studies on the involvement of the yeast abc transporter pdr18 in … · yeast abc transporter pdr18...

57
Studies on the involvement of the yeast ABC transporter Pdr18 in acetic acid tolerance at temperatures in the range 30-40°C Catarina Alexandra Simões Prata Thesis to obtain the Master of Science Degree in Biotechnology Supervisor: Prof. Dr. Isabel Maria de Sá Correia Leite de Almeida Co-supervisor: Dr. Margarida Isabel Rosa Bento Palma Examination Committee Chairperson: Prof. Dr. Arsénio do Carmo Sales Mendes Fialho Supervisor: Prof. Dr. Isabel Maria de Sá-Correia Leite de Almeida Member of the Committee: Prof. Dr. Miguel Nobre Parreira Cacho Teixeira November 2014

Upload: hatu

Post on 08-Nov-2018

213 views

Category:

Documents


0 download

TRANSCRIPT

Studies on the involvement of the yeast ABC transporter Pdr18 in acetic acid tolerance

at temperatures in the range 30-40°C

Catarina Alexandra Simões Prata

Thesis to obtain the Master of Science Degree in

Biotechnology

Supervisor: Prof. Dr. Isabel Maria de Sá Correia Leite de Almeida

Co-supervisor: Dr. Margarida Isabel Rosa Bento Palma

Examination Committee

Chairperson: Prof. Dr. Arsénio do Carmo Sales Mendes Fialho

Supervisor: Prof. Dr. Isabel Maria de Sá-Correia Leite de Almeida

Member of the Committee: Prof. Dr. Miguel Nobre Parreira Cacho Teixeira

November 2014

i

Acknowledgments

Firstly, I would like to acknowledge Professor Isabel Sá-Correia, as my supervisor and head of

Biological Sciences Research Group (BSRG) of the Institute for Biotechnology and Bioengineering

(IBB), Instituto Superior Técnico (IST), for giving me the opportunity to join BSRG, where this work

was developed. I am most thankful for her guidance, as well as for her prompt availability and

willingness to support me during the development of this thesis.

I would like to thank Dr. Margarida Palma for all the help throughout this work and for being always

available for me. Thank you for sharing your knowledge and experience with me.

This work was financially supported by “Fundação para a Ciência e a Tecnologia” (FCT): contract

ERA-IB/0002/2010 (INTACT – Integral Engineering of Acetic Acid Tolerance in Yeast), in the frame of

ERA-NET Industrial Biotechnology.

The following personal acknowledgements will be addressed in Portuguese:

Gostaria de agradecer a todos os membros do BSRG pelo excelente companheirismo e pelo

ambiente saudável e de entreajuda propício ao desenvolvimento de trabalhos de qualidade, o que

sem dúvida contribuiu para a realização deste trabalho. Um agradecimento muito especial à

Margarida Palma por todo o seu apoio e amizade e, principalmente, pela partilha de conhecimentos e

paciência. Não podia deixar de agradecer à Cláudia Godinho, pela sua disponibilidade e ajuda na

execução deste trabalho.

Quero também agradecer aos meus colegas de mestrado e de laboratório: Diana, José, Nicole, Pedro

e Sara; pela companhia e partilha dos bons e maus momentos.

Um agradecimento especial ao Fábio, pelo seu apoio e compreensão, que me ajudou a ultrapassar

as dificuldades que surgiram ao longo deste percurso. Por último, mas não menos importante,

agradeço aos meus pais, avó e restante família e amigos, por toda a força e ajuda incondicional que

foram preponderantes para a conclusão deste trabalho. Muito obrigada por tudo.

Catarina Prata

ii

Abstract

The occurrence of multidrug resistance (MDR) frequently relies on the action of membrane

transporters presumably involved in the active efflux of unrelated cytotoxic compounds out of the cell.

Although this is a major concern in the treatment of cancer and infection diseases, the biological

activity of multidrug transporters has also advantages, such as the improvement of the ability of

industrial strains to tolerate multiple stresses of biotechnological relevance.

Pdr18 is a plasma membrane MDR transporter of the ATP-binding cassette (ABC) superfamily

described as playing a role in ergosterol incorporation in Saccharomyces cerevisiae plasma

membrane. In this thesis, the physiological role of Pdr18 in yeast tolerance to acetic acid was

examined. PDR18 expression was found to increase yeast tolerance to acetic acid, to contribute to

counteract the dissipation of plasma membrane potential and the permeabilization of plasma

membrane in yeast cells growing in the presence of inhibitory concentrations of this weak acid.

The effect of PDR18 expression in the growth curve of yeast cells in growth medium supplemented

with increasing concentrations of acetic acid at temperatures in the range 30-40°C was also examined.

Results suggest that: i) the presence of acetic acid leads to the decrease of the optimum and

maximum temperature for growth; ii) the minimum inhibitory concentrations (MIC) for acetic acid

decreases as temperatures increases; and iii) PDR18 expression leads to the increase of yeast

robustness to the conjugated effect of these environmental stresses and that Pdr18 exerts a protective

effect against the additive deleterious action of acetic acid and supraoptimal temperatures.

Keywords: Saccharomyces cerevisiae; multidrug resistance; drug efflux pumps; ATP-binding cassette (ABC); Pdr18; ergosterol; yeast response to stress; supraoptimal temperatures.

iii

Resumo

O mecanismo de aquisição de resistência a múltiplas drogas é frequentemente dependente da ação

de transportadores membranares presumivelmente envolvidos no efluxo de diferentes compostos

citotóxicos. Deste fenómeno biológico resultam problemas graves no tratamento de cancro e doenças

infecciosas, sendo o seu estudo considerado de elevado interesse. Contudo, a atividade de

transportadores de resistência a múltiplas drogas pode também representar vantagens, tal como a

capacidade de estirpes industriais desenvolverem tolerância a diversos stresses químicos de

interesse biotecnológico.

Pdr18 é um transportador de resistência a múltiplas drogas que pertence à superfamília ATP-binding

cassette (ABC). O seu papel foi descrito na incorporação de ergosterol na membrana plasmática da

levedura Saccharomyces cerevisiae. Neste trabalho foi estudado o papel fisiológico do Pdr18 na

tolerância da levedura ao ácido acético. A expressão do gene PDR18 demonstrou ser essencial para

a maior tolerância da levedura a ácido acético, contribuindo para contrariar o efeito do ácido acético

na dissipação do potencial de membrana e no aumento da permeabilidade da membrana plasmática

em células de levedura crescidas na presença de concentrações inibitórias de ácido acético.

O efeito da expressão do gene PDR18 foi estudado no crescimento de células de levedura a

temperaturas entre 30°C e 40°C, na presença de concentrações crescentes de ácido acético. Os

resultados obtidos sugerem que: i) a presença de ácido acético leva à diminuição das temperaturas

ótima e máxima de crescimento; ii) a concentração mínima inibitória de ácido acético diminui com o

aumento da temperatura; e iii) a expressão do gene PDR18 resulta no aumento da robustez das

estirpes em relação ao efeito conjugado de ambos os stresses, e que o Pdr18 exerce um efeito

protetor contra a ação nefasta aditiva do ácido acético e de temperaturas supra-ótimas.

Palavras-chave: Saccharomyces cerevisiae; resistência a múltiplas drogas; bombas de efluxo de drogas; ATP-binding cassette (ABC); Pdr18; ergosterol; resposta da levedura a stress; temperaturas supra-ótimas.

iv

Contents

Acknowledgments .................................................................................................................. i

Abstract .................................................................................................................................. ii

Resumo ................................................................................................................................. iii

List of Figures ........................................................................................................................ vi

List of Abbreviations ............................................................................................................ viii

1. Motivation and thesis outline .......................................................................................... 1

2. Introduction .................................................................................................................... 3

2.1. The multidrug resistance (MDR) phenomenon .........................................................3

2.1.1. Multidrug resistance transporters in yeast .........................................................5

2.1.1.1. The ATP-binding cassette (ABC) transporters ............................................5

2.1.1.2. The Pdr18 multidrug resistance ABC transporter .......................................7

2.2. Effects of acetic acid stress in Saccharomyces cerevisiae .......................................8

2.2.1. Acetic acid as a stress agent .............................................................................8

2.2.2. Mechanisms of acetic acid tolerance .................................................................9

2.3. Stress associated with growth at supraoptimal temperatures .................................12

2.3.1. Cellular mechanisms associated with thermotolerance in yeast ......................13

3. Materials and methods ................................................................................................. 15

3.1. Strains, plasmids and growth conditions .................................................................15

v

3.2. Acetic acid susceptibility .........................................................................................15

3.3. Assessment of plasma membrane potential ...........................................................16

3.3.1. [14C]-Methylamine uptake experiments ............................................................16

3.3.2. DiOC6(3) (3-3’-dihexyloxacarbocianine iodide) accumulation assay ...............16

3.4. Extracellular acidification curves promoted by yeast cells cultivated in the presence or absence of acetic acid ..................................................................................................17

4. Results ......................................................................................................................... 18

4.1. Role of the ABC transporter Pdr18 in yeast tolerance to acetic acid at temperatures in the range 30-40ºC .........................................................................................................18

4.1.1. The lack of PDR18 expression leads to increased yeast susceptibility to acetic acid .....................................................................................................................18

4.1.2. Inhibitory effect of acetic acid in yeast growth at temperatures in the range 30-40ºC ........................................................................................................................19

4.2. PDR18 expression is essential to counteract the decrease of plasma membrane potential in yeast cells cultivated with inhibitory concentrations of acetic acid ...................22

4.3. Effect of acetic acid and PDR18 expression on the active efflux of H+ ....................26

5. Discussion .................................................................................................................... 28

References .......................................................................................................................... 31

Annexes .............................................................................................................................. 41

6. Annex I ......................................................................................................................... 42

7. Annex II ........................................................................................................................ 46

8. Annex III ....................................................................................................................... 48

vi

List of Figures

Figure 2.1 - Schematic representation of MDR transporters in yeast. .....................................4

Figure 2.2 - Mechanistic model for the adaptive yeast response to weak acid-induced stress. .............................................................................................................................................10

Figure 4.1 - PDR18 expression leads to increased yeast tolerance to acetic acid. ...............18

Figure 4.2 – Effect of PDR18 expression in yeast acetic acid tolerance................................19

Figure 4.3 – S. cerevisiae BY4741 and the derived deletion mutant Δpdr18 growth at

supraoptimal temperatures. ..................................................................................................20

Figure 4.4 - Effect of acetic acid stress in the growth of S. cerevisiae BY4741 parental strain and derived deletion mutant Δpdr18 at 30°C. .......................................................................20

Figure 4.5 - Conjugated effect of acetic acid and supraoptimal temperatures in the growth of S. cerevisiae BY4741 parental strain and derived deletion mutant Δpdr18. ..........................21

Figure 4.6 – Minimum concentrations of acetic acid required to inhibit the growth of S. cerevisiae BY4741 parental strain and derived deletion mutant Δpdr18 at four different

temperatures (30, 36, 38 and 40°C). ....................................................................................22

Figure 4.7 - PDR18 gene expression is essential to counteract the dissipation of plasma membrane potential in yeast cells cultivated in the absence or presence of inhibitory concentrations of acetic acid (60 mM, pH 4.0). .....................................................................24

Figure 4.8 – Distribution of cell population membrane potential, obtained from fluorescence microscopy using DiOC6(3) of the parental strain BY4741 (A1) and derived deletion mutant strain BY4741_Δpdr18 (A2) harvested from a mid-exponential culture growth performed in

absence (white bars) or presence (black bars) of acetic acid (60 mM)..................................25

Figure 4.9 – Effect of PDR18 gene expression in the external acidification rate in cells cultivated under control conditions or in the presence of inhibitory concentrations of acetic acid. .....................................................................................................................................27

Figure A.1 - Culture OD600nm of Saccharomyces cerevisiae parental strain BY4741 and derived deletion mutant Δpdr18 grown at 30°C in the absence and in the presence of acetic acid.. ....................................................................................................................................42

vii

Figure A.2 - Culture OD600nm of Saccharomyces cerevisiae parental strain BY4741 and derived deletion mutant Δpdr18 grown at 36°C in the absence and in the presence of acetic

acid. .....................................................................................................................................43

Figure A.3 - Culture OD600nm of Saccharomyces cerevisiae parental strain BY4741 and derived deletion mutant Δpdr18 grown at 38°C in the absence and in the presence of acetic

acid. .....................................................................................................................................44

Figure A.4 - Culture OD600nm of Saccharomyces cerevisiae parental strain BY4741 and derived deletion mutant Δpdr18 grown at 40°C in the absence and in the presence of acetic

acid. .....................................................................................................................................45

Figure B.1 - Culture OD600nm of Saccharomyces cerevisiae parental strain BY4741 harbouring a PDR18 expression plasmid or the corresponding cloning vector pRS416 in the

absence and in the presence of acetic acid. .........................................................................46

Figure B.2 - Culture OD600nm of Saccharomyces cerevisiae derived deletion mutant Δpdr18 harbouring a PDR18 expression plasmid or the corresponding cloning vector pRS416 in the absence and in the presence of acetic acid. .........................................................................47

