwyer N, Winter E, Gustin M (1993) An osmosensing signal transduction pathway in yeast.

Science 259: 1760-1763.

168 References

Bro C, Regenberg B, Forster J, Nielsen J (2006) In silico aided metabolic engineering of Saccharomyces cerevisiae for

improved bioethanol production. Metab Eng 8: 102-111.

Brookfield JF (2012) Heritability. Curr Biol 22: R217-219.

Brown AD (1978) Compatible solutes and extreme water stress in eukaryotic micro-organisms. Adv Microb Physiol 17:

181-242.

Buchner H (1897) Die Bedeutung der aktiven löslichen Zellprodukte für den Chemismus der Zelle. Münchener,

Medizinische Wochenschrift 44: 299–302, 321–322.

Camarasa C, Sanchez I, Brial P, Bigey F, Dequin S (2011) Phenotypic landscape of Saccharomyces cerevisiae during

wine fermentation: evidence for origin-dependent metabolic traits. PLoS One 6: e25147.

Cambon B, Monteil V, Remize F, Camarasa C, Dequin S (2006) Effects of GPD1 overexpression in Saccharomyces

cerevisiae commercial wine yeast strains lacking ALD6 genes. Appl Environ Microbiol 72: 4688-4694.

Cameron DC, Tong IT (1993) Cellular and metabolic engineering. An overview. Appl Biochem Biotechnol 38: 105-140.

Cao L, Zhang A, Kong Q, Xu X, Josine TL, Chen X (2007) Overexpression of GLT1 in fps1DeltagpdDelta mutant for

optimum ethanol formation by Saccharomyces cerevisiae. Biomol Eng 24: 638-642.

Costenoble R, Valadi H, Gustafsson L, Niklasson C, Franzen CJ (2000) Microaerobic glycerol formation in

Saccharomyces cerevisiae. Yeast 16: 1483-1495.

Cot M, Loret MO, Francois J, Benbadis L (2007) Physiological behaviour of Saccharomyces cerevisiae in aerated fed-

batch fermentation for high level production of bioethanol. FEMS Yeast Res 7: 22-32.

Crabtree HG (1928) The carbohydrate metabolism of certain pathological overgrowths. Biochem J 22: 1289-1298.

Cronwright GR, Rohwer JM, Prior BA (2002) Metabolic control analysis of glycerol synthesis in Saccharomyces

cerevisiae. Appl Environ Microbiol 68: 4448-4456.

D'Amato ME, Ehrenreich L, Cloete K, Benjeddou M, Davison S (2010) Characterization of the highly discriminatory loci

DYS449, DYS481, DYS518, DYS612, DYS626, DYS644 and DYS710. Forensic Sci Int Genet 4: 104-110.

Darku ID, Richard TL (2001) Biofuels: Ethanol Producers. In eLS: John Wiley & Sons, Ltd.

de Nadal E, Alepuz PM, Posas F (2002) Dealing with osmostress through MAP kinase activation. EMBO Rep 3: 735-

740.

de Nadal E, Casadome L, Posas F (2003) Targeting the MEF2-like transcription factor Smp1 by the stress-activated Hog1

mitogen-activated protein kinase. Mol Cell Biol 23: 229-237.

de Nadal E, Posas F (2010) Multilayered control of gene expression by stress-activated protein kinases. EMBO J 29: 4-

13.

de Nadal E, Posas F (2011) Elongating under Stress. Genet Res Int 2011: 326286.

De Nadal E, Zapater M, Alepuz PM, Sumoy L, Mas G, Posas F (2004) The MAPK Hog1 recruits Rpd3 histone

deacetylase to activate osmoresponsive genes. Nature 427: 370-374.

Demogines A, Smith E, Kruglyak L, Alani E (2008) Identification and dissection of a complex DNA repair sensitivity

phenotype in Baker's yeast. PLoS Genet 4: e1000123.

Dequin S, Casaregola S (2011) The genomes of fermentative Saccharomyces. C R Biol 334: 687-693.

Deutschbauer AM, Davis RW (2005) Quantitative trait loci mapped to single-nucleotide resolution in yeast. Nat Genet

37: 1333-1340.

Devantier R, Scheithauer B, Villas-Boas SG, Pedersen S, Olsson L (2005) Metabolite profiling for analysis of yeast stress

response during very high gravity ethanol fermentations. Biotechnol Bioeng 90: 703-714.

References 169

Dihazi H, Kessler R, Eschrich K (2004) High osmolarity glycerol (HOG) pathway-induced phosphorylation and

activation of 6-phosphofructo-2-kinase are essential for glycerol accumulation and yeast cell proliferation under

hyperosmotic stress. J Biol Chem 279: 23961-23968.

Dodou E, Treisman R (1997) The Saccharomyces cerevisiae MADS-box transcription factor Rlm1 is a target for the

Mpk1 mitogen-activated protein kinase pathway. Mol Cell Biol 17: 1848-1859.

Donalies UE, Nguyen HT, Stahl U, Nevoigt E (2008) Improvement of Saccharomyces yeast strains used in brewing,

wine making and baking. Adv Biochem Eng Biotechnol 111: 67-98.

Dos Santos MA (1997) Energy analysis of crops used for producing ethanol and CO2 emissions. The International

Virtual Institute of Global Change (IVIG) availabe at http://wwwivigcoppeufrjbr.

Douglas Crabb W, Mitchinson C (1997) Enzymes involved in the processing of starch to sugars. Trends in biotechnology

15: 349-352.

Duitama J, Srivastava P, Mandoiu I (2012) Towards accurate detection and genotyping of expressed variants from whole

transcriptome sequencing data. BMC Genomics 13: S6.

Duitama J, Srivastava PK, Ma, x, ndoiu II (2011) Towards accurate detection and genotyping of expressed variants from

whole transcriptome sequencing data. In Computational Advances in Bio and Medical Sciences (ICCABS), 2011 IEEE 1st

International Conference on pp 87-92.

Dujon B (1996) The yeast genome project: what did we learn? Trends Genet 12: 263-270.

Duong CT, Strack L, Futschik M, Katou Y, Nakao Y, Fujimura T, Shirahige K, Kodama Y, Nevoigt E Identification of

Sc-type ILV6 as a target to reduce diacetyl formation in lager brewers' yeast. Metab Eng 13: 638-647.

Edgley M, Brown AD (1983) Yeast Water Relations: Physiological Changes Induced by Solute Stress in Saccharomyces

cerevisiae and Saccharomyces rouxii. Journal of General Microbiology 129: 3453-3463.

Eglinton JM, Heinrich AJ, Pollnitz AP, Langridge P, Henschke PA, de Barros Lopes M (2002) Decreasing acetic acid

accumulation by a glycerol overproducing strain of Saccharomyces cerevisiae by deleting the ALD6 aldehyde

dehydrogenase gene. Yeast 19: 295-301.

Ehrenreich IM, Torabi N, Jia Y, Kent J, Martis S, Shapiro JA, Gresham D, Caudy AA, Kruglyak L (2010) Dissection of

genetically complex traits with extremely large pools of yeast segregants. Nature 464: 1039-1042.

Ehsani M, Fernandez MR, Biosca JA, Julien A, Dequin S (2009) Engineering of 2,3-butanediol dehydrogenase to reduce

acetoin formation by glycerol-overproducing, low-alcohol Saccharomyces cerevisiae. Appl Environ Microbiol 75: 3196-

3205.

