177
THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: REGULATION OF PH HOMEOSTASIS AND SURVIVAL IN SACCHAROMYCES CEREVISIAE BY ROS DAMAGE AND SIGNALING by James Allen Baron A dissertation submitted to Johns Hopkins University in conformity with the requirements for the degree of Doctor of Philosophy Baltimore, Maryland October, 2013 © 2013 J. Allen Baron All Rights Reserved

THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: REGULATION OF PH HOMEOSTASIS AND SURVIVAL IN SACCHAROMYCES CEREVISIAE BY ROS DAMAGE

AND SIGNALING

by James Allen Baron

A dissertation submitted to Johns Hopkins University in conformity with the requirements for the degree of Doctor of Philosophy

Baltimore, Maryland October, 2013

© 2013 J. Allen Baron All Rights Reserved

Page 2: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

ii

ABSTRACT

pH homeostasis is intimately linked with the metabolic state of cells. The baker’s

yeast, Saccharomyces cerevisiae, is an excellent model of this as it readily switches from

fermentation to respiration contingent on carbon source availability. When grown in

abundant glucose, yeast obtain energy by fermentation and maintain a stable intracellular

pH over a variety of environmental pH’s by activating energetically expensive H+-

ATPases. When glucose is limited or absent, yeast switch to respiration, decrease the

activity of H+-ATPases to conserve energy, and pH homeostasis becomes highly

dependent on environmental pH and buffering by cellular metabolites. In this thesis, the

regulation of pH homeostasis by reactive oxygen species (ROS) is examined at these two

extremes of glucose abundance and glucose starvation using a variety of genetic and

biochemical techniques.

During long-term growth in the absence of glucose, yeast alternately alkalinize

and acidify their environment which affects growth and survival. We demonstrate that

mitochondrial ROS initiate the alkali-to-acid shift by inactivating the Fe-S cluster

enzymes aconitase and succinate dehydrogenase of the tricarboxylic acid (TCA) cycle

and that this shift is accelerated by deletion of the mitochondrial superoxide dismutase,

SOD2. Inhibition of the TCA cycle enzymes alters metabolite flux through the cycle and

leads to the buildup and secretion of the upstream metabolite acetic acid, which is

generated by the aldehyde dehydrogenase Ald4p. Acetic acid secretion acidifies the

environment without activating the plasma membrane H+-ATPase, Pma1p, and promotes

the growth of yeast under these conditions by providing a new carbon source for survival.

Page 3: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

iii

This work demonstrates that inhibition of Fe-S cluster enzymes by ROS can be beneficial

under conditions of glucose starvation.

In abundant glucose the principally cytosolic superoxide dismutase, Sod1p, is

required for maximal activation of Pma1p, although the mechanism was not fully

understood. In this work we discovered that deletion of SOD1 in combination with a

specific mutation in the C-terminal autoinhibitory region of Pma1p, Pma1-T912D, is

lethal to yeast cells. This lethality can be rescued by treatment with Mn-based

antioxidants and by hypoxia. Furthermore, we identified spontaneous second-site

mutations in PMA1 that reverse the aerobic lethality by activating the H+-ATPase. Thus,

lethality results from profound inhibition of essential Pma1p activity. Together these

results support a model in which Sod1p helps protect Pma1p from oxidative damage,

particularly in cases where auto-inhibition of Pma1p is disrupted, as with Pma1-T912D.

Thesis Advisor: Dr. Valeria Culotta

Thesis Readers: Dr. Michael Trush

Dr. Rajini Rao

Dr. Jiou Wang

Page 4: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

iv

ACKNOWLEDGEMENTS

The first line of a song released in 1965 by the Lettermen speaks an important

truth “No man is an island, no man stands alone.” This is especially true in the life

sciences where expertise, reagents, and ideas are constantly shared. With this in mind, I

acknowledge those who contributed during the course of my PhD studies to this thesis.

Apart from myself, Dr. Valeria Culotta has done the most to ensure both the

success of this work and of my PhD career. She provided a place to do research and a

project to work on when I was required to find a new lab in my 4th year of study. Her

expertise established the basis of my knowledge about reactive oxygen species and

metals in biology. Her advice and guidance in both writing and presenting has been

essential in developing this thesis and improving my technical communication skills.

Janice Chen worked with me as an undergraduate from 2011 - 2013 and

contributed to this work by producing strains and plasmids, measuring intracellular

metabolites, performing growth tests and investigating a number of side projects. Of

particular note, Janice helped sort out the role of acetate and succinate in the acid burst.

To acknowledge others who contributed to the experiments found in this thesis:

Kaitlin Laws initiated the studies of Chapter 2 as a rotation student and some of her data

is included in her thesis; David Corrigan did early work that contributed to Chapter 3;

Dr. Aaron Haeusler and Sabin Mulepati helped me use a french press and microfluidizer,

respectively; Paul Miller helped me freeze samples for TCA metabolite analysis; Dr.

Yana Sandlers, assistant director of the Biochemical Genetics Laboratory at KKI,

performed the stable-isotope dilution to measure TCA metabolites; Caelin Potts provided

sage advice about indirect immunofluorescence and use of the microscope; Dr. Yong-

Page 5: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

v

Qiang Zhang provided protocols for crude membrane isolation and Pma1p activity

assays; Dr. Brian Learn provided technical expertise and access to an ultracentrifuge

rotor required for crude membrane isolations; various members of the Culotta lab

including Christine Vasquez, Dr. Julie Gleason, Chynna Broxton, Cissy Li, Dr. Daphne

Aguirre, and Dr. Jeffry Leitch inoculated/collected cells and provided reagents, among

other things; Margaret Dayhoff-Brannigan provided access to a dissecting scope which

helped me to sort out which strains develop rho- mutations; lastly, Brendan Cormack

provided GFP which will be fused to Pma1p for future experiments.

Dr. Amit Reddi deserves special recognition. He was examining the regulation of

Yck1p by Sod1p while I was doing the experiments in Chapter 3. This proved to be very

advantageous for comparison and discussion of results. He also introduced me to protein

activity assays and pH calculations using pH indicator dyes (Chapter 2). His influence

and professional expertise has proved invaluable.

To acknowledge those whose support was essential to the work of my PhD and

the production of this thesis: Rob Jensen, Rajini Rao, and Jiou Wang served as my thesis

advisory committee; Sharon Warner and Geraldine Graziano guided my academic

pursuits and helped with arranging the thesis defense and committee; Michael Matunis

provided guidance when switching labs; Gabriel Brandt provided assistance with

topology maps; Martha Baron helped with organization and formatting of this thesis; and

Marsha Baron cared for our son so I could devote the time needed to writing.

Drs. Eric Grote and Valerie Olmo worked closely with me in my early PhD career

examining the role of calcium in cell-cell fusion. Their contributions were great and

cannot be properly enumerated here.

Page 6: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

vi

TABLE OF CONTENTS

Abstract ……………………………………………………………………………... ii

Acknowledgements …………………………………………………………………. iv

Table of Contents …………………………………………………………………… vi

List of Tables ……………………………………………………………………….. viii

List of Figures ………………………………………………………………………. ix

List of Abbreviations ……………………………………………………………….. xi

Chapter 1 – Introduction

Introduction to Reactive Oxygen Species …………………………………... 2

Sources of ROS ……………………………………………………………... 5

Targets of ROS ………………………………………………………........... 10

Antioxidant Enzymes ……………………………………………………..... 12

ROS in Signaling …………………………………………………………… 16

pH Homeostasis in the Baker’s Yeast S. cerevisiae …………………........... 18

The Research in This Thesis ………………………………………….......... 24

References …………………………………………………………….......... 30

Chapter 2 – Superoxide Triggers an Acid Burst in Saccharomyces cerevisiae to

Condition the Environment of Glucose-starved Cells

Introduction ………………………………………………………………… 61

Materials and Methods ……………………………………………………... 63

Results ……………………………………………………………………… 68

Discussion …………………………………………………………….......... 76

Page 7: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

vii

References …………………………………………………………….......... 96

Chapter 3 – Oxidative Stress and Lethality in S. cerevisiae with Mutations in SOD1

and the Proton ATPase PMA1

Introduction ………………………………………………………………… 108

Materials and Methods ……………………………………………………... 110

Results ……………………………………………………………………… 114

Discussion …………………………………………………………….......... 119

References …………………………………………………………….......... 136

Chapter 4 - Review and Future Directions

Review of Thesis Research ………………………………………………. 148

Future Directions …………………………………………………………… 152

References …………………………………………………………….......... 156

Curriculum Vitae …………………………………………………………………… 159

Page 8: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

viii

LIST OF TABLES

TABLE 2-1. Effects of TCA and ETC mutants on extracellular pH ........................ 79

TABLE 2-2. Contribution of acetic acid to pH of growth medium ……………….. 81

TABLE 3-1. Effect of Pma1 mutations on activity and C-terminal conformation ... 122

TABLE 3-2. BY4741 strains used in this study …………………………………... 123

TABLE 3-3. NY13 strains used in this study ……………………………………... 124

TABLE 3-4. Plasmids used in this study ………………………………………….. 125

Page 9: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

ix

LIST OF FIGURES

FIGURE 1-1. Topology model of the plasma membrane H+-ATPase, Pma1p ……. 26

FIGURE 1-2. Diagram of extracellular pH changes and ammonia signaling by yeast

during long-term growth in the absence of glucose ………………………………… 28

FIGURE 2-1. Yeast cells alkalize their environment under respiratory conditions .. 82

FIGURE 2-2. Acid production following the alkaline phase of yeast cell growth is

accelerated by sod2∆ mutations and eliminated by ald4∆ mutations ……………… 84

FIGURE 2-3. Role of mitochondrial SOD2 and Mn-antioxidants in preventing the acid

burst of respiratory cells ……………………………………………………………. 85

FIGURE 2-4. The redox cycler paraquat and sdh∆ induce a rapid acid burst under

respiratory conditions and are not additive ………………………………………… 86

FIGURE 2-5. Activities of mitochondrial Fe-S cluster enzymes in acid producing cells

……………………………………………………………………………………… 88

FIGURE 2-6. Acid production with low glucose does not correlate with activation of

Pma1p, the proton ATPase ………………………………………………………… 90

FIGURE 2-7. Cells produce high levels of extracellular acetate during the acid burst

……………………………………………………………………………………… 91

FIGURE 2-8. The mitochondrial Ald4p aldehyde dehydrogenase is required for the acid

burst ………………………………………………………………………………… 93

FIGURE 2-9. Medium conditioned during the acid burst can promote yeast growth

……………………………………………………………………………………... 94

Page 10: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

x

FIGURE 3-1. Effect of sod1∆ on Pma1 localization and growth of Pma1 mutants with

altered glucose regulation ………………………………………………………….. 126

FIGURE 3-2. Mutations in Pma1 stalk segment 5 that alter glucose regulation are not

synthetically lethal with sod1∆ …………………………………………………….. 128

FIGURE 3-3. T912D sod1∆ cells develop suppressors with secondary mutations in

PMA1 ………………………………………………………………………………. 129

FIGURE 3-4. The PMA1-A906G mutation increases Pma1p ATPase activity but does

not alter Pma1p localization or protein level ………………………………………. 131

FIGURE 3-5. T912D sod1∆ synthetic lethality is due to oxidative stress and not Yck1p

misregulation ………………………………………………………………………. 132

FIGURE 3-6. Pma1-T912D mutant cells are not sensitive to environmental oxidants or

hyperoxia …………………………………………………………………………... 134

FIGURE 4-1. Models depicting ROS regulation of pH homeostasis at the metabolic

extremes of glucose abundance and glucose starvation ……………………………. 150

Page 11: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

xi

LIST OF ABBREVIATIONS

A absorbance

ACO1 aconitase

AHP1 thioredoxin peroxidase

AKR1 palmitoyl transferase, palmitoylates Yck1p

Ala alanine

ALD aldehyde dehydrogenase

ALD2 minor cytosolic ALD, stress-induced

ALD3 minor cytosolic ALD, stress-induced

ALD4 major mitochondrial ALD, glucose-repressed

ALD5 minor mitochondrial ALD, constitutively expressed

ALD6 major cytosolic ALD, constitutively expressed

Asp asparagine

ATO1 putative acetate importer/ammonia exporter

ATO2 putative acetate importer/ammonia exporter

ATO3 putative ammonia exporter

ATP adenosine triphosphate

BCG bromocresol green

BCP bromocresol purple

BSA bovine serum albumin

BSD2 heavy metal homeostasis protein, deletion suppresses sod1Δ

phenotypes

CaCl2 calcium chloride

Page 12: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

xii

CCS copper chaperone for SOD1

CIT1 citrate synthase

CO2 carbon dioxide

CoA coenzyme A

CTT1 cytosolic catalase

Cu copper

Cys cysteine

DABCO 1,4-diazabicyclo[2.2.2]octane

DAPI 4’,6-diamidino-2-phenylindole

DCIP 2,6-Dichloroindophenol

DNA deoxyribonucleic acid

DTT dithiothreitol

DUOX dual oxidase

E redox potential

EDTA ethylenediaminetetraacetic acid

EGF epidermal growth factor

EGTA ethylene glycol-bis(2-aminoethylether)-N,N,N′,N′-tetraacetic

acid

ETC electron transport chain

FAD flavin adenine dinucleotide

Fe iron

Fe-S iron-sulfur

FGF-2 fibroblast growth factor-2

Page 13: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

xiii

FUM1 fumarase

GC-MS gas chromatography-mass spectrometry

GLR1 glutathione reductase

GPX1 glutathione peroxidase

H+ proton

H2O2 hydrogen peroxide

H2SO4 sulfuric acid

HCl hydrochloric acid

HO• hydroxyl radical

HO2• hydroperoxyl radical

HOCl hypochlorous acid

hrs hours

IDH1 isocitrate dehydrogenase

IMS intermembrane space

KCN potassium cyanide

KGD1 α-ketoglutarate dehydrogenase

Km Michaelis constant

KNO3 potassium nitrate

KPhos potassium phosphate

Leu leucine

LEU1 isopropylmalate dehydratase

LSC1 succinyl-CoA ligase

Lys lysine

Page 14: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

xiv

LYS4 homoaconitase

MDH1 malate dehydrogenase

MES 2-(N-morpholino)ethanesulfonic acid

Met methionine

MgSO4 magnesium sulfate

mins minutes

mito mitochondrial

Mn manganese

MnCl2 manganese chloride

MOPS 3-(N-morpholino)propanesulfonic acid

NAD+ (NADH) nicotinamide adenine dinucleotide

NADP+ (NADPH) nicotinamide adenine dinucleotide phosphate

NaN3 sodium azide

NHA1 yeast plasma membrane Na+/H+ antiporter

NHE1 mammalian Na+/H+ exchanger

NHX1 yeast endocytic Na+(K+)/H+ exchanger

Ni nickel

NOX NADPH oxidase

Nox2 phagocytic NOX enzyme, also referred to as gp91phox

O2 molecular oxygen

O2• −

superoxide

OD600 optical density, 600 nm

PBS phosphate-buffered saline

Page 15: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

xv

PCD programmed cell death

PCR polymerase chain reaction

PDGF platelet-derived growth factor

pHex extracellular pH

pHi intracellular pH

pKa acid dissociation constant

PMA1 major plasma membrane proton ATPase

PMA2 minor plasma membrane proton ATPase

PMR1 golgi Ca2+/Mn2+ P-type ATPase, deletion suppresses sod1Δ

phenotypes

PMSF phenylmethylsulfonyl fluoride

PQ paraquat

psi pounds per square inch

PTK2 yeast Ser/Thr kinase, phosphorylates Pma1p

PTP protein tyrosine phosphatase

RHO+ WT mitochondrial DNA

rho- mitochondrial DNA mutant

ROS reactive oxygen species

rpm revolutions per minute

S. cerevisiae Saccharomyces cerevisiae

SC synthetic complete

SD standard deviation

SDH succinate dehydrogenase

Page 16: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

xvi

SDS sodium dodecyl sulfate

SERCA sarcoplasmic/endoplasmic reticulum Ca2+-ATPase

SOD superoxide dismutase

SOD1 cytosolic Cu/Zn-SOD

SOD2 mitochondrial Mn-SOD

-SOH sulfenic acid

-SO2H sulfinic acid

-SO3H sulfonic acid

SOK2 yeast negative transcriptional regulator of pseudohyphal

differentiation

TCA tricarboxylic acid

TNF tumor necrosis factor

TNFR tumor necrosis factor receptor

TRR2 mitochondrial thioredoxin reductase

TSA1 thioredoxin peroxidase

TUF1/RAP1 positive transcriptional regulator of Pma1p

URA3 orotidine-5'-phosphate (OMP) decarboxylase

V-ATPase vacuolar proton ATPase

VEGF vascular endothelial growth factor

WT wild-type

YCK1 yeast casein kinase I homologue

YNO1 yeast NADPH oxidase

YP yeast extract, peptone

Page 17: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

xvii

YPD yeast extract, peptone, 2% dextrose

YPDE YPD + 15 mg/L ergosterol and 0.5% Tween 80

Zn zinc

Page 18: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

1

CHAPTER 1

Introduction

Page 19: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

2

1. Introduction to Reactive Oxygen Species

1.1. Definition of Reactive Oxygen Species

Reactive oxygen species (ROS) comprise a set of molecules/compounds derived

from molecular oxygen (O2) that have increased reactivity relative to O2 in its ground

state. This includes the three species derived from the consecutive univalent reduction of

O2 prior to the formation of water: superoxide (O2• −), hydrogen peroxide (H2O2), and

hydroxyl radical (HO•). These ROS are the focus of this thesis.

ROS also include O2 molecules in which the normally unpaired electrons of parallel spin

have spin-flipped (i.e. have antiparallel spin) referred to as singlet oxygen.

1.2. The Dual Nature of ROS

For more than 40 years, it has been clear that ROS are significant damaging

agents in cells [1]. Discovery of the wide array of cellular antioxidant defenses in nearly

all aerobic organisms, as well as, a number of strict anaerobes [2] and the production of

ROS by cells of the immune system to eliminate invading pathogens [3] underscores their

destructive nature. More recently, a second role for ROS in cell signaling has been

identified. This much younger field of study began with the discovery of enzymes in

cells outside of the immune system that purposely produce ROS in response to specific

stimuli and the pathogenic states that arise when ROS production is diminished or

eliminated [4]. Although many roles for ROS in signaling have been uncovered, we still

know very little. This dual nature of ROS as both damaging and signaling agents

requires cells to exquisitely balance ROS generation and removal to ensure their required

O2

• − O2 H2O2 HO

• H2O

e-

e-

2H+

e

-

e-

2H+

Page 20: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

3

role in signaling is achieved without causing significant damage to the cell. What

remains unclear is the line that divides this dual nature.

1.3. ROS in Human Health and Disease

Numerous human diseases have been linked to altered ROS status, including

cancer, diabetes, cardiovascular disease, neurologic disorders, and Down Syndrome to

name a few. The exact contribution of ROS to these diseases is not clearly understood

but there is plenty of evidence that suggests ROS play both damaging and signaling roles.

In the development of cancer, for example, high levels of ROS can be a driving factor by

causing damage to DNA [5,6]. At the same time, increased levels of ROS may activate

growth signaling pathways and enhance the production of factors that promote metastasis

[7]. Another interesting example is insulin signaling and the development of insulin

resistance, a factor contributing to both type 2 diabetes and obesity. In this case, ROS

appears to be required for normal insulin signaling, as well as, to contribute to insulin

resistance [8]. Normally, insulin stimulates a cascade of events that produce ROS, which

then amplify the cellular response to insulin. However, elevated levels of ROS may

contribute to insulin resistance by causing mitochondrial damage and dysfunction or by

activating signaling pathways that antagonize the cellular response to insulin.

1.4. Basic Properties of ROS

Although by definition ROS are all more reactive than O2, their properties are not

exactly the same. While some conclusions can be drawn about the reactivity of

superoxide, hydrogen peroxide, and hydroxyl radical with biomolecules (addressed

generally in Section 3), the direct targets of each may be difficult to ascertain because

these three ROS are readily interconverted in the cellular environment. To establish a

Page 21: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

4

foundation the basic properties of superoxide, hydrogen peroxide, and hydroxyl radical

will be introduced.

1.4.1. Superoxide

Superoxide (O2• −) is essentially molecular oxygen with an extra electron. As

such, it is a free radical and has a negative charge. With a single-electron redox potential

of E = +0.89 V it is a much stronger oxidant than triplet oxygen (E = -0.16 V) [9].

Because of its unpaired electron, it primarily participates in single-electron transfer

reactions which limits its direct reactivity with biomolecules. Its anionic nature also

limits its reactivity with negatively charged biomolecules and can prevent it from

crossing cellular membranes [10]. The protonated form of superoxide, hydroperoxyl

radical (HO2•), is not limited in this manner but, with a pKa ≈ 4.88, less than 1% of

superoxide is protonated in cells at neutral pH [11-13].

An important property of superoxide is that it can act as both a reductant and an

oxidant. This permits the disproportionation of superoxide to hydrogen peroxide and O2.

The rate of this reaction is dependent on pH due to charge repulsion. At neutral pH the

second-order rate constant is k ≈ 105 [13-15].

1.4.2. Hydrogen Peroxide

Hydrogen peroxide (H2O2), unlike the other ROS described here, is not a free

radical. It is not as strong of a single-electron oxidant as superoxide (E = +0.38 V) but is

a very strong two-electron oxidant (E = +1.349 V) [9]. It is may also react with

biomolecules by electrophilic attack.

O2

• − O2 H2O2 O2

• −

2H+

+ +

Page 22: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

5

With a pKa ≈ 11.6, hydrogen peroxide is primarily neutral in cells [16]. It is also

the most stable of the ROS, with a half-life measured in seconds to minutes [17,18], and

is able to cross biological membranes.

1.4.3. Hydroxyl Radical

Hydroxyl radical (HO•) is the strongest oxidant of the ROS described here (E =

+2.32 V) [9]. Furthermore, its reactivity is not limited by charge repulsion or kinetic

restraints. It is, therefore, very unstable and reactive. Many of the effects attributed to

ROS are likely to be caused by hydroxyl radical, as it is readily generated from hydrogen

peroxide and redox-active metals by Fenton & Fenton-like chemistry (see Section 2.3).