Figure C.1 - Conjugated effect of acetic acid and temperature in the growth of Saccharomyces cerevisiae parental strain BY4741 and derived deletion mutant Δpdr18. ....48

viii

List of Abbreviations

2,4-D 2,4-dichlorophenoxyacetic acid

ABC ATP Binding Cassette

ATP Adenosine triphosphate

EUROSCARF European Saccharomyces cerevisiae archive for functional analysis

HMF Hydroxymethyl furfural

HSR Heat shock response

MDR Multidrug resistance

MFS Major Facilitator Superfamily

MM Minimal growth medium

NBD Nucleotide Binding Domain

OD600nm Optical density at 600nm

PDR Pleiotropic Drug Resistance

PDRE Pdr-responsive elements

P-gp Mammalian Glicoprotein

pHext Extracellular pH

pHi Intracellular pH

pKa -log(Ka), where Ka is the acid dissociation constant

Rpm Rotations per minute

S. cerevisiae Saccharomyces cerevisiae

SSF Simultaneous saccharification and fermentation

TMD Transmembrane Domain

TMS Transmembrane-spanning Segments

YRRE Yrr-responsive element

1

1. Motivation and thesis outline

Bio-ethanol produced by yeast cells from renewable substrates is an ecological and renewable liquid

fuel whose production has increased during the last years and is expected to continue to grow, as an

alternative to fossil fuels (Hahn-Hagerdal et al., 2006). Lignocellulosic biomass is widely available,

relatively inexpensive, and can be non-competitive with food products, leading to a cost-effective,

reproducible and sustainable large-scale production of bio-ethanol (Greetham et al., 2014, Pereira et

al., 2011). In the preparation of lignocellulosic hydrolysates for bio-ethanol production, a pretreatment

step is needed to release all the sugar monomers from the lignocellulosic biomass and make them

available for fermentation (Laluce et al., 2012, Wimalasena et al., 2014), which leads to the

appearance of sugar degradation products, such as weak acids, furan derivatives and phenolic

compounds (Galbe and Zacchi, 2002). These products are inhibitors that affect both cell growth and

ethanol biosynthesis (Lu et al., 2012, Palmqvist et al., 1999). Acetic acid is the most common of the

many weak organic acids that occur in lignocellulose hydrolysates (Ingram et al., 1999), and is formed

when the acetyl groups present in lignin and linked to the hemicelluloses chains are released,

depending on the pre-treatment type and amount of hemicelluloses present in the raw material

(Ingram et al., 1999, Palmqvist et al., 1999). Furthermore, for the production of ethanol from

lignocelluloses, a process economically advantageous is the simultaneous saccharification and

fermentation (SSF) (Lu et al., 2012). This process is less affected by product inhibitors due to the use

of enzymatic hydrolysis; but has also disadvantages, due to the optimal temperature for cellulase

activity (45 - 50°C), which is higher compared to the yeast optimal temperature for fermentation

(~30°C) (Dogan et al., 2014), affecting yeast viability and growth (Eklund et al., 1997, Olofsson et al.,

2008, Pereira et al., 2011).

In order to be economically feasible, bio-ethanol production requires yeast strains with specific

desirable traits, namely tolerance to high temperatures, acetic acid and others inhibitors present in

hydrolysates of lignocellulosic materials, and the efficient production of ethanol with high production

yields (Benjaphokee et al., 2012). Multidrug resistance (MDR) is a biological phenomenon in which

organisms become less sensitive to multiple structurally and functionally unrelated cytotoxic

compounds (Balzi and Goffeau, 1994). Even though MDR represents an alarming clinical problem in

the treatment of cancer and infectious diseases (Balzi and Goffeau, 1994), the acquisition of tolerance

to multiple chemical stresses also leads to the improvement of strain robustness for industrial

processes (Sa-Correia et al., 2009). Pdr18 is a MDR plasma membrane transporter of the ATP-

binding cassette (ABC) superfamily and its physiological role in mediating ergosterol incorporation in

the yeast plasma membrane was previously proposed by Biological Sciences Research Group

(BSRG) (Cabrito et al., 2011). Pdr18 was described as a determinant of resistance to an extensive

variety of chemical stress agents, such as the herbicides 2,4-dichlorophenoxyacetic acid (2,4-D), the

agricultural fungicide mancozeb (Cabrito et al., 2011), and also to toxic concentrations of ethanol

(Teixeira et al., 2012), which represents an advantage in industrial bio-ethanol production processes.

In this thesis, we intend to scrutinize the involvement of the MDR plasma membrane transporter Pdr18

2

in yeast response to acetic acid stress and to another environmental challenge affecting plasma

membrane lipid organization: a range of supraoptimal temperatures. For this, growth experiments in

the presence of increasing concentrations of acetic acid at temperatures in the range 30-40°C were

performed. Given Pdr18 effect in the incorporation of ergosterol in the membrane, we were interested

in studying the physiological role of PDR18 expression in the yeast plasma membrane properties of

cells cultivated in the presence of inhibitory concentrations of acetic acid, such as membrane potential

and permeability. Overall, the results of this thesis are a contribution to the characterization of the

physiological role of Pdr18 as a MDR determinant.

3

2. Introduction

2.1. The multidrug resistance (MDR) phenomenon

Multidrug resistance (MDR) is a phenomenon in the study of cellular simultaneous resistance to a

number of unrelated citotoxic compounds (Balzi and Goffeau, 1994). The occurrence of MDR was

registered in a broad range of organisms, from bacteria to mammals (Sa-Correia et al., 2009), and the

main mechanism is the active transport of structurally and pharmacologically distinct drugs out of the

cell (Higgins, 2007). In medicine the emergence of this phenomenon in therapy has led to severe

difficulties in the treatment of tumors and several infectious diseases. In agriculture, multidrug

mechanisms have also negative outcomes in the control of resistance of plant pathogens to natural

plant defence toxins and common fungicides (Balzi and Goffeau, 1994). Nevertheless, the ability of

industrial strains to tolerate different chemical stresses represents an advantage arising from the MDR

phenomenon (Sa-Correia et al., 2009).

The most important MDR membrane transporters are members of the ATP-binding cassette (ABC)

superfamily or the major facilitator superfamily (MFS) (Figure 2.1) (Del Sorbo et al., 2000). The best

characterized are the ABC transporters, which are primary active transporters, as they are involved in

the active transport of solutes using ATP hydrolysis, as the energy source. In contrast, in MFS

transporters, transport is driven by the proton-motive force, being classified as secondary active

transporters (Pao et al., 1998, Del Sorbo et al., 2000).

The study of multidrug resistance in the yeast Saccharomyces cerevisiae became very important due

to the fact that some yeast species are involved in pathogenicity for men and other animals and in

plants (Balzi and Goffeau, 1994). Moreover, yeast is considered an universal, easy-to-manipulate

model system for the study of more complex eukaryotic cells (Sa-Correia and Tenreiro, 2002),

assisting in the understand of the mechanisms underlying their cytotoxic insults (Sa-Correia et al.,

2009).

4

Figure 2.1 - Schematic representation of MDR transporters in yeast. The simplified topologies

are as predicted by the Saccharomyces Genome Database (SGD) of the (a) multidrug

transporters of the ABC superfamily and (b) the MFS-MDR transporters from the 12-spanner DHA1

(light blue) and 14-spanner DHA2 (dark blue) families (From Sá-Correia et al., 2009).

5

2.1.1. Multidrug resistance transporters in yeast

2.1.1.1. The ATP-binding cassette (ABC) transporters

ATP-binding cassette (ABC) proteins represent one of the largest membrane transporter superfamily

(Jungwirth and Kuchler, 2006). ABC transporters structure is divided in four core domains: two

transmembrane domains (TMD) and two nucleotide binding domains (NBD). The nucleotide binding

domains are located in the cytosol and contain the ATP-binding cassette (Dean et al., 2001), as well

as a variable number of predicted transmembrane-spanning segments (TMS) (Higgins, 2001), which

determine different substrate specificities of individual ABC proteins (Jungwirth and Kuchler, 2006).

They are conserved in all living organisms and use ATP as a source of energy to translocate an

extensive variety of substrates across the plasma membrane (Del Sorbo et al., 2000). One of the roles

of these transporters is as importers, transporting nutrients and ions into the cell (Higgins, 1992).

Others are involved in the efflux of a wide variety of xenobiotic compounds, such as the human P-

glycoprotein (P-gp), which functions as an ATP-dependent efflux pump for drugs (Juranka et al., 1989,

Nourani et al., 1997) and is responsible for multidrug resistance in cancer cells (Balzi and Goffeau,

1994, Juranka et al., 1989). Nevertheless, besides ABC functions as transporters, they also might be

involved in the regulation of heterologous channels and other membrane proteins (Higgins, 1995).

In yeast, many protein transporters belonging to the ABC superfamily are localized at the plasma

membrane or of important organelles, such as Golgi, mitochondria, endoplasmic reticulum, etc, and

one considered to have a wide range of functions, like drug transport, ion homeostasis, heavy metal

detoxification and sterol homeostasis (Prasad et al., 2002). The multidrug-resistance phenotype in

yeast is also referred as pleiotropic drug resistance (PDR) (Balzi and Goffeau, 1994, Balzi and

Goffeau, 1995, Nourani et al., 1997) and frequently results from the overexpression of plasma

membrane pumps from the ABC superfamily (Wendler et al., 1997). They represent the first line of

defense and assist in yeast detoxification, essential for the drug resistance observed in the multidrug-

tolerant cells (Balzi and Goffeau, 1994, Del Sorbo et al., 2000).

MDR regulation entails a critical regulatory network of transcription factors, including the homologous

Zn(II)2Cys6 zinc cluster transcription factors Pdr1 and Pdr3, defined as the major regulators of the PDR

network (Balzi and Goffeau, 1995), and other transcription regulators, such as Pdr7, Pdr9, Yap1 and

Yrr1 (Balzi and Goffeau, 1995, Wendler et al., 1997, Zhang et al., 2001, Dexter et al., 1994). Pdr1 and

Pdr3 have a common sequence identity and drug resistance effects (Katzmann et al., 1994). They

form homo- and heterodimers, which bind to the same DNA element designated as Pdr1/Pdr3

response elements (PDRE) (Mamnun et al., 2002), and regulate the expression of SNQ2, YOR1,

PDR5 (ABC transporters genes) (Nourani et al., 1997) and YRR1 (Zhang et al., 2001). According to

Zhang et al., Yrr1 binds to Yrr1 responsive elements (YRRE), an element different than PDRE, leading

to an increased activation of Pdr1 and Pdr3, which suggests interactions between Pdr1, Pdr3, and

Yrr1 activities.

6

The main conceivable physiological function of the PDR network appeared to be detoxification, since

PDR transporters apparently export hundreds of structurally and functionally unrelated cytotoxic

compounds, and potentially toxic metabolites (Kolaczkowska and Goffeau, 1999, Jungwirth and

Kuchler, 2006). Consequently, this hypothesize active transport may result in hypertolerance to

several unrelated exogenous drugs or xenobiotics and cell protection from unwanted side effects of

endogenous toxic metabolites (Jungwirth and Kuchler, 2006, Gbelska et al., 2006).

Moreover, some MDR transporters of the ABC superfamily, regulated by the PDR network

transcription factors Pdr1/3, have been related with lipid bilayer homeostasis, lipid uptake and control

of membrane permeability (Jungwirth and Kuchler, 2006). These include the ABC transporters Pdr5,

Pdr10 and Pdr15 which may control the membrane lipid bilayer distribution in the plasma membrane

(Schuller et al., 2007, Rockwell et al., 2009). Aus1 and Pdr11 were also found to increase the

availability of sterol in the plasma membrane for esterification in the endoplasmic reticulum, thus,

contributing to the intracellular sterol transport (Li and Prinz, 2004). Also, Yor1 together with Pdr5 may

be involved in the transport of phospholipids through the plasma membrane (Decottignies et al.,

1998). All these facts indicate that PDR pathway has a role in phospholipid and sterol composition,

controlling in part the distribution of lipid components in inner and outer leaflets of the plasma

membrane (Gulshan and Moye-Rowley, 2007). Furthermore, in S. cerevisiae plasma membrane, high

levels of ergosterol are present (Eisenkolb et al., 2002, Gulshan and Moye-Rowley, 2007). The

association of sphingolipids with ergosterol is crucial to form lipid rafts, a membrane microdomain with

important functions in the stable delivery of specific proteins, such as Pma1, to the plasma membrane

and signal transduction (Bagnat et al., 2000, Dickson and Lester, 2002). These high levels of

ergosterol and sphingolipids may be intrinsically involved in the protection of the cell against

environmental stress conditions, because of their role in the reduction of the membrane permeability

(Eisenkolb et al., 2002). This membrane lipids biosynthesis may also be directly correlated with the

PDR pathway. Therefore, the PDR pathway is required not only for the regulation of the membrane

transporters activity, but also for the lipid composition of cell membranes, which exerts an important

role in the activity of membrane embedded proteins (Gulshan and Moye-Rowley, 2007).

7

2.1.1.2. The Pdr18 multidrug resistance ABC transporter

Pdr18 is a protein plasma membrane transporter of the ABC superfamily found to be related with

plasma membrane sterol incorporation (Cabrito et al., 2011). Its physiological role in plasma

membrane sterol composition and membrane potential may affect transport across cell membrane and

drug partition between the cell interior and the extracellular medium, which was proposed to contribute

to the multidrug resistance phenotype related with its increased expression (Cabrito et al., 2011).

Pdr18 was described to determine yeast resistance to chemical stress agents, including the herbicides

2,4-dichlorophenoxyacetic acid (2,4-D) and barban, the agricultural fungicide mancozeb and cadmium,

copper, manganese and zinc (Cabrito et al., 2011). Also, PDR18 expression displays a role in

decreasing ethanol-induced plasma membrane permeabilization and reducing intracellular ethanol

concentration, which results in increased yeast tolerance to toxic concentrations of ethanol (Teixeira et

al., 2012). The effect of Pdr18 in the incorporation of ergosterol in the yeast plasma membrane

(Cabrito et al., 2011) can be explored in the construction of more robust strains for improvement of

high gravity alcoholic fermentations.