Eriksson P, Andre L, Ansell R, Blomberg A, Adler L (1995) Cloning and characterization of GPD2, a second gene

encoding sn-glycerol 3-phosphate dehydrogenase (NAD+) in Saccharomyces cerevisiae, and its comparison with GPD1.

Mol Microbiol 17: 95-107.

EU (2003) DIRECTIVE 2003/30/EC OF THE EUROPEAN PARLIAMENT AND OF THE COUNCIL of 8 May 2003

on the promotion of the use of biofuels or other renewable fuels for transport. Official Journal of the European Union.

F.O.Licht's (2013) World Ethanol & Biofuels Report. 11.

FAO (2006) Food Outlook Global Market Analysis.

FAO (2008) The State of Food and Agriculture; Biofuels: prospects, risks and opportunities.

Farrell AE, Plevin RJ, Turner BT, Jones AD, O'Hare M, Kammen DM (2006) Ethanol can contribute to energy and

environmental goals. Science 311: 506-508.

Ferreira C, van Voorst F, Martins A, Neves L, Oliveira R, Kielland-Brandt MC, Lucas C, Brandt A (2005) A member of

the sugar transporter family, Stl1p is the glycerol/H+ symporter in Saccharomyces cerevisiae. Mol Biol Cell 16: 2068-

2076.

Ferrigno P, Posas F, Koepp D, Saito H, Silver PA (1998) Regulated nucleo/cytoplasmic exchange of HOG1 MAPK

requires the importin beta homologs NMD5 and XPO1. EMBO J 17: 5606-5614.

170 References

Fischer CR, Alper H, Nevoigt E, Jensen KL, Stephanopoulos G (2006) Response to Hammer et al.: Tuning genetic

control--importance of thorough promoter characterization versus generating promoter diversity. Trends Biotechnol 24:

55-56.

Fischer CR, Klein-Marcuschamer D, Stephanopoulos G (2008) Selection and optimization of microbial hosts for biofuels

production. Metab Eng 10: 295-304.

Fischer E (1894) Synthesen in der Zuckergruppe II. Berichte der deutschen chemischen Gesellschaft 27: 3189-3232.

Fischer E, Thierfelder H (1894) Verhalten der verschiedenen Zucker gegen reine Hefen. Ber Dtsch Chem Ges 27: 2031–

2037.

Forsburg SL, Nurse P (1991) Cell cycle regulation in the yeasts Saccharomyces cerevisiae and Schizosaccharomyces

pombe. Annu Rev Cell Biol 7: 227-256.

Galbe M, Sassner P, Wingren A, Zacchi G (2007) Process engineering economics of bioethanol production. Adv Biochem

Eng Biotechnol 108: 303-327.

Galibert F, Alexandraki D, Baur A, Boles E, Chalwatzis N, Chuat JC, Coster F, Cziepluch C, De Haan M, Domdey H,

Durand P, Entian KD, Gatius M, Goffeau A, Grivell LA, Hennemann A, Herbert CJ, Heumann K, Hilger F, Hollenberg

CP, Huang ME, Jacq C, Jauniaux JC, Katsoulou C, Karpfinger-Hartl L, et al. (1996) Complete nucleotide sequence of

Saccharomyces cerevisiae chromosome X. Embo J 15: 2031-2049.

Gancedo C, Gancedo JM, Sols A (1968) Glycerol metabolism in yeasts. Pathways of utilization and production. Eur J

Biochem 5: 165-172.

Gardner N, Rodrigue N, Champagne CP (1993) Combined effects of sulfites, temperature, and agitation time on

production of glycerol in grape juice by Saccharomyces cerevisiae. Appl Environ Microbiol 59: 2022-2028.

Gasch AP, Spellman PT, Kao CM, Carmel-Harel O, Eisen MB, Storz G, Botstein D, Brown PO (2000) Genomic

expression programs in the response of yeast cells to environmental changes. Mol Biol Cell 11: 4241-4257.

Gay-Lussac JL (1810) Extrait d'une mémoire sur la fermentation. Ann Chim 76: 245–259.

Geertman JM, van Dijken JP, Pronk JT (2006) Engineering NADH metabolism in Saccharomyces cerevisiae: formate as

an electron donor for glycerol production by anaerobic, glucose-limited chemostat cultures. FEMS Yeast Res 6: 1193-

1203.

Geertman JM, van Maris AJ, van Dijken JP, Pronk JT (2006) Physiological and genetic engineering of cytosolic redox

metabolism in Saccharomyces cerevisiae for improved glycerol production. Metab Eng 8: 532-542.

Geijer C, Ahmadpour D, Palmgren M, Filipsson C, Klein DM, Tamas MJ, Hohmann S, Lindkvist-Petersson K (2012)

Yeast aquaglyceroporins use the transmembrane core to restrict glycerol transport. J Biol Chem 287: 23562-23570.

Gervais P, Beney L (2001) Osmotic mass transfer in the yeast Saccharomyces cerevisiae. Cell Mol Biol (Noisy-le-grand)

47: 831-839.

Giaever G, Chu AM, Ni L, Connelly C, Riles L, Veronneau S, Dow S, Lucau-Danila A, Anderson K, Andre B, Arkin

AP, Astromoff A, El-Bakkoury M, Bangham R, Benito R, Brachat S, Campanaro S, Curtiss M, Davis K, Deutschbauer

A, Entian KD, Flaherty P, Foury F, Garfinkel DJ, Gerstein M, Gotte D, Guldener U, Hegemann JH, Hempel S, Herman

Z, Jaramillo DF, Kelly DE, Kelly SL, Kotter P, LaBonte D, Lamb DC, Lan N, Liang H, Liao H, Liu L, Luo C, Lussier M,

Mao R, Menard P, Ooi SL, Revuelta JL, Roberts CJ, Rose M, Ross-Macdonald P, Scherens B, Schimmack G, Shafer B,

Shoemaker DD, Sookhai-Mahadeo S, Storms RK, Strathern JN, Valle G, Voet M, Volckaert G, Wang CY, Ward TR,

Wilhelmy J, Winzeler EA, Yang Y, Yen G, Youngman E, Yu K, Bussey H, Boeke JD, Snyder M, Philippsen P, Davis

RW, Johnston M (2002) Functional profiling of the Saccharomyces cerevisiae genome. Nature 418: 387-391.

Gibson DG, Glass JI, Lartigue C, Noskov VN, Chuang R-Y, Algire MA, Benders GA, Montague MG, Ma L, Moodie

MM, Merryman C, Vashee S, Krishnakumar R, Assad-Garcia N, Andrews-Pfannkoch C, Denisova EA, Young L, Qi Z-

Q, Segall-Shapiro TH, Calvey CH, Parmar PP, Hutchison CA, Smith HO, Venter JC (2010) Creation of a Bacterial Cell

Controlled by a Chemically Synthesized Genome. Science 329: 52-56.

Gietz D, St Jean A, Woods RA, Schiestl RH (1992) Improved method for high efficiency transformation of intact yeast

cells. Nucleic Acids Res 20: 1425.

References 171

Gietz RD, Schiestl RH (1991) Applications of high efficiency lithium acetate transformation of intact yeast cells using

single-stranded nucleic acids as carrier. Yeast 7: 253-263.