2. Sources of ROS

ROS are produced in cells by enzymes, redox-active metals, and by a variety of

environmental/exogenous sources. Three significant sources of cellular ROS have been

implicated in this thesis: the mitochondria, primarily as a by-product of the electron

transport chain (ETC) [19], the NADPH oxidases, and metal-catalyzed ROS.

2.1. Mitochondrial ETC

The ETC is widely believed to be a large producer of ROS in cells. It is

comprised of four redox-active, multi-subunit complexes that collectively couple electron

transfer to H+ transport across the inner mitochondrial membrane to generate an

electrochemical gradient. The final step in the ETC is a 2-electron transfer to the ultimate

electron acceptor O2 along with protons to generate water [20,21]. All other steps in the

ETC catalyze single-electron transfers [21]. For this reason, the primary ROS produced

by the ETC is superoxide.

Page 23: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

6

Examining the redox potentials of all reactions in the ETC and comparing them

with that of O2 permits identification of the complexes with the potential for superoxide

production. The redox potential of O2 is estimated to be E = -130 mV [21,22]. Only

complexes I (E = -20 to -380 mV) and III (E = -400 to 700 mV) have redox potentials

lower than that of O2 and are able to reduce it [21]. Complexes II (E = -90 to 140 mV)

and IV (E = 210 to 815 mV) do not [21]. A large body of experimental evidence

validates complex I and III as ROS producers [23-32]. The baker’s yeast,

Saccharomyces cerevisiae, used in these studies does not have a true, multisubunit

complex I [33]. Therefore, complex III is likely to be the major ROS producer of the

ETC in yeast.

It is worth noting that the ETC, more particularly complex III, can produce

superoxide on both sides of the inner mitochondrial membrane [23-27]. Superoxide

produced in the intermembrane space (IMS) may also escape to the cytoplasm, with

voltage dependent anion channels implicated in its transport [34]. A small fraction could

also exit mitochondria in the protonated hydroperoxyl radical form. Therefore, the ROS

produced by the ETC are not strictly limited to affects in the mitochondria but can also

affect the remainder of the cell.

A variety of factors influence the quantity of ROS produced by the ETC. While

some of these are exogenous in origin (e.g. inhibitors such as rotenone or antimycin A)

[24,28,29,32,35], others are genetic (e.g. mutations in complex II) [36-39] or metabolic in

nature (e.g. NADH/NAD+ ratio, carbon source utilization) [28,30-32,35]. We examine

the effects of mitochondrial ROS under specific metabolic conditions in Chapter 2.

2.2. NADPH Oxidases

Page 24: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

7

The NADPH oxidases (NOX) are a family of multi-subunit enzymes that utilize

electrons from NADPH to produce superoxide. At their core, the catalytic subunit of all

NOX enzymes has six membrane-spanning domains, two hemes, and conserved

NADPH- and FAD-binding sites [40,41]. Some members have additional

transmembrane domains and/or calcium-binding EF-hand motifs [40]. The dual oxidases

(DUOX) also have a peroxidase-like domain and hydrogen peroxide appears to be the

primary ROS they produce [40].

NOX enzymes have solidified our acceptance of the dual nature of ROS. They

create ROS as damaging agents during innate immunity and in non-immune cells can

regulate a whole host of signaling pathways [40,42]. A NOX enzyme was also recently

identified in yeast with potential implications for this thesis (see Sections 2.2.2 and 5.3;

also Chapter 3) [43].

2.2.1. Nox2 - ROS Produced During the Mammalian Immune Response

The first NOX enzyme discovered, Nox2 (also referred to as gp91phox), was

identified in phagocytic cells where it generates the ROS responsible for destroying

invading pathogens, referred to as the “respiratory burst”. Individuals with defects in

Nox2 subunits develop chronic granulomatous disease and exhibit an inability to

eliminate certain microbial and fungal infections [41,44]. The ROS produced by Nox2

can be converted to other damaging species, such as the hypochlorous acid (HOCl; i.e.

the active agent in bleach) formed by myeloperoxidase [45-47].

2.2.2. Nox Enzymes in Non-immune Cells and Yeast

Some time after discovery of Nox2, other mammalian NOX enzymes were

identified. These enzymes are found in tissues throughout the body, including the

Page 25: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

8

vascular smooth muscle, kidneys, gastrointestinal tract, inner ear, thyroid, testis, and

uterus [40,48]. While some may be involved in immunity, these enzymes are primarily

involved in cellular signaling pathways (see Section 5.1 for an example) [40,48].

The yeast NADPH oxidase, Yno1p (also known as Aim14p), was identified by

Rinnerthaler and colleagues only recently [43]. Yno1p, formerly, was believed to be a

ferric reductase based on homology but it has no ferric reductase activity and is, in fact, a

bona fide NOX enzyme [43,49]. Though little is known about its cellular role, Yno1p

has been implicated in regulating actin filaments, cell death, and glucose and oxygen

signaling [43,50]. Its role in glucose and oxygen sensing will be addressed in Section

5.3.

2.3. Metal-catalyzed ROS

Transition metals, particularly Fe and Cu, are both essential and toxic to

eukaryotic organisms. Because they are able to redox cycle they are important producers

of cellular ROS, catalyzing the conversion of hydrogen peroxide to the significantly more

reactive hydroxyl radical by Fenton and Fenton-like chemistry [51-53]:

To prevent the formation of hydroxyl radical by these metals their availability is strictly

regulated in cells. However, as will be discussed in Section 3.1, ROS themselves can

increase bioreactive levels of Fe.

2.4. Other Enzymatic Sources

Many other enzymes in cells are known to generate ROS. These enzymes are

located throughout mitochondria, peroxisomes, the endoplasmic reticulum, and the

cytosol, and include but are certainly not limited to cytochrome p450s, various oxidases,

H2O2 + Fe2+

/Cu+

HO• + HO

− + Fe

3+/Cu

2+

Page 26: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

9

glycerol 3-phosphate dehydrogenase, pyruvate and α-ketoglutarate dehydrogenases, and

the electron-transfer flavoprotein of fatty acid beta-oxidation [29,35,54-59]. These will

not be reviewed in detail.

2.5. Environmental/Exogenous Sources

Environmental stimuli that produce ROS include ionizing radiation [60],

ultraviolet light [61], and exogenous toxicants [62,63]. These environmental sources are

pervasive and the ROS they produce are not always neutralized by cells. Two exogenous

ROS producers that are particularly relevant to this thesis include paraquat and

menadione.

Paraquat is a widely used herbicide, killing plants by accepting electrons from

photosynthetic machinery and subsequently transferring them to O2, effectively

regenerating the compound and producing superoxide [64]. In non-photosynthetic

organisms, the cellular reductant NADPH and the ETC can reduce paraquat to incite

superoxide production [63,65-73]. For this reason, mitochondrial damage due to

superoxide production appears to be the primary cause of paraquat toxicity but depletion

of cellular NADPH may also be important [65,67-69,74,75].

Menadione is a synthetic compound that has been used as a supplement based on

its vitamin K activity but is currently banned in the United States due to concerns about

toxicity. Like paraquat, menadione is a redox cycling compound that generates

superoxide [63,76-78]. However, it may also induce toxicity by forming glutathione

conjugates leading to diminished glutathione levels in cells [79-81]. For this reason,

menadione and paraquat may have different effects on yeast cells [82].

Page 27: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

10

3. Targets of ROS

ROS can react with essentially all classes of biomolecules: DNA, lipids,

carbohydrates, and proteins. The consequences of ROS production can therefore, be

extremely detrimental to life. Since the work of this thesis focuses on the effects of ROS

on proteins, this section will be restricted to protein modifications that result from ROS.

These modifications include damage to Fe-S cluster cofactors, cysteine oxidation,

methionine sulfoxide formation, and protein carbonylation.

3.1. Damage to Fe-S Clusters

Fe-S clusters are redox-active cofactors composed of two to four Fe atoms

bridged by inorganic sulfur and are usually ligated to proteins by cysteine residues. The

reaction of superoxide with Fe-S clusters occurs with a second order rate constant ≈ 106-

107 M-1 s-1 [83-86], faster than the spontaneous disproportionation rate of superoxide (≈

105 M-1 s-1). Fe-S clusters are widely regarded as the primary targets of superoxide but

they are also inactivated at appreciable rates by hydrogen peroxide (≈ 103 M-1 s-1) [83-

85,87].

A classic Fe-S cluster protein inactivated by ROS is aconitase. Aconitase (Aco1p

in yeast) functions in the tricarboxylic acid (TCA) cycle and catalyzes the reversible

isomerization of citrate to isocitrate, forming cis-aconitate as an intermediate. It is one of

a class of [4Fe-4S] enzymes especially susceptible to ROS inactivation because one of

the Fe atoms is solvent exposed [83-85,87]. ROS oxidize this Fe atom resulting in cluster

destabilization and subsequent release of Fe2+ [86].

O2

• − + 2H

+ [4Fe – 4S]

2+ [4Fe – 4S]

3+ H2O2 + +

[4Fe – 4S] 3+

[3Fe – 4S] 1+

+ Fe2+

Page 28: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

11

Though Fe-S cluster oxidation results in protein inactivation, it is reversible,

largely because Fe-S clusters can be repaired and re-assembled [88-91]. However,

release of the reduced Fe from the cluster may catalyze the formation of hydroxyl radical

by Fenton chemistry and result in significant secondary, irreversible damage to

surrounding biomolecules [92-94].

Fe-S cluster enzymes susceptible to ROS inactivation exist in both the

mitochondrial matrix and the cytoplasm and are involved in biosynthesis of various

amino acids (including Leu, Met, Lys), the TCA cycle (aconitase, succinate

dehydrogenase) and the ETC, among others [95-99]. We revisit superoxide inactivation

of the mitochondrial Fe-S cluster proteins, aconitase and succinate dehydrogenase (SDH),

in Chapter 2 of this thesis and show surprisingly, that their inactivation can actually be

beneficial in glucose starvation conditions.

3.2. Other ROS Targets

Cysteine oxidation, methionine sulfoxide formation, and carbonyl formation are

more general markers of oxidative damage that affect a large number of proteins

[11,16,100].

Protein cysteine residues can be oxidized by ROS to sulfenic acid (-SOH).

Frequently, the sulfenic acid derivative reacts with another cysteine to form a disulfide

bond [16]. These cysteine modifications are reversible and evidence is mounting that

they are important in signaling. A very important class of targets are the protein tyrosine

phosphatases, which have a catalytic cysteine that is inactivated by hydrogen peroxide

(see Section 5.1) [101]. Successive oxidation of sulfenic acid by ROS generates sulfinic

Page 29: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

12

(-SO2H) and sulfonic acid (-SO3H) residues that are generally considered irreversible

[102,103].

Methionine residues in proteins are oxidized by ROS to methionine sulfoxide

[104,105]. This modification is enzymatically reversible by NADPH-requiring

methionine sulfoxide reductases [106,107] and may both activate and inactivate enzymes

[104], indicating it may not strictly be damaging but may also play a role in signaling.

Protein carbonylation is essentially non-specific damage to proteins [11,108].

Carbonyls are formed by ROS oxidation of arginine, lysine, proline, cysteine, threonine,

leucine, and histidine side-chains [108]. Carbonyls are irreversible and are often used as

biomarkers for oxidative damage [108,109].

4. Antioxidant Enzymes

The antioxidant defenses of cells are designed to prevent the accumulation of

ROS species to damaging levels and to repair damage when it occurs. They must also

carefully balance the level of ROS to allow signaling. The antioxidant enzymes used by

cells include superoxide dismutases, catalases, peroxidases, and the general reducing

factors glutathione and thioredoxin.

4.1. Superoxide Dismutases

Superoxide dismutases (SODs) are the first line of defense. They catalyze the

disproportionation of superoxide to hydrogen peroxide and O2 using Cu, Fe, Mn, or Ni as

a catalytic metal cofactor [110,111].

O2

• − + Cu

2+/Mn

3+ Cu

+/Mn

2+ + O2

O2

• − + Cu

+/Mn

2+ + 2H

+ Cu

2+/Mn

3+ + H2O2

Page 30: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

13

Although superoxide spontaneously disproportionates at a fast rate (105 M-1 s-1 see

Section 1.4.1), SOD-catalyzed disproportionation occurs with rate constants in the 109 M-

1 s-1 range [15,112,113]. Based on homology, there are three major types of SODs

frequently referred to by the metal cofactors they use: Cu/Zn-SOD, Fe- or Mn-SOD, and

Ni-SOD [110]. Most eukaryotes have at least one largely cytosolic Cu/Zn-SOD and one

mitochondrial matrix Mn-SOD.

4.1.1. Cu/Zn-SOD

The largely cytosolic Cu/Zn-SOD (known in many organisms as SOD1) is a

homodimer where each subunit contains a Zn that serves as structural support, a catalytic

Cu, and a conserved and essential disulfide bond [110,111]. Free Cu levels in cells are

generally extremely low [114]. For this reason, most eukaryotes have a dedicated

chaperone known as CCS that delivers Cu and catalyzes disulfide bond formation in an

oxygen-dependent manner [114-119]. In baker’s yeast, this copper chaperone is

absolutely essential for activation of the Cu/Zn-SOD, known as Sod1p [120].

Cu/Zn-SOD is found primarily in the cytoplasm and mitochondrial IMS, though it

has been found in other locations throughout the cell [121-123]. In studies that have been

done in baker’s yeast and in mammals, the import of Cu/Zn-SOD1 into the IMS is

dependent on its interaction with CCS and a disulfide relay system [121,124-127]. IMS

Cu/Zn-SOD1 is important for scavenging superoxide released by the ETC on the outside

of the inner membrane [124]. Outside of the mitochondria, Cu/Zn-SOD protects

cytosolic proteins from oxidation and can also operate in cell signaling through NOX-

dependent pathways (see Section 5.1 and 5.3).

Page 31: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

14

The importance of Cu/Zn-SOD is underscored by its abundance (≈90% of the

total SOD activity [128,129]) and the phenotypes that develop in sod1 null mice and

yeast. Sod1-/- mice develop normally and survive into adulthood but exhibit, among

other things, reduced lifespan, increased incidence of liver cancer, motor neuropathy, loss

of muscle mass, and reduced female fertility [130-136].

sod1Δ null mutants of baker’s yeast exhibit a plethora of phenotypes resulting

from superoxide damage. The activities of the Fe-S cluster proteins aconitase (Aco1p),

homoaconitase (Lys4p), and isopropylmalate dehydratase (Leu1p) are all reduced, and

the Lys4p and Leu1p defects result in aerobic auxotrophies for lysine and leucine, or a

dependency on these amino acids for aerobic growth [137]. Interestingly, Lys4p and

Aco1p are both mitochondrial matrix proteins which suggests Sod1p somehow provides

cross-compartment protection against superoxide. sod1Δ null yeast also exhibit a

dependence on exogenous methionine for aerobic growth, a defect resulting from reduced

levels of NADPH needed for methionine biosynthesis [99]. The mutants are additionally

sensitive to environmental oxidants including hyperoxia and the redox-cycling compound

paraquat [99,128,138], have defects in vacuolar morphology [139], and exhibit a high

mutation rate [140].

Deletion of SOD1 also has effects on signaling in yeast. For example, Sod1p is

required for stabilization of the Ca2+-activated phosphatase calcineurin [141]. More

recently, Sod1p has been shown to play a role in regulating the switch between

respiration and fermentation [50,140]. In this case, Sod1p stabilizes a key regulator of

glucose repression, the yeast casein kinase I homologue Yck1p, in oxygen- and glucose-

Page 32: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

15

dependent manners. Details on this signaling pathway involving Sod1p and Yck1p are

presented below in Section 5.3.

4.1.2. Mn-SOD

The mitochondrial Mn-SOD (also known SOD2 in many eukaryotes) is a

homotetramer with a single Mn per subunit that localizes to the mitochondrial matrix

where it plays an important role in detoxifying superoxide produced within the

mitochondrial matrix [142,143]. Sod2-/- mice are neonatal lethal due to dilated

cardiomyopathy and possible central nervous system defects [144,145]. At the molecular

level these mice exhibit defects in the TCA cycle and ETC enzymes aconitase, SDH, and

complex I, as well as, extensive DNA damage [96].

Unlike Sod2-/- mice, the phenotypes of sod2Δ null yeast appear relatively minor.

SOD2 deletion in the baker’s yeast S. cerevisiae causes increased sensitivity to hyperoxia

and paraquat, reduced chronological lifespan, aconitase deficiency, and a hypermutation

rate but has no apparent effect on growth rate [98,146]. Interestingly, during long-term

periods of glucose starvation, sod2Δ null yeast are unable to enter a quiescent state

characterized by alkalinization of the extracellular environment [147]. This phenotype is

described in more detail in Section 6.4.2 and forms the basis of work in Chapter 2 of this

thesis.

4.2. Catalases, Peroxidases, Glutathione and Thioredoxin

Catalases and peroxidases are similar in that they both catalyze the decomposition

of hydrogen peroxide [103]. However, they differ in that catalases use heme to strictly

catalyze hydrogen peroxide decomposition to water and oxygen, whereas the peroxidase

enzymes use various cofactors to detoxify hydrogen peroxide or alkyl peroxides and to

Page 33: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

16

produce a variety of other compounds [148]. Two general classes of peroxidases that

function in ROS scavenging include the glutathione and thioredoxin peroxidases, which

use glutathione and thioredoxin as cofactors, respectively [103,149]. These two cofactors

are also the primary regulators of the reducing environment of cells.

5. ROS in Signaling

The roles identified for ROS in signaling are increasing rapidly. For the sake of

brevity we will focus on three examples involving superoxide dismutase enzymes. The

first two demonstrate how regulation of the cytoplasmic Cu/Zn-SOD1 and mitochondrial

matrix Mn-SOD2 proteins can influence ROS signaling, respectively. The third is a

specific signaling pathway involving Sod1p in yeast that is directly pertinent to Chapter 3

of this thesis.

5.1. SOD1 Regulation of Growth Factor Signaling

Stimulation of growth factor tyrosine kinase receptors has been broadly shown to

induce ROS production by NOX enzymes [4,100]. The downstream result of this ROS

production is the reversible inactivation of protein tyrosine phosphatases (PTPs) which

amplifies the phosphorylation signals induced during growth factor signaling [101].

Juarez and colleagues demonstrated in mammalian cell culture systems that the NOX-

produced superoxide during growth factor signaling is converted to hydrogen peroxide by

SOD1 [150]. This SOD1-produced peroxide can then inactivate the PTPs for a variety of

tyrosine kinase receptors including those for EGF, PDGF, FGF-2 and VEGF.

Page 34: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

17

5.2. Mitochondrial ROS Stimulation by TNF-α, an Apoptotic Signal

Tumor necrosis factor (TNF)-α is an initiating factor of programmed cell death

(PCD). It binds to the TNF receptor which results in a wide variety of downstream

signaling, eventually resulting in cell death [151]. One of the pathways activated by

TNFR stimulates ROS production by the ETC which in turn activates the c-Jun NH2-

terminal kinase initiating a positive feedback loop that enhances cell death [151,152]. In

support of the role of mitochondrial ROS in PCD, NF-kB mediated activation of the

SOD2 gene was shown to inhibit PCD [151]. The exact mechanism by which

mitochondrial ROS contribute to PCD has not been completely elucidated. In Chapter 2

of this thesis, we examine the effects of mitochondrial ROS and SOD2 in yeast under

conditions of glucose starvation and demonstrate that ROS have a beneficial effect.

5.3. ROS Regulation of Yck1p in Yeast

In the budding yeast S. cerevisiae, fermentation is preferred over respiration even

with abundant oxygen, similar to the Warburg effect of cancers and other rapidly dividing

cells. This repression of respiration is mediated in part by a signaling casein kinase at the

cell surface, Yck1p [153]. Recent work in our lab has shown that Yck1p-control of

glucose repression requires active Sod1p enzyme [50]. sod1∆ mutants respire with high

glucose and this correlates with loss of the Yck1p polypeptide. Examination of the

underlying mechanism demonstrated that Sod1p stabilizes Yck1p in an oxygen- and

glucose-dependent manner. Sod1p binds directly to Yck1p and the peroxide produced by

Sod1p protects Yck1p from degradation. Oxygen and glucose fermentation conditions

are needed to provide the superoxide substrate for Sod1p, which appears to be derived

from the aforementioned Yno1p NADPH oxidase (see Section 2.2.2) as opposed to the

Page 35: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

18

mitochondria. When Yck1p is stabilized by Sod1p and peroxide, the casein kinase

triggers downstream signaling of many events favoring fermentation over respiration,

including activation of the plasma membrane H+-ATPase, Pma1p. In fact, sod1∆ mutants

show a decrease in Pma1p activity with high glucose, consistent with the requirement for

Sod1p in stabilizing Yck1p. However, other possibilities for Sod1p control of Pma1p

could not be excluded including those involving oxidative stress protection. The

mechanism by which Sod1p maintains essential Pma1p activity is examined in more

detail in Chapter 3 of this thesis.

6. pH Homeostasis in the Baker’s Yeast S. cerevisiae

Regulating pH is essential for life. To survive and function, cells must maintain

their internal pH (pHi) within strict tolerances to ensure the proper functioning of

enzymes and maintain electrochemical gradients required for the uptake of nutrients and

energy generation. In general, this is accomplished by the combined action of buffering

and proton/metabolite transport. An excellent conceptual review of pH regulation was

presented by Walter Boron [154].

6.1. Metabolism and pH in Baker’s Yeast

The baker’s yeast Saccharomyces cerevisiae provides an attractive model system

to study pH homeostasis as it relates to metabolism. When glucose is abundant, the

organism prefers to ferment rather than respire, and fermentation results in acidification

of the extracellular environment. Conversely, when glucose has been depleted or cells

grow with non-fermentable carbon sources, the organism respires and if anything, the pH

of the extracellular environment (pHex) is elevated [155-157]. The effects of carbon

Page 36: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

19

source on pH homeostasis in yeast are a major topic of Chapters 2 and 3 of this thesis.

Thus, this section of the introduction focuses on mechanisms of pH control in baker’s

yeast.