8

2.2. Effects of acetic acid stress in Saccharomyces cerevisiae

2.2.1. Acetic acid as a stress agent

Acetic acid is a by-product of the alcoholic fermentation carried out by S. cerevisiae. S. cerevisiae may

use acetic acid as a carbon source by the activity of the anapleurotic glyoxylate cycle and

gluconeogenesis, which pathways are regulated by glucose repression (Gancedo, 1998). In the

presence of glucose there is a repression of these metabolic pathways, leading to the accumulation of

acetic acid in the growth medium (Casal et al., 1996). During alcoholic fermentation by yeast, acetic

acid concentration increases, and the concentrations of ethanol attained and other toxic metabolites

produced result in growth and fermentation rate inhibition (Mira et al., 2010b, Graves et al., 2006).

Acetic acid is also one of the inhibitors of the microbial fermentation of lignocellulosic hydrolysates,

non-feedstock substrates considered important alternative for a sustainable production of bioethanol

(Lu et al., 2012).

In addition, organic acids, such as acetic acid, are weak acids used in food industry as preservatives

against microbial spoilage (Fernandes et al., 2005, Ullah et al., 2012). Although these food

preservatives inhibit most of yeast and of mould species, a number of these microorganisms have

developed the ability to grow at low pH in the presence of high acid concentrations leading to food

spoilage (Stratford et al., 2013). The antimicrobial effect of carboxylic acids is mostly determined by

their chemical properties, in particular, hydrophobicity, volatility, and pKa. Acetic acid toxicity is strongly

dependent on the pH of the medium, as it is a weak monocarboxylic acid with a pKa of 4.76 (Mira et

al., 2010c) and at neutral pH, acid acetic as well as other carboxylic acids are completely dissociated.

Thus, the antimicrobial potential of acetic acid is mainly at low pH values (below pKa), where the

protonated form predominates, since in its undissociated form, acetic acid has biophysical properties

that enable it to passively diffuse into the cell through the lipid bilayer (Piper et al., 2001). The

undissociated acid, being uncharged, readily diffuses across the plasma membrane lipid bilayer and

dissociates in the higher pH environment close to neutrality of the cytosol, resulting in the

accumulation of protons and acetate counter-ion (CH3COO-) in the interior (Piper et al., 2001). This

negatively charged form will accumulate intracellularly to very high levels due to its inability to diffuse

out of the cell. Consequently, intracellular pH decreases, affecting internal pH homeostasis and many

cellular processes such as glycolytic enzymes (Pampulha and Loureiro-Dias, 1990), acetic-acid-

mediated inhibition of NADH dehydrogenase (Zhao et al., 2008) and other metabolic processes (Krebs

et al., 1983, Piper et al., 2001). In addition, the intracellular acidification may also lead to the

generation of high turgor pressure, as well as the production of free radical that will induce severe

oxidative stress (Piper et al., 2001).

9

2.2.2. Mechanisms of acetic acid tolerance

Several mechanisms of response to acetic acid in S. cerevisiae are part of the global yeast response

to weak acids in general. After the addition of a weak acid to the culture medium, yeast cells usually

exit the cell cycle and enter a long period of stasis. During a period of several hours, which depends

on the weak acid concentration and the level of growth inhibition, a range of molecular responses are

activated and functional changes take place in yeast (Mira et al., 2010c). This stress response acts in

order to counteract weak acid effects, reducing the accumulation of high, potentially toxic levels of

acids within yeast cells (Piper et al., 2001). Therefore, in order to avoid the dissipation of plasma

membrane potential induced by lipophilic weak acids, and maintain the internal pH within physiological

values (Mira et al., 2010c), protons released are pumped out of the cell by the activation of the H+-

ATPase Pma1 protein (PM-H+-ATPase) by acetic acid (Figure 2.2) (Carmelo et al., 1996) and other

weak acids (Viegas and Sa-Correia, 1991, Holyoak et al., 1996). Consistent with the importance of

this cell response, a lowered expression of the gene encoding the major H+-ATPase (PMA1) leads to

the increase of the sensitivity of cells to weak acids (Holyoak et al., 1996), confirming its role in weak

acid adaptation. In addition, PMA2 gene is homologous to the PMA1 gene and also encodes a H+-

ATPase identical to the one encoded by PMA1 gene. However, while PMA1 gene is highly expressed,

PMA2 gene is expressed at a much lower level and is considered not essential (Fernandes and Sa-

Correia, 2001).

Furthermore, the H+-ATPase in the vacuolar membrane (V-ATPase) is also necessary for intracellular

pH homeostasis under weak acid stress, also revealing an increased activity in response to numerous

weak acids (Serrano et al., 1986, Fernandes et al., 2003, Makrantoni et al., 2007, Carmelo et al.,

1996). This V-ATPase pumps protons into the lumen of the vacuole, contributing to the recovery of

more physiological cytosolic pH and counteracting the acid-induced dissipation of the transmembrane

potential across the vacuolar membrane (Desmoucelles et al., 2002, Fernandes et al., 2003,

Makrantoni et al., 2007, Mira et al., 2009). Nevertheless, the active expulsion of weak acid anions from

the cell interior is energetically expensive (Mira et al., 2010c) and for yeast adaptation to weak acids,

specifically to acetic acid, the restriction of the passive diffusion and re-entrance of the undissociated

form of this weak acid is essential. One of the mechanisms proposed to reduce the diffusion rate of

weak acids is the reduction in porosity of the cell wall structure and the reconfiguration of membrane

lipid composition (Figure 2.2), which leads to a decrease of the internal concentration of weak acids

(Mira et al., 2009).

10

Figure 2.2 - Mechanistic model for the adaptive yeast response to weak acid-induced stress.

Representation of the activity of H+-ATPases in the plasma and vacuolar membranes for the

intracellular pH (pHi) recovery and reconfiguration of the cellular envelope (From Mira et al. 2010c).

In addition, the plasma membrane ABC transporter Pdr12, which is related to multidrug resistance

(MDR), has also a role in the reduction of the intracellular pool of the weak acid counter-ions in S.

cerevisiae (Piper et al., 1998, Holyoak et al., 1996). PDR12 is strongly induced by weak acid

preservatives, such as sorbate, benzoate, and certain other moderately lipophilic carboxylate

compounds, but not by high levels of acetate and organic alcohols (Hatzixanthis et al., 2003). Its

induction is essential for the development of weak acid tolerance and growth in the presence of food

preservatives, as it functions as an efflux pump, exporting the organic acid anion from the cell (Piper et

al., 1998, Piper et al., 2001).

Another response to acetic acid in yeast is the activation of the TOR (Target-of-Rapamycin) pathway

(Almeida et al., 2009), a regulatory system of response to nutrient starvation, such as the reduced

concentration of amino acids in weak acid challenged cells. This mechanism was proposed to

counteract nutrient limitation conditions induced by weak acid stress (Teixeira et al., 2005, Almeida et

al., 2009). Weak acid-stressed yeast cells show also modifications on their carbohydrate metabolism,

by the up-regulation of several genes encoding enzymes of the Krebs cycle and of glycolysis in

response to acetic acid (Mira et al., 2010a), such as phosphofructokinase (Pfk2) and fructose 1,6-

bisphosphate aldolase (Fba1) (Almeida et al., 2009). This response is related with the severe

11

depletion of ATP observed in these yeast cells and may be caused by the inhibition of the activity of

glycolytic enzymes (Krebs et al., 1983, Holyoak et al., 1996). In addition, several genes involved in

ATP synthesis were identified as determinants of resistance to multiple weak acids, including acetic

acid (Kawahata et al., 2006, Mira et al., 2010a). This is related with the demonstrated activation of

energy-consuming defense mechanisms, such as PM-H+-ATPase, V-H

+-ATPase and drug efflux

pumps (Mira et al., 2010c).

12

2.3. Stress associated with growth at supraoptimal temperatures

Yeasts growing at supraoptimal temperatures are exposed to heat stress, which has negative effects

at the level of several cellular processes, such as the inhibition of cell division, problems in protein

homeostasis because of protein aggregation, and membrane lipids disorganization (Verghese et al.,

2012). Several studies have demonstrated a great influence of temperature on the viability of yeast

(van Uden and Madeira-Lopes, 1970). Usually, there is a reduction in the growth rate of yeast cell

population within the superoptimal temperature range, the temperatures between the optimum and the

maximum temperature for growth (van Uden and Madeira-Lopes, 1970, Van Uden and Madeira-

Lopes, 1976). S. cerevisiae is a mesophilic yeast, and when growing within its superoptimal

temperature range (from about 36°C to 44°C) (van Uden, 1984), simultaneous growth and death

occur, displaying an associative profile (van Uden and Madeira-Lopes, 1970). Nevertheless, a

dissociative profile in other yeasts was also proposed, in which temperature range of exponential

growth was separated from the temperature range of thermal death (Madeira-Lopes and van Uden,

1979). Moreover, the specific growth and death rates depend on the temperature (Simoes-Mendes et

al., 1978) and, the presence of antimicrobial agents affect negatively the temperature profiles of

growth. Regarding acetic acid, Ramos and Madeira-Lopes (1990) demonstrated that acetic acid has

negative effects in the temperature profile of growth with ethanol of S. cerevisiae. In the absence of

ethanol, increasing concentrations of acetic acid narrowed the temperature range of growth, by the

progressive increase of the minimum temperature and the reduction the maximum temperature. Acetic

acid also enhanced yeast thermal death, as well as decreased tolerance to ethanol, since the

temperature range in the presence of ethanol became shorten. Nonetheless, acetic acid did not alter

S. cerevisiae associative growth profile (Ramos and Madeira-Lopes, 1990).

Membranes are affected by lipophilic compounds and by high temperatures, resulting in the damage

of membrane proteins and increase of membrane non-specific permeability and modifications in

fluidity, due to changes in the lipid organization (Salgueiro et al., 1988, Dogan et al., 2014).

Temperature effect in plasma membrane affects yeast growth and cell viability, by decreasing

plasma membrane H+-ATPase activity (Piper, 1995) and others transport systems involved in nutrient

transport (Viegas et al., 1995, Salgueiro et al., 1988), as well as by changing the fatty acid and sterol

composition of cell membranes (Beltran et al., 2008). Heat stress has also consequences in

mitochondrial electron transport chain in S. cerevisiae, leading to oxidative stress (Morano et al.,

2012). In addition, membrane fluidity is also strongly influenced by the ratio, composition, and

structure of sterols that are found in membranes (Caspeta et al., 2014) and changes in fluidity may

also have as consequences, alterations in ion transport (Verghese et al., 2012). Indeed, (Cyert, 2003)

demonstrated that cells lacking calcineurin activity are highly stress sensitive, which might be related

with the fact that the Ca2+

regulated protein phosphatase calcineurin function is the upregulation of

genes involved in several mechanisms, in response to stress, such as the biosynthesis of membrane

lipids and ergosterol (Verghese et al., 2012).

13

Despite the detrimental effects associated with supraoptimal temperature in yeasts, fermentation

temperatures higher than 35°C have several advantages, namely the reduction in the fermentation

costs because of the reducing cooling costs between pretreatment and fermentation, the reduction of

the risk of contamination input and lower energy required for ethanol separation (easy recovery of

ethanol) (Rajoka et al., 2005). Specifically in SSF, temperature for enzymatic hydrolysis in

saccharification (45-50°C) is usually higher than the optimal temperature of fermentation processes

(around 30°C) (Eklund, 1995, Olofsson et al., 2008, Dogan et al., 2014), which represents a

disadvantage of this process. Nevertheless, the use of yeast strains with higher thermotolerance,

capable of growth and of remaining active during fermentation at higher temperatures, similar to

optimal cellulase and hemicellulase activities (Lu et al., 2012, Pereira et al., 2011), could be an

approach to increase ethanol yield. Therefore, thermotolerance is undoubtedly an important issue for

these fermentation processes.

2.3.1. Cellular mechanisms associated with thermotolerance in yeast

At temperatures above 36-37°C, yeast cells activate a protective transcriptional program, the heat

shock response (HSR). Under these conditions, heat shock proteins (HSP) are expressed resulting in

the modification of other components of cells physiology, including membrane composition and

carbohydrate flux (Morano et al., 2012), in order to increase yeast thermotolerance (Yamamoto et al.,

2008). Several Hsp proteins work as molecular chaperones in the synthesis, folding, trafficking,

maturation and degradation of proteins (Mayer and Bukau, 1998, Bukau and Horwich, 1998).

In the HSR, the transcription of HSP genes with protective functions in heat-challenged yeast cells is

regulated by Hsf1, the primary regulator of HSR that is important for the recovery from a brief

exposure to extreme temperature, and the transcription factors Msn2 and Msn4, essential for long-

term survival at high temperatures (Morano et al., 2012, Yamamoto et al., 2008). Hsf1, Msn2 and

Msn4 are positive regulators of Hsp104, an antistress chaperone with a biological role in cellular

protein homeostasis (Yamamoto et al., 2008). This chaperone is essential for thermotolerance and cell

survival, since the expression of Hsp104, as well as other Hsp is required for the refolding of heat-

denatured proteins in cells recovering from severe heat shock (Parsell et al., 1994, Yamamoto et al.,

2008). Hence, thermotolerance requires an accurate equilibrium in the amounts and activities of

several cellular components before and after severe heat shock (Yamamoto et al., 2008), which is

achieved by changes in gene expression co-regulated by Hsf1, Msn2 and Msn4 (Morano et al., 2012).

Although heat shock proteins have received the major attention in the subject of thermotolerance,

there are alternative pathways to promote cell survival under temperature stress that are independent

of heat-shock protein synthesis (De Virgilio et al., 1994). Disaccharide trehalose capacities to stabilize

proteins and repress aggregation of misfolded proteins provide evidences for its role in

thermotolerance, for the maintenance of the structural integrity of cell membrane (De Virgilio et al.,

14

1994, Verghese et al., 2012). In addition, since Hsp104 chaperone role is similar to trehalose in high

temperature, both Hsp104 and trehalose may have complementary but not overlapping functions in

thermotolerance (Elliott et al., 1996).