Gietz RD, Schiestl RH (1991) Applications of high efficiency lithium acetate transformation of intact yeast cells using

single-stranded nucleic acids as carrier. Yeast 7: 253-263.

Gietz RD, Schiestl RH (2007) High-efficiency yeast transformation using the LiAc/SS carrier DNA/PEG method. Nat

Protoc 2: 31-34.

Goffeau A, Barrell BG, Bussey H, Davis RW, Dujon B, Feldmann H, Galibert F, Hoheisel JD, Jacq C, Johnston M, Louis

EJ, Mewes HW, Murakami Y, Philippsen P, Tettelin H, Oliver SG (1996) Life with 6000 genes. Science 274: 546-&.

Grant WD (2004) Life at low water activity. Philos Trans R Soc Lond B Biol Sci 359: 1249-1266; discussion 1266-1247.

Guadalupe Medina V, Almering MJ, van Maris AJ, Pronk JT (2010) Elimination of glycerol production in anaerobic

cultures of a Saccharomyces cerevisiae strain engineered to use acetic acid as an electron acceptor. Appl Environ

Microbiol 76: 190-195.

Gueldener U, Heinisch J, Koehler GJ, Voss D, Hegemann JH (2002) A second set of loxP marker cassettes for Cre-

mediated multiple gene knockouts in budding yeast. Nucleic Acids Res 30: e23.

Guo W, Adams V, Mason J, McCabe ER (1997) Identification of a ferritin light chain pseudogene near the glycerol

kinase locus in Xp21 by cDNA amplification for identification of genomic expressed sequences. Biochem Mol Med 60:

169-173.

Guo W, Lovell RS, Zhang YH, Huang BL, Burris TP, Craigen WJ, McCabe ER (1996) Ahch, the mouse homologue of

DAX1: cloning, characterization and synteny with GyK, the glycerol kinase locus. Gene 178: 31-34.

Guo W, Worley K, Adams V, Mason J, Sylvester-Jackson D, Zhang YH, Towbin JA, Fogt DD, Madu S, Wheeler DA, et

al. (1993) Genomic scanning for expressed sequences in Xp21 identifies the glycerol kinase gene. Nat Genet 4: 367-372.

Guo X, Huang Z, Szoka FC (2004) Improved preparation of PEG-diortho ester-diacyl glycerol conjugates. Methods

Enzymol 387: 147-152.

Guo X, Zhuge B, Zhuge J (2002) [Research progress on the glycerol kinase]. Wei Sheng Wu Xue Bao 42: 510-513.

Guo ZP, Zhang L, Ding ZY, Shi GY (2011) Minimization of glycerol synthesis in industrial ethanol yeast without

influencing its fermentation performance. Metab Eng 13: 49-59.

Guo ZP, Zhang L, Ding ZY, Wang ZX, Shi GY (2009) Interruption of glycerol pathway in industrial alcoholic yeasts to

improve the ethanol production. Appl Microbiol Biotechnol 82: 287-292.

Guo ZP, Zhang L, Ding ZY, Wang ZX, Shi GY (2010) Improving ethanol productivity by modification of glycolytic

redox factor generation in glycerol-3-phosphate dehydrogenase mutants of an industrial ethanol yeast. J Ind Microbiol

Biotechnol.

Guo ZP, Zhang L, Ding ZY, Wang ZX, Shi GY (2011) Improving ethanol productivity by modification of glycolytic

redox factor generation in glycerol-3-phosphate dehydrogenase mutants of an industrial ethanol yeast. J Ind Microbiol

Biotechnol 38: 935-943.

Harden A, Young WJ (1905) The influence of phosphates on the fermentation of glucose by yeast-juice: preliminary

communication. Proc Chem Soc (London) 21: 189–191.

Hermann BG, Patel M (2007) Today's and tomorrow's bio-based bulk chemicals from white biotechnology: a techno-

economic analysis. Appl Biochem Biotechnol 136: 361-388.

Heux S, Sablayrolles JM, Cachon R, Dequin S (2006) Engineering a Saccharomyces cerevisiae wine yeast that exhibits

reduced ethanol production during fermentation under controlled microoxygenation conditions. Appl Environ Microbiol

72: 5822-5828.

Hoffman CS, Winston F (1987) A ten-minute DNA preparation from yeast efficiently releases autonomous plasmids for

transformation of Escherichia coli. Gene 57: 267-272.

172 References

Hohmann S (2002) Osmotic stress signaling and osmoadaptation in yeasts. Microbiol Mol Biol Rev 66: 300-372.

Hohmann S (2009) Control of high osmolarity signalling in the yeast Saccharomyces cerevisiae. FEBS Lett 583: 4025-

4029.

Holst B, Lunde C, Lages F, Oliveira R, Lucas C, Kielland-Brandt MC (2000) GUP1 and its close homologue GUP2,

encoding multimembrane-spanning proteins involved in active glycerol uptake in Saccharomyces cerevisiae. Mol

Microbiol 37: 108-124.

Homer N, Merriman B, Nelson SF (2009) BFAST: an alignment tool for large scale genome resequencing. PLoS One 4:

e7767.

Hong KK, Nielsen J (2012) Metabolic engineering of Saccharomyces cerevisiae: a key cell factory platform for future

biorefineries. Cell Mol Life Sci 69: 2671-2690.

Hubmann G, Guillouet S, Nevoigt E (2011) Gpd1 and Gpd2 fine-tuning for sustainable reduction of glycerol formation in

Saccharomyces cerevisiae. Appl Environ Microbiol 77: 5857-5867.

Huxley C, Green ED, Dunham I (1990) Rapid assessment of S. cerevisiae mating type by PCR. Trends Genet 6: 236.

Jacques KA, Lyons TP, Kelsall DR (eds) (2003) The alcohol textbook - a reference for the beverage, fuel and industrial

alcohol industries. Nottingham: Nottingham University Press.

Jain VK, Divol B, Prior BA, Bauer FF (2011) Elimination of glycerol and replacement with alternative products in

ethanol fermentation by Saccharomyces cerevisiae. J Ind Microbiol Biotechnol 38: 1427-1435.

Johnston CG, Aust SD (1994) Detection of Phanerochaete chrysosporium in soil by PCR and restriction enzyme analysis.

Appl Environ Microbiol 60: 2350-2354.

Jones AM, Ingledew WM (1994) Fuel alcohol production: appraisal of nitrogenous yeast foods for very high gravity

wheat mash fermentation. Process Biochemistry 29: 483-488.

Jung JY, Kim TY, Ng CY, Oh MK (2012) Characterization of GCY1 in Saccharomyces cerevisiae by metabolic

profiling. J Appl Microbiol 113: 1468-1478.

Jung S, Marelli M, Rachubinski RA, Goodlett DR, Aitchison JD (2010) Dynamic changes in the subcellular distribution

of Gpd1p in response to cell stress. J Biol Chem 285: 6739-6749.

Kelsall DR, Lyons TP (2003) Grain dry milling and cooking procedures: extracting sugars in preparation for

fermentation. In The Alcohol Textbook - A reference for the beverage, fuel and industrial alcohol industries, 4th ed., pp

9-21. Nottingham: Nottingham University Press.