S. cerevisiae relies on proton transport for both pHi regulation and the generation

of electrochemical gradients that drive secondary transport. Because of this, yeast cells

grow optimally in environments with low extracellular pH (pHex 3.5-6.0), though they are

able to grow in nutrient-rich media ranging from pHex 2.5-8.5 [158,159]. Since S.

cerevisiae has a strong preference for fermentation as its primary mode of energy

production, copious amounts of intracellular acid equivalents, including organic acids and

H+ are continuously generated [160]. Despite this acid production and the constant influx

of H+ driving nutrient uptake, yeast are able to maintain steady pHi ≈ 7.2 under

fermentative conditions independent from pHex between 3.0 and 7.5 [161], indicating a

significant pHi regulatory capacity. The primary machinery yeast employ to prevent pHi

acidification includes the plasma membrane H+-ATPase, Pma1p, and the vacuolar H+-

ATPase (V-ATPase). Yeast also export organic acids such as acetate generated from

fermentative metabolism to prevent their buildup in the cell [162-164]. The combined

action of Pma1p and organic acid secretion acidifies the extracellular environment, which

maintains a pH ≈ 4-5, an ideal range for yeast growth.

6.2. The Major pH Regulator of Yeast, Pma1p

Plasma membrane H+-ATPases, found throughout the fungal and plant kingdoms,

pump protons out of the cell and, therefore, serve as major regulators of pHi [165]. They

also generate an electrochemical H+ gradient that drives secondary nutrient uptake [166].

Yeast have two plasma membrane H+-ATPases designated Pma1p and Pma2p. Due to

Page 37: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

20

low expression, Pma2p is not a significant regulator of pHi [167]. Pma1p, on the other

hand, is one of the most abundant proteins in the plasma membrane [168]. Its activity

correlates with cellular growth rate and the transporter is essential for life [169,170].

Pma1p, like all plasma membrane H+-ATPases, is a P-type ATPase. This family

includes the well-known sarcoplasmic/endoplasmic reticulum Ca2+-ATPase (SERCA)

and the plasma membrane Na+/K+-ATPase. Thus, understanding Pma1p regulation and

function may shed light on other important members of this family.

Structurally Pma1p is an integral membrane protein comprised of 10 α-helical

transmembrane segments with both the N- and C-termini located in the cytoplasm (see

Fig. 1-1). The large central portion of the protein between membrane segments 4 and 5 is

located in the cytoplasm and constitutes the ATPase domain. The C-terminus of the

protein contains an α-helix that acts as an inhibitory domain by binding to the ATPase

domain, presumably through interactions involving the extended α-helix of membrane

segment 5, referred to as stalk segment 5 [171-174]. Interactions between the C-terminal

inhibitory domain and the ATPase domain are particularly important under glucose

control as described below in Section 6.4.1. The function of the cytosolic N-terminus of

Pma1p is largely unknown.

Pma1p is not only essential, it is also a major consumer of cellular ATP [175].

Therefore, this proton ATPase must fall under tight control. The PMA1 gene is

transcriptionally regulated by carbon source through the TUF1/RAP1 transcription factor

[170,176]. Post-translationally, Pma1p can be regulated by phosphorylation,

ubiquitination and protein degradation, calcium signaling, glucose sensing, and glycolysis

[50,177-182]. Pma1p activity can also be controlled at the level of trafficking through

Page 38: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

21

the secretory pathway. For example, deletion of essential subunits from the V-ATPase

results in mis-localization of Pma1p to the vacuolar membrane [183]. Numerous

conditions have been shown to enhance Pma1p activity which are at best poorly

understood. These include intracellular acidification, nitrogen starvation, weak acid

stress, and ethanol stress [184-187]. By comparison, the response of Pma1p to glucose

availability has been studied extensively and will be reviewed in Sections 6.4.1 below.

6.3. Other Transport Systems for pH Homeostasis

Along with Pma1p, a number of other ATPases contribute to cellular pH in

baker’s yeast. Foremost among these is the V-ATPase. The V-ATPase indirectly

regulates pHi by affecting Pma1p localization, as mentioned previously [183]. More

importantly, the V-ATPase directly regulates pHi by pumping H+ from the cytosol into

the vacuole where acidification is critical for vacuolar function and for providing an

electrochemical gradient for vacuolar membrane transport [160,183]. The Na+(K+)/H+

exchangers, Nhx1p (in the endosomal pathway) and Nha1p (on the plasma membrane)

are also important in pHi regulation. Deletion of NHX1 causes intracellular acidification

and defects in vesicle trafficking out of endosomal compartments [188]. Deletion of

NHA1, on the other hand, results in intracellular alkalinization [188,189].

6.4. Effect of Carbon Source on Regulation of pH Homeostasis in Yeast

As mentioned earlier, S. cerevisiae prefers fermentation over respiration to obtain

energy. Because of this, the presence of the fermentable carbon source glucose

significantly effects yeast biology, including pH homeostasis.

Page 39: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

22

6.4.1. Glucose Control of pH

When added to media, glucose induces a rapid decrease in pHi followed by an

increase in pHi to a resting pH of 7.2 [1]. The initial decrease is attributed to acid

production during glycolysis, while the subsequent increase in pHi is attributed to

activation of the proton ATPases, Pma1p and the V-ATPase. Glucose enhances the H+-

pumped out of the cytosol and into the vacuole by increasing the number of V-ATPase

complexes assembled on the vacuolar membrane [190,191]. At the same time, glucose

increases the Vmax of ATP hydrolysis and Km for ATP of the cell surface proton ATPase

Pma1p, in part by inducing phosphorylation of its C-terminal autoinhibitory domain at

residues S899, S911, and T912 (Fig. 1-1) [192-194]. S899 is phosphorylated by the

kinase PTK2 and phosphorylation at this residue appears to regulate the decrease in Km

[194,195]. Successive phosphorylation of T912 followed by S911 is responsible for the

Vmax increase [192,194]. The kinases responsible for phosphorylating these residues have

not yet been identified. Together the increased activity of Pma1p and the V-ATPase are

able to maintain pHi in this state until glucose is consumed.

As the fermentation of glucose progresses numerous organic acids are generated.

The exact acids and their quantity can vary with growth conditions but the primary

organic acids secreted are acetic and succinic acid [164,196-198]. Carbonic acid may

also accumulate as the CO2 released absorbs protons in the acidic environment. Along

with the H+ pumped out by Pma1p these organic acids lower and maintain pHex ≈ 4-5, an

optimal condition for yeast growth [164]. We examine the secretion and contribution of

both acetic and succinic acids to pHex in Chapter 2 of this thesis.

Page 40: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

23

6.4.2. Effects of Glucose Starvation on pH

When cells become starved for glucose or are grown in non-fermentable carbon

sources that promote respiration (e.g., glycerol or ethanol), pHi and pHex are dramatically

altered. The most significant difference is that yeast cannot maintain a stable pHi

independent of pHex [158]. In fact, pHi under these conditions slowly equilibrates with

pHex [158,161]. The primary cause for this is the reduced level of H+-pumping by Pma1p

and V-ATPase activity, essentially making the buffering within cells a primary factor in

regulating pHi. This buffering comprises the sum total of the buffering capacity of all the

weak acid/base pairs in a cell, including proteins, peptides, organic and inorganic

metabolites.

When extracellular glucose has largely been consumed following a long period of

fermentation, cells shift to a period of growth dependent on respiration (known as the

diauxic shift) [199]. The high levels of ethanol and organic acids such as acetate that

filled the extracellular environment during fermentation begin to enter the cell [199-201].

The uptake of acetic acid into the cell is believed to be toxic [186,196,202,203].

However, in studies reported in Chapter 2, we provide evidence that acetate can actually

be beneficial with prolonged periods of nutrient starvation.

On the other hand, if cells are continuously cultured in non-fermentable carbon

sources or in media containing very low glucose, the environment is not acidified. Instead

cells alkalinize their environment. This occurs because cells consume low levels of

organic acids in the environment (such as acetate) as a carbon source while also

producing weak bases such as ammonia and bicarbonate [155,157,204]. Media

alkalinization slows growth but prevents pHi from dropping too low [157,158,161].

Page 41: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

24

Interestingly, S. cerevisiae colonies starved for glucose over long periods

alternately acidify and alkalinize their environment (Fig. 1-2A) [204,205]. The acidic

phases are thought to correspond to cell growth and the alkaline to a resting phase. In

depth studies by Palkova and colleagues have demonstrated that the alkaline phase of

growth coincides with release of volatile ammonia which then serves as a signal to

neighboring colonies to synchronize growth phase (Fig. 1-2B) [204-207]. As

neighboring colonies synchronize, the ammonia concentrations between them

preferentially increase over that on their periphery. This predominantly limits growth at

the colony interface preventing colonies from growing into one another and has been

proposed as a strategy yeast use to prevent colony competition for nutrients.

Palkova and co-workers have investigated the mechanism of this alternating

acidification and alkalinization using a yeast genetic approach. They identified a number

of genes which when deleted appear to prevent or limit entry into the alkaline phase,

including the putative ammonia and/or acetate transporters, ATO1, ATO2, and ATO3; the

transcriptional regulator SOK2, which represses pseudohyphal growth; the cytosolic

catalase, CTT1, and of great interest to us, the mitochondrial Mn-SOD, SOD2

[147,205,208]. The mechanism by which the mitochondrial matrix Sod2p controlled

extracellular pH was not known, and is the focus of work presented in Chapter 2.

7. The Research in This Thesis

In this thesis, we examine the role of ROS in regulating pH homeostasis of the

baker’s yeast, S. cerevisiae, at two metabolic extremes: that of glucose abundance, when

Page 42: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

25

cells prefer to ferment, and that of glucose starvation, when cells obtain energy by

respiration.

The work in Chapter 2 addresses ROS regulation of pH under glucose starvation.

Specifically, it explores the mechanism by which yeast cells grown over long periods in

the absence of glucose shift from alkalinizing their surrounding environment to

producing acid and addresses why deletion of the mitochondrial Mn-SOD, SOD2, affects

this transition. We demonstrate that media acidification is initiated by ROS in the

mitochondrial matrix, which can be further enhanced by deletion of SOD2. These ROS

inactivate the Fe-S cluster enzymes, aconitase and SDH, which inhibits metabolite flow

through the TCA cycle leading to a significant build up and secretion of acetic acid by the

mitochondrial aldehyde dehydrogenase, Ald4p. This not only acidifies the extracellular

environment but provides a new carbon source for neighboring yeast cells that can

enhance growth under these conditions.

The work in Chapter 3 examines the role of cytosolic Sod1p in ensuring

activation of the plasma membrane H+-ATPase, Pma1p, during growth in abundant

glucose. During the course of these studies, we uncovered a surprising genetic

interaction between PMA1 and SOD1. Namely, a specific mutation in the C-terminal

autoinhibitory domain of Pma1p, Pma1-T912D, is lethal in sod1∆ cells, but not in SOD1+

cells. By exploring the mechanism underlying this “synthetic lethality” [209,210] we

provide evidence that Sod1p guarantees maximal activation of Pma1p by preventing

oxidative damage. This is likely in addition to its role in signaling via stabilizing the

Pma1p-regulatory kinase, Yck1p, as described previously [50].

Page 43: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

26

FIGURE 1-1. Topology model of the plasma membrane H+-ATPase, Pma1p.

A topology map of Pma1p based on the previously published map of Lecchi, et al. [174]

with zigzag lines representing α-helices. The N- and C-termini are indicated with an “N”

and “C,” respectively and the 10 transmembrane α-helices are numbered from M1 to

M10. The ATPase domain of Pma1p is indicated by the grey dashed box with the

approximate binding site for ATP designated. The auto-inhibitory C-terminal α-helix

binds to the ATPase domain in the absence of glucose, presumably through interactions

involving the extended α-helix of membrane segment 5 (indicated by grey arrow) [171-

174]. This inhibition is relieved by glucose-induced phosphorylation of residues S899,

S911, and T912 found in the auto-inhibitory α-helix. The approximate location of these

residues is specified.

Page 44: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

27

Page 45: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

28

FIGURE 1-2. Diagram of extracellular pH changes and ammonia signaling by yeast

during long-term growth in the absence of glucose.

(A) Over a period of days to weeks on nonfermentable (i.e. respiratory) carbon sources,

yeast cells alternately alkalinize (purple color) and acidify (yellow color) their

extracellular environment [204]. The alkalinization slows cell growth while acidification

enhances growth. (B) The alkaline period is marked by production of volatile ammonia

(purple arrows). If colonies are not synchronized the ammonia produced by one causes

the neighboring colonies to enter the alkaline phase as well. Once synchronized, volatile

ammonia accumulates with the highest concentration between colonies (purple cloud).

This preferentially limits growth at the colony interface over that on the periphery

preventing colony competition for nutrients. (C) The mechanism by which cells enter the

acid phase and the effect this phase has on cells was previously unknown (indicated by

question marks). This is the subject of Chapter 2 of this thesis.

Page 46: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

29

Page 47: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

30

8. Reference List

[1] H.A. Saltzman, I. Fridovich, Oxygen toxicity. Introduction to a protective enzyme:

superoxide dismutase, Circulation. 48 (1973) 921-923.

[2] A.L. Brioukhanov, R.K. Thauer, A.I. Netrusov, Catalase and Superoxide Dismutase

in the Cells of Strictly Anaerobic Microorganisms, Microbiology. 71 (2002) 281-285.

[3] B.M. Babior, Oxygen-Dependent Microbial Killing by Phagocytes, N. Engl. J. Med.

298 (1978) 659-668.

[4] D.I. Brown, K.K. Griendling, Nox proteins in signal transduction, Free Radical

Biol. Med. 47 (2009) 1239-1253.

[5] Y. Li, M.A. Trush, Reactive oxygen-dependent DNA damage resulting from the

oxidation of phenolic compounds by a copper-redox cycle mechanism, Cancer Res. 54

(1994) 1895s-1898s.

[6] M.S. Cooke, M.D. Evans, M. Dizdaroglu, J. Lunec, Oxidative DNA

damage: mechanisms, mutation, and disease. FASEB J. 17 (2003) 1195-1214.

[7] N.S. Brown, R. Bicknell, Hypoxia and oxidative stress in breast cancer. Oxidative

stress: its effects on the growth, metastatic potential and response to therapy of breast

cancer, Breast Cancer Res. 3 (2001) 323-327.

[8] T. Tiganis, Reactive oxygen species and insulin resistance: the good, the bad and the

ugly, Trends Pharmacol. Sci. 32 (2011) 82-89.

Page 48: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

31

[9] P.M. Wood, The potential diagram for oxygen at pH 7, Biochem. J. 253 (1988) 287-

289.

[10] M. Takahashi, K. Asada, Superoxide anion permeability of phospholipid membranes

and chloroplast thylakoids, Arch. Biochem. Biophys. 226 (1983) 558-566.

[11] J.A. Imlay, PATHWAYS OF OXIDATIVE DAMAGE, Annu. Rev. Microbiol. 57

(2003) 395-418.

[12] I. Fridovich, Fundamental Aspects of Reactive Oxygen Species, or What's the

Matter with Oxygen? Ann. N. Y. Acad. Sci. 893 (1999) 13-18.

[13] D. Behar, G. Czapski, J. Rabani, L.M. Dorfman, H.A. Schwarz, Acid dissociation

constant and decay kinetics of the perhydroxyl radical, J. Phys. Chem. 74 (1970) 3209-

3213.

[14] B.H.J. Bielski, A.O. Allen, Mechanism of the disproportionation of superoxide

radicals, J. Phys. Chem. 81 (1977) 1048-1050.

[15] S. Marklund, Spectrophotometric study of spontaneous disproportionation of

superoxide anion radical and sensitive direct assay for superoxide dismutase. J. Biol.

Chem. 251 (1976) 7504-7507.

[16] C.C. Winterbourn, The biological chemistry of hydrogen peroxide, Methods

Enzymol. 528 (2013) 3-25.

Page 49: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

32

[17] M.L. Barnard, S. Matalon, Mechanisms of extracellular reactive oxygen species

injury to the pulmonary microvasculature, J. Appl. Physiol. 72 (1992) 1724-1729.

[18] R. Dringen, B. Hamprecht, Involvement of glutathione peroxidase and catalase in

the disposal of exogenous hydrogen peroxide by cultured astroglial cells, Brain Res. 759

(1997) 67-75.

[19] M.D. Brand, The sites and topology of mitochondrial superoxide production, Exp.

Gerontol. 45 (2010) 466-472.

[20] D.A. Proshlyakov, M.A. Pressler, G.T. Babcock, Dioxygen activation and bond

cleavage by mixed-valence cytochrome c oxidase, Proc. Natl. Acad. Sci. U. S. A. 95

(1998) 8020-8025.

[21] F. Muller, The nature and mechanism of superoxide production by the electron

transport chain: Its relevance to aging, J. Am. Aging Assoc. 23 (2000) 227-253.

[22] J. Petlicki, d.V. van, The equilibrium between the oxidation of hydrogen peroxide by

oxygen and the dismutation of peroxyl or superoxide radicals in aqueous solutions in

contact with oxygen, J. Chem. Soc. , Faraday Trans. 94 (1998) 2763-2767.

[23] D. Han, E. Williams, E. Cadenas, Mitochondrial respiratory chain-dependent

generation of superoxide anion and its release into the intermembrane space, Biochem. J.

353 (2001) 411-416.

Page 50: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

33

[24] D. Han, R. Canali, D. Rettori, N. Kaplowitz, Effect of glutathione depletion on sites

and topology of superoxide and hydrogen peroxide production in mitochondria, Mol.

Pharmacol. 64 (2003) 1136-1144.

[25] S. Miwa, M.D. Brand, The topology of superoxide production by complex III and

glycerol 3-phosphate dehydrogenase in Drosophila mitochondria, Biochim. Biophys.

Acta. 1709 (2005) 214-219.

[26] J. St-Pierre, J.A. Buckingham, S.J. Roebuck, M.D. Brand, Topology of superoxide

production from different sites in the mitochondrial electron transport chain, J. Biol.

Chem. 277 (2002) 44784-44790.

[27] F.L. Muller, Y. Liu, H. Van Remmen, Complex III Releases Superoxide to Both

Sides of the Inner Mitochondrial Membrane, J. Biol. Chem. 279 (2004) 49064-49073.

[28] R. Hansford, B. Hogue, V. Mildaziene, Dependence of H2O2 Formation by Rat

Heart Mitochondria on Substrate Availability and Donor Age, J. Bioenerg. Biomembr. 29

(1997) 89-95.

[29] L.K. Kwong, R.S. Sohal, Substrate and Site Specificity of Hydrogen Peroxide

Generation in Mouse Mitochondria, Arch. Biochem. Biophys. 350 (1998) 118-126.

[30] Y. Liu, G. Fiskum, D. Schubert, Generation of reactive oxygen species by the

mitochondrial electron transport chain, J. Neurochem. 80 (2002) 780-787.

[31] F.L. Muller, Y. Liu, M.A. Abdul-Ghani, M.S. Lustgarten, A. Bhattacharya, Y.C.

Jang, H. Van Remmen, High rates of superoxide production in skeletal-muscle

Page 51: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

34

mitochondria respiring on both complex I- and complex II-linked substrates, Biochem. J.

409 (2008) 491-499.

[32] E.B. Tahara, F.D. Navarete, A.J. Kowaltowski, Tissue-, substrate-, and site-specific

characteristics of mitochondrial reactive oxygen species generation, Free Radic. Biol.

Med. 46 (2009) 1283-1297.

[33] C.A. Marres, S. de Vries, L.A. Grivell, Isolation and inactivation of the nuclear gene

encoding the rotenone-insensitive internal NADH: ubiquinone oxidoreductase of

mitochondria from Saccharomyces cerevisiae, Eur. J. Biochem. 195 (1991) 857-862.

[34] D. Han, F. Antunes, R. Canali, D. Rettori, E. Cadenas, Voltage-dependent Anion

Channels Control the Release of the Superoxide Anion from Mitochondria to Cytosol, J.

Biol. Chem. 278 (2003) 5557-5563.

[35] S. Miwa, J. St-Pierre, L. Partridge, M.D. Brand, Superoxide and hydrogen peroxide

production by Drosophila mitochondria, Free Radic. Biol. Med. 35 (2003) 938-948.

[36] N. Ishii, M. Fujii, P.S. Hartman, M. Tsuda, K. Yasuda, N. Senoo-Matsuda, S.

Yanase, D. Ayusawa, K. Suzuki, A mutation in succinate dehydrogenase cytochrome b

causes oxidative stress and ageing in nematodes, Nature. 394 (1998) 694-697.

[37] J. Huang, B.D. Lemire, Mutations in the C. elegans Succinate Dehydrogenase Iron–

Sulfur Subunit Promote Superoxide Generation and Premature Aging, J. Mol. Biol. 387

(2009) 559-569.

Page 52: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

35

[38] C. Bardella, P.J. Pollard, I. Tomlinson, SDH mutations in cancer,

Biochim. Biophys. Acta, Bioenerg. 1807 (2011) 1432.

[39] S.S.W. Szeto, S.N. Reinke, B.D. Sykes, B.D. Lemire, Ubiquinone-binding Site

Mutations in the Saccharomyces cerevisiae Succinate Dehydrogenase Generate

Superoxide and Lead to the Accumulation of Succinate, J. Biol. Chem. 282 (2007)

27518-27526.

[40] M. Geiszt, NADPH oxidases: new kids on the block, Cardiovasc. Res. 71 (2006)

289-299.

[41] M.T. Quinn, K.A. Gauss, Structure and regulation of the neutrophil respiratory burst

oxidase: comparison with nonphagocyte oxidases, J. Leukoc. Biol. 76 (2004) 760-781.

[42] B. Rada, T.L. Leto, Oxidative innate immune defenses by Nox/Duox family

NADPH oxidases, Contrib. Microbiol. 15 (2008) 164-187.

[43] M. Rinnerthaler, S. Buttner, P. Laun, G. Heeren, T.K. Felder, H. Klinger, M.

Weinberger, K. Stolze, T. Grousl, J. Hasek, O. Benada, I. Frydlova, A. Klocker, B.

Simon-Nobbe, B. Jansko, H. Breitenbach-Koller, T. Eisenberg, C.W. Gourlay, F. Madeo,

W.C. Burhans, M. Breitenbach, Yno1p/Aim14p, a NADPH-oxidase ortholog, controls

extramitochondrial reactive oxygen species generation, apoptosis, and actin cable

formation in yeast, Proc. Natl. Acad. Sci. U. S. A. 109 (2012) 8658-8663.

[44] P.G. Heyworth, A.R. Cross, J.T. Curnutte, Chronic granulomatous disease, Curr.

Opin. Immunol. 15 (2003) 578-584.