Membrane responses to changes in temperature are also strictly dependent on the degree of

unsaturation of fatty acids of the membrane lipids (Carratu et al., 1996). Indeed, membrane fatty-acyl

composition in the yeast membrane changes with temperature: lower temperatures result in a more

unsaturated fatty-acyl composition of the membrane (Beltran et al., 2008). Moreover, sphingolipids

identified as long-chain bases (LCBs) are lipid mediators and manage intracellular protein aggregation

and translation in heat response to protect cells (Verghese et al., 2012). In fact, a strain lacking the

capacity to synthesize LCBs is heat shock sensitive, and this phenotype may be reversed by genetic

or chemical supplementation with sphingolipids, confirming these molecules role in thermotolerance

(Jenkins et al., 1997, Verghese et al., 2012).

For the maintenance of membrane fluidity and functionality, which are negatively affected at

supraoptimal temperatures (Verghese et al., 2012), yeast cells undergo changes in membrane lipid

composition (Suutari et al., 1990). Ergosterol is the most abundant sterol in the yeast plasma

membrane (Zinser et al., 1993) and studies on sterol auxotrophs demonstrated that ergosterol exerts

an effect on the biosynthesis of the different phospholipid species (Parks and Casey, 1995). Indeed,

ergosterol strongly determines the dynamic characteristics of the lipids and proteins in the yeast

membranes (Beney and Gervais, 2001). In addition, the degree of lipid saturation and the presence of

ergosterol as modulators of membrane fluidity in the S. cerevisiae plasma membrane were proposed

as a determinant factor to stress tolerance, independently of heat shock proteins and trehalose (Swan

and Watson, 1998). Thus, cells survival is strongly related with the role of ergosterol as a stabilizer of

yeast plasma membrane (Swan and Watson, 1998), affecting membrane rigidity, fluidity, and

permeability (Parks and Casey, 1995).

15

3. Materials and methods

3.1. Strains, plasmids and growth conditions

Saccharomyces cerevisiae parental strain BY4741 (MATa, his3Δ1, leu2Δ0, met15Δ0, ura3Δ0) and the

derived deletion mutant strain BY4741_Δpdr18 used in this work were obtained from the

EUROSCARF collection. The plasmid pRS416_PDR18, expressing the PDR18 gene from its natural

promoter, and the corresponding cloning vector, pRS416, were both obtained from the EUROSCARF

collection and were used for phenotypic complementation tests.

Cells were batch-cultured at 30°C with orbital agitation (250 rpm) in liquid minimal growth medium

(MM4) containing 1.7 g/L yeast nitrogen base without amino acids (Difco), 20 g/L glucose (Merck),

2.65 g/L (NH4)2SO4 (Panreac AppliChem), 20 mg/l L-methionine (Merck), 20 mg/L L-histidine (Merck),

60 mg/L L-leucine (Sigma) and 20 mg/L L-uracil (Sigma), adjusted to pH 4.0 with HCl. Cells

harbouring pRS416 or derived plasmids were grown in the same medium but without uracil (MM4-U

medium) and were preserved in MM4-U. Yeast cells were supplemented with 30% glycerol, and

stored at −80°C.

3.2. Acetic acid susceptibility

The susceptibility of the parental strain BY4741 and the derived deletion mutant Δpdr18 to acetic acid

was assessed as follows. Yeast cells were cultivated until mid-exponential phase (OD600nm=0.5±0.05)

in liquid medium MM4 (BY4741 and Δpdr18), adjusted to pH 4.0, at 30°C, and then re-inoculated by

filtration (Membrane filters white, 0.2 μm, WhatmanTM

, ME24/21ST) at an OD600nm of 0.1±0.05, in

100 mL Erlenmeyer flasks containing 50 mL of fresh medium, either or not supplemented with 60 mM

of acetic acid. Growth was followed by measuring culture OD600nm during batch cultivation at 30ºC with

an orbital agitation of 250 rpm.

In order to study the simultaneous effect of acetic acid and supraoptimal temperatures in yeast cell

growth, cell suspensions cultivated until mid-exponential phase (OD600nm=0.5±0.05) in liquid medium

MM4 (BY4741 and Δpdr18) or MM4-U (BY4741 and Δpdr18 harbouring the plasmid pRS416_PDR18

and the corresponding cloning vector, pRS416), adjusted to pH 4.0, at 30°C, were used to re-inoculate

by filtration (Membrane filters white, 0.2 μm, WhatmanTM

, ME24/21ST) at an OD600nm of 0.1±0.05

series of 100 mL Erlenmeyer flasks prepared with 50 mL liquid MM4 or MM4-U supplemented or not

(control) with progressively higher concentration of acetic acid (35-120 mM). Erlenmeyer flasks were

incubated with orbital agitation (250 rpm) at four different temperatures: 30, 36, 38 and 40°C, and

growth was followed by measuring culture OD600nm. In this work, minimum inhibitory concentration

(MIC) of acetic acid value was also assessed. MIC value was defined as the lowest concentration of

acetic acid at which no growth was detected after 10 days of incubation.

16

3.3. Assessment of plasma membrane potential

To compare the plasma membrane potential of the parental strain BY4741 and the derived deletion

mutant Δpdr18, two complementary methods were used: the [14

C]-Methylamine uptake assay

(Hoeberichts et al., 2010, Teixeira et al., 2011) and the DiOC6(3) (3-3’-dihexyloxacarbocianine iodide)

accumulation assay (Madrid et al., 1998).

3.3.1. [14C]-Methylamine uptake experiments

The uptake of [14

C]-methylamine in the parental strain BY4741 and the mutant strain Δpdr18 was

estimated as described previously (Hoeberichts et al., 2010, Teixeira et al., 2011). The parental and

mutant strains were grown in liquid MM4 medium (pH 4.0) until an OD600nm=0.5±0.1, and then re-

inoculated at an OD600nm of 0.1±0.05 into fresh liquid MM4 medium (pH 4.0) in the absence or

presence of acetic acid (60 mM). Cells grown until mid-exponential phase were harvested by

centrifugation (8600g, 5 minutes, at 4°C) and washed three times with deionized water and

resuspended in deionized water. Cell suspension (0.1 ml) was added to 0.9 ml of MES glucose buffer

[10 mM MES, 0.1 mM MgCl2 and 20 g/L glucose (pH 4.0)]. After 5 minutes of incubation at 30°C, 3 µl

[14

C]-methylamine (Biotrend) was added to a final concentration of 55 mM. At the indicated times (0, 5,

10, 15 and 20 minutes), samples (200 μl) were diluted in 10 ml cold water to stop the transport

reaction, filtered through pre-wetted glass microfiber filters (25 mm; Filter-Lab, MFV3) and washed

four times with 5 ml cold water. Subsequently, filters were put into tubes containing 7 mL scintillation

liquid (Ultima Gold™, Beckman) and the radioactivity was measured in a Beckman LS 5000TD

scintillation counter. Non-specific [14

C]-methylamine adsorption to the filters and to the cells (less than

5 % of the total bound radioactivity) was assessed and taken into consideration. Relative levels of

[14

C]-methylamine are the average of at least three independent experiments and the error bars

correspond to standard deviation. In order to confirm the significance of the differences between cell

populations after 20 minutes of incubation in MES glucose (30°C), we performed a one-way analysis

of variance for the repeated independent measures. All the comparisons between BY4741 and Δpdr18

under different conditions were done using an unpaired t test. Statistical significance was established

at P < 0.05.

3.3.2. DiOC6(3) (3-3’-dihexyloxacarbocianine iodide) accumulation assay

Plasma membrane potential was estimated using fluorescence microscopy. In order to perform the

DiOC6(3) accumulation assay, cells were harvested as described above (Section 3.3.1) and

resuspended in MES/glucose buffer [10 mM MES, 0.1 mM MgCl2 and 20 g/L glucose (pH 4.0)]. Cell

suspensions were supplemented with DiOC6(3) (Molecular Probes) at a final concentration of 0.25 nM

and incubated in the dark for 30 minutes at 30°C with orbital agitation (250 rpm). After centrifugation

(8600g during 5 minutes, at 4°C), cells were washed three times and immediately observed with a

Zeiss Axioplan microscope equipped with adequate epifluorescence filters (Zeiss BP450-490 and

17

Zeiss LP520). Fluorescence emission was collected using a CCD (charge-coupled device) camera

(Cool SNAPFX, Roper Scientific Photometrics) and MetaMorph 3.5 was used for image analysis.

Fluorescence images were background-corrected and the intensity values emitted by each individual

cell were measured pixel-by-pixel in a minimum of 80 cells per experiment. Fluorescence intensity

levels given by the software were expressed as a percentage of the value obtained for BY4741 cells in

control conditions. Only living cells were analyzed and dead cells were differentiated by bright-field

analysis of cell morphology. Levels of fluorescence are means for at least three independent

experiments and error bars represent standard deviation. To confirm the significance of the

differences between cell populations we performed a one-way analysis of variance for repeated

independent measures, as explained above (section 3.3.1).

3.4. Extracellular acidification curves promoted by yeast cells cultivated in the

presence or absence of acetic acid

The parental strain BY4741 and the derived deletion mutant Δpdr18 were grown in liquid MM4

medium (pH 4.0) until an OD600nm=0.5±0.05, and then re-inoculated at an OD600nm of 0.1±0.05 into

fresh liquid MM4 medium (pH 4.0) in the absence or presence of acetic acid (60 mM). BY4741 and

Δpdr18 strains grown until mid-exponential phase were collected by filtration (Membrane filters white,

0.2 μm, WhatmanTM

, ME24/21ST), washed twice with deionized water, and incubated in a sorbitol

solution (20 g/L, Sigma-Aldrich®) for 30 minutes at 30°C, with orbital agitation of 250 rpm. After this

incubation, deenergized cells were then harvested by filtration and resuspended in distilled water, in

order to obtain a dense cell suspension (OD600nm of 20.0 ± 2.0). The acidification assays were carried

out in a water-jacketed cell of 10 mL capacity, at 30°C, containing 5 mL of the cellular suspension

described above. The pH of the resulting suspension was adjusted to 4.0±0.1, and 1 ml of 100 g/L

glucose (pH 4.0) was added (to obtain a final concentration of 20 g/L). The external medium pH was

followed during 10 minutes by potentiometry using a pH microelectrode (Metrohm 6.0204.000)

attached to a pH meter (Metrohm 605). Comparisons of the pHextracellular variation between conditions

tested were perfomed using a one-way analysis of variance for repeated independent measures.

Experiments were done in triplicate. The comparisons between BY4741 and Δpdr18 cells under

different conditions were done using an unpaired t test and the statistical significance was established

at P < 0.05.

18

4. Results

4.1. Role of the ABC transporter Pdr18 in yeast tolerance to acetic acid at

temperatures in the range 30-40ºC

4.1.1. The lack of PDR18 expression leads to increased yeast susceptibility to acetic

acid

The deletion mutant Δpdr18 strain was found to be less tolerant to acetic acid stress (60 mM, pH 4.0),

compared to the parental strain BY4741. In fact, an extended lag-phase (approximately 40 hours,

compared to 10 hours for the parental strain) was observed for Δpdr18 cells when cultivated at 30ºC in

liquid minimal medium (pH 4.0) supplemented with 60 mM acetic acid (Figure 4.1). In addition, yeast

cell population from both strains attained a reduced final biomass value when cultivated under this

stress condition, and this reduction was more significant for the Δpdr18 culture (Figure 4.1).

The higher susceptibility of the deletion mutant Δpdr18 to acetic acid was also confirmed by the

determination of the minimum inhibitory concentration (MIC) of acetic acid, which corresponds to the

concentration at which a complete inhibition of each yeast strain growth is observed (Figure 4.2).

Hence, BY4741 cells were able to grow at acetic acid concentration of 95 mM and unable to cope with

acetic acid at 100 mM. Cells lacking PDR18 expression were unable to grow in the presence of 75

mM acetic acid (Figure 4.2). Furthermore, the expression of PDR18 from a centromeric plasmid in

Δpdr18 was found to partially rescue acetic acid susceptibility phenotype exhibited by this mutant,

since the acetic acid concentration in which there was no detectable growth of the

Δpdr18+pRS416_PDR18 strain, expressing the PDR18 gene from its natural promoter, was higher

than MIC value attained by the deletion mutant Δpdr18 cells harbouring the empty plasmid pRS416

(Figure 4.2). Also, BY4741 cells harbouring the plasmid pRS416_PDR18 were found to reach a MIC

value of 110 mM, being able to grow in the presence of higher acetic acid levels than the parental

strain BY4741 (Figure 4.2).

Figure 4.1 - PDR18 expression leads to increased yeast tolerance to acetic acid. Growth curves

of S. cerevisiae parental strain BY4741 and the derived deletion mutant Δpdr18 strain at 30ºC. Cell

suspensions of BY4741 (○, ●) and Δpdr18 (□, ■), grown in MM4 liquid medium (pH 4.0) without acetic

acid supplementation were used to inoculate MM4 liquid medium in the absence (open symbols) or

presence (closed symbols) of 60 mM acetic acid. Growth curves were followed measuring culture

OD600nm and are representative of at least three independent growth experiments.

19

Figure 4.2 – Effect of PDR18 expression in yeast acetic acid tolerance. Comparison of the acetic

acid concentration values which completely inhibit growth of BY4741 and the deletion mutant Δpdr18

strains, harbouring the PDR18 expression plasmid (pRS16_PDR18) or the corresponding cloning

vector pRS416. Minimum inhibitory concentrations (MIC) were determined by lack of detectable

growth after 240 hours of incubation at 30ºC. Values are representative of three independent

experiments that gave rise to the same results. Values result from data presented in Annexes I and II.