Kleijn RJ, Geertman JM, Nfor BK, Ras C, Schipper D, Pronk JT, Heijnen JJ, van Maris AJ, van Winden WA (2007)

Metabolic flux analysis of a glycerol-overproducing Saccharomyces cerevisiae strain based on GC-MS, LC-MS and

NMR-derived C-labelling data. FEMS Yeast Res 7: 216-231.

Koizumi T (2003) The Brazilian ethanol programme: impacts on world ethanol and sugar markets. FAO comodity and

trade policy research working paper No.1.

Kong QX, Cao LM, Zhang AL, Chen X (2007) Overexpressing GLT1 in gpd1Delta mutant to improve the production of

ethanol of Saccharomyces cerevisiae. Appl Microbiol Biotechnol 73: 1382-1386.

Kong QX, Gu JG, Cao LM, Zhang AL, Chen X, Zhao XM (2006) Improved production of ethanol by deleting FPS1 and

over-expressing GLT1 in Saccharomyces cerevisiae. Biotechnol Lett 28: 2033-2038.

Kong QX, Zhang AL, Cao LM, Chen X (2007) Over-expressing GLT1 in a gpd2Delta mutant of Saccharomyces

cerevisiae to improve ethanol production. Appl Microbiol Biotechnol 75: 1361-1366.

Kotaka A, Sahara H, Kondo A, Ueda M, Hata Y (2009) Efficient generation of recessive traits in diploid sake yeast by

targeted gene disruption and loss of heterozygosity. Appl Microbiol Biotechnol 82: 387-395.

Kuhn C, Petelenz E, Nordlander B, Schaber J, Hohmann S, Klipp E (2008) Exploring the impact of osmoadaptation on

glycolysis using time-varying response-coefficients. Genome Inform 20: 77-90.

References 173

Kumar P, Barrett DM, Delwiche MJ, Stroeve P (2009) Methods for Pretreatment of Lignocellulosic Biomass for Efficient

Hydrolysis and Biofuel Production. Industrial & Engineering Chemistry Research 48: 3713-3729.

Lages F, Silva-Graca M, Lucas C (1999) Active glycerol uptake is a mechanism underlying halotolerance in yeasts: a

study of 42 species. Microbiology 145 ( Pt 9): 2577-2585.

Larsson K, Ansell R, Eriksson P, Adler L (1993) A gene encoding sn-glycerol 3-phosphate dehydrogenase (NAD+)

complements an osmosensitive mutant of Saccharomyces cerevisiae. Mol Microbiol 10: 1101-1111.

Lavoisier AL (1789) Traité Élémentaire de Chimie. Paris: Cuchet [Translated by R Kerr (1790) as Elements of

Chemistry Edinburgh: Creech].

Liden G, Walfridsson M, Ansell R, Anderlund M, Adler L, Hahn-Hagerdal B (1996) A glycerol-3-phosphate

dehydrogenase-deficient mutant of Saccharomyces cerevisiae expressing the heterologous XYL1 gene. Appl Environ

Microbiol 62: 3894-3896.

Lin SJ, Guarente L (2003) Nicotinamide adenine dinucleotide, a metabolic regulator of transcription, longevity and

disease. Curr Opin Cell Biol 15: 241-246.

Lin Y, Tanaka S (2006) Ethanol fermentation from biomass resources: current state and prospects. Appl Microbiol

Biotechnol 69: 627-642.

Luttik MA, Overkamp KM, Kotter P, de Vries S, van Dijken JP, Pronk JT (1998) The Saccharomyces cerevisiae NDE1

and NDE2 genes encode separate mitochondrial NADH dehydrogenases catalyzing the oxidation of cytosolic NADH. J

Biol Chem 273: 24529-24534.

Luyten K, Albertyn J, Skibbe WF, Prior BA, Ramos J, Thevelein JM, Hohmann S (1995) Fps1, a yeast member of the

MIP family of channel proteins, is a facilitator for glycerol uptake and efflux and is inactive under osmotic stress. Embo J

14: 1360-1371.

Mabee WE (2007) Policy options to support biofuel production. Adv Biochem Eng Biotechnol 108: 329-357.

Madson PW, Monceaux DA (1999) Fuel ethanol production. In The Alcohol Textbook. A Reference for the Beverage,

Fuel and Industrial Alcohol Industries, Jacques KA, Lyons TP, Kelsall DR (eds), third ed. edn, pp 257–268. Nottingham,

UK: Nottingham University Press.

Maeda T, Takekawa M, Saito H (1995) Activation of yeast PBS2 MAPKK by MAPKKKs or by binding of an SH3-

containing osmosensor. Science 269: 554-558.

Maeda T, Watanabe Y, Kunitomo H, Yamamoto M (1994) Cloning of the pka1 gene encoding the catalytic subunit of the

cAMP-dependent protein kinase in Schizosaccharomyces pombe. J Biol Chem 269: 9632-9637.

Maeda T, Wurgler-Murphy SM, Saito H (1994) A two-component system that regulates an osmosensing MAP kinase

cascade in yeast. Nature 369: 242-245.

Marchler-Bauer A, Anderson JB, Chitsaz F, Derbyshire MK, DeWeese-Scott C, Fong JH, Geer LY, Geer RC, Gonzales

NR, Gwadz M, He S, Hurwitz DI, Jackson JD, Ke Z, Lanczycki CJ, Liebert CA, Liu C, Lu F, Lu S, Marchler GH,

Mullokandov M, Song JS, Tasneem A, Thanki N, Yamashita RA, Zhang D, Zhang N, Bryant SH (2009) CDD: specific

functional annotation with the Conserved Domain Database. Nucleic Acids Research 37: D205-D210.

MarketLine (2012) Alcoholic Drinks: Global Industry Guide. http://wwwjust-drinkscom/market-research/alcoholic-

drinks-global-industry-guide_id138428aspx.

Martinez-Montanes F, Pascual-Ahuir A, Proft M (2010) Toward a genomic view of the gene expression program

regulated by osmostress in yeast. OMICS 14: 619-627.

Martinez-Pastor MT, Marchler G, Schuller C, Marchler-Bauer A, Ruis H, Estruch F (1996) The Saccharomyces

cerevisiae zinc finger proteins Msn2p and Msn4p are required for transcriptional induction through the stress response

element (STRE). EMBO J 15: 2227-2235.

Mettetal JT, Muzzey D, Gomez-Uribe C, van Oudenaarden A (2008) The frequency dependence of osmo-adaptation in

Saccharomyces cerevisiae. Science 319: 482-484.

Mobini-Dehkordi M, Nahvi I, Zarkesh-Esfahani H, Ghaedi K, Tavassoli M, Akada R (2008) Isolation of a novel mutant

174 References

strain of Saccharomyces cerevisiae by an ethyl methane sulfonate-induced mutagenesis approach as a high producer of

bioethanol. J Biosci Bioeng 105: 403-408.

Molin M, Norbeck J, Blomberg A (2003) Dihydroxyacetone Kinases in Saccharomyces cerevisiaeAre Involved in

Detoxification of Dihydroxyacetone. Journal of Biological Chemistry 278: 1415-1423.

Mollapour M, Piper PW (2006) Hog1p mitogen-activated protein kinase determines acetic acid resistance in

Saccharomyces cerevisiae. FEMS Yeast Res 6: 1274-1280.