Page 53: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

36

[45] S.J. Klebanoff, Myeloperoxidase: friend and foe, J. Leukocyte Biol. 77 (2005) 598-

625.

[46] J.E. Harrison, J. Schultz, Studies on the chlorinating activity of myeloperoxidase. J.

Biol. Chem. 251 (1976) 1371-1374.

[47] J.P. Gaut, G.C. Yeh, H.D. Tran, J. Byun, J.P. Henderson, G.M. Richter, M. Brennan,

A.J. Lusis, A. Belaaouaj, R.S. Hotchkiss, J.W. Heinecke, Neutrophils employ the

myeloperoxidase system to generate antimicrobial brominating and chlorinating oxidants

during sepsis, Proc. Natl. Acad. Sci. U. S. A. 98 (2001) 11961-11966.

[48] J.D. Lambeth, NOX enzymes and the biology of reactive oxygen, Nat. Rev.

Immunol. 4 (2004) 181-189.

[49] E. Georgatsou, D. Alexandraki, Regulated expression of the Saccharomyces

cerevisiae Fre1p/Fre2p Fe/Cu reductase related genes, Yeast. 15 (1999) 573-584.

[50] A. Reddi, V. Culotta, SOD1 Integrates Signals from Oxygen and Glucose to Repress

Respiration, Cell. 152 (2013) 224-235.

[51] J. Prousek, Fenton chemistry in biology and medicine, Pure Appl. Chem. 79 (2007)

2325-2338.

[52] S. Goldstein, D. Meyerstein, G. Czapski, The Fenton reagents, Free Radical

Biol. Med. 15 (1993) 435-445.

Page 54: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

37

[53] B. Halliwell, J.M.C. Gutteridge, Biologically relevant metal ion-dependent hydroxyl

radical generation An update, FEBS Lett. 307 (1992) 108-112.

[54] M. Schrader, H.D. Fahimi, Mammalian peroxisomes and reactive oxygen species,

Histochem. Cell Biol. 122 (2004) 383-393.

[55] N.A. Bonekamp, A. Völkl, H.D. Fahimi, M. Schrader, Reactive oxygen species and

peroxisomes: Struggling for balance, Biofactors. 35 (2009) 346-355.

[56] S. Puntarulo, A.I. Cederbaum, Production of Reactive Oxygen Species by

Microsomes Enriched in Specific Human Cytochrome P450 Enzymes, Free Radical

Biol. Med. 24 (1998) 1324-1330.

[57] A.A. Starkov, G. Fiskum, C. Chinopoulos, B.J. Lorenzo, S.E. Browne, M.S. Patel,

M.F. Beal, Mitochondrial alpha-ketoglutarate dehydrogenase complex generates reactive

oxygen species, J. Neurosci. 24 (2004) 7779-7788.

[58] J.V. Rodrigues, C.M. Gomes, Mechanism of superoxide and hydrogen peroxide

generation by human electron-transfer flavoprotein and pathological variants, Free

Radical Biol. Med. 53 (2012) 12-19.

[59] D.F. Lewis, Oxidative stress: the role of cytochromes P450 in oxygen activation,

J. Chem. Technol. Biotechnol. 77 (2002) 1095-1100.

[60] P.A. Riley, Free Radicals in Biology: Oxidative Stress and the Effects of Ionizing

Radiation, Int. J. Radiat. Biol. 65 (1994) 27-33.

Page 55: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

38

[61] K. Scharffetter-Kochanek, M. Wlaschek, P. Brenneisen, M. Schauen, R.

Blaudschun, J. Wenk, UV-induced reactive oxygen species in photocarcinogenesis and

photoaging, Biol. Chem. 378 (1997) 1247-1257.

[62] J.L. Bolton, M.A. Trush, T.M. Penning, G. Dryhurst, T.J. Monks, Role of Quinones

in Toxicology, Chem. Res. Toxicol. 13 (2000) 135-160.

[63] H.M. Hassan, I. Fridovich, Intracellular production of superoxide radical and of

hydrogen peroxide by redox active compounds, Arch. Biochem. Biophys. 196 (1979)

385-395.

[64] T. Fujii, E. Yokoyama, K. Inoue, H. Sakurai, The sites of electron donation of

Photosystem I to methyl viologen, Biochim. Biophys. Acta, Bioenerg. 1015 (1990) 41-

48.

[65] J.S. Bus, J.E. Gibson, Paraquat: model for oxidant-initiated toxicity, Environ. Health

Perspect. 55 (1984) 37-46.

[66] M. Comporti, Three models of free radical-induced cell injury, Chem. Biol. Interact.

72 (1989) 1-56.

[67] P.R. Castello, D.A. Drechsel, M. Patel, Mitochondria Are a Major Source of

Paraquat-induced Reactive Oxygen Species Production in the Brain, J. Biol. Chem. 282

(2007) 14186-14193.

[68] M. Blaszczynski, J. Litwinska, D. Zaborowska, T. Bilinski, The role of respiratory

chain in paraquat toxicity in yeast, Acta Microbiol. Pol. 34 (1985) 243-254.

Page 56: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

39

[69] S.I. Liochev, I. Fridovich, Paraquat diaphorases in Escherichia coli, Free Radical

Biol. Med. 16 (1994) 555-559.

[70] H.M. Hassan, I. Fridovich, Paraquat and Escherichia coli. Mechanism of production

of extracellular superoxide radical, J. Biol. Chem. 254 (1979) 10846-10852.

[71] M.A. Trush, E.G. Mimnaugh, E. Ginsburg, T.E. Gram, In vitro stimulation by

paraquat of reactive oxygen-mediated lipid peroxidation in rat lung microsomes, Toxicol.

Appl. Pharmacol. 60 (1981) 279-286.

[72] H.M. Cocheme, M.P. Murphy, Complex I is the major site of mitochondrial

superoxide production by paraquat, J. Biol. Chem. 283 (2008) 1786-1798.

[73] H.M. Cochemé, M.P. Murphy, Chapter 22 The Uptake and Interactions of the Redox

Cycler Paraquat with Mitochondria, in: William S. Allison and Immo E. Scheffler (Ed.),

Methods in Enzymology, Academic Press, 2009, pp. 395-417.

[74] J.A.F. Vicente, F. Peixoto, M.L. Lopes, V.M.C. Madeira, Differential sensitivities of

plant and animal mitochondria to the herbicide paraquat, J. Biochem. Mol. Toxicol. 15

(2001) 322-330.

[75] C.M. Palmeira, A.J. Moreno, V.M.C. Madeira, Mitochondrial bioenergetics is

affected by the herbicide paraquat, Biochim. Biophys. Acta, Bioenerg. 1229 (1995) 187-

192.

[76] Y. Nosoh, J. Kajioka, M. Itoh, Effect of menadione on the electron transport

pathway of yeast mitochondria, Arch. Biochem. Biophys. 127 (1968) 1-6.

Page 57: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

40

[77] D.J. Jamieson, S.L. Rivers, D.W.S. Stephen, Analysis of Saccharomyces cerevisiae

proteins induced by peroxide and superoxide stress, Microbiology. 140 (1994) 3277-

3283.

[78] M. Chaput, J. Brygier, Y. Lion, A. Sels, Potentiation of oxygen toxicity by

menadione in Saccharomyces cerevisiae, Biochimie. 65 (1983) 501-512.

[79] W.C. McAmis, R.C. Schaeffer Jr., J.W. Baynes, M.B. Wolf, Menadione causes

endothelial barrier failure by a direct effect on intracellular thiols, independent of reactive

oxidant production, Biochim. Biophys. Acta, Mol. Cell Res. 1641 (2003) 43-53.

[80] R. Z?dzinski, A. Fortuniak, G. Bartosz, T. Bilinski, M. Grey, Menadione toxicity in

Saccharomyces cerevisiae cells: Activation by conjugation with glutathione, IUBMB

Life. 44 (1998) 747-759.

[81] J. Mauzeroll, A.J. Bard, Scanning electrochemical microscopy of menadione-

glutathione conjugate export from yeast cells, Proc. Natl. Acad. Sci. U. S. A. 101 (2004)

7862-7867.

[82] M.A. Wallace, S. Bailey, J.M. Fukuto, J.S. Valentine, E.B. Gralla, Induction of

Phenotypes Resembling CuZn-Superoxide Dismutase Deletion in Wild-Type Yeast Cells:

An in Vivo Assay for the Role of Superoxide in the Toxicity of Redox-Cycling

Compounds, Chem. Res. Toxicol. 18 (2005) 1279-1286.

[83] D.H. Flint, M.H. Emptage, Dihydroxy-Acid Dehydratase: Isolation, Characterization

as Fe-S Proteins, and Sensitivity to Inactivation by Oxygen Radicals, , 1990.

Page 58: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

41

[84] P.R. Gardner, I. Fridovich, Superoxide sensitivity of the Escherichia coli aconitase.

J. Biol. Chem. 266 (1991) 19328-19333.

[85] S.I. Liochev, I. Fridovich, Modulation of the Fumarases of Escherichia coli in

Response to Oxidative Stress, Arch. Biochem. Biophys. 301 (1993) 379-384.

[86] D.H. Flint, J.F. Tuminello, M.H. Emptage, The inactivation of Fe-S cluster

containing hydro-lyases by superoxide, J. Biol. Chem. 268 (1993) 22369-22376.

[87] D.H. Flint, J.F. Tuminello, M.H. Emptage, The inactivation of Fe-S cluster

containing hydro-lyases by superoxide. J. Biol. Chem. 268 (1993) 22369-22376.

[88] P.R. Gardner, Superoxide-driven aconitase FE-S center cycling, Biosci. Rep. 17

(1997) 33-42.

[89] P.R. Gardner, I. Fridovich, Inactivation-reactivation of aconitase in Escherichia coli.

A sensitive measure of superoxide radical. J. Biol. Chem. 267 (1992) 8757-8763.

[90] R. Lill, R. Dutkiewicz, H. Elsässer, A. Hausmann, D.J.A. Netz, A.J. Pierik, O.

Stehling, E. Urzica, U. Mühlenhoff, Mechanisms of iron–sulfur protein maturation in

mitochondria, cytosol and nucleus of eukaryotes, Biochim. Biophys. Acta, Mol. Cell

Res. 1763 (2006) 652-667.

[91] O. Djaman, F.W. Outten, J.A. Imlay, Repair of Oxidized Iron-Sulfur Clusters in

Escherichia coli, J. Biol. Chem. 279 (2004) 44590-44599.

Page 59: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

42

[92] S.I. Liochev, I. Fridovich, The role of O2.− in the production of HO.: in vitro and in

vivo, Free Radical Biol. Med. 16 (1994) 29-33.

[93] L. Benov, How superoxide radical damages the cell, Protoplasma. 217 (2001) 33-36.

[94] K. Keyer, J. Imlay, Superoxide accelerates DNA damage by elevating free-

iron levels, Proc. Natl. Acad. Sci. U. S. A. 93 (1996) 13635-13640.

[95] C.S. Powell, R.M. Jackson, Mitochondrial complex I, aconitase, and succinate

dehydrogenase during hypoxia-reoxygenation: modulation of enzyme activities by

MnSOD, Am. J. Physiol. - Lung Cell. Mol. Physiol. 285 (2003) L189-L198.

[96] S. Melov, P. Coskun, M. Patel, R. Tuinstra, B. Cottrell, A.S. Jun, T.H. Zastawny, M.

Dizdaroglu, S.I. Goodman, T. Huang, H. Miziorko, C.J. Epstein, D.C. Wallace,

Mitochondrial disease in superoxide dismutase 2 mutant mice,

Proc. Natl. Acad. Sci. U. S. A. 96 (1999) 846-851.

[97] Y. Zhang, O. Marcillat, C. Giulivi, L. Ernster, K.J. Davies, The oxidative

inactivation of mitochondrial electron transport chain components and ATPase. J. Biol.

Chem. 265 (1990) 16330-16336.

[98] M.A. Wallace, L.L. Liou, J. Martins, M.H. Clement, S. Bailey, V.D. Longo, J.S.

Valentine, E.B. Gralla, Superoxide inhibits 4Fe-4S cluster enzymes involved in amino

acid biosynthesis. Cross-compartment protection by CuZn-superoxide dismutase, J. Biol.

Chem. 279 (2004) 32055-32062.

Page 60: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

43

[99] K.H. Slekar, D.J. Kosman, V.C. Culotta, The Yeast Copper/Zinc Superoxide

Dismutase and the Pentose Phosphate Pathway Play Overlapping Roles in Oxidative

Stress Protection, J. Biol. Chem. 271 (1996) 28831-28836.

[100] V.J. Thannickal, B.L. Fanburg, Reactive oxygen species in cell signaling,

Am. J. Physiol. : Lung Cell. Mol. Physiol. 279 (2000) L1005-L1028.

[101] A. Salmeen, D. Barford, Functions and mechanisms of redox regulation of

cysteine-based phosphatases, Antioxid. Redox Signal. 7 (2005) 560-577.

[102] H.A. Woo, W. Jeong, T. Chang, K.J. Park, S.J. Park, J.S. Yang, S.G. Rhee,

Reduction of Cysteine Sulfinic Acid by Sulfiredoxin Is Specific to 2-Cys Peroxiredoxins,

J. Biol. Chem. 280 (2005) 3125-3128.

[103] S.G. Rhee, K.S. Yang, S.W. Kang, H.A. Woo, T.S. Chang, Controlled elimination

of intracellular H(2)O(2): regulation of peroxiredoxin, catalase, and glutathione

peroxidase via post-translational modification, Antioxid. Redox Signal. 7 (2005) 619-

626.

[104] W. Vogt, Oxidation of methionyl residues in proteins: Tools, targets, and reversal,

Free Radical Biol. Med. 18 (1995) 93-105.

[105] N. Brot, H. Weissbach, Biochemistry and physiological role of methionine

sulfoxide residues in proteins, Arch. Biochem. Biophys. 223 (1983) 271-281.

Page 61: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

44

[106] J. Moskovitz, B.S. Berlett, J.M. Poston, E.R. Stadtman, The yeast peptide-

methionine sulfoxide reductase functions as an antioxidant in vivo,

Proc. Natl. Acad. Sci. U. S. A. 94 (1997) 9585-9589.

[107] H. Kim, V.N. Gladyshev, Methionine Sulfoxide Reduction in Mammals:

Characterization of Methionine-R-Sulfoxide Reductases, Mol. Biol. Cell. 15 (2004)

1055-1064.

[108] E.R. Stadtman, R.L. Levine, Free radical-mediated oxidation of free amino acids

and amino acid residues in proteins, Amino Acids. 25 (2003) 207-218.

[109] I. Dalle-Donne, R. Rossi, D. Giustarini, A. Milzani, R. Colombo, Protein carbonyl

groups as biomarkers of oxidative stress, Clinica Chimica Acta. 329 (2003) 23-38.

[110] V.C. Culotta, M. Yang, T.V. O'Halloran, Activation of superoxide dismutases:

putting the metal to the pedal, Biochim. Biophys. Acta. 1763 (2006) 747-758.

[111] J.M. McCord, I. Fridovich, Superoxide

Dismutase: An enzymic function for erythrocuprein (hemocuprein). J. Biol. Chem. 244

(1969) 6049-6055.

[112] E.M. Fielden, P.B. Roberts, R.C. Bray, D.J. Lowe, G.N. Mautner, G. Rotilio, L.

Calabrese, Mechanism of action of superoxide dismutase from pulse radiolysis and

electron paramagnetic resonance. Evidence that only half the active sites function in

catalysis, Biochem. J. 139 (1974) 49-60.

Page 62: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

45

[113] H.J. Forman, I. Fridovich, Superoxide dismutase: a comparison of rate constants,

Arch. Biochem. Biophys. 158 (1973) 396-400.

[114] T.D. Rae, P.J. Schmidt, R.A. Pufahl, V.C. Culotta, T.V. O'Halloran, Undetectable

intracellular free copper: the requirement of a copper chaperone for superoxide

dismutase, Science. 284 (1999) 805-808.

[115] P.C. Wong, D. Waggoner, J.R. Subramaniam, L. Tessarollo, T.B. Bartnikas, V.C.

Culotta, D.L. Price, J. Rothstein, J.D. Gitlin, Copper chaperone for superoxide dismutase

is essential to activate mammalian Cu/Zn superoxide dismutase, Proc. Natl. Acad. Sci. U.

S. A. 97 (2000) 2886-2891.

[116] J.M. Leitch, L.T. Jensen, S.D. Bouldin, C.E. Outten, P.J. Hart, V.C. Culotta,

Activation of Cu,Zn-superoxide dismutase in the absence of oxygen and the copper

chaperone CCS, J. Biol. Chem. 284 (2009) 21863-21871.

[117] M.C. Carroll, C.E. Outten, J.B. Proescher, L. Rosenfeld, W.H. Watson, L.J.

Whitson, P.J. Hart, L.T. Jensen, V.C. Culotta, The Effects of Glutaredoxin and Copper

Activation Pathways on the Disulfide and Stability of Cu,Zn Superoxide Dismutase, J.

Biol. Chem. 281 (2006) 28648-28656.

[118] A.L. Lamb, A.S. Torres, T.V. O'Halloran, A.C. Rosenzweig, Heterodimer

Formation between Superoxide Dismutase and Its Copper Chaperone†, Biochemistry

(N. Y. ). 39 (2000) 14720-14727.

Page 63: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

46

[119] N.M. Brown, A.S. Torres, P.E. Doan, T.V. O'Halloran, Oxygen and the copper

chaperone CCS regulate posttranslational activation of Cu,Zn superoxide dismutase,

Proc. Natl. Acad. Sci. U. S. A. 101 (2004) 5518-5523.

[120] V.C. Culotta, L.W. Klomp, J. Strain, R.L. Casareno, B. Krems, J.D. Gitlin, The

copper chaperone for superoxide dismutase, J. Biol. Chem. 272 (1997) 23469-23472.

[121] R.A. Weisiger, I. Fridovich, Mitochondrial Superoxide Dismutase: SITE OF

SYNTHESIS AND INTRAMITOCHONDRIAL LOCALIZATION, J. Biol. Chem. 248

(1973) 4793-4796.

[122] J. Lindenau, H. Noack, H. Possel, K. Asayama, G. Wolf, Cellular distribution of

superoxide dismutases in the rat CNS, Glia. 29 (2000) 25-34.

[123] L.Y. Chang, J.W. Slot, H.J. Geuze, J.D. Crapo, Molecular immunocytochemistry of

the CuZn superoxide dismutase in rat hepatocytes. J. Cell Biol. 107 (1988) 2169-2179.

[124] L.A. Sturtz, K. Diekert, L.T. Jensen, R. Lill, V.C. Culotta, A fraction of yeast

Cu,Zn-superoxide dismutase and its metallochaperone, CCS, localize to the

intermembrane space of mitochondria—A physiological role for SOD1 in guarding

against mitochondrial oxidative damage, J. Biol. Chem. 276 (2001) 38084-38089.

[125] L.S. Field, Y. Furukawa, T.V. O'Halloran, V.C. Culotta, Factors controlling the

uptake of yeast copper/zinc superoxide dismutase into mitochondria, J. Biol. Chem. 278

(2003) 28052-28059.

Page 64: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

47

[126] A. Okado-Matsumoto, I. Fridovich, Amyotrophic lateral sclerosis: A proposed

mechanism, Proc. Natl. Acad. Sci. U. S. A. 99 (2002) 9010-9014.

[127] A. Varabyova, U. Topf, P. Kwiatkowska, L. Wrobel, M. Kaus-Drobek, A.

Chacinska, Mia40 and MINOS act in parallel with Ccs1 in the biogenesis of

mitochondrial Sod1, FEBS Journal. (2013) n/a-n/a.

[128] E.B. Gralla, D.J. Kosman, Molecular Genetics of Superoxide Dismutases in Yeasts

and Related Fungi, in: J.G. Scandalios, T.R.F. Wright (Eds.), Advances in Genetics,

Academic Press, 1992, pp. 251-319.

[129] A.R. Reddi, V.C. Culotta, Regulation of manganese antioxidants by nutrient

sensing pathways in Saccharomyces cerevisiae, Genetics. 189 (2011) 1261-1270.

[130] A.G. Reaume, J.L. Elliott, E.K. Hoffman, N.W. Kowall, R.J. Ferrante, D.F. Siwek,

H.M. Wilcox, D.G. Flood, M.F. Beal, R.H. Brown Jr, R.W. Scott, W.D. Snider, Motor

neurons in Cu/Zn superoxide dismutase-deficient mice develop normally but exhibit

enhanced cell death after axonal injury, Nat. Genet. 13 (1996) 43-47.

[131] S. Elchuri, T.D. Oberley, W. Qi, R.S. Eisenstein, L. Jackson Roberts, H. Van

Remmen, C.J. Epstein, T.T. Huang, CuZnSOD deficiency leads to persistent and

widespread oxidative damage and hepatocarcinogenesis later in life, Oncogene. 24 (2005)

367-380.

Page 65: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

48

[132] D.G. Flood, A.G. Reaume, J.A. Gruner, E.K. Hoffman, J.D. Hirsch, Y. Lin, K.S.

Dorfman, R.W. Scott, Hindlimb Motor Neurons Require Cu/Zn Superoxide Dismutase

for Maintenance of Neuromuscular Junctions, Am. J. Pathol. 155 (1999) 663-672.

[133] F.L. Muller, W. Song, Y. Liu, A. Chaudhuri, S. Pieke-Dahl, R. Strong, T. Huang,

C.J. Epstein, L.J. Roberts II, M. Csete, J.A. Faulkner, H. Van Remmen, Absence of CuZn

superoxide dismutase leads to elevated oxidative stress and acceleration of age-dependent

skeletal muscle atrophy, Free Radical Biol. Med. 40 (2006) 1993-2004.

[134] L.R. Fischer, A. Igoudjil, J. Magrané, Y. Li, J.M. Hansen, G. Manfredi, J.D. Glass,

SOD1 targeted to the mitochondrial intermembrane space prevents motor neuropathy in

the Sod1 knockout mouse, Brain. 134 (2011) 196-209.

[135] M.M. Matzuk, L. Dionne, Q. Guo, T.R. Kumar, R.M. Lebovitz, Ovarian Function

in Superoxide Dismutase 1 and 2 Knockout Mice, Endocrinology. 139 (1998) 4008-4011.