4.1.2. Inhibitory effect of acetic acid in yeast growth at temperatures in the range 30-

40ºC

Yeast cells were cultivated at 30°C until mid-exponential phase, and then, these cells were used to

inoculate series of 100 mL Erlenmeyers flasks with 50 mL of growth liquid medium MM4

supplemented or not (control) with increasing concentrations of acetic acid and incubated at the

temperature range 30-40°C (section 3.2; Materials and Methods). Additionally, the minimum acetic

acid concentrations required to inhibit growth of S. cerevisiae BY4741 parental strain and derived

deletion mutant Δpdr18 were determined in the temperature range 30-40°C.

Results obtained demonstrate a marked reduction in the biomass concentration at the stationary

phase of growth of the cultivation of cells devoid of PDR18, specially when incubated at 40°C (Figure

4.3). Indeed, PDR18 expression was found to lead to a higher biomass concentration at the stationary

phase of growth when the yeast cells, pre-cultivated at 30°C, were incubated at temperatures 38°C

and 40°C (Figure 4.3). Regarding acetic acid tolerance at different temperatures, an increase in lag

phase duration and a reduction of the biomass concentration values at the stationary phase of growth

were observed as the growth temperature increases from 30ºC to 40ºC, for the same acetic acid

concentrations (Figure 4.4; Figure 4.5). This decrease in tolerance to acetic acid with increasing

growth temperatures up to 40°C was more pronounced in deletion mutant cells Δpdr18. In fact, lower

concentrations of acetic acid had to be tested, because cells lacking Pdr18 were not able to grow in

the presence of acetic acid 55 mM at 38°C and 40 mM at 40°C (Figure 4.5 and Figure 4.6).

20

PDR18 expression was seen to contribute to increase yeast tolerance not only to acetic acid, but also

to the conjugated effect of this stress with supraoptimal temperatures (Figure 4.4 and Figure 4.5). In

general, data obtained in this study indicate that Pdr18 exert a protective effect against the deleterious

additive action of acetic acid and supraoptimal temperatures.

Figure 4.3 – S. cerevisiae BY4741 and the derived deletion mutant Δpdr18 growth at

supraoptimal temperatures. Cell suspensions of BY4741 and Δpdr18 pre-cultivated at 30°C until

mid-exponential phase (OD600nm = 0.5±0.05) in MM4 liquid medium (pH 4.0) were used to re-inoculate

fresh MM4 liquid medium (pH 4.0), at an OD600nm = 0.1±0.05. Culture OD600nm obtained during 240

hours of incubation at 30°C (in blue ■), 36°C (in orange ■), 38°C (in light blue ■) and 40°C (in red ■)

are shown. Values result from data presented in Figures A.1 to A.4 from Annex I and are

representative of three independent growth experiments.

Figure 4.4 - Effect of acetic acid stress in the growth of S. cerevisiae BY4741 parental strain

and derived deletion mutant Δpdr18 at 30°C. Culture OD600nm obtained during 240 hours of

incubation at 30°C in liquid MM4 supplemented with increasing concentrations of acetic acid, at pH 4.0

are shown. Values result from data presented in Figures A.1 from Annex I and are representative of

three independent growth experiments.

Colors in the figure correspond to the following acetic acid concentrations:

21

Figure 4.5 - Conjugated effect of acetic acid and supraoptimal temperatures in the growth of S.

cerevisiae BY4741 parental strain and derived deletion mutant Δpdr18. Culture OD600nm obtained

during 240 hours of incubation at 36°C, 38°C and 40°C in liquid MM4 at pH 4.0, supplemented with

increasing concentrations of acetic acid, are shown. Values result from data presented in Figures A.2

to A.4 from Annex I and are representative of three independent growth experiments.

Colors in the figure correspond to the following acetic acid concentrations:

22

Figure 4.6 – Minimum concentrations of acetic acid required to inhibit the growth of S.

cerevisiae BY4741 parental strain and derived deletion mutant Δpdr18 at four different

temperatures (30, 36, 38 and 40°C). Values were determined by lack of detectable growth after 240

hours of incubation at temperatures in the range 30-40°C. Values result from data presented in Figure

A.1 to A.4 and Figure C.1, from Annexes I and III, respectively, and are representative of three

independent experiments.

4.2. PDR18 expression is essential to counteract the decrease of plasma membrane

potential in yeast cells cultivated with inhibitory concentrations of acetic acid

In order to understand the effect of Pdr18 on yeast tolerance to acetic acid, we compared the

membrane potential of S. cerevisiae BY4741 and derived deletion mutant Δpdr18 cells harvested in

mid-exponential phase of growth in the presence or absence of 60 mM acetic acid. Two methods were

used to estimate plasma membrane potential: the [14

C]-methylamine uptake assay (Figure 4.7) and

the DiOC6(3) (3-3’-dihexyloxacarbocianine iodide accumulation assay (Figure 4.8).

In the first technique, yeast plasma membrane potential was assessed using radioactively labelled

methylammonium, which is a non-metabolizable ammonium analogue and its uptake is dependent on

the plasma membrane potential (Mulet et al., 1999). The second method used is based on the

accumulation of DiOC6(3) probe, which also relies on yeast plasma membrane potential. Regarding

this microscopy technique, an increase in fluorescence denotes higher level of plasma membrane

polarization (Madrid et al., 1998). Hence, the effect of PDR18 gene deletion on plasma membrane

potential was evaluated in cells previously cultivated in liquid media supplemented or not with acetic

acid (60 mM, pH 4.0) (Figure 4.7 and Figure 4.8). In control conditions, a depolarization of the plasma

membrane of cells lacking PDR18 expression, when compared to the parental strain, was detected

using both methods (Figure 4.7 A1 and A2 and Figure 4.8 A1 and A2), as previously described

(Cabrito et al., 2011). In [14

C]-methylamine uptake assays, this decrease in membrane potential of

0.0

20.0

40.0

60.0

80.0

100.0

30 36

38 40

Aceti

c a

cid

co

ncen

trati

on

(m

M)

Temperature of growth (ºC)

BY4741_Δpdr18

BY4741

23

Δpdr18 cells was considered statistically not significant (Figure 4.7 B). However, in DiOC6(3)

accumulation assays, this decrease in membrane potential of deletion mutant Δpdr18 cells was more

pronounced and the difference from the membrane potential observed in BY4741 cells was

considered extremely significant (P < 0.0001) (Figure 4.8 B).

In yeast cells grown in the presence of inhibitory concentration of acetic acid, methylammonium

accumulation assay demonstrated that PDR18 expression lead to a significantly higher plasma

membrane potential (P < 0.05), than the observed in the same cells in the absence of stress (Figure

4.7 A1 and B). In fluorescence microscopy results, this difference in plasma membrane potential

between BY4741 cells grown in the absence and in the presence of acetic acid was seen to be

extremely significant (P < 0.0001) (Figure 4.8 A1 and B). On the contrary, in cells devoid of the PDR18

gene cultivated in the presence of inhibitory concentrations of acetic acid, there is a decrease in

membrane potential (Figure 4.7 A2 and B and Figure 4.8 A2 and B). This difference in membrane

potential between cells expressing or not the PDR18 gene, cultivated in the presence of acetic acid,

was found to be extremely significant (P < 0.0001) in DiOC6(3) accumulation assays (Figure 4.8 B)

and very significant (P < 0.01) in [14

C]-methylamine uptake assays (Figure 4.7 B). Therefore, results

obtained show that Pdr18 is responsible for an increase in membrane potential of yeast cells grown in

the presence of 60 mM acetic acid (Figures 4.7 and 4.8), in order to counteract the dissipation of the

plasma membrane potential caused by acetic acid.

Although [14

C]-methylamine uptake assays have shown differences less significant than in DiOC6(3)

accumulation assays, results obtained in both methods have a similar tendency regarding changes in

plasma membrane potential. In DiOC6(3) accumulation assays, significant differences were displayed

in the statistical tests. This is probably due to the differences in the two approaches, as in

fluorescence microscopy each cell is considered an individual experiment and it is thus, possible to

notice differences within the yeast population. At the same time, in [14

C]-methylamine uptake assay,

the population is considered as a whole and an average value is obtained for all yeast cells in the

population. Indeed, the distribution of fluorescence among a population of BY4741 and deletion

mutant Δpdr18 cells loaded with DiOC6(3) in control conditions was shown to be well described by

a Gaussian distribution (Figure 4.8 A1 and A2). Nevertheless, BY4741 cell population harvested from

a mid-exponential phase culture carried out in the presence of acetic acid, proved to be equally

distributed among the different fluorescence intensity classes, which is indicative of a highly

heterogeneous population (Figure 4.8 A1). For the deletion mutant Δpdr18 cells, a Gaussian

distribution of cell population with a translation of the average intensity to lower values of fluorescence

was registered (Figure 4.8 A2), suggesting that the lack of PDR18 leads to a decrease of membrane

potential in cells cultivated in the presence of acetic acid (Figure 4.8 A2).

24

Figure 4.7 - PDR18 gene expression is essential to counteract the dissipation of plasma

membrane potential in yeast cells cultivated in the presence of inhibitory concentrations of

acetic acid (60 mM, pH 4.0). A: Time-course accumulation of [14

C]-methylammonium during

incubation at 30ºC of S. cerevisiae parental strain BY4741 (A1) and derived deletion mutant strain

BY4741_Δpdr18 (A2) yeast cells in MES buffer with glucose (20 g/L) supplemented with the

radiolabelled methylammonium. These yeast cells were cultivated until mid-exponential phase in the

absence (open symbols) or presence (closed symbols) of 60 mM acetic acid, at 30ºC, as described in

section 3.3.1 . Relative levels of [14

C]-methylammonium were assessed during 20 minutes, as

described in section 3.3.1, and are the average of at least three independent experiments. The error

bars represent standard deviation. B: Accumulation of [14

C]-methylammonium in BY4741 and deletion

mutant Δpdr18 cells after 20 minutes, as previously described. Values are the average of at least

three independent experiments and the error bars represent standard deviation. Ns indicates P > 0.05;

a single asterisk (*) indicates P < 0.05; double asterisks (**) indicate P < 0.01.

25

Figure 4.8 – Distribution of cell population membrane potential, obtained from fluorescence

microscopy using DiOC6(3) of the parental strain BY4741 (A1) and derived deletion mutant strain

BY4741_Δpdr18 (A2) harvested from a mid-exponential culture growth performed in absence (white

bars) or presence (black bars) of acetic acid (60 mM). B: Comparison of the average membrane

potential using the fluorescent probe DiOC6(3). Values of membrane potential are set as the

percentage of the value obtained for BY4741 cells in control conditions and are means for at least

three independent experiments. Error bars represent the correspondent standard deviation. Quadruple

asterisks (****) indicate P < 0.0001.

26

4.3. Effect of acetic acid and PDR18 expression on the active efflux of H+

Yeast cells under weak acid stress conditions undergo a strong activation of plasma membrane H+-

ATPase in the plasma membrane, resulting in the efflux of the excess amounts of the protons to the

outside, counteracting the toxic effects of weak acids (Mira et al., 2010c). Extracellular acidification

depends on the plasma membrane H+-ATPase proton pumping activity and on the plasma membrane

permeability, which influences the passive diffusion of protons into yeast cells (Rosa and Sá-Correia,

1994). In order to examine the effect of PDR18 expression in the extracellular acidification of cells

grown until mid-exponential phase in liquid media supplemented or not (control conditions) with acetic

acid (60 mM, pH 4.0), the final extracellular pH and the extracellular acidification rate obtained were

compared between each strain at different experimental conditions (Figure 4.9). Results show that

yeast cells devoid of the PDR18 gene do not have an external acidification significantly different from

the one exhibithed by the parental strain BY4741, when cultivated in the absence of acetic acid

(pHextracellular ≈ 3.2 for Δpdr18 and pHextracellular ≈ 2.9 for BY4741) (Figure 4.9 A1 and A2). When yeast

cells were cultivated in the presence of an inhibitory concentration of acetic acid (60 mM, pH 4.0),

although differences regarding final extracellular pH, as well as external acidification rate, between

cells expressing or not the PDR18 gene were observed, statistical tests considered them as not

significant (Figure 4.9 B and C).

Deletion mutant Δpdr18 cells cultivated in the presence of acetic acid exhibited a significant decrease

(P < 0.001) in the acidification rate and, consequently, an increase in the final extracellular pH

(pHextracellular ≈ 3.5) (Figure 4.9 B and C), compared to the same cells grown under control conditions

(Figure 4.9 A2). Furthermore, differences in the extracellular acidification between cells devoid of

PDR18 cultivated in control conditions and in the presence of acetic acid were more significant, than

among cells expressing PDR18 gene. This is probably due to the previous results showing that yeast

cell plasma membrane permeabilization caused by acetic acid is extremely pronounced in cells

lacking PDR18 expression (Godinho et al., unpublished results). Nevertheless, a significant reduction

of the rate of extracellular acidification (P < 0.05) was observed in BY4741 cells previously cultivated

in the presence of acetic acid (Figure 4.9 A1 and C), which led to an increase in the less acidic final

pHextracellular of 3.2, rather than the 2.9 obtained when cells were harvested from control conditions

(Figure 4.9 A1).