Moon TS, Lou C, Tamsir A, Stanton BC, Voigt CA (2012) Genetic programs constructed from layered logic gates in

single cells. Nature 491: 249-253.

Nadal-Ribelles M, Conde N, Flores O, Gonzalez-Vallinas J, Eyras E, Orozco M, de Nadal E, Posas F (2012) Hog1

bypasses stress-mediated down-regulation of transcription by RNA polymerase II redistribution and chromatin

remodeling. Genome Biol 13: R106.

Neuberg C, Reinfurth E (1918) Natürliche und erzwungene Glycerinbildung bei der alkoholischen Gärung. . BiochemZ

92: 32.

Neves L, Lages F, Lucas C (2004) New insights on glycerol transport in Saccharomyces cerevisiae. FEBS Lett 565: 160-

162.

Nevoigt E (2008) Progress in metabolic engineering of Saccharomyces cerevisiae. Microbiol Mol Biol Rev 72: 379-412.

Nevoigt E, Fischer C, Mucha O, Matthaus F, Stahl U, Stephanopoulos G (2007) Engineering promoter regulation.

Biotechnol Bioeng 96: 550-558.

Nevoigt E, Kohnke J, Fischer CR, Alper H, Stahl U, Stephanopoulos G (2006) Engineering of promoter replacement

cassettes for fine-tuning of gene expression in Saccharomyces cerevisiae. Appl Environ Microbiol 72: 5266-5273.

Nevoigt E, Pilger R, Mast-Gerlach E, Schmidt U, Freihammer S, Eschenbrenner M, Garbe L, Stahl U (2002) Genetic

engineering of brewing yeast to reduce the content of ethanol in beer. FEMS Yeast Res 2: 225-232.

Nevoigt E, Stahl U (1996) Reduced pyruvate decarboxylase and increased glycerol-3-phosphate dehydrogenase [NAD+]

levels enhance glycerol production in Saccharomyces cerevisiae. Yeast 12: 1331-1337.

Nevoigt E, Stahl U (1997) Osmoregulation and glycerol metabolism in the yeast Saccharomyces cerevisiae. FEMS

Microbiol Rev 21: 231-241.

Nguyen HT, Dieterich A, Athenstaedt K, Truong NH, Stahl U, Nevoigt E (2004) Engineering of Saccharomyces

cerevisiae for the production of L-glycerol 3-phosphate. Metab Eng 6: 155-163.

Nguyen HT, Nevoigt E (2009) Engineering of Saccharomyces cerevisiae for the production of dihydroxyacetone (DHA)

from sugars: a proof of concept. Metab Eng 11: 335-346.

Nicolaou SA, Gaida SM, Papoutsakis ET (2010) A comparative view of metabolite and substrate stress and tolerance in

microbial bioprocessing: From biofuels and chemicals, to biocatalysis and bioremediation. Metab Eng 12: 307-331.

Nissen TL, Anderlund M, Nielsen J, Villadsen J, Kielland-Brandt MC (2001) Expression of a cytoplasmic

transhydrogenase in Saccharomyces cerevisiae results in formation of 2-oxoglutarate due to depletion of the NADPH

pool. Yeast 18: 19-32.

Nissen TL, Hamann CW, Kielland-Brandt MC, Nielsen J, Villadsen J (2000) Anaerobic and aerobic batch cultivations of

Saccharomyces cerevisiae mutants impaired in glycerol synthesis. Yeast 16: 463-474.

Nissen TL, Kielland-Brandt MC, Nielsen J, Villadsen J (2000) Optimization of ethanol production in Saccharomyces

cerevisiae by metabolic engineering of the ammonium assimilation. Metab Eng 2: 69-77.

Norbeck J, Blomberg A (1997) Metabolic and regulatory changes associated with growth of Saccharomyces cerevisiae in

1.4 M NaCl. Evidence for osmotic induction of glycerol dissimilation via the dihydroxyacetone pathway. J Biol Chem

272: 5544-5554.

Norbeck J, Pahlman AK, Akhtar N, Blomberg A, Adler L (1996) Purification and characterization of two isoenzymes of

DL-glycerol-3-phosphatase from Saccharomyces cerevisiae. Identification of the corresponding GPP1 and GPP2 genes

References 175

and evidence for osmotic regulation of Gpp2p expression by the osmosensing mitogen-activated protein kinase signal

transduction pathway. J Biol Chem 271: 13875-13881.

Olesen K, Franke Johannesen P, Hoffmann L, Bech Sorensen S, Gjermansen C, Hansen J (2000) The pYC plasmids, a

series of cassette-based yeast plasmid vectors providing means of counter-selection. Yeast 16: 1035-1043.

Olesen K, Franke Johannesen P, Hoffmann L, Bech Sorensen S, Gjermansen C, Hansen J (2000) The pYC plasmids, a

series of cassette-based yeast plasmid vectors providing means of counter-selection. Yeast 16: 1035-1043.

O'Rourke SM, Herskowitz I (2002) A third osmosensing branch in Saccharomyces cerevisiae requires the Msb2 protein

and functions in parallel with the Sho1 branch. Mol Cell Biol 22: 4739-4749.

O'Rourke SM, Herskowitz I (2004) Unique and redundant roles for HOG MAPK pathway components as revealed by

whole-genome expression analysis. Mol Biol Cell 15: 532-542.

O'Rourke SM, Herskowitz I, O'Shea EK (2002) Yeast go the whole HOG for the hyperosmotic response. Trends Genet

18: 405-412.

Oud B, van Maris AJ, Daran JM, Pronk JT (2012) Genome-wide analytical approaches for reverse metabolic engineering

of industrially relevant phenotypes in yeast. FEMS Yeast Res 12: 183-196.

Overkamp KM, Bakker BM, Kotter P, Luttik MA, Van Dijken JP, Pronk JT (2002) Metabolic engineering of glycerol

production in Saccharomyces cerevisiae. Appl Environ Microbiol 68: 2814-2821.

Pagliardini J, Hubmann G, Alfenore S, Nevoigt E, Bideaux C, Guillouet SE (2013) The metabolic costs of improving

ethanol yield by reducing glycerol formation capacity under anaerobic conditions in Saccharomyces cerevisiae. Microb

Cell Fact 12: 29.

Pagliardini J, Hubmann G, Bideaux C, Alfenore S, Nevoigt E, Guillouet SE (2010) Quantitative evaluation of yeast's

requirement for glycerol formation in very high ethanol performance fed-batch process. Microb Cell Fact 9: 36.

Pahlman AK, Granath K, Ansell R, Hohmann S, Adler L (2001) The yeast glycerol 3-phosphatases Gpp1p and Gpp2p

are required for glycerol biosynthesis and differentially involved in the cellular responses to osmotic, anaerobic, and

oxidative stress. J Biol Chem 276: 3555-3563.

Parts L, Cubillos FA, Warringer J, Jain K, Salinas F, Bumpstead SJ, Molin M, Zia A, Simpson JT, Quail MA, Moses A,

Louis EJ, Durbin R, Liti G (2011) Revealing the genetic structure of a trait by sequencing a population under selection.

Genome Res 21: 1131-1138.

Pascual-Ahuir A, Serrano R, Proft M (2001) The Sko1p repressor and Gcn4p activator antagonistically modulate stress-

regulated transcription in Saccharomyces cerevisiae. Mol Cell Biol 21: 16-25.