[136] Y. Ho, M. Gargano, J. Cao, R.T. Bronson, I. Heimler, R.J. Hutz, Reduced Fertility

in Female Mice Lacking Copper-Zinc Superoxide Dismutase, J. Biol. Chem. 273 (1998)

7765-7769.

[137] M.A. Wallace, L. Liou, J. Martins, M.H.S. Clement, S. Bailey, V.D. Longo, J.S.

Valentine, E.B. Gralla, Superoxide Inhibits 4Fe-4S Cluster Enzymes Involved in Amino

Acid Biosynthesis. Cross-compartment protection by CuZn-superoxide dismutase, J.

Biol. Chem. 279 (2004) 32055-32062.

Page 66: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

49

[138] C.E. Outten, R.L. Falk, V.C. Culotta, Cellular factors required for protection from

hyperoxia toxicity in Saccharomyces cerevisiae, Biochem. J. 388 (2005) 93-101.

[139] L.B. Corson, J. Folmer, J.J. Strain, V.C. Culotta, D.W. Cleveland, Oxidative stress

and iron are implicated in fragmenting vacuoles of Saccharomyces cerevisiae lacking

Cu,Zn-superoxide dismutase, J. Biol. Chem. 274 (1999) 27590-27596.

[140] E.B. Gralla, J.S. Valentine, Null mutants of Saccharomyces cerevisiae Cu,Zn

superoxide dismutase: characterization and spontaneous mutation rates, J. Bacteriol. 173

(1991) 5918-5920.

[141] X. Wang, V.C. Culotta, C.B. Klee, Superoxide dismutase protects calcineurin from

inactivation, Nature. 383 (1996) 434-437.

[142] E. Luk, M. Yang, L.T. Jensen, Y. Bourbonnais, V.C. Culotta, Manganese

Activation of Superoxide Dismutase 2 in the Mitochondria of Saccharomyces cerevisiae,

J. Biol. Chem. 280 (2005) 22715-22720.

[143] E. Luk, M. Carroll, M. Baker, V.C. Culotta, Manganese activation of superoxide

dismutase 2 in Saccharomyces cerevisiae requires MTM1, a member of the mitochondrial

carrier family, Proc. Natl. Acad. Sci. U. S. A. 100 (2003) 10353-10357.

[144] Y. Li, T.T. Huang, E.J. Carlson, S. Melov, P.C. Ursell, J.L. Olson, L.J. Noble, M.P.

Yoshimura, C. Berger, P.H. Chan, D.C. Wallace, C.J. Epstein, Dilated cardiomyopathy

and neonatal lethality in mutant mice lacking manganese superoxide dismutase, Nat.

Genet. 11 (1995) 376-381.

Page 67: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

50

[145] R.M. Lebovitz, H. Zhang, H. Vogel, J. Cartwright, L. Dionne, N. Lu, S. Huang,

M.M. Matzuk, Neurodegeneration, myocardial injury, and perinatal death in

mitochondrial superoxide dismutase-deficient mice, Proc. Natl. Acad. Sci. U. S. A. 93

(1996) 9782-9787.

[146] X.F. Liu, I. Elashvili, E.B. Gralla, J.S. Valentine, P. Lapinskas, V.C. Culotta, Yeast

lacking superoxide dismutase. Isolation of genetic suppressors, J. Biol. Chem. 267

(1992) 18298-18302.

[147] M. Čáp, L. Váchová, Z. Palková, Yeast Colony Survival Depends on Metabolic

Adaptation and Cell Differentiation Rather Than on Stress Defense, J. Biol. Chem. 284

(2009) 32572-32581.

[148] L. Flohe, F. Ursini, Peroxidase: a term of many meanings, Antioxid. Redox Signal.

10 (2008) 1485-1490.

[149] M.B. Toledano, A. Delaunay-Moisan, C.E. Outten, A. Igbaria, Functions and

cellular compartmentation of the thioredoxin and glutathione pathways in yeast,

Antioxid. Redox Signal. 18 (2013) 1699-1711.

[150] J.C. Juarez, M. Manuia, M.E. Burnett, O. Betancourt, B. Boivin, D.E. Shaw, N.K.

Tonks, A.P. Mazar, F. Doñate, Superoxide dismutase 1 (SOD1) is essential for H2O2-

mediated oxidation and inactivation of phosphatases in growth factor signaling,

Proc. Natl. Acad. Sci. U. S. A. 105 (2008) 7147-7152.

Page 68: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

51

[151] H. Shen, S. Pervaiz, TNF receptor superfamily-induced cell death: redox-dependent

execution, FASEB J. 20 (2006) 1589-1598.

[152] J.J. Ventura, P. Cogswell, R.A. Flavell, A.S. Baldwin Jr, R.J. Davis, JNK

potentiates TNF-stimulated necrosis by increasing the production of cytotoxic reactive

oxygen species, Genes Dev. 18 (2004) 2905-2915.

[153] H. Moriya, M. Johnston, Glucose sensing and signaling in Saccharomyces

cerevisiae through the Rgt2 glucose sensor and casein kinase I,

Proc. Natl. Acad. Sci. U. S. A. 101 (2004) 1572-1577.

[154] W.F. Boron, Regulation of intracellular pH, Adv. Physiol. Ed. 28 (2004) 160-179.

[155] E. Bogonez, A. Machado, J. Satrustegui, Ammonia accumulation in acetate-

growing yeast, Biochim. Biophys. Acta. 733 (1983) 234-241.

[156] K. Ohkuni, M. Hayashi, I. Yamashita, Bicarbonate-mediated social communication

stimulates meiosis and sporulation of Saccharomyces cerevisiae, Yeast. 14 (1998) 623-

631.

[157] M. Hayashi, K. Ohkuni, I. Yamashita, Control of division arrest and entry into

meiosis by extracellular alkalisation in Saccharomyces cerevisiae, Yeast. 14 (1998) 905-

913.

[158] R. Orij, S. Brul, G.J. Smits, Intracellular pH is a tightly controlled signal in yeast,

Biochim. Biophys. Acta, Gen. Subj. 1810 (2011) 933.

Page 69: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

52

[159] T.M. Matthews, C. Webb, Culture Systems, in: M. Tuite, S. Oliver (Eds.), Springer

US, 1991, pp. 249-282.

[160] E. Bignell, Chapter 1 - The Molecular Basis of pH Sensing, Signaling, and

Homeostasis in Fungi, in: Anonymous Advances in Applied Microbiology, Academic

Press, 2012, pp. 1-18.

[161] R. Orij, J. Postmus, A. Ter Beek, S. Brul, G.J. Smits, In vivo measurement of

cytosolic and mitochondrial pH using a pH-sensitive GFP derivative in Saccharomyces

cerevisiae reveals a relation between intracellular pH and growth, Microbiology. 155

(2009) 268-278.

[162] K. Sigler, A. Knotková, J. Páca, M. Wurst, Extrusion of metabolites from baker’s

yeast during glucose-induced acidification, Folia Microbiol. 25 (1980) 311-317.

[163] K. Sigler, A. Knotková, A. Kotyk, Factors governing substrate-induced generation

and extrusion of protons in the yeast Saccharomyces cerevisiae,

Biochim. Biophys. Acta, Biomembr. 643 (1981) 572-582.

[164] K. Sigler, A. Kotyk, A. Knotková, M. Opekarová, Processes involved in the

creation of buffering capacity and in substrate-induced proton extrusion in the yeast

Saccharomyces cerevisiae, Biochim. Biophys. Acta, Biomembr. 643 (1981) 583-592.

[165] M.G. Palmgren, K.B. Axelsen, Evolution of P-type ATPases,

Biochim. Biophys. Acta, Bioenerg. 1365 (1998) 37-45.

Page 70: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

53

[166] F. Portillo, Regulation of plasma membrane H+-ATPase in fungi and plants,

Biochim. Biophys. Acta, Rev. Biomembr. 1469 (2000) 31-42.

[167] A.R. Fernandes, I. Sá-Correia, Transcription patterns of PMA1 and PMA2 genes

and activity of plasma membrane H+-ATPase in Saccharomyces cerevisiae during

diauxic growth and stationary phase, Yeast. 20 (2003) 207-219.

[168] P. Morsomme, C.W. Slayman, A. Goffeau, Mutagenic study of the structure,

function and biogenesis of the yeast plasma membrane H+-ATPase,

Biochim. Biophys. Acta, Rev. Biomembr. 1469 (2000) 133-157.

[169] F. Portillo, R. Serrano, Growth control strength and active site of yeast plasma

membrane ATPase studied by site-directed mutagenesis, Eur. J. Biochem. 186 (1989)

501-507.

[170] R. Rao, D. Drummond-Barbosa, C.W. Slayman, Transcriptional regulation by

glucose of the yeast PMA1 gene encoding the plasma membrane H(+)-ATPase, Yeast. 9

(1993) 1075-1084.

[171] M. Miranda, K.E. Allen, J.P. Pardo, C.W. Slayman, Stalk segment 5 of the yeast

plasma membrane H+-ATPase: mutational evidence for a role in glucose regulation,

J. Biol. Chem. 276 (2001) 22485-22490.

[172] F. Portillo, I.F. de Larrinoa, R. Serrano, Deletion analysis of yeast plasma

membrane H+ -ATPase and identification of a regulatory domain at the carboxyl-

terminus, FEBS Lett. 247 (1989) 381-385.

Page 71: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

54

[173] R. Serrano, F. Portillo, Catalytic and regulatory sites of yeast plasma membrane

H+-ATPase studied by directed mutagenesis, Biochim. Biophys. Acta, Bioenerg. 1018

(1990) 195-199.

[174] S. Lecchi, K.E. Allen, J.P. Pardo, A.B. Mason, C.W. Slayman, Conformational

changes of yeast plasma membrane H(+)-ATPase during activation by glucose: role of

threonine-912 in the carboxy-terminal tail, Biochemistry. 44 (2005) 16624-16632.

[175] R. Rao, C.W. Slayman, Plasma membrane and related ATPases. in: R. Brambl, G.

Marzluf (Eds.), The Mycota, Springer-Verlag, Berlin, Germany, 1996, pp. 29-56.

[176] E. Capieaux, M.L. Vignais, A. Sentenac, A. Goffeau, The yeast H+-ATPase gene is

controlled by the promoter binding factor TUF. J. Biol. Chem. 264 (1989) 7437-7446.

[177] M.A.A. Souza, M.J. Trópia, R.L. Brandão, New aspects of the glucose activation of

the H+-ATPase in the yeast Saccharomyces cerevisiae, Microbiology. 147 (2001) 2849-

2855.

[178] J.L. Withee, R. Sen, M.S. Cyert, Ion Tolerance of Saccharomyces cerevisiae

Lacking the Ca2+/CaM-Dependent Phosphatase (Calcineurin) Is Improved by Mutations

in URE2 or PMA1, Genetics. 149 (1998) 865-878.

[179] N. de la Fuente, A.M. Maldonado, F. Portillo, Glucose activation of the yeast

plasma membrane H+-ATPase requires the ubiquitin–proteasome proteolytic pathway,

FEBS Lett. 411 (1997) 308-312.

Page 72: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

55

[180] J. Kolarov, J. Kulpa, M. Baijot, A. Goffeau, Characterization of a protein serine

kinase from yeast plasma membrane. J. Biol. Chem. 263 (1988) 10613-10619.

[181] E. Estrada, P. Agostinis, J.R. Vandenheede, J. Goris, W. Merlevede, J. François, A.

Goffeau, M. Ghislain, Phosphorylation of Yeast Plasma Membrane H+-ATPase by

Casein Kinase I, J. Biol. Chem. 271 (1996) 32064-32072.

[182] A. Goossens, N. de la Fuente, J. Forment, R. Serrano, F. Portillo, Regulation of

Yeast H+-ATPase by Protein Kinases Belonging to a Family Dedicated to Activation of

Plasma Membrane Transporters, Mol. Cell. Biol. 20 (2000) 7654-7661.

[183] G.A. Martínez-Muñoz, P. Kane, Vacuolar and Plasma Membrane Proton Pumps

Collaborate to Achieve Cytosolic pH Homeostasis in Yeast, J. Biol. Chem. 283 (2008)

20309-20319.

[184] B. Benito, F. Portillo, R. Lagunas, In vivo activation of the yeast plasma membrane

ATPase during nitrogen starvation Identification of the regulatory domain that controls

activation, FEBS Lett. 300 (1992) 271-274.

[185] G.A. Monteiro, I. Sá-Correia, In vivo activation of yeast plasma membrane H+-

ATPase by ethanol: effect on the kinetic parameters and involvement of the carboxyl-

terminus regulatory domain, Biochim. Biophys. Acta, Biomembr. 1370 (1998) 310-316.

[186] M. Stratford, G. Nebe-von-Caron, H. Steels, M. Novodvorska, J. Ueckert, D.B.

Archer, Weak-acid preservatives: pH and proton movements in the yeast Saccharomyces

cerevisiae, Int. J. Food Microbiol. 161 (2013) 164-171.

Page 73: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

56

[187] R. Serrano, In vivo glucose activation of the yeast plasma membrane ATPase,

FEBS Lett. 156 (1983) 11-14.

[188] C.L. Brett, D.N. Tukaye, S. Mukherjee, R. Rao, The Yeast Endosomal

Na+(K+)/H+ Exchanger Nhx1 Regulates Cellular pH to Control Vesicle Trafficking,

Mol. Biol. Cell. 16 (2005) 1396-1405.

[189] H. Sychrova, J. Ramirez, A. Pena, Involvement of Nha1 antiporter in regulation of

intracellular pH in Saccharomyces cerevisiae, FEMS Microbiol. Lett. 171 (1999) 167-

172.

[190] P.M. Kane, Disassembly and Reassembly of the Yeast Vacuolar H+-ATPase in

Vivo, J. Biol. Chem. 270 (1995) 17025-17032.

[191] K.J. Parra, P.M. Kane, Reversible Association between the V1and V0 Domains of

Yeast Vacuolar H+-ATPase Is an Unconventional Glucose-Induced Effect,

Mol. Cell. Biol. 18 (1998) 7064-7074.

[192] S. Lecchi, C.J. Nelson, K.E. Allen, D.L. Swaney, K.L. Thompson, J.J. Coon, M.R.

Sussman, C.W. Slayman, Tandem phosphorylation of Ser-911 and Thr-912 at the C

terminus of yeast plasma membrane H+-ATPase leads to glucose-dependent activation,

J. Biol. Chem. 282 (2007) 35471-35481.

[193] F. Portillo, P. Eraso, R. Serrano, Analysis of the regulatory domain of yeast plasma

membrane H+-ATPase by directed mutagenesis and intragenic suppression, FEBS Lett.

287 (1991) 71-74.

Page 74: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

57

[194] P. Eraso, F. Portillo, Molecular mechanism of regulation of yeast plasma

membrane H(+)-ATPase by glucose. Interaction between domains and identification of

new regulatory sites. J. Biol. Chem. 269 (1994) 10393-10399.

[195] P. Eraso, M.J. Mazón, F. Portillo, Yeast protein kinase Ptk2 localizes at the plasma

membrane and phosphorylates in vitro the C-terminal peptide of the H+-ATPase,

Biochim. Biophys. Acta, Biomembr. 1758 (2006) 164-170.

[196] C.R. Burtner, C.J. Murakami, B.K. Kennedy, M. Kaeberlein, A molecular

mechanism of chronological aging in yeast, Cell Cycle. 8 (2009) 1256-1270.

[197] W.C. Burhans, M. Weinberger, Acetic acid effects on aging in budding yeast: are

they relevant to aging in higher eukaryotes? Cell Cycle. 8 (2009) 2300-2302.

[198] S. Yoshida, A. Yokoyama, Identification and characterization of genes related to

the production of organic acids in yeast, J. Biosci. Bioeng. 113 (2012) 556-561.

[199] V. Haurie, M. Perrot, T. Mini, P. Jenö, F. Sagliocco, H. Boucherie, The

Transcriptional Activator Cat8p Provides a Major Contribution to the Reprogramming of

Carbon Metabolism during the Diauxic Shift inSaccharomyces cerevisiae, J. Biol. Chem.

276 (2001) 76-85.

[200] M. Casal, H. Cardoso, C. Leao, Mechanisms regulating the transport of acetic acid

in Saccharomyces cerevisiae, Microbiology. 142 (1996) 1385-1390.

Page 75: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

58

[201] S. Paiva, F. Devaux, S. Barbosa, C. Jacq, M. Casal, Ady2p is essential for the

acetate permease activity in the yeast Saccharomyces cerevisiae, Yeast. 21 (2004) 201-

210.

[202] A. Ullah, R. Orij, S. Brul, G.J. Smits, Quantitative Analysis of the Modes of

Growth Inhibition by Weak Organic Acids in Saccharomyces cerevisiae,

Appl. Environ. Microbiol. 78 (2012) 8377-8387.

[203] P. Ludovico, M.J. Sousa, M.T. Silva, C. Leão, M. Côrte-Real, Saccharomyces

cerevisiae commits to a programmed cell death process in response to acetic acid,

Microbiology. 147 (2001) 2409-2415.

[204] Z. Palkova, B. Janderova, J. Gabriel, B. Zikanova, M. Pospisek, J. Forstova,

Ammonia mediates communication between yeast colonies, Nature. 390 (1997) 532-536.

[205] Z. Palkova, F. Devaux, M. Ricicova, L. Minarikova, S. Le Crom, C. Jacq,

Ammonia Pulses and Metabolic Oscillations Guide Yeast Colony Development, Mol.

Biol. Cell. 13 (2002) 3901-3914.

[206] Z. Palková, L. Váchová, Ammonia signaling in yeast colony formation, in: Kwang

W. Jeon (Ed.), International Review of Cytology, Academic Press, 2003, pp. 229-272.

[207] Z. Palkova, J. Forstova, Yeast colonies synchronise their growth and development,

J. Cell. Sci. 113 (2000) 1923-1928.

Page 76: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

59

[208] L. Váchová, F. Devaux, H. Kučerová, M. Řičicová, C. Jacq, Z. Palková, Sok2p

Transcription Factor Is Involved in Adaptive Program Relevant for Long Term Survival

of Saccharomyces cerevisiae Colonies, J. Biol. Chem. 279 (2004) 37973-37981.

[209] C. Boone, H. Bussey, B.J. Andrews, Exploring genetic interactions and networks

with yeast, Nat. Rev. Genet. 8 (2007) 437-449.

[210] J.L. Hartman 4th, B. Garvik, L. Hartwell, Principles for the buffering of genetic

variation, Science. 291 (2001) 1001-1004.

Page 77: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

60

CHAPTER 2

Superoxide Triggers an Acid Burst in Saccharomyces

cerevisiae to Condition the Environment of

Glucose-starved Cells

This research was originally published in the Journal of Biological Chemistry.

J.A. Baron, K.M. Laws, J.S. Chen, V.C. Culotta. Superoxide triggers an acid burst in

Saccharomyces cerevisiae to condition the environment of glucose-starved cells. J. Biol.

Chem. 2013; 288: 4557-4566. © the American Society for Biochemistry and Molecular

Biology.

Page 78: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

61

1. Introduction

Cells exist and thrive in environments that are both supportive of life and

potentially hostile. The environment is in a constant state of change and the onset of

stress conditions, such as nutrient deprivation and exposure to toxins, require appropriate

adaptation strategies. Cells have evolved a variety of mechanisms to appropriately sense

and respond to variations in the environment through induction of stress pathways and

adjustments in metabolism. Yet in some cases, cells respond by modifying the

microenvironment itself to enhance its suitability for their fitness and survival. The

pathogen Plasmodium falciparum, for example, is known to secrete a variety of proteins

that enhance the permeability of the red blood cell membrane to increase the availability

of nutrients [1]. Cancer cells are particularly adept at altering microenvironments. During

the process of invasion, these cells secrete a variety of metalloproteases [2] and also

acidify their surrounding environment to inhibit the toxic activity of immune cells and

chemotherapies, and promote migration to more nutrient-rich environments [3,4]. To

make nutrients more accessible, the bakers' yeast Saccharomyces cerevisiae secretes

phospholipases [5], a variety of enzymes that break down sugars [6], and phosphatases

[7].

To S. cerevisiae, the pH of the environment is critical for regulating growth and

survival [8-10]. Cells growing on the preferred carbon source glucose acidify the

surrounding media by activating the proton ATPase, Pma1p [11], and secreting organic

acids [12,13]. This acidic environment (≈ pH 4-5) helps to maintain the electrochemical

gradient across the plasma membrane and drive uptake of many essential nutrients [14].

On the other hand, when yeast cells grow on the less preferred non-fermentable carbon

Page 79: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

62

sources or are starved for glucose, they alkalize their surrounding environment, which

slows growth but enhances survival under stress [8,15,16]. This period of alkalization

correlates with increases in volatile ammonia that serves as a signaling molecule to

synchronize colony growth [15-18]. Interestingly, following prolonged growth on the

non-fermentable carbon source glycerol, yeast cells can switch from this alkalization

phase to producing acid [17,19,20]. The mechanism or rationale for this sudden “acid

burst” has been unknown. In previous studies completed by Palkova's group [17,19,20],

the production of ammonia during the alkalization phase was reportedly defective in S.

cerevisiae cells lacking the mitochondrial matrix manganese-containing superoxide

dismutase, SOD2 (or yeast Sod2p). Sod2p was suggested to affect ammonia signaling,

but through an unknown pathway [19,21].

Superoxide dismutases (SODs) represent a family of metalloenzymes responsible

for detoxifying superoxide radicals generated as a by-product of aerobic metabolism.

Most eukaryotes, including S. cerevisiae, contain both a Mn-SOD2 in the mitochondrial

matrix as well as a Cu/Zn-containing SOD14 that is largely cytosolic but also found in

the mitochondrial intermembrane space [22,23]. The presence of SODs on both sides of

the mitochondrial inner membrane collectively protect against superoxide generated by

the electron transport chain. The importance of SOD2 in guarding against mitochondrial

oxidative damage has been underscored in various model organisms for SOD2

deficiency. SOD2-/- mice are neonatal lethal and early death in this case is thought to

arise from severe oxidative damage to the mitochondrial respiratory chain [24,25]. Loss

of SOD2 also induces early mortality in Drosophila adults, associated with damage to the

mitochondrial respiratory chain and TCA (tricarboxylic acid) cycle enzymes [26]. In the

Page 80: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

63

bakers' yeast S. cerevisiae, cells lacking Sod2p are viable but are highly sensitive to

hyperoxia and redox cycling agents [27-29], defects ascribed to increased superoxide

damage to mitochondrial components. However, the means by which loss of yeast

mitochondrial matrix Sod2p can control extracellular pH as described earlier [19] is not

clear, particularly because the superoxide substrate for Sod2p should not cross the

mitochondrial membranes in significant quantities [30-33].