27

Figure 4.9 – Effect of PDR18 gene expression in the external acidification rate in cells

cultivated under control conditions or in the presence of inhibitory concentrations of acetic

acid. A: The extracellular acidification of energized cells was assessed for cells of the parental strain

BY4741 (A1) and the derived deletion mutant Δpdr18 (A2) harvested from mid-exponential culture

grown in the absence (open symbols) or presence (closed symbols) of acetic acid (60 mM, pH 4.0), at

30ºC. B: Final extracellular pH after the 10 minutes following the addition of glucose. C: External

acidification rate during the 10 minutes after addition of glucose. Values are the average of three

independent experiments and error bars represent the correspondent standard deviation. Ns indicates

P > 0.05; a single asterisk (*) indicates P < 0.05; double asterisks (**) indicate P < 0.01; and quadruple

asterisks (****) indicate P < 0.0001.

28

5. Discussion

The involvement of the Saccharomyces cerevisiae multidrug resistance (MDR) ABC transporter Pdr18

in yeast tolerance to the conjugated effect of acetic acid and supraoptimal temperatures was

examined. Results indicate that PDR18 expression contributes to increased tolerance of S. cerevisiae

BY4741 strain to acetic acid, being part in the global response that allow yeast to overcome acetic

acid induced stress. Based on the relevance of PDR18 gene expression in acetic acid stress

tolerance, and considering the role of ergosterol as a stabilizer of membrane components in heat and

ethanol tolerance (Swan and Watson, 1998), the effect of temperature in yeast growth in the presence

of inhibitory concentrations of acetic acid was evaluated in the parental S. cerevisiae BY4741 strain

and in the derived deletion mutant Δpdr18. Our results showed that yeast growth of both the parental

and the derived deletion mutant Δpdr18 strains at temperatures in the range 30-40°C could be

arrested by acetic acid, although at significantly different concentrations. Furthermore, when yeast

cells were incubated at the supraoptimal temperatures 38°C and 40°C, in the presence of increasing

concentrations of acetic acid, a marked decrease in tolerance towards this weak acid in the deletion

mutant Δpdr18 was observed. Indeed, PDR18 deletion proved to lead to extended lag-phase period

and reduced values of biomass concentration at the stationary phase of growth, when compared to

the same parameters in the parental strain BY4741. These may be indicative of Pdr18 contribution to

counteract the additive deleterious effect of this weak acid and supraoptimal temperatures on the

organization and functionality of biological membranes that lead to increased permeability and

modifications in fluidity (Salgueiro et al., 1988, Verghese et al., 2012). Survival depends on the

protection of the optimal level of the plasma membrane fluidity (Turk et al., 2011). In fact, ergosterol

incorporation in the yeast plasma membrane, in which Pdr18 is described to be involved (Cabrito et

al., 2011), is a mechanism that contributes to the maintenance of adequate fluidity and physiological

functionality of plasma membrane (Beney and Gervais, 2001, Dupont et al., 2011, Klose et al., 2012).

Therefore, these results are consistent with the need of ergosterol as a stabilizer of yeast plasma

membrane (Swan and Watson, 1998), confirming Pdr18 role in the yeast protective response to these

environmental stresses.

Furthermore, a recent study hypothesizes the decrease of sterol functions at high temperature and

observes the presence of “bended” fecosterol, rather than “flat” ergosterol in the membrane of S.

cerevisiae strains that developed thermotolerance (Caspeta et al., 2014). Therefore, since membrane

fluidity is influenced by the ratio, composition, and structure of sterols that are found in membranes

(Caspeta et al., 2014), a comparison of the relative amounts of ergosterol and other sterols in yeast

plasma membrane of the BY4741 strain and the derived deletion mutant Δpdr18 both cultivated in the

absence or presence of acetic acid at supraoptimal temperatures is proposed as future work, in order

to elucidate the specific function of Pdr18 in response to these conjugated stresses.

In order to clarify the physiological role of Pdr18 in the protective response against acetic acid stress,

the effect of PDR18 expression in yeast plasma membrane potential of cells grown in the presence of

inhibitory concentrations of acetic acid was examined. To achieve more confident results, two

29

methods were used to compare the differences in plasma membrane potential between parental

BY4741 cells and deletion mutant BY4741_Δpdr18. Results from both techniques are in agreement

and demonstrated that in cells cultivated with 60 mM acetic acid, PDR18 expression lead to an

increase of yeast plasma membrane potential, counteracting the proposed effect of this weak acid in

the dissipation of plasma membrane potential (Mira et al., 2010c). The deletion of PDR18 gene leads

to a reduction of the plasma membrane potential in cells cultivated with acetic acid, which is

hypothesized to be related with the fact that in cells lacking Pdr18, acetic acid-induced stress was

found to increase membrane permeabilization (Godinho et al., unpublished results). Given that

ergosterol plays an essential role in the normal physiological function of the plasma membrane by

regulating its fluidity and the delicate balance among other membrane components such as lipids and

proteins (Aguilera et al., 2006), membrane depolarization and increased permeabilization observed in

Δpdr18 cells may be related with low levels of ergosterol in the plasma membrane, corroborating

Pdr18 contribution against acetic acid stress in yeast. In addition, since cells used in these

experiments were previously cultivated in the presence of acetic acid (60 mM) and membrane

potential measurements were performed in the absence of stress and in the presence of glucose,

these results indicate that when acetic acid is present in the growth medium, a proper modification of

plasma membrane composition might have occurred in cells expressing PDR18, resulting in a less

harmful plasma membrane permeabilization. Thus, these results suggest that the proposed

modifications in plasma membrane lipid configuration and cell wall structure of cells in response to

weak acids (Mira et al., 2010c, Viegas et al., 2005, Simoes et al., 2006) may also involve PDR18

expression. Therefore, we hypothesize that Pdr18 involvement in ergosterol incorporation in the

membrane may contribute to the remodeling of yeast plasma membrane, leading to the reduction of

the permeabilizing effect of acetic acid and, consequently, preventing its harmful effects, as a

protective response to acetic acid stress.

Additionally, results obtained in plasma membrane potential assays using fluorescence microscopy

demonstrated that BY4741 and derived deletion mutant Δpdr18 strains cultivated in the absence of

acetic acid, displayed a Gaussian distribution of cell population. However, BY4741 cell population

presented an equally distribution among the different fluorescence intensity classes, when harvested

from a mid-exponential phase culture carried out in the presence of acetic acid. This distribution

showed that some cells displayed higher levels of fluorescence, and thus increased plasma

membrane potential; indicative of the proposed plasma membrane remodeling in cells expressing

PDR18, and decreased membrane permeabilization (Godinho et al., unpublished results). Cells

lacking Pdr18 cultivated in the presence of acetic acid proved to have a Gaussian distribution of cell

population, as in control conditions, although with a higher number of cells with lower membrane

potential, which is correlated with the increased membrane permeabilization of these cells (Godinho et

al., unpublished results). These differences in cell population distribution corroborate the protective

role of Pdr18 against acetic acid deleterious effects in yeast plasma membrane.

Moreover, since Pdr18 was proposed to have a role in reducing plasma membrane permeabilization,

as a response to acetic acid-induced stress, the effect of PDR18 expression in the extracellular

30

acidification of cells cultivated under control conditions and in the presence of inhibitory concentrations

of acetic acid, was compared. Extracellular acidification in deletion mutant Δpdr18 cells grown under

acetic acid stress conditions was significantly lower, than Δpdr18 cells grown in the absence of stress,

as expected, regarding plasma membrane potential results. This confirms that in these cells, there is

an increase of the rate of passive protons influx, which cannot be counterbalanced by the ATPase

activity, because of the enhanced permeabilizing effect of acetic acid in the plasma membrane of cells

devoid of Pdr18 (Godinho et al., unpublished results). In cells expressing the PDR18 gene, previously

cultivated in the presence of acetic acid, a reduction in the rate of extracellular acidification was

observed which is also mainly related with the passive proton influx, due to the permeabilization of the

plasma membrane caused by acetic acid (Godinho et al., unpublished results). In this case, results

obtained in extracellular acidification do not illustrate the proposed effect of PDR18 expression in the

reduction of the plasma membrane permeabilization in cells cultivated in the presence of acetic acid

observed in plasma membrane potential assays, which may be explained by limitations associated

with the experimental methodology used. Since it is an in vivo method, there are other phenomenon

and mechanisms occurring in the cell, despite the in vivo activity of the H+-ATPase and the passive H

+

influx through the plasma membrane that were roughly evaluated by the extracellular acidification

curves. In fact, this result may also be related with different availability of ATP in yeast cells, despite of

the activation of H+-ATPase, disabling the proton pumping activity, and, consequently, the acidification

of the extracellular media. Nevertheless, although statistical tests displayed differences not significant

in the final extracellular pH and extracellular acidification rate between cells expressing or not PDR18,

cultivated in the presence of acetic acid, the deletion of PDR18 was found to lead to a lower

extracellular acidification rate and a higher final extracellular pH, than cells expressing PDR18, which

demonstrate that these results have a similar tendency as those obtained in membrane potential

assays.

Altogether, results support the hypothesis that Pdr18 may be essential for the maintenance of the

membrane function and to prevent membrane permeabilization in yeast cells growing in the presence

of inhibitory concentrations of acetic acid. This work contributes to improve our understanding of the

ABC transporter Pdr18 biological function in the context of multidrug resistance, in the eukaryotic

model S. cerevisiae. The importance of the combined tolerance to both of the stress factors examined

is relevant for the improvement of the fermentation performance envisaging more efficient production

of bioethanol using lignocellulosic hydrolysates.

31

References

AGUILERA, F., PEINADO, R. A., MILLAN, C., ORTEGA, J. M. & MAURICIO, J. C. 2006. Relationship between ethanol tolerance, H+ -ATPase activity and the lipid composition of the plasma membrane in different wine yeast strains. Int J Food Microbiol, 110, 34-42.

ALMEIDA, B., OHLMEIER, S., ALMEIDA, A. J., MADEO, F., LEAO, C., RODRIGUES, F. & LUDOVICO, P. 2009. Yeast protein expression profile during acetic acid-induced apoptosis indicates causal involvement of the TOR pathway. Proteomics, 9, 720-32.

BAGNAT, M., KERANEN, S., SHEVCHENKO, A., SHEVCHENKO, A. & SIMONS, K. 2000. Lipid rafts function in biosynthetic delivery of proteins to the cell surface in yeast. Proc Natl Acad Sci U S A, 97, 3254-9.

BALZI, E. & GOFFEAU, A. 1994. Genetics and biochemistry of yeast multidrug resistance. Biochim Biophys Acta, 1187, 152-62.

BALZI, E. & GOFFEAU, A. 1995. Yeast multidrug resistance: the PDR network. J Bioenerg Biomembr, 27, 71-6.

BELTRAN, G., NOVO, M., GUILLAMON, J. M., MAS, A. & ROZES, N. 2008. Effect of fermentation temperature and culture media on the yeast lipid composition and wine volatile compounds. Int J Food Microbiol, 121, 169-77.

BENEY, L. & GERVAIS, P. 2001. Influence of the fluidity of the membrane on the response of microorganisms to environmental stresses. Appl Microbiol Biotechnol, 57, 34-42.

BUKAU, B. & HORWICH, A. L. 1998. The Hsp70 and Hsp60 chaperone machines. Cell, 92, 351-66.

CABRITO, T. R., TEIXEIRA, M. C., SINGH, A., PRASAD, R. & SA-CORREIA, I. 2011. The yeast ABC transporter Pdr18 (ORF YNR070w) controls plasma membrane sterol composition, playing a role in multidrug resistance. Biochem J, 440, 195-202.

CARMELO, V., BOGAERTS, P. & SA-CORREIA, I. 1996. Activity of plasma membrane H+-ATPase and expression of PMA1 and PMA2 genes in Saccharomyces cerevisiae cells grown at optimal and low pH. Arch Microbiol, 166, 315-20.

CARRATU, L., FRANCESCHELLI, S., PARDINI, C. L., KOBAYASHI, G. S., HORVATH, I., VIGH, L. & MARESCA, B. 1996. Membrane lipid perturbation modifies the set point of the temperature of heat shock response in yeast. Proc Natl Acad Sci U S A, 93, 3870-5.

32

CASAL, M., CARDOSO, H. & LEAO, C. 1996. Mechanisms regulating the transport of acetic acid in Saccharomyces cerevisiae. Microbiology, 142 ( Pt 6), 1385-90.

CASPETA, L., CHEN, Y., GHIACI, P., FEIZI, A., BUSKOV, S., HALLSTROM, B. M., PETRANOVIC, D. & NIELSEN, J. 2014. Biofuels. Altered sterol composition renders yeast thermotolerant. Science, 346, 75-8.

CYERT, M. S. 2003. Calcineurin signaling in Saccharomyces cerevisiae: how yeast go crazy in response to stress. Biochem Biophys Res Commun, 311, 1143-50.

DE VIRGILIO, C., HOTTIGER, T., DOMINGUEZ, J., BOLLER, T. & WIEMKEN, A. 1994. The role of trehalose synthesis for the acquisition of thermotolerance in yeast. I. Genetic evidence that trehalose is a thermoprotectant. Eur J Biochem, 219, 179-86.

DECOTTIGNIES, A., GRANT, A. M., NICHOLS, J. W., DE WET, H., MCINTOSH, D. B. & GOFFEAU, A. 1998. ATPase and multidrug transport activities of the overexpressed yeast ABC protein Yor1p. J Biol Chem, 273, 12612-22.

DEL SORBO, G., SCHOONBEEK, H. & DE WAARD, M. A. 2000. Fungal transporters involved in efflux of natural toxic compounds and fungicides. Fungal Genet Biol, 30, 1-15.

DESMOUCELLES, C., PINSON, B., SAINT-MARC, C. & DAIGNAN-FORNIER, B. 2002. Screening the yeast "disruptome" for mutants affecting resistance to the immunosuppressive drug, mycophenolic acid. J Biol Chem, 277, 27036-44.