Pasteur L (1857) Mémoire sur la fermentation alcoolique. Compt Rend 45: 1032–1036.

Pasteur ML (1858) Production constante de glycérine dans la fermentation alcoolique. C R Acad Sci 46: 857.

Pavlik P, Simon M, Schuster T, Ruis H (1993) The glycerol kinase (GUT1) gene of Saccharomyces cerevisiae: cloning

and characterization. Curr Genet 24: 21-25.

Pecori Giraldi F, Pesce S, Maroni P, Pagliardini L, Lasio G, Losa M, Cavagnini F (2010) Inhibitory effect of prepro-

thyrotrophin-releasing hormone (178-199) on adrenocorticotrophic hormone secretion by human corticotroph tumours. J

Neuroendocrinol 22: 294-300.

Peralta-Yahya PP, Keasling JD (2010) Advanced biofuel production in microbes. Biotechnol J 5: 147-162.

Peralta-Yahya PP, Zhang F, del Cardayre SB, Keasling JD (2012) Microbial engineering for the production of advanced

biofuels. Nature 488: 320-328.

Petelenz-Kurdziel E, Eriksson E, Smedh M, Beck C, Hohmann S, Goksor M (2011) Quantification of cell volume

changes upon hyperosmotic stress in Saccharomyces cerevisiae. Integr Biol (Camb) 3: 1120-1126.

Popp A, Nguyen HT, Boulahya K, Bideaux C, Alfenore S, Guillouet SE, Nevoigt E (2008) Fermentative production of L-

glycerol 3-phosphate utilizing a Saccharomyces cerevisiae strain with an engineered glycerol biosynthetic pathway.

Biotechnol Bioeng 100: 497-505.

176 References

Posas F, Chambers JR, Heyman JA, Hoeffler JP, de Nadal E, Arino J (2000) The transcriptional response of yeast to

saline stress. J Biol Chem 275: 17249-17255.

Posas F, Wurgler-Murphy SM, Maeda T, Witten EA, Thai TC, Saito H (1996) Yeast HOG1 MAP kinase cascade is

regulated by a multistep phosphorelay mechanism in the SLN1-YPD1-SSK1 "two-component" osmosensor. Cell 86:

865-875.

Proft M, Mas G, de Nadal E, Vendrell A, Noriega N, Struhl K, Posas F (2006) The stress-activated Hog1 kinase is a

selective transcriptional elongation factor for genes responding to osmotic stress. Mol Cell 23: 241-250.

Proft M, Pascual-Ahuir A, de Nadal E, Arino J, Serrano R, Posas F (2001) Regulation of the Sko1 transcriptional

repressor by the Hog1 MAP kinase in response to osmotic stress. EMBO J 20: 1123-1133.

Proft M, Struhl K (2002) Hog1 kinase converts the Sko1-Cyc8-Tup1 repressor complex into an activator that recruits

SAGA and SWI/SNF in response to osmotic stress. Mol Cell 9: 1307-1317.

Proft M, Struhl K (2004) MAP kinase-mediated stress relief that precedes and regulates the timing of transcriptional

induction. Cell 118: 351-361.

Pronk JT (2002) Auxotrophic yeast strains in fundamental and applied research. Appl Environ Microbiol 68: 2095-2100.

Raitt DC, Posas F, Saito H (2000) Yeast Cdc42 GTPase and Ste20 PAK-like kinase regulate Sho1-dependent activation

of the Hog1 MAPK pathway. EMBO J 19: 4623-4631.

Ralser M, Wamelink MM, Kowald A, Gerisch B, Heeren G, Struys EA, Klipp E, Jakobs C, Breitenbach M, Lehrach H,

Krobitsch S (2007) Dynamic rerouting of the carbohydrate flux is key to counteracting oxidative stress. J Biol 6: 10.

Remize F, Barnavon L, Dequin S (2001) Glycerol export and glycerol-3-phosphate dehydrogenase, but not glycerol

phosphatase, are rate limiting for glycerol production in Saccharomyces cerevisiae. Metab Eng 3: 301-312.

Remize F, Cambon B, Barnavon L, Dequin S (2003) Glycerol formation during wine fermentation is mainly linked to

Gpd1p and is only partially controlled by the HOG pathway. Yeast 20: 1243-1253.

Remize F, Roustan JL, Sablayrolles JM, Barre P, Dequin S (1999) Glycerol overproduction by engineered

saccharomyces cerevisiae wine yeast strains leads to substantial changes in By-product formation and to a stimulation of

fermentation rate in stationary phase. Appl Environ Microbiol 65: 143-149.

Rep M, Krantz M, Thevelein JM, Hohmann S (2000) The transcriptional response of Saccharomyces cerevisiae to

osmotic shock. Hot1p and Msn2p/Msn4p are required for the induction of subsets of high osmolarity glycerol pathway-

dependent genes. J Biol Chem 275: 8290-8300.

Rep M, Proft M, Remize F, Tamas M, Serrano R, Thevelein JM, Hohmann S (2001) The Saccharomyces cerevisiae

Sko1p transcription factor mediates HOG pathway-dependent osmotic regulation of a set of genes encoding enzymes

implicated in protection from oxidative damage. Mol Microbiol 40: 1067-1083.

Rep M, Reiser V, Gartner U, Thevelein JM, Hohmann S, Ammerer G, Ruis H (1999) Osmotic stress-induced gene

expression in Saccharomyces cerevisiae requires Msn1p and the novel nuclear factor Hot1p. Mol Cell Biol 19: 5474-

5485.

Rigoulet M, Aguilaniu H, Averet N, Bunoust O, Camougrand N, Grandier-Vazeille X, Larsson C, Pahlman IL, Manon S,

Gustafsson L (2004) Organization and regulation of the cytosolic NADH metabolism in the yeast Saccharomyces

cerevisiae. Mol Cell Biochem 256-257: 73-81.

Ronnow B, Kielland-Brandt MC (1993) GUT2, a gene for mitochondrial glycerol 3-phosphate dehydrogenase of

Saccharomyces cerevisiae. Yeast 9: 1121-1130.

Saerens SM, Duong CT, Nevoigt E Genetic improvement of brewer's yeast: current state, perspectives and limits. Appl

Microbiol Biotechnol 86: 1195-1212.

Sahara H, Kotaka A, Kondo A, Ueda M, Hata Y (2009) Using promoter replacement and selection for loss of

heterozygosity to generate an industrially applicable sake yeast strain that homozygously overproduces isoamyl acetate. J

Biosci Bioeng 108: 359-364.

References 177

Sambrook J, Maniatis T, Fritsch EF (1989) Molecular cloning : a laboratory manual, 2nd edn. Cold Spring Harbor,

N.Y.: Cold Spring Harbor Laboratory.

Sanchez-Gonzalez Y, Cameleyre X, Molina-Jouve C, Goma G, Alfenore S (2008) Dynamic microbial response under

ethanol stress to monitor Saccharomyces cerevisiae activity in different initial physiological states. Bioprocess Biosyst

Eng.

Sanger F, Coulson AR (1975) A rapid method for determining sequences in DNA by primed synthesis with DNA

polymerase. J Mol Biol 94: 441-448.