Here, we describe a new connection between mitochondrial superoxide and the

control of environmental pH. Specifically we find that the acid burst produced by cells

during long-term respiratory conditions can be induced by mitochondrial superoxide.

With either mutation in yeast SOD2 or by treatment with the redox cycler paraquat, the

acid burst is accelerated in yeast cells starved for glucose. Moreover, the acid burst is

eliminated through Mn-antioxidants that act as SOD mimics and remove intracellular

superoxide. We provide evidence that superoxide damage to Fe-S enzymes in the TCA

cycle results in massive production of acetate involving the mitochondrial aldehyde

dehydrogenase Ald4p. The concomitant acetate burst during nutrient starvation provides

a new carbon source to enhance cell growth during long-term nutrient deprivation.

2. Materials and Methods

2.1. Yeast Strains

The yeast strains in this study were all derived from the parent strain BY4741

(MATa, his3Δ1, leu2Δ0, LYS2, met15Δ0, ura3Δ0). Strains with single gene deletions

were obtained from the MATa haploid deletion library (Open Biosystems) and verified by

sequencing in earlier studies [27] or in the current studies: sod2Δ, sod1Δ, sdh2Δ, sdh4Δ,

Page 81: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

64

cyt1Δ, cox7Δ, cit1Δ, idh1Δ, kgd1Δ, lsc1Δ, fum1Δ, mdh1Δ, ald2Δ, ald3Δ, ald4Δ, ald5Δ,

ald6Δ, pmr1Δ. Strains LJ111 (aco1::LEU2) (gift of L. Jensen), CO217 (coq1::LEU2)

[34], and LJ109 (rho−) [35] were described previously. Strains JAB075

(ald4::kanMX, sod2::URA3) and AR148 (pmr1::kanMX, sod2::URA3; gift of A. Reddi)

were created by transformation of the aforementioned ald4Δ andpmr1Δ strains,

respectively, with the sod2Δ::URA3 plasmid, pGSOD2 as described [36]. Strain JAB069

(ald4::kanMX, sdh4::HIS3) was created by transformation ofald4Δ with a PCR-

based sdh4::HIS3 deletion cassette generated by amplifyingHIS3 from pRS403 [37] using

as primers: forward primer, 5′-

ACGCTTTCGACTTTCTTCCTACGCGCTTTATAATAGCTATGGCGGCATCAGAG

CAGATTG-3′; reverse primer, 5′-

GTTACATGACCGAACAAATGATTCGTGGTGATTTATCTACGTTTACAATTTCC

TGATGCG-3′. Transformations were performed by the standard lithium acetate

procedure [38].

2.2. Culture Conditions and pH Measurements

To examine the effects of yeast colony growth on extracellular pH, solid growth

medium containing 3% glycerol (or where indicated, 2% glucose), 1% yeast extract,

0.01% bromocresol purple (BCP; Sigma, B5880), and 2% bacto-agar was prepared

precisely as described by Palkova and co-workers [15], except supplemental CaCl2 was

generally omitted. pH was typically adjusted to 5.75 with HCl prior to autoclaving but

could range from 4.35 (enhance detection of media alkalization) to 6.5 (enhance detection

of media acidification) without altering the cell alkaline and acid phases. When needed,

the specified concentrations of paraquat (MP Biomedicals) or MnCl2 were added

Page 82: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

65

immediately before pouring. In all cases, extracellular pH was monitored using “giant

colony” growth as prescribed by Palkova et al. [15] where 2 × 105 cells in 10 μl were

spotted onto plates and incubated at 30 °C. Images were taken with a Sony Cybershot

DSC-F828 on the days indicated. Cell viability measurements were obtained by removing

cells at the designated times and measuring colony-forming units on YPD (1% yeast

extract, 2% peptone, 2% glucose) averaged over 3 giant colonies as described (20).

Results were normalized to WT at day 4. WT and sod2Δ strains retain full viability

through day 8 and at day 11 exhibited 92 (S.D. ± 8) and 89% (S.D. ± 11) viability,

respectively.

For monitoring effects of yeast growth under low glucose conditions, cells were

inoculated at an initial A600 = 0.05 and grown shaking at 220 rpm, at 30 °C for 24 hrs in a

liquid media consisting of 1% yeast extract and 0.2% glucose, adjusted to pH 5.5 with

HCl. To determine the extracellular pH of these cultures, a colorimetric assay based on

the pH-sensitive dyes BCP and bromocresol green (BCG; Sigma, 114359) was

developed. Briefly, cells were removed by centrifugation at 17,000 × g for 1 min and the

media was collected (90 μl each) in triplicate and applied to a 96-well plate. Water

(blank), 0.1% BCP, or 0.1% BCG (10 μl) were added and absorbance was measured

at A430, A589, A514, and A616 in a Synergy HT Multi-Mode Microplate Reader (BioTek). pH

was calculated by subtracting the absorbance of the blank at each wavelength and using

the following empirically derived equations:

For BCP (pH ≥ 5.3),

pH = 5.852 + 1.081 × LOG �𝐴589𝐴430

Page 83: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

66

For BCG (pH < 5.3),

pH = 4.590 − LOG�5.963

�𝐴616𝐴514� − 0.045

− 1�

Plates were imaged on an HP Scanjet 8200 scanner. For conditioned media tests,

cultures from low glucose conditions as described earlier were filter-sterilized (Millipore,

Millex-GV, 0.22 μm) to remove cells and the conditioned medium collected was then

used for inoculation of WT cells at A600 = 0.1 followed by aerobic growth for the

indicated time points.

2.3. Metabolite Measurements

To measure glucose, acetate, and succinate concentrations in media, conditioned

medium was collected from cells grown in low glucose as described earlier and cells

were removed by centrifugation. Measurements were performed with a Synergy HT

Multi-Mode Microplate Reader using the QuantiChrom Glucose Assay Kit (BioAssay

Systems), the Acetic Acid kit (R-Biopharm), and the Succinic Acid kit (R-Biopharm),

respectively, according to the manufacturers' specifications. To adapt the Acetic Acid and

Succinic Acid kits to a plate reader format, one-tenth the recommended volumes were

used and each reading was path length corrected using the standard correction of the plate

reader: (A977 −A900)/0.18.

Intracellular acetate measurements were performed in the same manner as those

for media, except cells were first lysed by glass bead homogenization in 20 mM

potassium phosphate, 1.2 M sorbitol, 1.0 mM PMSF, pH 7.4, and clarified by

centrifugation at 17,000 × g for 10 mins. After determining the total protein by Bradford

assay, lysates were heated 15 mins at 85 °C and the supernatant was collected for analysis

Page 84: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

67

after a second centrifugation. Intracellular acetate concentrations were normalized to total

protein.

To measure intracellular metabolites of the TCA cycle, cells grown as described

earlier in low glucose media were harvested at A600 = 1-1.2 and washed in PBS. Cells

were resuspended in PBS at a ratio of 5 μl of PBS/A600 cells and lysed by glass bead

homogenization as described [39]. Lysates were clarified by centrifugation at 17,000

× g for 10 mins and total protein was determined by Bradford assay. Duplicate samples

were analyzed by stable-isotope dilution GC-MS by the Kennedy Krieger Institute

Biochemical Genetics Laboratory.

2.4. Enzymatic Assays

For measurements of aconitase and succinate dehydrogenase (SDH) activity, cells

were collected from low glucose liquid media (10 ml) after 24 hrs growth by

centrifugation or scraped from glycerol plates containing BCP (2-9 × 108 cells collected

from strains plated as in Fig. 2-1A) on the days indicated and stored at -20 °C. Cells were

then resuspended in lysis buffer consisting of 20 mM potassium phosphate,

1.2 M sorbitol, 1.0 mM PMSF, 0.1% Triton X-100, pH 7.4, at a ratio of 10 μl of buffer/2

× 107 cells and lysed by glass bead homogenization. Protein concentrations were

determined by Bradford assay and enzymatic assay measurements were performed with

10-50 μg of lysate protein per reaction. Aconitase activity was measured as described

[39] in 250 μl of total reaction volume and calculated as the nanomoles of cis-aconitate

consumed per mg of protein per min using the Beer-Lambert law and an extinction

coefficient of 4.88 mM−1 cm−1. SDH activity was measured with the procedure of

Ackrell et al. [40], adapted to a 200-μl 96-well plate format. Briefly, SDH in whole cell

Page 85: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

68

lysates was activated at 30 °C for 10 mins in the reaction mixture without electron

acceptors/dyes (178 μl volume; final concentration after addition of dyes = 20

mM potassium phosphate, 20 mM succinate or malonate, 1 mM KCN, pH 7.4). After 10

mins, pre-warmed dyes (22 μl volume; final concentration in complete reaction mix = 1

mM phenazine methosulfate (Sigma), 0.05 mM 2,6-dichlorindophenol (Sigma)) were

added to initiate the reaction. SDH activity was calculated as the malonate-sensitive

nanomoles of 2,6-dichlorindophenol reduced per mg of protein per minute using an

extinction coefficient of 21 mM−1 cm−1.

For Pma1p ATPase activity, cells were grown in 1% yeast extract with 2 or 0.2%

glucose, pH 5.5 to an A600 = 0.75-0.9. Plasma membranes were isolated [41] and Pma1p

ATPase activity was measured by assaying orthovanadate-sensitive release of inorganic

phosphate precisely as described by Monk et al. [42] and normalized to the activity of

WT cells grown in 0.2% glucose.

3. Results

3.1. A Role for Superoxide and SOD2 in Controlling the Acid Burst of Respiring Cells

Extracellular pH in yeast growth medium is readily monitored by pH indicators

such as BCP and BCG. When grown in high glucose, fermenting yeast cells rapidly

acidify the medium (Fig. 2-1A top), but when grown on non-fermentable carbon sources,

the respiring yeast cells initially elevate the extracellular pH (Fig. 2-1A,

middle and bottom). As prescribed by Čáp et al. [19], this alkalization is observed on

glycerol medium supplemented with 30 mM calcium (Fig. 2-1A, middle), yet is also seen

without added calcium (Fig. 2-1A, bottom); hence all subsequent studies were conducted

Page 86: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

69

in the absence of calcium supplements. Alkalinization of the growth medium is not

unique to non-fermentable carbon sources, but is also observed in liquid cultures with

low glucose, conditions where cells switch from fermentation to respiration [8] (Fig. 2-

1B). Cells grown in 0.2% glucose rapidly deplete their glucose (Fig. 2-1C) and the pH of

the extracellular medium rises from 5.5 to ≥7.0 within 24 hrs (Fig. 2-1B).

Following long-term growth on glycerol containing medium, S. cerevisiae cells

can enter an acid phase as previously described, which we refer to herein as the acid burst

(Fig. 2-2) [17,19,20]. We observe that the initiation of the acid burst is substantially

accelerated in strains lacking Mn-SOD2, the mitochondrial matrix manganese-containing

SOD. The sod2Δ null cells show no defect in the initial alkalization of the medium (Fig.

2-2, Fig. 2-3A); however their onset of entry into the acid phase is accelerated compared

with WT strains (Fig. 2-2). This sod2Δ impact on extracellular pH appears specific to

respiratory conditions, as sod2Δ cells are indistinguishable from WT cells in acid

production under fermentative conditions with high glucose (Fig. 2-3A, right). Moreover,

the early acid phase of sod2Δ cells under respiratory conditions is not due to accelerated

death of these cells, because our viability tests (see Materials and Methods) estimate that

WT andsod2Δ both retain ≈ 90% viability after 11 days on glycerol.

The premature acid burst appears specific to sod2Δ strains and was not observed

in other yeast mutants we tested that control oxidative stress resistance. For example,

deletion of genes required for glutathione and thioredoxin reduction in the mitochondria

(GLR1 [27] and TRR2 [43]) did not show the same accelerated acid burst of sod2Δ cells,

nor did deletion of the thioredoxin and glutathione peroxidases AHP1 [44], TSA1 [45],

and GPX1 [46] (data not shown). We also observed that loss of SOD1, encoding the

Page 87: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

70

cytosolic and mitochondrial intermembrane space Cu/Zn-SOD, does not induce a

premature acid burst (Fig. 2-3A). The effects on extracellular pH appear specific to loss

of SOD activity on the matrix side of the mitochondrial inner membrane.

In addition to SOD enzymes, yeast cells have the capacity to remove superoxide

through the action of small non-proteinaceous complexes of manganese such as Mn-

phosphate. These redox active compounds of manganese, so called Mn-antioxidants, are

effective SOD mimics in vitro and in yeast cells in vivo [47-53]. Mn-antioxidants are

elevated when yeast cells accumulate high manganese through either supplementation of

manganese salts to the growth medium or by mutations in the pmr1Δ Golgi pump for

manganese [49-51]. We observed that in both cases, the Mn-antioxidants prevented the

premature acid burst of sod2Δ strains (Fig. 2-3, B and C left). Moreover, boosting levels

of Mn-antioxidants in WT cells completely prevented cells from exiting the alkaline

phase and the acid burst was eliminated (Fig. 2-3C, right). These studies together with

results from sod2Δ cells strongly indicate that the transition to acid phase following long-

term respiratory conditions is triggered by superoxide.

Redox cycling agents such as paraquat can induce production of cellular

superoxide, including mitochondrial matrix superoxide [54-56]. As seen in Fig. 2-4A,

paraquat treatment dramatically accelerated the transition to the acid burst in glycerol and

within 24 hrs, both WT and sod2Δ cells treated with paraquat were producing copious

levels of acid. Much less paraquat was required in the case of sod2Δ cells (Fig. 2-4A),

consistent with the anticipated high superoxide in these mutants that cannot catalytically

remove mitochondrial matrix superoxide. In addition to effects on glycerol plates,

paraquat treatment also caused cells grown in low glucose to produce acid rather than

Page 88: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

71

base (Fig. 2-4B). The paraquat-induced burst of acid was not due to cell death; in fact

total cell growth over 24 hrs was slowed by only 25-30% with doses of paraquat that

induced maximal acidification (Fig. 2-4C), which translates to an overall reduction in

growth rate of only ≈ 15%. These studies support the notion that acid production during

respiratory conditions can be induced by superoxide, specifically the mitochondrial

superoxide relevant to Sod2p.

3.2. Fe-S Cluster Enzymes of the TCA Cycle as Targets of Superoxide during the Acid

Burst

Superoxide can disrupt Fe-S clusters [57,58], and we tested whether the

mitochondrial Fe-S proteins SDH and aconitase were inhibited in cells producing acid. In

WT cells, both aconitase and SDH are affected by paraquat. Aconitase is inhibited even

by low (50 μM) levels of paraquat when cells are not producing acid, but maximal levels

of SDH inhibition requires very high (800 μM) paraquat, conditions where WT cells

produce acid (Fig. 2-5A). In sod2Δ cells, aconitase is constitutively low without paraquat

consistent with previous studies [58,59], and the main effect of 50 μM paraquat is

inhibition of SDH activity (Fig. 2-5A). Hence with WT and sod2Δ cells, acid production

is associated with losses in both aconitase and SDH. The same trends were apparent with

cells grown for extended periods on glycerol without paraquat: WT cells exhibited

reductions in both aconitase and SDH, whereas sod2Δ cells are constitutively low in

aconitase and lose SDH activity over time (Fig. 2-5B).

We tested whether the degree of SDH and aconitase inhibition observed in vitro

reflected disruptions in the TCA cycle, in vivo. TCA cycle metabolites were measured

from cells grown in low glucose and treated with paraquat. As controls, mutants of SDH

Page 89: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

72

and aconitase were used. Yeast sdh4Δ mutants have high succinate but low citrate,

whereas aco1Δ mutants lacking aconitase have high citrate and low succinate (Fig. 2-

5C). WT cells treated with 800 μM paraquat exhibited the elevation in succinate typical

of sdh mutants, but did not lose citrate (Fig. 2-5C), consistent with the double inhibition

of SDH and aconitase (as in Fig. 2-5A). In sod2Δ null cells treated with 50 μM paraquat,

citrate was elevated consistent with losses in aconitase, but there was no loss of succinate

(Fig. 2-5C), again consistent with TCA blockages in both SDH and aconitase.

The correlation between the acid burst and TCA cycle inhibition prompted us to

test whether mutations in TCA cycle genes were by themselves able to induce an acid

burst. As seen in Fig. 2-4D, the aco1Δ and sdh2Δ mutants lacking aconitase and SDH

showed acid production in low glucose, analogous to what was seen with paraquat

treatment. Moreover, there was no additive effect of paraquat, indicating that paraquat

and loss of the TCA cycle operate in the same pathway to induce acid. Acid production is

not specific to disruptions at aconitase and SDH, as mutations in citrate synthase (cit1Δ),

α-ketoglutarate dehydrogenase (kgd1Δ), and fumarase (fum1Δ) all resulted in acid

production (Table 2-1). Disruptions in 5 of 8 steps in the TCA cycle mimicked the effects

of paraquat and induced acid production in low glucose. Only deletions in isocitrate

dehydrogenase (idh1Δ), succinyl-CoA ligase (lsc1Δ), and malate dehydrogenase

(mdh1Δ) resulted in little to no acid production (Table 2-1).

Because the TCA cycle is required for electron transport chain functioning, we

addressed whether blockages in the electron transport chain also lead to the acid burst. As

seen in Table 2-1, mutations that inhibit the electron transport chain, other than SDH loss,

including deletions affecting ubiquinone, complex III, or complex IV, and rho− mutations

Page 90: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

73

lacking mitochondrial DNA, did not rapidly acidify the medium. Thus, it is not a general

defect in mitochondrial function that leads to acid production in low glucose, but specific

disruptions in the TCA cycle.

3.3. The Acid Burst Results from Acetate Production by the Aldehyde Dehydrogenase

Ald4p

What is the source of the acid burst with low glucose? In the case of high glucose,

yeast acidification of the environment involves activation of a cell surface proton ATPase

Pma1p, as well as cellular export of organic acids [11-13]. As seen in Fig. 2-6A, cells

grown in low glucose have extremely low levels of Pma1p activity compared with cells

grown in high glucose. However, Pma1p activity remained low in the TCA cycle mutants

that produce acid (Fig. 2-6, B and C). We therefore examined effects of organic acids.

Previous studies have shown that Fe-S cluster assembly or delivery defects result in

extracellular acidification, marked by secretion of various organic acid mixtures [60-62]

and, in yeast, sdh3 and sdh4 mutants secrete succinate [62]. However, sdh mutants grown

under glucose starvation conditions exhibit only small increases in extracellular succinate

(Fig. 2-6D). The same is true with WT cells treated with paraquat (Fig. 2-6D). These

small increases in extracellular succinate are not likely to account for the large changes in

extracellular pH, and other acid producing cells (e.g. cit1Δ, aco1Δ, kdg1Δ, and fum1Δ)

do not elevate extracellular succinate (Fig. 2-6D). Another organic acid more highly

elevated in sdh mutants is acetate [63], which feeds into the TCA cycle via acetyl-CoA.

WT yeast cells grown in high glucose produce abundant acetate [13] but with glucose-

starved WT cells, extracellular acetate is virtually undetectable (Fig. 2-7A). Remarkably,

when glucose-starved WT cells are treated with paraquat, extracellular acetate levels rise

Page 91: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

74

dramatically to levels comparable with that seen with high glucose (Fig. 2-7, A and B).

The high expulsion of acetate into the growth medium was also seen in sod2Δ cells

treated with low paraquat (Fig. 2-7B) and with all the acid producing TCA cycle mutants

(sdh2Δ, cit1Δ, aco1Δ, kgd1Δ, and fum1Δ) (Fig. 2-7D). In every case where cells make

acid with low glucose, there is ≥100-fold higher levels of extracellular acetate in the

growth medium than is seen with control cells not producing acid (Fig. 2-7, A, B, and D,

and Table 2-2). Intracellular acetate is also elevated, albeit not to the same degree as

extracellular acetate (Fig. 2-7, C and E). The high level of extracellular acetate can easily

account for the acid burst observed, as spiking the growth medium with levels of acetic

acid produced in paraquat-treated cells (0.276 μg/μl) was sufficient to lower the pH to

roughly the same degree (Table 2-2).

The acetate that normally enters the TCA cycle can be derived from the cytosolic

and/or mitochondrial pyruvate dehydrogenase bypass pathways or from the hydrolysis of

acetyl-CoA. In the pyruvate dehydrogenase bypass pathways, pyruvate is converted to

acetaldehyde, which in turn, is converted to acetate via the action of aldehyde

dehydrogenase (ALD) enzymes [64]. S. cerevisiae expresses 5 aldehyde dehydrogenase

isoforms [65] and we tested if mutants in any of these would prevent the acid burst. Of

these, deletion of the major mitochondrial isoform Ald4p most significantly inhibited

paraquat-induced acidification of the growth medium with low glucose (Fig. 2-8A). Loss

of Ald4p also attenuated the constitutive acidification of low glucose medium by sdh4Δ

mutants (Fig. 2-8B). Consistent with these reversals of acidification, we observed

corresponding decreases in acetate production when ald4Δ is disrupted in paraquat-

treated cells and in sdh4Δ mutants grown with low glucose (Fig. 2-7F). The effect

Page 92: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

75

of ald4Δ mutations was also examined in cells grown under long-term glycerol

conditions. As seen in Fig. 2-8C, ald4Δ mutations prevented the early acid burst of sod2Δ

cells on glycerol. Moreover, loss of ALD4 in the background of SOD2+ cells completely

prevented the acid burst with long-term respiratory conditions, and the extracellular pH

continued to rise over time with ald4Δ mutants (Fig. 2-2). Under all cases studied, the

acid burst produced by cells under glucose starvation is dependent on mitochondrial

pyruvate dehydrogenase bypass enzyme, Ald4p.

3.4. Conditioning the Growth Medium by the Acid Burst

What is the significance of the Acid Burst? As one possibility, cells might

produce copious amounts of acetate to provide a carbon source for neighboring cells

under conditions of long-term carbon source starvation. To determine whether acetate

production and/or media acidification could enhance growth of yeast cells, we collected

“conditioned” growth media that was conditioned by cells grown 24 hrs in low glucose.