DEXTER, D., MOYE-ROWLEY, W. S., WU, A. L. & GOLIN, J. 1994. Mutations in the yeast PDR3, PDR4, PDR7 and PDR9 pleiotropic (multiple) drug resistance loci affect the transcript level of an ATP binding cassette transporter encoding gene, PDR5. Genetics, 136, 505-15.

DICKSON, R. C. & LESTER, R. L. 2002. Sphingolipid functions in Saccharomyces cerevisiae. Biochim Biophys Acta, 1583, 13-25.

DOGAN, A., DEMIRCI, S., AYTEKIN, A. O. & SAHIN, F. 2014. Improvements of Tolerance to Stress Conditions by Genetic Engineering in Saccharomyces Cerevisiae during Ethanol Production. Appl Biochem Biotechnol.

DUPONT, S., BENEY, L., FERREIRA, T. & GERVAIS, P. 2011. Nature of sterols affects plasma membrane behavior and yeast survival during dehydration. Biochim Biophys Acta, 1808, 1520-8.

EISENKOLB, M., ZENZMAIER, C., LEITNER, E. & SCHNEITER, R. 2002. A specific structural requirement for ergosterol in long-chain fatty acid synthesis mutants important for maintaining raft domains in yeast. Mol Biol Cell, 13, 4414-28.

33

EKLUND, R. Z., G. 1995. Simultaneous saccharification and fermentation of steam-pretreated willow. Enzyme and Microbial Technology

EKLUND, S. A., PITTMAN, J. L. & SMITH, R. C. 1997. Trends in dental care among insured Americans: 1980 to 1995. J Am Dent Assoc, 128, 171-8.

ELLIOTT, B., HALTIWANGER, R. S. & FUTCHER, B. 1996. Synergy between trehalose and Hsp104 for thermotolerance in Saccharomyces cerevisiae. Genetics, 144, 923-33.

FERNANDES, A. R., DURAO, P. J., SANTOS, P. M. & SA-CORREIA, I. 2003. Activation and significance of vacuolar H+-ATPase in Saccharomyces cerevisiae adaptation and resistance to the herbicide 2,4-dichlorophenoxyacetic acid. Biochem Biophys Res Commun, 312, 1317-24.

FERNANDES, A. R., MIRA, N. P., VARGAS, R. C., CANELHAS, I. & SA-CORREIA, I. 2005. Saccharomyces cerevisiae adaptation to weak acids involves the transcription factor Haa1p and Haa1p-regulated genes. Biochem Biophys Res Commun, 337, 95-103.

FERNANDES, A. R. & SA-CORREIA, I. 2001. The activity of plasma membrane H(+)-ATPase is strongly stimulated during Saccharomyces cerevisiae adaptation to growth under high copper stress, accompanying intracellular acidification. Yeast, 18, 511-21.

GALBE, M. & ZACCHI, G. 2002. A review of the production of ethanol from softwood. Appl Microbiol Biotechnol, 59, 618-28.

GANCEDO, J. M. 1998. Yeast carbon catabolite repression. Microbiol Mol Biol Rev, 62, 334-61.

GBELSKA, Y., KRIJGER, J. J. & BREUNIG, K. D. 2006. Evolution of gene families: the multidrug resistance transporter genes in five related yeast species. FEMS Yeast Res, 6, 345-55.

GRAVES, T., NARENDRANATH, N. V., DAWSON, K. & POWER, R. 2006. Effect of pH and lactic or acetic acid on ethanol productivity by Saccharomyces cerevisiae in corn mash. J Ind Microbiol Biotechnol, 33, 469-74.

GREETHAM, D., WIMALASENA, T., KERRUISH, D. W., BRINDLEY, S., IBBETT, R. N., LINFORTH, R. L., TUCKER, G., PHISTER, T. G. & SMART, K. A. 2014. Development of a phenotypic assay for characterisation of ethanologenic yeast strain sensitivity to inhibitors released from lignocellulosic feedstocks. J Ind Microbiol Biotechnol, 41, 931-45.

GULSHAN, K. & MOYE-ROWLEY, W. S. 2007. Multidrug resistance in fungi. Eukaryot Cell, 6, 1933-42.

HAHN-HAGERDAL, B., GALBE, M., GORWA-GRAUSLUND, M. F., LIDEN, G. & ZACCHI, G. 2006. Bio-ethanol--the fuel of tomorrow from the residues of today. Trends Biotechnol, 24, 549-56.

34

HATZIXANTHIS, K., MOLLAPOUR, M., SEYMOUR, I., BAUER, B. E., KRAPF, G., SCHULLER, C., KUCHLER, K. & PIPER, P. W. 2003. Moderately lipophilic carboxylate compounds are the selective inducers of the Saccharomyces cerevisiae Pdr12p ATP-binding cassette transporter. Yeast, 20, 575-85.

HIGGINS, C. F. 1992. ABC transporters: from microorganisms to man. Annu Rev Cell Biol, 8, 67-113.

HIGGINS, C. F. 1995. The ABC of channel regulation. Cell, 82, 693-6.

HIGGINS, C. F. 2001. ABC transporters: physiology, structure and mechanism--an overview. Res Microbiol, 152, 205-10.

HIGGINS, C. F. 2007. Multiple molecular mechanisms for multidrug resistance transporters. Nature, 446, 749-57.

HOEBERICHTS, F. A., PEREZ-VALLE, J., MONTESINOS, C., MULET, J. M., PLANES, M. D., HUESO, G., YENUSH, L., SHARMA, S. C. & SERRANO, R. 2010. The role of K(+) and H(+) transport systems during glucose- and H(2)O(2)-induced cell death in Saccharomyces cerevisiae. Yeast, 27, 713-25.

HOLYOAK, C. D., STRATFORD, M., MCMULLIN, Z., COLE, M. B., CRIMMINS, K., BROWN, A. J. & COOTE, P. J. 1996. Activity of the plasma membrane H(+)-ATPase and optimal glycolytic flux are required for rapid adaptation and growth of Saccharomyces cerevisiae in the presence of the weak-acid preservative sorbic acid. Appl Environ Microbiol, 62, 3158-64.

INGRAM, L. O., ALDRICH, H. C., BORGES, A. C., CAUSEY, T. B., MARTINEZ, A., MORALES, F., SALEH, A., UNDERWOOD, S. A., YOMANO, L. P., YORK, S. W., ZALDIVAR, J. & ZHOU, S. 1999. Enteric bacterial catalysts for fuel ethanol production. Biotechnol Prog, 15, 855-66.

JENKINS, G. M., RICHARDS, A., WAHL, T., MAO, C., OBEID, L. & HANNUN, Y. 1997. Involvement of yeast sphingolipids in the heat stress response of Saccharomyces cerevisiae. J Biol Chem, 272, 32566-72.

JUNGWIRTH, H. & KUCHLER, K. 2006. Yeast ABC transporters-- a tale of sex, stress, drugs and aging. FEBS Lett, 580, 1131-8.

JURANKA, P. F., ZASTAWNY, R. L. & LING, V. 1989. P-glycoprotein: multidrug-resistance and a superfamily of membrane-associated transport proteins. FASEB J, 3, 2583-92.

KATZMANN, D. J., BURNETT, P. E., GOLIN, J., MAHE, Y. & MOYE-ROWLEY, W. S. 1994. Transcriptional control of the yeast PDR5 gene by the PDR3 gene product. Mol Cell Biol, 14, 4653-61.

35

KAWAHATA, M., MASAKI, K., FUJII, T. & IEFUJI, H. 2006. Yeast genes involved in response to lactic acid and acetic acid: acidic conditions caused by the organic acids in Saccharomyces cerevisiae cultures induce expression of intracellular metal metabolism genes regulated by Aft1p. FEMS Yeast Res, 6, 924-36.

KLOSE, C., SURMA, M. A., GERL, M. J., MEYENHOFER, F., SHEVCHENKO, A. & SIMONS, K. 2012. Flexibility of a eukaryotic lipidome--insights from yeast lipidomics. PLoS One, 7, e35063.

KOLACZKOWSKA, A. & GOFFEAU, A. 1999. Regulation of pleiotropic drug resistance in yeast. Drug Resist Updat, 2, 403-414.

KREBS, H. A., WIGGINS, D., STUBBS, M., SOLS, A. & BEDOYA, F. 1983. Studies on the mechanism of the antifungal action of benzoate. Biochem J, 214, 657-63.

LALUCE, C., SCHENBERG, A. C., GALLARDO, J. C., CORADELLO, L. F. & POMBEIRO-SPONCHIADO, S. R. 2012. Advances and developments in strategies to improve strains of Saccharomyces cerevisiae and processes to obtain the lignocellulosic ethanol--a review. Appl Biochem Biotechnol, 166, 1908-26.

LI, Y. & PRINZ, W. A. 2004. ATP-binding cassette (ABC) transporters mediate nonvesicular, raft-modulated sterol movement from the plasma membrane to the endoplasmic reticulum. J Biol Chem, 279, 45226-34.

LU, Y., CHENG, Y. F., HE, X. P., GUO, X. N. & ZHANG, B. R. 2012. Improvement of robustness and ethanol production of ethanologenic Saccharomyces cerevisiae under co-stress of heat and inhibitors. J Ind Microbiol Biotechnol, 39, 73-80.

MADEIRA-LOPES, A. & VAN UDEN, N. 1979. Thermal association and dissociation in thermosensitive mutants of Saccharomyces cerevisiae. Z Allg Mikrobiol, 19, 303-5.

MADRID, R., GOMEZ, M. J., RAMOS, J. & RODRIGUEZ-NAVARRO, A. 1998. Ectopic Potassium Uptake in trk1 trk2 Mutants ofSaccharomyces cerevisiae Correlates with a Highly Hyperpolarized Membrane Potential. Journal of Biological Chemistry, 273, 14838-14844.

MAKRANTONI, V., DENNISON, P., STARK, M. J. & COOTE, P. J. 2007. A novel role for the yeast protein kinase Dbf2p in vacuolar H+-ATPase function and sorbic acid stress tolerance. Microbiology, 153, 4016-26.

MAMNUN, Y. M., PANDJAITAN, R., MAHE, Y., DELAHODDE, A. & KUCHLER, K. 2002. The yeast zinc finger regulators Pdr1p and Pdr3p control pleiotropic drug resistance (PDR) as homo- and heterodimers in vivo. Mol Microbiol, 46, 1429-40.

36

MAYER, M. P. & BUKAU, B. 1998. Hsp70 chaperone systems: diversity of cellular functions and mechanism of action. Biol Chem, 379, 261-8.

MIRA, N. P., BECKER, J. D. & SA-CORREIA, I. 2010a. Genomic expression program involving the Haa1p-regulon in Saccharomyces cerevisiae response to acetic acid. OMICS, 14, 587-601.

MIRA, N. P., LOURENCO, A. B., FERNANDES, A. R., BECKER, J. D. & SA-CORREIA, I. 2009. The RIM101 pathway has a role in Saccharomyces cerevisiae adaptive response and resistance to propionic acid and other weak acids. FEMS Yeast Res, 9, 202-16.

MIRA, N. P., PALMA, M., GUERREIRO, J. F. & SA-CORREIA, I. 2010b. Genome-wide identification of Saccharomyces cerevisiae genes required for tolerance to acetic acid. Microb Cell Fact, 9, 79.

MIRA, N. P., TEIXEIRA, M. C. & SA-CORREIA, I. 2010c. Adaptive response and tolerance to weak acids in Saccharomyces cerevisiae: a genome-wide view. OMICS, 14, 525-40.

MORANO, K. A., GRANT, C. M. & MOYE-ROWLEY, W. S. 2012. The response to heat shock and oxidative stress in Saccharomyces cerevisiae. Genetics, 190, 1157-95.

MULET, J. M., LEUBE, M. P., KRON, S. J., RIOS, G., FINK, G. R. & SERRANO, R. 1999. A novel mechanism of ion homeostasis and salt tolerance in yeast: the Hal4 and Hal5 protein kinases modulate the Trk1-Trk2 potassium transporter. Mol Cell Biol, 19, 3328-37.

NOURANI, A., WESOLOWSKI-LOUVEL, M., DELAVEAU, T., JACQ, C. & DELAHODDE, A. 1997. Multiple-drug-resistance phenomenon in the yeast Saccharomyces cerevisiae: involvement of two hexose transporters. Mol Cell Biol, 17, 5453-60.

OLOFSSON, K., BERTILSSON, M. & LIDEN, G. 2008. A short review on SSF - an interesting process option for ethanol production from lignocellulosic feedstocks. Biotechnol Biofuels, 1, 7.

PALMQVIST, E., GRAGE, H., MEINANDER, N. Q. & HAHN-HAGERDAL, B. 1999. Main and interaction effects of acetic acid, furfural, and p-hydroxybenzoic acid on growth and ethanol productivity of yeasts. Biotechnol Bioeng, 63, 46-55.

PAMPULHA, M. E., LOUREIRO-DIAS, M. C. 1990. Activity of glycolytic enzymes of Saccharomyces cerevisiae in the presence of acetic acid. Applied Microbiology and Biotechnology 34, 375-380.

PAO, S. S., PAULSEN, I. T. & SAIER, M. H., JR. 1998. Major facilitator superfamily. Microbiol Mol Biol Rev, 62, 1-34.

37

PARKS, L. W. & CASEY, W. M. 1995. Physiological implications of sterol biosynthesis in yeast. Annu Rev Microbiol, 49, 95-116.

PARSELL, D. A., KOWAL, A. S., SINGER, M. A. & LINDQUIST, S. 1994. Protein disaggregation mediated by heat-shock protein Hsp104. Nature, 372, 475-8.