Sato N, Kawahara H, Toh-e A, Maeda T (2003) Phosphorelay-regulated degradation of the yeast Ssk1p response

regulator by the ubiquitin-proteasome system. Mol Cell Biol 23: 6662-6671.

Schmidtke LM, Blackman JW, Agboola SO (2012) Production technologies for reduced alcoholic wines. J Food Sci 77:

R25-41.

Schmidtke LM, Clark AC, Scollary GR (2011) Micro-oxygenation of red wine: techniques, applications, and outcomes.

Crit Rev Food Sci Nutr 51: 115-131.

Schuller D, Casal M (2005) The use of genetically modified Saccharomyces cerevisiae strains in the wine industry. Appl

Microbiol Biotechnol 68: 292-304.

Schuller D, Casal M (2005) The use of genetically modified Saccharomyces cerevisiae strains in the wine industry. Appl

Microbiol Biotechnol 68: 292-304.

Schwann T (1837) Vorläufige Mittheilung, bettreffend Versuche über die Weingährung und Fäulniss. Ann Phys Chem

41: 184–193.

Schwartz MA, Madhani HD (2004) Principles of MAP kinase signaling specificity in Saccharomyces cerevisiae. Annu

Rev Genet 38: 725-748.

Shepherd A, Piper PW (2010) The Fps1p aquaglyceroporin facilitates the use of small aliphatic amides as a nitrogen

source by amidase-expressing yeasts. FEMS Yeast Res 10: 527-534.

Sherman F, Hicks J (1991) Micromanipulation and dissection of asci. Methods Enzymol 194: 21-37.

Sicard D, Legras JL (2011) Bread, beer and wine: yeast domestication in the Saccharomyces sensu stricto complex. C R

Biol 334: 229-236.

Siderius M, Van Wuytswinkel O, Reijenga KA, Kelders M, Mager WH (2000) The control of intracellular glycerol in

Saccharomyces cerevisiae influences osmotic stress response and resistance to increased temperature. Mol Microbiol 36:

1381-1390.

Sinha H, David L, Pascon RC, Clauder-Munster S, Krishnakumar S, Nguyen M, Shi G, Dean J, Davis RW, Oefner PJ,

McCusker JH, Steinmetz LM (2008) Sequential elimination of major-effect contributors identifies additional quantitative

trait loci conditioning high-temperature growth in yeast. Genetics 180: 1661-1670.

Sprague GF, Cronan JE (1977) Isolation and characterization of Saccharomyces cerevisiae mutants defective in glycerol

catabolism. Journal of Bacteriology 129: 1335-1342.

Steinmetz LM, Scharfe C, Deutschbauer AM, Mokranjac D, Herman ZS, Jones T, Chu AM, Giaever G, Prokisch H,

Oefner PJ, Davis RW (2002) Systematic screen for human disease genes in yeast. Nat Genet 31: 400-404.

Steinmetz LM, Sinha H, Richards DR, Spiegelman JI, Oefner PJ, McCusker JH, Davis RW (2002) Dissecting the

architecture of a quantitative trait locus in yeast. Nature 416: 326-330.

Stephanopoulos G (2007) Challenges in engineering microbes for biofuels production. Science 315: 801-804.

Stephanopoulos G, Sinskey AJ (1993) Metabolic engineering methodologies and future prospects. Trends in

biotechnology 11: 392-396.

Sutherland FC, Lages F, Lucas C, Luyten K, Albertyn J, Hohmann S, Prior BA, Kilian SG (1997) Characteristics of

Fps1-dependent and -independent glycerol transport in Saccharomyces cerevisiae. J Bacteriol 179: 7790-7795.

178 References

Swinnen S, Schaerlaekens K, Pais T, Claesen J, Hubmann G, Yang Y, Demeke M, Foulquie-Moreno MR, Goovaerts A,

Souvereyns K, Clement L, Dumortier F, Thevelein JM (2012) Identification of novel causative genes determining the

complex trait of high ethanol tolerance in yeast using pooled-segregant whole-genome sequence analysis. Genome

Research.

Swinnen S, Thevelein JM, Nevoigt E (2012) Genetic mapping of quantitative phenotypic traits in Saccharomyces

cerevisiae. FEMS Yeast Res 12: 215-227.

Tamas MJ, Luyten K, Sutherland FC, Hernandez A, Albertyn J, Valadi H, Li H, Prior BA, Kilian SG, Ramos J,

Gustafsson L, Thevelein JM, Hohmann S (1999) Fps1p controls the accumulation and release of the compatible solute

glycerol in yeast osmoregulation. Mol Microbiol 31: 1087-1104.

Teige M, Scheikl E, Reiser V, Ruis H, Ammerer G (2001) Rck2, a member of the calmodulin-protein kinase family, links

protein synthesis to high osmolarity MAP kinase signaling in budding yeast. Proc Natl Acad Sci U S A 98: 5625-5630.

Thomas KC, Ingledew WM (1990) Fuel alcohol production: effects of free amino nitrogen on fermentation of very-high-

gravity wheat mashes. Appl Environ Microbiol 56: 2046-2050.

Thorsen M, Di Y, Tangemo C, Morillas M, Ahmadpour D, Van der Does C, Wagner A, Johansson E, Boman J, Posas F,

Wysocki R, Tamas MJ (2006) The MAPK Hog1p modulates Fps1p-dependent arsenite uptake and tolerance in yeast.

Mol Biol Cell 17: 4400-4410.

Treco DA, Winston F (2001) Growth and Manipulation of Yeast. In Current Protocols in Molecular Biology: John Wiley

& Sons, Inc.

Tyo KE, Nevoigt E, Stephanopoulos G Directed evolution of promoters and tandem gene arrays for customizing RNA

synthesis rates and regulation. Methods Enzymol 497: 135-155.

USA (2007) ENERGY INDEPENDENCE AND SECURITY ACT 110th Congress Public Law 110-140.

USDA (2012) Feed Grains: Yearbook tables. United States Departement of Agriculture, available at:

http://wwwersusdagov/data-products/feed-grains-database/feed-grains-yearbook-tablesaspx.

Valadi A, Granath K, Gustafsson L, Adler L (2004) Distinct intracellular localization of Gpd1p and Gpd2p, the two yeast

isoforms of NAD+-dependent glycerol-3-phosphate dehydrogenase, explains their different contributions to redox-driven

glycerol production. J Biol Chem 279: 39677-39685.

Valadi H, Larsson C, Gustafsson L (1998) Improved ethanol production by glycerol-3-phosphate dehydrogenase mutants

of Saccharomyces cerevisiae. Appl Microbiol Biotechnol 50: 434-439.

Valadi H, Valadi A, Ansell R, Gustafsson L, Adler L, Norbeck J, Blomberg A (2004) NADH-reductive stress in

Saccharomyces cerevisiae induces the expression of the minor isoform of glyceraldehyde-3-phosphate dehydrogenase

(TDH1). Curr Genet 45: 90-95.