Low acetate/high pH medium was obtained from cultures of WT and ald4Δ cells, and

high acetate/low pH medium was obtained from cultures of sdh4Δ cells and from sod2Δ

cells treated with 50 μM paraquat, a low dose that is non-toxic to WT cells (Fig. 2-4C).

After removal of cells by filtration, these various conditioned media were used to

inoculate WT cells and growth was monitored over 10 hrs. As seen in Fig. 2-9A, the high

acetate/low pH media of sdh4Δ and paraquat-treated sod2Δ cells supported strong growth

of WT cells, whereas the low acetate/high pH media from WT cells or from ald4Δ

mutants was poorly effective at supporting growth. To determine whether acetate alone

accounts for these differences, we supplemented the low acetate medium from WT

and ald4Δ cells with levels of glacial acetic acid (0.20-0.22 μg/μl) equivalent to that in

Page 93: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

76

cultures of sdh4Δ and paraquat-treated sod2Δ cells. This supplementation of acetic acid

was sufficient to eliminate differences in the various conditioned media, and all four

media types supported growth of WT cells (Fig. 2-9B). Thus, the production of acetate

can enhance growth of yeast cells under glucose starvation. This event triggered by

superoxide disruptions in the TCA cycle may enable the organism to withstand long-term

periods of starvation.

4. Discussion

Herein, we describe a mechanism by which the S. cerevisiae yeast modifies its

extracellular environment during prolonged nutrient deprivation. Under long-term

colonization on respiratory carbon sources, yeast cells shift from alkalization to

producing acid and we demonstrate here that this acid burst is dependent on the acetate-

producing Ald4p aldehyde dehydrogenase. Media conditioned by Ald4-acetate is indeed

able to promote yeast cell growth under glucose starvation conditions. Our findings are

consistent with a model in which the acid burst is produced in response to disruption of

the TCA cycle by superoxide (Fig. 2-9C).

Perhaps the most intriguing finding of this work is the apparent role of

mitochondrial matrix superoxide in conditioning the extracellular environment.

Mitochondrial superoxide is not likely to cross mitochondrial membranes in significant

quantities [30-33] and yet it can influence extracellular pH by modulating organic acid

production. In support of this effect of mitochondrial superoxide, we observed that the

acid burst under long-term glycerol conditions is accelerated by sod2Δ mutations lacking

mitochondrial matrix Mn-SOD, but not by sod1Δ mutations affecting the largely

Page 94: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

77

cytosolic Cu/Zn-SOD1. Previously, the loss of Sod2p was shown to reduce volatile

ammonia production and prevent transition to the alkali phase under long-term glycerol

conditions but the mechanism was unknown [19]. This may be explained by the early

acid burst of sod2Δ cells which by lowering pH would trap a significant portion of

ammonia in the non-volatile ammonium form (pKa ≈ 9.25). Alternatively, Sod2p may

have separable roles in regulating extracellular pH through ammonia and acetate

production.

In further support of the role of superoxide in triggering the acid burst, we

observe that acid production under both glycerol and low glucose conditions is

dramatically accelerated when WT cells are treated with the redox cycler paraquat;

conversely, the acid burst with long-term glycerol conditions is eliminated by Mn-

antioxidants, chemical mimics of SOD enzymes that can remove superoxide in yeast [47-

53]. Superoxide, more specifically mitochondrial superoxide, represents an ideal

candidate to initiate the acid burst. Key Fe-S cluster enzymes in the TCA cycle (i.e. SDH

and aconitase) are sensitive to superoxide inactivation [24-26,57,58], and we show here

that disruption in the TCA cycle can cause build-up of acetate derived from

mitochondrial Ald4p. Although multiple reports have established a role for ROS in cell

signaling and regulation [66-69], the inactivation of TCA cycle enzymes by

mitochondrial superoxide was up until now considered a detrimental outcome of

oxidative damage. This is, perhaps, the first example where superoxide inactivation of

mitochondrial enzymes can be beneficial, particularly under the stress conditions of

nutrient deprivation.

Page 95: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

78

Superoxide is a natural by-product of respiration, yet in WT cells the acid burst

only occurs after prolonged incubations under respiratory conditions (as in Fig. 2-2). The

products of superoxide damage appear to accumulate slowly in these cells, as we observe

a gradual loss in SDH activity over prolonged respiratory growth (Fig. 2-5B). In sod2Δ

cells the combined effects of this slow loss in SDH together with the already low

aconitase of these cells can explain the acceleration in the acid burst observed. Once SDH

and/or aconitase activity have been sufficiently inhibited to disrupt TCA cycle

functioning, the Ald4p acid burst occurs. In the simplest model, the blockage in the TCA

cycle itself prevents mitochondrial consumption of Ald4-acetate, and the accumulated

metabolite is expelled from the cell. Ald4 production of acetate might also be enhanced

in a futile attempt to correct the TCA cycle defect. It is noteworthy that blockages in 5 of

8 steps in the TCA cycle resulted in an acetate burst, including losses in SDH and

aconitase, as well as citrate dehydrogenase, α-ketoglutarate dehydrogenase, and

fumarase. It is possible that mutants for the remaining three steps (idh1Δ, mdh1Δ,

and lsc1Δ [70,71]) do not yield the same disruption in TCA cycle functioning.

Although high acetate has previously been associated with increased

chronological aging in yeast [8,9], these previous studies were conducted with fermenting

yeast cells with abundant glucose. In contrast, our studies with respiring yeast (as in Fig.

2-9) indicate that high acetate can be beneficial in environments with scarce nutrients.

Overall, these studies provide the first line of evidence for mitochondrial superoxide and

the TCA cycle in conditioning the extracellular environment under glucose starvation

conditions.

Page 96: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

79

TABLE 2-1. Effects of TCA and ETC mutants on extracellular pH.

pH was measured as described in Materials and Methods after the indicated strains were

grown 24 hrs in 0.2% glucose growth medium. The average pH of triplicate samples is

denoted with standard deviation in parentheses. For WT + PQ, WT cells were grown in

media containing 800 μM paraquat. TCA = tricarboxylic acid cycle, ETC = electron

transport chain, mito = mitochondrial

Page 97: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

ET

C m

utan

ts

T

CA

mut

ants

Sa

mpl

e D

escr

iptio

n pH

Sam

ple

Des

crip

tion

pH

No

cells

5.43

(± 0

.00)

No

cells

5.53

(± 0

.00)

W

T

6.90

(± 0

.03)

WT

7.

17 (±

0.0

5)

WT

+ PQ

4.67

(± 0

.03)

WT

+ PQ

4.59

(± 0

.03)

sd

h2∆

Com

plex

II

4.47

(± 0

.02)

sdh2∆

Succ

inat

e de

hydr

ogen

ase

4.54

(± 0

.05)

sd

h4∆

4.47

(± 0

.02)

sdh4∆

4.55

(± 0

.04)

co

q1∆

Coe

nzym

e Q

5.

19 (±

0.0

1)

ci

t1∆

Citr

ate

synt

hase

4.

54 (±

0.0

2)

cyt1∆

Com

plex

III

5.21

(± 0

.03)

aco1∆

Aco

nita

se

4.64

(± 0

.02)

co

x7∆

Com

plex

IV

5.20

(± 0

.03)

idh1∆

Isoc

itrat

e de

hydr

ogen

ase

5.46

(± 0

.10)

rh

o−

No

mito

DN

A

5.17

(± 0

.03)

kgd1∆

α-ke

togl

utar

ate

dehy

drog

enas

e 4.

72 (±

0.0

1)

lsc1∆

Succ

inyl

-CoA

liga

se

6.42

(± 0

.08)

fu

m1∆

Fu

mar

ase

4.71

(± 0

.01)

m

dh1∆

M

alat

e de

hydr

ogen

ase

7.14

(± 0

.07)

80

Page 98: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

81

TABLE 2-2. Contribution of acetic acid to pH of growth medium.

Acetic acid and pH levels were measured as described in Materials and Methods in 0.1%

glucose growth medium (starting pH 5.55) that was cultured where indicated (WT cells)

for 24 hrs with WT yeast grown ± paraquat. Supplementation of medium with 0.25 µg/µl

acetic acid is sufficient to drop pH to levels seen in medium from paraquat treated WT

cells.

Sample Acetate (μg/μl) pH

WT cells undetectable 6.46 WT cells + 800 μM PQ 0.276 4.87 No cells 0.026 5.55 No cells + 0.25 μg/μl

glacial acetic acid 0.278 4.93

Page 99: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

82

FIGURE 2-1. Yeast cells alkalize their environment under respiratory conditions.

A) “Giant colonies” [15] of WT BY4741 yeast cells were obtained by spotting 2x105

cells onto the indicated growth medium containing the pH indicator BCP and as carbon

source either 2% glucose or 3% glycerol. Where indicated, medium was supplemented

with 30 mM CaCl2. Cells were photographed on the second day of growth. (B,C) WT

yeast cells were inoculated at OD600 = 0.05 in a liquid media containing 0.2% glucose

and following specific time points, cells were removed and extracellular medium

analyzed for either glucose or pH. These data were provided by K.M. Laws. (B) Media

pH was determined following 24 hrs of growth using BCP and BCG indicator dyes.

Shown is a photograph of media color change with the pH indicators and the

corresponding pH value calculated as outlined in Materials and Methods. Numerical

results represent the averages of triplicate analyses where SD ≤0.03. “Cells: -” = growth

media alone. (C) Extracellular glucose was determined at the indicated time points as

described in Materials and Methods. Results represent the averages of triplicate samples

where error bars denote SD. Illustrated in 2-1A bottom and 2-1B right are the gradient of

color changes with BCP and BCG as pH ranges from acidic (BCP yellow; BCG green) to

basic (BCP purple; BCG blue).

Page 100: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

83

Page 101: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

84

FIGURE 2-2. Acid production following the alkaline phase of yeast cell growth is

accelerated by sod2∆ mutations and eliminated by ald4∆ mutations.

The indicated yeast strains were analyzed for acid and base production following the

designated times of incubation on glycerol containing plates (no calcium) as described in

Fig. 2-1A.

Page 102: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

85

FIGURE 2-3. Role of mitochondrial SOD2 and Mn-antioxidants in preventing the

acid burst of respiratory cells.

Giant colonies of the indicated strains were plated under either respiratory, 3% glycerol

(A left, B-C), or fermenting, 2% glucose (A right), conditions as in Fig. 2-1A and

photographed on the designated days of incubation (part A, B) or on day 11 (sod2∆) or

day 16 (WT) (part C). (C) Where indicated, plates were supplemented with 0.5 mM

MnCl2, a concentration that is non-toxic but sufficient to maximize intracellular levels of

Mn-antioxidants that remove superoxide [47-53]. Data in parts A and B were provided

by K.M. Laws.

Page 103: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

86

FIGURE 2-4. The redox cycler paraquat and sdh∆ induce a rapid acid burst under

respiratory conditions and are not additive.

The indicated yeast strains were grown either on glycerol containing medium for 2 days

(A), or in low glucose liquid medium for 24 hrs (B, C, D) as in Fig. 2-1. Where

indicated, the growth medium was supplemented with the designated concentrations of

paraquat (PQ). (B, C, open triangles and D) pH was calculated in triplicate as in Fig. 2-

1B where values in parentheses (B,D) denote SD. (C) Cell growth (black circles) was

measured turbidimetrically at an optical density of 600 nm (OD600) and plotted as a

function of control cell growth with no paraquat.

Page 104: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

87

Page 105: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

88

FIGURE 2-5. Activities of mitochondrial Fe-S cluster enzymes in acid producing

cells.

The designated yeast strains were grown for 24 hrs in 0.2% glucose medium that was

supplemented with either 800 µM (WT) or 50 µM (sod2∆) paraquat. (A) Extracellular

pH shown in left column was measured in triplicate samples as in Fig. 2-1B, and whole

cell lysates were prepared and assayed for aconitase and SDH activity as described in

Materials and Methods. Units are represented as either nmol cis-aconitate converted /

mg protein /min (aconitase), or nmol DCIP reduced / mg protein / min (SDH); values

represent the average of triplicate samples; error bars = SD. (B) The indicated

intracellular metabolites were measured by gas chromatography mass spectrometry as

described in Materials and Methods. Values represent averages of duplicate samples;

error bars = SD.

Page 106: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

89

Page 107: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

90

FIGURE 2-6. Acid production with low glucose does not correlate with activation of

Pma1p, the proton ATPase.

The indicated yeast strains were grown in either 2% (A) or 0.2% (A-C) glucose to mid

log phase, and Pma1p activity was measured in membranes by orthovanadate-sensitive

ATPase activity as described in Materials and Methods. Values are shown normalized to

WT cells in 0.2% glucose, and represent the average of duplicate measurements; error

bars = SD. (C) “Base” and “Acid” reflects the corresponding effect of the yeast strain on

extracellular medium in 0.2% glucose.

Page 108: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

91

FIGURE 2-7. Cells produce high levels of extracellular acetate during the acid

burst.

Extracellular acetate/acetic acid was measured in the growth medium of WT cells (A) or

the indicated strains grown in either 2% (A) or 0.2% (A-C) glucose for 24 hrs as

described in Materials and Methods. Asterisks donate the particular conditions and

strains in which cells acidify the growth medium. “Media” = levels of acetate in the

starting growth medium with no cells. “+PQ” = either 50 µM (sod2∆ cells) or 800 µM

(all other strains) paraquat added during growth. Results represent the averages of

triplicate samples; error bars = SD.

Page 109: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

92

Page 110: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

93

FIGURE 2-8. The mitochondrial Ald4p aldehyde dehydrogenase is required for the

acid burst.

The indicated strains were grown either in low glucose liquid medium (A and B) or on

glycerol containing plates (C) and pH monitored as in Fig. 2-1. (A, B) pH values

represent the results of duplicate (A) or triplicate (B) samples where SD ≤0.12. +PQ =

cultures supplemented with 800 µM paraquat.

Page 111: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

94

FIGURE 2-9. Medium conditioned during the acid burst can promote yeast growth.

Growth medium was collected from the indicated cultures following 24 hrs incubation in

0.2% glucose medium. This cell-free “conditioned medium” was used to inoculate WT

cultures of yeast as described in Materials and Methods and growth monitored

turbidimetrically at 600 nm. Results represent the averages of triplicates cultures; error

bars = SD. In some cases, the error was too small to be seen by error bars. (A)

Conditioned medium was used as isolated from cells. (B) The levels of acetic

acid/acetate in all four types of conditioned medium were normalized by supplementing

0.20-0.22 μg/μl glacial acetic acid to the medium conditioned by WT and ald4∆ cells.

(C) A model for how mitochondrial superoxide produced during long-term respiratory

conditions can lead to the acid burst and conditioning of the extracellular environment.

Superoxide can damage the Fe-S enzymes aconitase and SDH in the mitochondrial

matrix and the concomitant loss of TCA cycle activity induces massive production of

acetate from the mitochondrial Ald4p aldehyde dehydrogenase. The high levels of

acetate expelled into the growth medium conditions the extracellular environment to

promote growth of neighboring yeast cells.

Page 112: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

95

Page 113: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

96

5. Reference List

[1] A. Mbengue, X.Y. Yam, C. Braun-Breton, Human erythrocyte remodelling during

Plasmodium falciparum malaria parasite growth and egress, Br. J. Haematol. 157 (2012)

171-179.

[2] P. Friedl, S. Alexander, Cancer Invasion and the Microenvironment: Plasticity and

Reciprocity, Cell. 147 (2011) 992.

[3] R. Martinez-Zaguilan, E.A. Seftor, R.E. Seftor, Y.W. Chu, R.J. Gillies, M.J. Hendrix,

Acidic pH enhances the invasive behavior of human melanoma cells, Clin. Exp.

Metastasis. 14 (1996) 176-186.

[4] F. Hirschhaeuser, U.G.A. Sattler, W. Mueller-Klieser, Lactate: A Metabolic Key

Player in Cancer, Cancer Research. 71 (2011) 6921-6925.

[5] O. Merkel, M. Fido, J.A. Mayr, H. Prüger, F. Raab, G. Zandonella, S.D. Kohlwein, F.

Paltauf, Characterization and Function in Vivo of Two Novel Phospholipases

B/Lysophospholipases from Saccharomyces cerevisiae, J. Biol. Chem. 274 (1999) 28121-

28127.

[6] M. Jules, V. Guillou, J. François, J. Parrou, Two Distinct Pathways for Trehalose

Assimilation in the Yeast Saccharomyces cerevisiae, Appl. Environ. Microbiol. 70 (2004)

2771-2778.

[7] K. Vogel, A. Hinnen, The yeast phosphatase system, Mol. Microbiol. 4 (1990) 2013-

2017.

Page 114: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

97

[8] C.R. Burtner, C.J. Murakami, B.K. Kennedy, M. Kaeberlein, A molecular mechanism

of chronological aging in yeast, Cell Cycle. 8 (2009) 1256-1270.

[9] C.R. Burtner, C.J. Murakami, B. Olsen, B.K. Kennedy, M. Kaeberlein, A genomic

analysis of chronological longevity factors in budding yeast, Cell Cycle. 10 (2011) 1385-

1396.

[10] M. Hayashi, K. Ohkuni, I. Yamashita, Control of division arrest and entry into

meiosis by extracellular alkalisation in Saccharomyces cerevisiae, Yeast. 14 (1998) 905-

913.

[11] F. Portillo, Regulation of plasma membrane H+-ATPase in fungi and plants,

Biochim. Biophys. Acta, Rev. Biomembr. 1469 (2000) 31-42.

[12] K. Sigler, A. Kotyk, A. Knotková, M. Opekarová, Processes involved in the creation

of buffering capacity and in substrate-induced proton extrusion in the yeast

Saccharomyces cerevisiae, Biochim. Biophys. Acta, Biomembr. 643 (1981) 583-592.

[13] W.C. Burhans, M. Weinberger, Acetic acid effects on aging in budding yeast: are

they relevant to aging in higher eukaryotes? Cell Cycle. 8 (2009) 2300-2302.

[14] M.E. van der Rest, A.H. Kamminga, A. Nakano, Y. Anraku, B. Poolman, W.N.

Konings, The plasma membrane of Saccharomyces cerevisiae: structure, function, and

biogenesis, Microbiol. Rev. 59 (1995) 304-322.

[15] Z. Palkova, B. Janderova, J. Gabriel, B. Zikanova, M. Pospisek, J. Forstova,

Ammonia mediates communication between yeast colonies, Nature. 390 (1997) 532-536.

Page 115: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

98

[16] Z. Palková, L. Váchová, Ammonia signaling in yeast colony formation, in: Kwang

W. Jeon (Ed.), International Review of Cytology, Academic Press, 2003, pp. 229-272.

[17] Z. Palkova, J. Forstova, Yeast colonies synchronise their growth and development, J.

Cell. Sci. 113 (2000) 1923-1928.

[18] Z. Palkova, F. Devaux, M. Ricicova, L. Minarikova, S. Le Crom, C. Jacq, Ammonia

Pulses and Metabolic Oscillations Guide Yeast Colony Development, Mol. Biol. Cell. 13

(2002) 3901-3914.

[19] M. Čáp, L. Váchová, Z. Palková, Yeast Colony Survival Depends on Metabolic

Adaptation and Cell Differentiation Rather Than on Stress Defense, J. Biol. Chem. 284

(2009) 32572-32581.

[20] Z. Palková, L. Váchová, D. Gášková, H. Kučerová, Synchronous plasma membrane

electrochemical potential oscillations during yeast colony development and aging, Mol.

Membr. Biol. 26 (2009) 228-235.

[21] M. Čáp, L. Váchová, Z. Palková, How to survive within a yeast colony: Change

metabolism or cope with stress? Commun. Integr. Biol. 3 (2010) 198-200.

[22] L.A. Sturtz, K. Diekert, L.T. Jensen, R. Lill, V.C. Culotta, A fraction of yeast

Cu,Zn-superoxide dismutase and its metallochaperone, CCS, localize to the

intermembrane space of mitochondria—A physiological role for SOD1 in guarding

against mitochondrial oxidative damage, J. Biol. Chem. 276 (2001) 38084-38089.

Page 116: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

99

[23] R.A. Weisiger, I. Fridovich, Mitochondrial Superoxide Dismutase: SITE OF

SYNTHESIS AND INTRAMITOCHONDRIAL LOCALIZATION, J. Biol. Chem. 248

(1973) 4793-4796.

[24] S. Melov, P. Coskun, M. Patel, R. Tuinstra, B. Cottrell, A.S. Jun, T.H. Zastawny, M.

Dizdaroglu, S.I. Goodman, T. Huang, H. Miziorko, C.J. Epstein, D.C. Wallace,

Mitochondrial disease in superoxide dismutase 2 mutant mice,

Proc. Natl. Acad. Sci. U. S. A. 96 (1999) 846-851.

[25] Y. Li, T.T. Huang, E.J. Carlson, S. Melov, P.C. Ursell, J.L. Olson, L.J. Noble, M.P.

Yoshimura, C. Berger, P.H. Chan, D.C. Wallace, C.J. Epstein, Dilated cardiomyopathy

and neonatal lethality in mutant mice lacking manganese superoxide dismutase, Nat.

Genet. 11 (1995) 376-381.

[26] K. Kirby, J. Hu, A.J. Hilliker, J.P. Phillips, RNA interference-mediated silencing of

Sod2 in Drosophila leads to early adult-onset mortality and elevated endogenous

oxidative stress, Proc. Natl. Acad. Sci. U. S. A. 99 (2002) 16162-16167.

[27] C.E. Outten, R.L. Falk, V.C. Culotta, Cellular factors required for protection from

hyperoxia toxicity in Saccharomyces cerevisiae, Biochem. J. 388 (2005) 93-101.

[28] D.M. Guidot, J.M. McCord, R.M. Wright, J.E. Repine, Absence of electron transport

(Rho 0 state) restores growth of a manganese-superoxide dismutase-deficient

Saccharomyces cerevisiae in hyperoxia. Evidence for electron transport as a major source

of superoxide generation in vivo. J. Biol. Chem. 268 (1993) 26699-26703.

Page 117: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

100

[29] A.B. Demir, A. Koc, Assessment of chronological lifespan dependent molecular

damages in yeast lacking mitochondrial antioxidant genes,

Biochem. Biophys. Res. Commun. 400 (2010) 106.

[30] J.A. Imlay, PATHWAYS OF OXIDATIVE DAMAGE, Annu. Rev. Microbiol. 57

(2003) 395-418.