PEREIRA, F. B., GUIMARAES, P. M., GOMES, D. G., MIRA, N. P., TEIXEIRA, M. C., SA-CORREIA, I. & DOMINGUES, L. 2011. Identification of candidate genes for yeast engineering to improve bioethanol production in very high gravity and lignocellulosic biomass industrial fermentations. Biotechnol Biofuels, 4, 57.

PIPER, P., CALDERON, C. O., HATZIXANTHIS, K. & MOLLAPOUR, M. 2001. Weak acid adaptation: the stress response that confers yeasts with resistance to organic acid food preservatives. Microbiology, 147, 2635-42.

PIPER, P., MAHE, Y., THOMPSON, S., PANDJAITAN, R., HOLYOAK, C., EGNER, R., MUHLBAUER, M., COOTE, P. & KUCHLER, K. 1998. The pdr12 ABC transporter is required for the development of weak organic acid resistance in yeast. EMBO J, 17, 4257-65.

PIPER, P. W. 1995. The heat shock and ethanol stress responses of yeast exhibit extensive similarity and functional overlap. FEMS Microbiol Lett, 134, 121-7.

PRASAD, R., PANWAR, S. L. & SMRITI 2002. Drug resistance in yeasts--an emerging scenario. Adv Microb Physiol, 46, 155-201.

RAJOKA, M. I., FERHAN, M. & KHALID, A. M. 2005. Kinetics and thermodynamics of ethanol production by a thermotolerant mutant of Saccharomyces cerevisiae in a microprocessor-controlled bioreactor. Lett Appl Microbiol, 40, 316-21.

RAMOS, M. T., MADEIRA-LOPES, A. 1990. Effects of acetic acid on the temperature profile of ethanol tolerance in Saccharomyces cerevisiae. Biotechnology Letters, 12, 229-234.

ROCKWELL, N. C., WOLFGER, H., KUCHLER, K. & THORNER, J. 2009. ABC transporter Pdr10 regulates the membrane microenvironment of Pdr12 in Saccharomyces cerevisiae. J Membr Biol, 229, 27-52.

ROSA, M. F., SÁ-CORREIA, I. 1994. Limitations to the use of extracellular acidification for the assessment of plasma membrane H+-ATPase activity and ethanol tolerance in yeasts. Enzyme Microb Technol, 16, 5.

SA-CORREIA, I., DOS SANTOS, S. C., TEIXEIRA, M. C., CABRITO, T. R. & MIRA, N. P. 2009. Drug:H+ antiporters in chemical stress response in yeast. Trends Microbiol, 17, 22-31.

38

SA-CORREIA, I. & TENREIRO, S. 2002. The multidrug resistance transporters of the major facilitator superfamily, 6 years after disclosure of Saccharomyces cerevisiae genome sequence. J Biotechnol, 98, 215-26.

SALGUEIRO, S. P., SA-CORREIA, I. & NOVAIS, J. M. 1988. Ethanol-Induced Leakage in Saccharomyces cerevisiae: Kinetics and Relationship to Yeast Ethanol Tolerance and Alcohol Fermentation Productivity. Appl Environ Microbiol, 54, 903-9.

SCHULLER, C., MAMNUN, Y. M., WOLFGER, H., ROCKWELL, N., THORNER, J. & KUCHLER, K. 2007. Membrane-active compounds activate the transcription factors Pdr1 and Pdr3 connecting pleiotropic drug resistance and membrane lipid homeostasis in saccharomyces cerevisiae. Mol Biol Cell, 18, 4932-44.

SERRANO, R., KIELLAND-BRANDT, M. C. & FINK, G. R. 1986. Yeast plasma membrane ATPase is essential for growth and has homology with (Na+ + K+), K+- and Ca2+-ATPases. Nature, 319, 689-93.

SIMOES-MENDES, B., MADEIRA-LOPES, A. & VAN UDEN, N. 1978. Kinetics of petite mutation and thermal death in Saccharomyces cerevisiae growing at superoptimal temperatures. Z Allg Mikrobiol, 18, 275-9.

SIMOES, T., MIRA, N. P., FERNANDES, A. R. & SA-CORREIA, I. 2006. The SPI1 gene, encoding a glycosylphosphatidylinositol-anchored cell wall protein, plays a prominent role in the development of yeast resistance to lipophilic weak-acid food preservatives. Appl Environ Microbiol, 72, 7168-75.

STRATFORD, M., STEELS, H., NEBE-VON-CARON, G., NOVODVORSKA, M., HAYER, K. & ARCHER, D. B. 2013. Extreme resistance to weak-acid preservatives in the spoilage yeast Zygosaccharomyces bailii. Int J Food Microbiol, 166, 126-34.

SUUTARI, M., LIUKKONEN, K. & LAAKSO, S. 1990. Temperature adaptation in yeasts: the role of fatty acids. J Gen Microbiol, 136, 1469-74.

SWAN, T. M. & WATSON, K. 1998. Stress tolerance in a yeast sterol auxotroph: role of ergosterol, heat shock proteins and trehalose. FEMS Microbiol Lett, 169, 191-7.

TEIXEIRA, M. C., CABRITO, T. R., HANIF, Z. M., VARGAS, R. C., TENREIRO, S. & SA-CORREIA, I. 2011. Yeast response and tolerance to polyamine toxicity involving the drug : H+ antiporter Qdr3 and the transcription factors Yap1 and Gcn4. Microbiology, 157, 945-56.

TEIXEIRA, M. C., GODINHO, C. P., CABRITO, T. R., MIRA, N. P. & SA-CORREIA, I. 2012. Increased expression of the yeast multidrug resistance ABC transporter Pdr18 leads to increased

39

ethanol tolerance and ethanol production in high gravity alcoholic fermentation. Microb Cell Fact, 11, 98.

TEIXEIRA, M. C., SANTOS, P. M., FERNANDES, A. R. & SA-CORREIA, I. 2005. A proteome analysis of the yeast response to the herbicide 2,4-dichlorophenoxyacetic acid. Proteomics, 5, 1889-901.

TURK, M., PLEMENITAS, A. & GUNDE-CIMERMAN, N. 2011. Extremophilic yeasts: plasma-membrane fluidity as determinant of stress tolerance. Fungal Biol, 115, 950-8.

ULLAH, A., ORIJ, R., BRUL, S. & SMITS, G. J. 2012. Quantitative analysis of the modes of growth inhibition by weak organic acids in Saccharomyces cerevisiae. Appl Environ Microbiol, 78, 8377-87.

VAN UDEN, N. 1984. Temperature profiles of yeasts. Adv Microb Physiol, 25, 195-251.

VAN UDEN, N. & MADEIRA-LOPES, A. 1970. Concurrent exponential growth and death of cell populations of Saccharomyces cerevisiae at superoptimal growth temperatures. Z Allg Mikrobiol, 10, 515-26.

VAN UDEN, N. & MADEIRA-LOPES, A. 1976. Yield and maintenance relations of yeast growth in the chemostat at superoptimal temperatures. Biotechnol Bioeng, 18, 791-804.

VERGHESE, J., ABRAMS, J., WANG, Y. & MORANO, K. A. 2012. Biology of the heat shock response and protein chaperones: budding yeast (Saccharomyces cerevisiae) as a model system. Microbiol Mol Biol Rev, 76, 115-58.

VIEGAS, C. A., CABRAL, M. G., TEIXEIRA, M. C., NEUMANN, G., HEIPIEPER, H. J. & SA-CORREIA, I. 2005. Yeast adaptation to 2,4-dichlorophenoxyacetic acid involves increased membrane fatty acid saturation degree and decreased OLE1 transcription. Biochem Biophys Res Commun, 330, 271-8.

VIEGAS, C. A. & SA-CORREIA, I. 1991. Activation of plasma membrane ATPase of Saccharomyces cerevisiae by octanoic acid. J Gen Microbiol, 137, 645-51.

VIEGAS, C. A., SEBASTIAO, P. B., NUNES, A. G. & SA-CORREIA, I. 1995. Activation of plasma membrane H(+)-ATPase and expression of PMA1 and PMA2 genes in Saccharomyces cerevisiae cells grown at supraoptimal temperatures. Appl Environ Microbiol, 61, 1904-9.

WENDLER, F., BERGLER, H., PRUTEJ, K., JUNGWIRTH, H., ZISSER, G., KUCHLER, K. & HOGENAUER, G. 1997. Diazaborine resistance in the yeast Saccharomyces cerevisiae reveals a link between YAP1 and the pleiotropic drug resistance genes PDR1 and PDR3. J Biol Chem, 272, 27091-8.

40

WIMALASENA, T. T., GREETHAM, D., MARVIN, M. E., LITI, G., CHANDELIA, Y., HART, A., LOUIS, E. J., PHISTER, T. G., TUCKER, G. A. & SMART, K. A. 2014. Phenotypic characterisation of Saccharomyces spp. yeast for tolerance to stresses encountered during fermentation of lignocellulosic residues to produce bioethanol. Microb Cell Fact, 13, 47.

YAMAMOTO, N., MAEDA, Y., IKEDA, A. & SAKURAI, H. 2008. Regulation of thermotolerance by stress-induced transcription factors in Saccharomyces cerevisiae. Eukaryot Cell, 7, 783-90.

ZHANG, X., CUI, Z., MIYAKAWA, T. & MOYE-ROWLEY, W. S. 2001. Cross-talk between transcriptional regulators of multidrug resistance in Saccharomyces cerevisiae. J Biol Chem, 276, 8812-9.

ZHAO, J., WANG, Z., WANG, M., HE, Q. & ZHANG, H. 2008. The inhibition of Saccharomyces cerevisiae cells by acetic acid quantified by electrochemistry and fluorescence. Bioelectrochemistry, 72, 117-21.

ZINSER, E., PALTAUF, F. & DAUM, G. 1993. Sterol composition of yeast organelle membranes and subcellular distribution of enzymes involved in sterol metabolism. J Bacteriol, 175, 2853-8.

41

Annexes

42

6. Annex I

Acetic acid concentrations required to inhibit growth of Saccharomyces cerevisiae

parental strain BY4741 and the derived deletion mutant strain Δpdr18 incubated at

temperatures in the range 30-40°C.

Figure 6.1 - Culture OD600nm of Saccharomyces cerevisiae parental strain BY4741 and derived

deletion mutant Δpdr18 grown at 30°C in the absence and in the presence of acetic acid. Cell

suspensions of BY4741 and Δpdr18 grown at 30°C were used to inoculate MM4 liquid medium (pH

4.0) supplemented or not with increasing concentrations of acetic acid (50 – 120 mM). Growth at 30°C

was followed during 240 hours.

43

Figure 6.2 - Culture OD600nm of Saccharomyces cerevisiae parental strain BY4741 and derived

deletion mutant Δpdr18 grown at 36°C in the absence and in the presence of acetic acid. Cell

suspensions of BY4741 and Δpdr18 grown at 30°C were used to inoculate MM4 liquid medium (pH

4.0) supplemented or not with increasing concentrations of acetic acid (50 – 120 mM). Growth at 36°C

was followed during 240 hours.

44

Figure 6.3 - Culture OD600nm of Saccharomyces cerevisiae parental strain BY4741 and derived

deletion mutant Δpdr18 grown at 38°C in the absence and in the presence of acetic acid. Cell

suspensions of BY4741 and Δpdr18 grown at 30°C were used to inoculate MM4 liquid medium (pH

4.0) supplemented or not with increasing concentrations of acetic acid (50 – 110 mM). Growth at 38°C

was followed during 240 hours.

45

Figure 6.4 - Culture OD600nm of Saccharomyces cerevisiae parental strain BY4741 and derived

deletion mutant Δpdr18 grown at 40°C in the absence and in the presence of acetic acid. Cell

suspensions of BY4741 and Δpdr18 grown at 30°C were used to inoculate MM4 liquid medium (pH

4.0) supplemented or not with increasing concentrations of acetic acid (35 – 90 mM). Growth at 40°C

was followed during 240 hours.

46

7. Annex II

Acetic acid concentrations required to inhibit growth at 30°C of Saccharomyces cerevisiae

parental strain BY4741 and the derived deletion mutant strain Δpdr18 harbouring a PDR18

expression plasmid or the corresponding cloning vector pRS416.

Figure 7.1 - Culture OD600nm of Saccharomyces cerevisiae parental strain BY4741

harbouring a PDR18 expression plasmid or the corresponding cloning vector pRS416 in

the absence and in the presence of acetic acid. Cell suspensions grown at 30°C were used

to inoculate MM4 liquid medium (pH 4.0) supplemented or not with increasing concentrations

of acetic acid (30 – 110 mM). Growth at 30°C was followed during 240 hours

47

Figure 7.2 - Culture OD600nm of Saccharomyces cerevisiae derived deletion mutant Δpdr18

harbouring a PDR18 expression plasmid or the corresponding cloning vector pRS416 in the

absence and in the presence of acetic acid. Cell suspensions grown at 30°C were used to inoculate

MM4 liquid medium (pH 4.0) supplemented or not with increasing concentrations of acetic acid

(30 – 110 mM). Growth at 30°C was followed during 240 hours.

48

8. Annex III

Tolerance of Saccharomyces cerevisiae parental strain BY4741 and derived deletion

mutant Δpdr18 to acetic acid.

Figure 8.1 - Conjugated effect of acetic acid and temperature in the growth of Saccharomyces

cerevisiae parental strain BY4741 and derived deletion mutant Δpdr18. Cell suspensions of

BY4741 and Δpdr18 grown at 30°C were used to inoculate MM4 liquid medium (pH 4.0) supplemented

or not with acetic acid. Minimum inhibitory concentration (MIC) values attained are shown during 240

hours of incubation at 30°C, 36°C, 38°C and 40°C in minimal medium supplemented with different

acetic acid concentrations (35 - 120 mM, at pH 4.0). Values presented in Y axis are the lowest

concentration (mM) of acetic acid to completely inhibit growth.