Van Aelst L, Hohmann S, Zimmermann FK, Jans AW, Thevelein JM (1991) A yeast homologue of the bovine lens fibre

MIP gene family complements the growth defect of a Saccharomyces cerevisiae mutant on fermentable sugars but not its

defect in glucose-induced RAS-mediated cAMP signalling. EMBO J 10: 2095-2104.

van Dijken JP, Bauer J, Brambilla L, Duboc P, Francois JM, Gancedo C, Giuseppin ML, Heijnen JJ, Hoare M, Lange

HC, Madden EA, Niederberger P, Nielsen J, Parrou JL, Petit T, Porro D, Reuss M, van Riel N, Rizzi M, Steensma HY,

Verrips CT, Vindelov J, Pronk JT (2000) An interlaboratory comparison of physiological and genetic properties of four

Saccharomyces cerevisiae strains. Enzyme Microb Technol 26: 706-714.

van Dijken JP, Scheffers WA (1986) Redox balances in the metabolism of sugars by yeasts. FEMS Microbiology Letters

32: 199-224.

Van Dyk JS, Pletschke BI (2012) A review of lignocellulose bioconversion using enzymatic hydrolysis and synergistic

cooperation between enzymes—Factors affecting enzymes, conversion and synergy. Biotechnology Advances 30: 1458-

1480.

van Hoek P, de Hulster E, van Dijken JP, Pronk JT (2000) Fermentative capacity in high-cell-density fed-batch cultures

of baker's yeast. Biotechnol Bioeng 68: 517-523.

References 179

van Hoek P, van Dijken JP, Pronk JT (2000) Regulation of fermentative capacity and levels of glycolytic enzymes in

chemostat cultures of Saccharomyces cerevisiae. Enzyme Microb Technol 26: 724-736.

Verduyn C, Postma E, Scheffers WA, Van Dijken JP (1992) Effect of benzoic acid on metabolic fluxes in yeasts: a

continuous-culture study on the regulation of respiration and alcoholic fermentation. Yeast 8: 501-517.

Vilela-Moura A, Schuller D, Mendes-Faia A, Silva RD, Chaves SR, Sousa MJ, Corte-Real M (2011) The impact of

acetate metabolism on yeast fermentative performance and wine quality: reduction of volatile acidity of grape musts and

wines. Appl Microbiol Biotechnol 89: 271-280.

Wang DI, Cooney C. L., Demain A. L., Dunnill P., Humphrey A., M. L (1979) Fermentation and enzyme technology.

New York.

Wang XD, Zhang XE, Guo YC, Zhang ZP, Cao ZA, Zhou YF (2009) Characterization of glycerol dehydratase expressed

by fusing its alpha- and beta-subunits. Biotechnol Lett 31: 711-717.

Wang ZX, Zhuge J, Fang H, Prior BA (2001) Glycerol production by microbial fermentation: a review. Biotechnol Adv

19: 201-223.

Warburg O (1932) Das sauerstoffübertragende Ferment der Atmung. Angewandte Chemie 45: 1-6.

Warmka J, Hanneman J, Lee J, Amin D, Ota I (2001) Ptc1, a type 2C Ser/Thr phosphatase, inactivates the HOG pathway

by dephosphorylating the mitogen-activated protein kinase Hog1. Mol Cell Biol 21: 51-60.

Westfall PJ, Ballon DR, Thorner J (2004) When the stress of your environment makes you go HOG wild. Science 306:

1511-1512.

Westfall PJ, Patterson JC, Chen RE, Thorner J (2008) Stress resistance and signal fidelity independent of nuclear MAPK

function. Proc Natl Acad Sci U S A 105: 12212-12217.

Williamson DH, Lund P, Krebs HA (1967) The redox state of free nicotinamide-adenine dinucleotide in the cytoplasm

and mitochondria of rat liver. Biochem J 103: 514-527.

Winzeler EA, Richards DR, Conway AR, Goldstein AL, Kalman S, McCullough MJ, McCusker JH, Stevens DA,

Wodicka L, Lockhart DJ, Davis RW (1998) Direct allelic variation scanning of the yeast genome. Science 281: 1194-

1197.

Winzeler EA, Richards DR, Conway AR, Goldstein AL, Kalman S, McCullough MJ, McCusker JH, Stevens DA,

Wodicka L, Lockhart DJ, Davis RW (1998) Direct allelic variation scanning of the yeast genome. Science 281: 1194-

1197.

Winzeler EA, Shoemaker DD, Astromoff A, Liang H, Anderson K, Andre B, Bangham R, Benito R, Boeke JD, Bussey

H, Chu AM, Connelly C, Davis K, Dietrich F, Dow SW, El Bakkoury M, Foury F, Friend SH, Gentalen E, Giaever G,

Hegemann JH, Jones T, Laub M, Liao H, Liebundguth N, Lockhart DJ, Lucau-Danila A, Lussier M, M'Rabet N, Menard

P, Mittmann M, Pai C, Rebischung C, Revuelta JL, Riles L, Roberts CJ, Ross-MacDonald P, Scherens B, Snyder M,

Sookhai-Mahadeo S, Storms RK, Veronneau S, Voet M, Volckaert G, Ward TR, Wysocki R, Yen GS, Yu K,

Zimmermann K, Philippsen P, Johnston M, Davis RW (1999) Functional characterization of the S. cerevisiae genome by

gene deletion and parallel analysis. Science 285: 901-906.

Wojda I, Alonso-Monge R, Bebelman JP, Mager WH, Siderius M (2003) Response to high osmotic conditions and

elevated temperature in Saccharomyces cerevisiae is controlled by intracellular glycerol and involves coordinate activity

of MAP kinase pathways. Microbiology 149: 1193-1204.

Wurgler-Murphy SM, Maeda T, Witten EA, Saito H (1997) Regulation of the Saccharomyces cerevisiae HOG1 mitogen-

activated protein kinase by the PTP2 and PTP3 protein tyrosine phosphatases. Mol Cell Biol 17: 1289-1297.

Yadav VG, De Mey M, Lim CG, Ajikumar PK, Stephanopoulos G (2012) The future of metabolic engineering and

synthetic biology: towards a systematic practice. Metab Eng 14: 233-241.

Yale J, Bohnert HJ (2001) Transcript expression in Saccharomyces cerevisiae at high salinity. J Biol Chem 276: 15996-

16007.

Ye GM, Chen C, Huang S, Han DD, Guo JH, Wan B, Yu L (2005) Cloning and characterization a novel human 1-acyl-

sn-glycerol-3-phosphate acyltransferase gene AGPAT7. DNA Seq 16: 386-390.

180 References

Zarrinpar A, Bhattacharyya RP, Nittler MP, Lim WA (2004) Sho1 and Pbs2 act as coscaffolds linking components in the

yeast high osmolarity MAP kinase pathway. Mol Cell 14: 825-832.

Zhang A, Kong Q, Cao L, Chen X (2007) Effect of FPS1 deletion on the fermentation properties of Saccharomyces

cerevisiae. Lett Appl Microbiol 44: 212-217.

Zhang L, Tang Y, Guo ZP, Ding ZY, Shi GY (2011) Improving the ethanol yield by reducing glycerol formation using

cofactor regulation in Saccharomyces cerevisiae. Biotechnol Lett 33: 1375-1380.

Zhao S, Douglas NW, Heine MJ, Williams GM, Winther-Larsen HC, Meaden PG (1994) The STL1 gene of

Saccharomyces cerevisiae is predicted to encode a sugar transporter-like protein. Gene 146: 215-219.

Zwietering MH, Jongenburger I, Rombouts FM, van 't Riet K (1990) Modeling of the bacterial growth curve. Appl

Environ Microbiol 56: 1875-1881.