[31] I. Fridovich, Fundamental Aspects of Reactive Oxygen Species, or What's the

Matter with Oxygen? Ann. N. Y. Acad. Sci. 893 (1999) 13-18.

[32] D. Behar, G. Czapski, J. Rabani, L.M. Dorfman, H.A. Schwarz, Acid dissociation

constant and decay kinetics of the perhydroxyl radical, J. Phys. Chem. 74 (1970) 3209-

3213.

[33] R.A. Gus'kova, I.I. Ivanov, V.K. Kol'tover, V.V. Akhobadze, A.B. Rubin,

Permeability of bilayer lipid membranes for superoxide (O2-.) radicals,

Biochim. Biophys. Acta. 778 (1984) 579-585.

[34] C.E. Outten, V.C. Culotta, A novel NADH kinase is the mitochondrial source of

NADPH in Saccharomyces cerevisiae, EMBO J. 22 (2003) 2015-2024.

[35] E. Luk, M. Carroll, M. Baker, V.C. Culotta, Manganese activation of superoxide

dismutase 2 in Saccharomyces cerevisiae requires MTM1, a member of the mitochondrial

carrier family, Proc. Natl. Acad. Sci. U. S. A. 100 (2003) 10353-10357.

[36] V.C. Culotta, L.W. Klomp, J. Strain, R.L. Casareno, B. Krems, J.D. Gitlin, The

copper chaperone for superoxide dismutase, J. Biol. Chem. 272 (1997) 23469-23472.

Page 118: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

101

[37] R.S. Sikorski, P. Hieter, A System of Shuttle Vectors and Yeast Host Strains

Designed for Efficient Manipulation of DNA in Saccharomyces cerevisiae, Genetics. 122

(1989) 19-27.

[38] R.H. Schiestl, R.D. Gietz, High efficiency transformation of intact yeast cells using

single stranded nucleic acids as a carrier, Curr. Genet. 16 (1989) 339-346.

[39] J.E. Gleason, D.J. Corrigan, J.E. Cox, A.R. Reddi, L.A. McGinnis, V.C. Culotta,

Analysis of hypoxia and hypoxia-like states through metabolite profiling, PLoS One. 6

(2011) e24741.

[40] B.A.C. Ackrell, E.B. Kearney, T.P. Singer, [47] Mammalian succinate

dehydrogenase, in: Sidney Fleischer and Lester Packer (Ed.), Methods in Enzymology,

Academic Press, 1978, pp. 466-483.

[41] D.S. Perlin, S.L. Harris, D. Seto-Young, J.E. Haber, Defective H(+)-ATPase of

hygromycin B-resistant pma1 mutants from Saccharomyces cerevisiae. J. Biol. Chem.

264 (1989) 21857-21864.

[42] B.C. Monk, M.B. Kurtz, J.A. Marrinan, D.S. Perlin, Cloning and characterization of

the plasma membrane H(+)-ATPase from Candida albicans. J. Bacteriol. 173 (1991)

6826-6836.

[43] J.R. Pedrajas, E. Kosmidou, A. Miranda-Vizuete, J. Gustafsson, A.P.H. Wright, G.

Spyrou, Identification and Functional Characterization of a Novel Mitochondrial

Thioredoxin System in Saccharomyces cerevisiae, J. Biol. Chem. 274 (1999) 6366-6373.

Page 119: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

102

[44] J. Lee, D. Spector, C. Godon, J. Labarre, M.B. Toledano, A New Antioxidant with

Alkyl Hydroperoxide Defense Properties in Yeast, J. Biol. Chem. 274 (1999) 4537-4544.

[45] H.Z. Chae, S.J. Chung, S.G. Rhee, Thioredoxin-dependent peroxide reductase from

yeast. J. Biol. Chem. 269 (1994) 27670-27678.

[46] Y. Inoue, T. Matsuda, K. Sugiyama, S. Izawa, A. Kimura, Genetic Analysis of

Glutathione Peroxidase in Oxidative Stress Response of Saccharomyces cerevisiae, J.

Biol. Chem. 274 (1999) 27002-27009.

[47] J.D. Aguirre, V.C. Culotta, Battles with iron: manganese in oxidative stress

protection, J. Biol. Chem. 287 (2012) 13541-13548.

[48] A.R. Reddi, V.C. Culotta, Regulation of manganese antioxidants by nutrient sensing

pathways in Saccharomyces cerevisiae, Genetics. 189 (2011) 1261-1270.

[49] A.R. Reddi, L.T. Jensen, A. Naranuntarat, L. Rosenfeld, E. Leung, R. Shah, V.C.

Culotta, The overlapping roles of manganese and Cu/Zn SOD in oxidative stress

protection, Free Radical Biol. Med. 46 (2009) 154-162.

[50] R.J. Sanchez, C. Srinivasan, W.H. Munroe, M.A. Wallace, J. Martins, T.Y. Kao, K.

Le, E.B. Gralla, J.S. Valentine, Exogenous manganous ion at millimolar levels rescues all

known dioxygen-sensitive phenotypes of yeast lacking CuZnSOD, J. Biol. Inorg. Chem.

10 (2005) 913-923.

Page 120: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

103

[51] P. Lapinskas, K. Cunningham, X. Liu, G. Fink, V. Culotta, Mutations in PMR1

suppress oxidative damage in yeast cells lacking superoxide dismutase, Mol. Cell. Biol.

15 (1995) 1382-1388.

[52] K. Barnese, E.B. Gralla, D.E. Cabelli, J. Selverstone Valentine, Manganous

Phosphate Acts as a Superoxide Dismutase, J. Am. Chem. Soc. 130 (2008) 4604-4606.

[53] K. Barnese, E.B. Gralla, J.S. Valentine, D.E. Cabelli, Biologically relevant

mechanism for catalytic superoxide removal by simple manganese compounds, Proc.

Natl. Acad. Sci. U. S. A. 109 (2012) 6892-6897.

[54] H.M. Cocheme, M.P. Murphy, Complex I is the major site of mitochondrial

superoxide production by paraquat, J. Biol. Chem. 283 (2008) 1786-1798.

[55] H.M. Cochemé, M.P. Murphy, Chapter 22 The Uptake and Interactions of the Redox

Cycler Paraquat with Mitochondria, in: William S. Allison and Immo E. Scheffler (Ed.),

Methods in Enzymology, Academic Press, 2009, pp. 395-417.

[56] M. Blaszczynski, J. Litwinska, D. Zaborowska, T. Bilinski, The role of respiratory

chain in paraquat toxicity in yeast, Acta Microbiol. Pol. 34 (1985) 243-254.

[57] K. Keyer, J. Imlay, Superoxide accelerates DNA damage by elevating free-

iron levels, Proc. Natl. Acad. Sci. U. S. A. 93 (1996) 13635-13640.

[58] M.A. Wallace, L. Liou, J. Martins, M.H.S. Clement, S. Bailey, V.D. Longo, J.S.

Valentine, E.B. Gralla, Superoxide Inhibits 4Fe-4S Cluster Enzymes Involved in Amino

Page 121: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

104

Acid Biosynthesis. Cross-compartment protection by CuZn-superoxide dismutase, J.

Biol. Chem. 279 (2004) 32055-32062.

[59] J. Strain, C.R. Lorenz, J. Bode, S. Garland, G.A. Smolen, D.T. Ta, L.E. Vickery,

V.C. Culotta, Suppressors of Superoxide Dismutase (SOD1) Deficiency in

Saccharomyces cerevisiae. IDENTIFICATION OF PROTEINS PREDICTED TO

MEDIATE IRON-SULFUR CLUSTER ASSEMBLY, J. Biol. Chem. 273 (1998)

31138-31144.

[60] A. Navarro-Sastre, F. Tort, O. Stehling, M. Uzarska, J. Arranz, M. del Toro, M. .

Labayru, J. Landa, A. Font, J. Garcia-Villoria, B. Merinero, M. Ugarte, L. Gutierrez-

Solana, J. Campistol, A. Garcia-Cazorla, J. Vaquerizo, E. Riudor, P. Briones, O. Elpeleg,

A. Ribes, R. Lill, A Fatal Mitochondrial Disease Is Associated with Defective NFU1

Function in the Maturation of a Subset of Mitochondrial Fe-S Proteins,

Am. J. Hum. Genet. 89 (2011) 656-667.

[61] A.D. Sheftel, O. Stehling, A.J. Pierik, D.J.A. Netz, S. Kerscher, H. Elsässer, I.

Wittig, J. Balk, U. Brandt, R. Lill, Human Ind1, an Iron-Sulfur Cluster Assembly Factor

for Respiratory Complex I, Mol. Cell. Biol. 29 (2009) 6059-6073.

[62] S.S.W. Szeto, S.N. Reinke, B.D. Sykes, B.D. Lemire, Ubiquinone-binding Site

Mutations in the Saccharomyces cerevisiae Succinate Dehydrogenase Generate

Superoxide and Lead to the Accumulation of Succinate, J. Biol. Chem. 282 (2007)

27518-27526.

Page 122: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

105

[63] S.S.W. Szeto, S.N. Reinke, B.D. Sykes, B.D. Lemire, Mutations in the

Saccharomyces cerevisiae Succinate Dehydrogenase Result in Distinct Metabolic

Phenotypes Revealed Through 1H NMR-Based Metabolic Footprinting, J. Proteome

Res. 9 (2010) 6729-6739.

[64] S. Boubekeur, O. Bunoust, N. Camougrand, M. Castroviejo, M. Rigoulet, B. Guérin,

A Mitochondrial Pyruvate Dehydrogenase Bypass in the Yeast Saccharomyces

cerevisiae, J. Biol. Chem. 274 (1999) 21044-21048.

[65] J.P. Navarro-Aviño, R. Prasad, V.J. Miralles, R.M. Benito, R. Serrano, A proposal

for nomenclature of aldehyde dehydrogenases in Saccharomyces cerevisiae and

characterization of the stress-inducible ALD2 and ALD3 genes, Yeast. 15 (1999) 829-

842.

[66] R. Mittler, S. Vanderauwera, N. Suzuki, G. Miller, V.B. Tognetti, K. Vandepoele,

M. Gollery, V. Shulaev, F. Van Breusegem, ROS signaling: the new wave? Trends Plant

Sci. 16 (2011) 300-309.

[67] T. Meng, T. Fukada, N.K. Tonks, Reversible Oxidation and Inactivation of Protein

Tyrosine Phosphatases In Vivo, Mol. Cell. 9 (2002) 387-399.

[68] B.L. Prosser, C.W. Ward, W.J. Lederer, X-ROS Signaling: Rapid Mechano-Chemo

Transduction in Heart, Science. 333 (2011) 1440-1445.

[69] B. D'Autreaux, M.B. Toledano, ROS as signalling molecules: mechanisms that

generate specificity in ROS homeostasis, Nat. Rev. Mol. Cell Biol. 8 (2007) 813-824.

Page 123: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

106

[70] M.T. McCammon, Mutants of Saccharomyces cerevisiae with Defects in Acetate

Metabolism: Isolation and Characterization of Acn(-) Mutants, Genetics. 144 (1996) 57-

69.

[71] B. Sumegi, M.T. McCammon, A.D. Sherry, D.A. Keys, L. McAlister-Henn, P.A.

Srere, Metabolism of [3-13C]pyruvate in TCA cycle mutants of yeast, Biochemistry (N.

Y. ). 31 (1992) 8720-8725.

Page 124: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

107

CHAPTER 3

Oxidative Stress and Lethality in S. cerevisiae with

Mutations in SOD1 and the Proton ATPase PMA1

Page 125: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

108

1. Introduction

The plasma membrane H+-ATPase is found in species throughout the fungal and

plant kingdoms and regulates intracellular pH and the membrane electrochemical

potential required for secondary nutrient uptake [1]. In the yeast Saccharomyces

cerevisiae, the major plasma membrane H+-ATPase, Pma1p, is essential for life and its

activity correlates directly with growth [2,3]. Furthermore, Pma1p activity significantly

affects viability under a variety of stress conditions [4], making it a key regulator of cell

survival and growth.

Pma1p itself is regulated by a variety of conditions with the most notable and well

studied being glucose availability [5-8]. In response to glucose, Pma1p is both

transcriptionally up-regulated [3,9] and post-translationally activated [8]. However,

given the high expression of Pma1p even in the absence of glucose [3] and its long half-

life [10,11], post-translational regulation plays a more significant role in modulating its

activity. This regulation depends primarily on an alpha-helix in the C-terminus of the

protein which binds to the ATPase domain and inhibits Pma1p activity in the absence of

glucose [12-14]. The addition of glucose stimulates phosphorylation of three residues

(Ser-899, Ser-911, Thr-912) in the C-terminal inhibitory domain triggering its release

from the ATPase domain coincident with an increase in the Vmax for ATP hydrolysis and

decrease in the Km for ATP [14-17]. Analysis of mutations at these residues, either to

Ala to prevent, or Asp to mimic phosphorylation, has suggested a unique role for each

phosphorylation site in glucose-induced Pma1p activation (see Table 3-1).

Recently our lab demonstrated that glucose activation of S. cerevisiae Pma1p is

inhibited by mutations in SOD1 encoding the largely cytoplasmic Cu/Zn superoxide

Page 126: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

109

dismutase [18]. The mechanism was not fully understood. Sod1p catalyzes the

disproportionation of superoxide to hydrogen peroxide and molecular oxygen and is thus

best known for its role in cellular antioxidant defense [19-22]. However, Sod1p can also

function in cell signaling independent of its role in oxidative stress protection. For

example, in S. cerevisiae the peroxide produced by Sod1p can stabilize the casein kinase

I homologue needed for glucose signaling, Yck1p [18]. Pma1p is one known target of

Yck1p [18,23,24], suggesting that the aforementioned effect of sod1∆ on Pma1p may

involve loss of Yck1p. However, we could not exclude the possible contribution of

oxidative stress protection in Sod1p control of Pma1p.

The connection between Sod1p and Pma1p was explored further in the current

study using mutants of Pma1p known to disrupt glucose control and/or ATPase activity.

These Pma1p mutations targeted the phosphorylation sites at the C-terminus, as well as,

potential interacting residues in the ATPase domain. Out of 14 Pma1p alleles examined,

one variant exhibited a striking “synthetic” interaction [25,26] with sod1∆. Specifically,

the T912D substitution in the C-terminal inhibitory domain of Pma1p was synthetically

lethal with the sod1∆ null mutation, i.e. the double mutant was not viable in air. This

synthetic lethality was reversed by second-site suppressor mutations in Pma1p that

hyperactivate the ATPase, indicating the lethality results from profound loss in activity of

the essential Pma1p proton ATPase. The synthetic lethality was also reversed under

anaerobic conditions and by antioxidant treatments that reverse oxidative damage of

sod1∆ strains but do not rescue the sod1∆ loss of Yck1p. Together, our data support a

model in which Sod1p maintains Pma1p activity through both oxidative stress-dependent

and -independent (Yck1p signaling) pathways.

Page 127: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

110

2. Materials and Methods

2.1. Yeast Growth, Yeast Strains and Plasmids

S. cerevisiae yeast strains were maintained at 30°C either in an enriched YPD

(yeast extract, peptone, 2% dextrose) or a minimal synthetic complete (SC) medium

devoid of lysine where indicated. Anaerobic growth was carried out using the GasPak

EZ Anaerobe Container System (Becton Dickinson) in medium supplemented with 15

mg/L ergosterol and 0.5% Tween 80 (YPDE). For plate-based growth tests, 104, 103, and

102 cells were spotted onto the designated medium and growth proceeded at 30°C for

three to five days. For liquid growth assays, 106 cells/ml were inoculated into YPD with

or without the concentrations of drug indicated (Fig. 3-6A and B) and grown shaking at

30°C for 18 hrs (paraquat) or 15 hrs (menadione) after which cell growth was determined

turbidimetrically at 600 nm using a spectrophotometer.

The S. cerevisiae strains used in this study are either of the BY4741 (MATa

his3∆1 leu2∆0 met15∆0 ura3∆0) or NY13 (MATa ura3–52) backgrounds and can be

found in Tables 3-2 and 3-3. Pma1p mutants in the NY13 background were generous

gifts of Ken Allen and Carolyn Slayman and where indicated, sod1∆ mutations were

introduced using the sod1::KanMX4 plasmid, pJAB002, as described [27]. The Pma1-

T912D sod1∆ strain, JAB020, is only viable under anaerobic conditions, but aerobically

viable suppressors with intragenic mutations in PMA1 spontaneously appeared during

yeast transformation procedures (e.g., A906G, strain JAB142) or could be isolated at a

frequency of 105 by plating 107 cells for two days in air on YPD media (e.g., T540A,

strain JAB124; L586V, strain JAB125 and F666L, strain JAB126). Site specific single or

double mutations in PMA1 were introduced in BY4741 or the isogenic sod∆1::KanMX or

Page 128: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

111

akr1∆::KanMX (Open Biosystems) strains by gene replacement using derivatives of the

URA3 vector pGW201 vector expressing specific alleles of PMA1. To create the Pma1-

T912D/A906G sod1∆ strain, JAB140, the sod1::LEU2 plasmid, pKS1, was introduced as

described previously [28] into strain JC20. It is noteworthy that in the BY4741

background, the single Pma1-T912A (JC02) and Pma1-T912D (JC03) strains developed

rho- mutations at a very high frequency, but not in the background of the isogenic sod1∆

or akr1∆ strains. Tests for rho- mutations were conducted by plating cells on YP

medium containing 3% glycerol rather than 2% glucose and incubating at 5% oxygen that

is sufficient to drive mitochondrial respiration but is not lethal to the oxygen-sensitive

T912D sod1∆ strain.

The plasmids used in this study are outlined in Table 3-4. The vectors for

generating site-specific substitutions in PMA1 were all from the phagemid pGW201 and

contained PMA1 mutations introduced by Quikchange site-directed mutagenesis

(Stratagene). pGW201 [29] contains a PMA1 cassette harboring URA3 at the 3′ non-

coding end that is liberated by HindIII digestion. After transformation into yeast,

successful introduction of the desired pma1 allele was verified by sequencing of the C-

terminal PMA1 coding region. The sod1:LEU2 and sod1::KanMX vectors were as

previously described [27,28].

2.2. Biochemical and Microscopy Analyses

For western blots of Pma1p, cells were grown at 30°C in YPD to mid-log phase.

Cells were harvested, washed in water, and resuspended in buffer comprised of 20 mM

potassium phosphate, 1.2 M sorbitol, 1% Triton X-100, 1.0 mM PMSF, 1:100 protease

inhibitor cocktail (P8340, Sigma), pH 7.4. Cells were lysed in the presence of zirconium

Page 129: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

112

oxide beads by vortexing twice for 2 mins at 50 Hz in a TissueLyzer LT (Qiagen).

Lysates were clarified by centrifugation and 20 μg total protein was then prepared and

loaded onto NuPAGE Novex 4-12% Bis-Tris Gels (Life Technologies) according to the

manufacturer’s instructions, except lysates were incubated at room temperature for 15

mins. Gel electrophoresis was carried out in MOPS buffer according to the

manufacturer’s instructions and then transferred to a nitrocellulose membrane with the

iBlot Gel Transfer Device (Life Technologies) using the standard settings. Membranes

were probed with anti-Pma1p primary (1:5000, Abcam, ab4645) and Alexa Fluor 680

Donkey Anti-Mouse IgG secondary (1:5,000, Life Technologies, A10038).

For Pma1p activity assays, a culture of 4 x 106 cells was seeded into 1L YPD and

grown to mid-log phase at 30°C. Cells were washed once with water and pellets stored at

-80°C until ready for membrane isolation. Plasma membranes were isolated essentially

as described [18,30,31]. Briefly, pellets were resuspended in 25 ml ice cold

homogenization buffer (50 mM Tris-HCl, pH 7.5, 0.3 M sucrose, 5 mM EDTA, 1 mM

EGTA, 5 mg/ml BSA, 2 mM DTT, 1 mM PMSF, 1:100 protease inhibitor cocktail) and

passed twice through a French press at ≈ 20,000 psi for lysis. Lysates were clarified by

centrifugation at 3,500×g for 5 mins and then at 14,000×g for 10 mins. Membranes were

then pelleted by centrifugation at 200,000×g, washed once by dounce homogenization in

a wash buffer containing 10 mM Tris-HCl, pH 7.0, 1 mM EGTA, 10% glycerol, 1 mM

PMSF, 1:100 protease inhibitor cocktail and harvested again by centrifugation as above.

Membranes were resuspended in 1 ml wash buffer by dounce homogenization and stored

at -80°C. After thawing, protein concentrations were estimated using the Bradford assay

and Pma1p activity was determined by the orthovanadate-sensitive phosphate release

Page 130: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

113

from ATP following published protocols [18,31,32]. Basically, 10-50 μg purified

membrane protein in 20 μl membrane wash buffer was added to 120 μl Pma1p activity

buffer (50 mM MES-Tris, pH 6.0, 15 mM MgSO4, 50 mM KNO3, 0.2 mM ammonium

molybdate, 5 mM NaN3, 3 mM ATP) with or without 100 μM orthovanadate and

incubated for 10 mins at room temperature in a 96-well plate. The reaction was stopped

by the addition of 140 μl of the developing reagent containing 1% SDS, 100 mM sodium

molybdate, 0.6 M H2SO4, and 0.8% ascorbate. The A620 was recorded within 5 mins after

addition of the developing reagent and activity was subsequently normalized to WT.

Indirect immunofluorescence was performed largely as described [33,34]. Briefly,

cells were grown in YPD to OD600 = 1.0 and then fixed with 4.4% formaldehyde added

directly to the shaking culture for 30 mins. The cells were then washed once with 100

mM potassium phosphate, pH 6.5 (KPhos buffer) and then fixed a second time overnight

in KPhos buffer with 4% formaldehyde. Cells were washed once in KPhos buffer and

spheroplasts were generated by incubation in 200 mM Tris-HCl, pH 8.0, 20 mM EDTA,

1% 2-mercaptoethanol for 10 mins, followed by a second incubation in 1.2 M sorbitol,

100 mM potassium phosphate, pH 6.5 with 1.5 mg/ml Zymolyase 20T for 30-60 mins.

Cells were washed once in 1.2 M sorbitol and permeabilized in 1.2 M sorbitol, 2% SDS

for 2 mins. Cells were then washed twice in 1.2 M sorbitol, allowed to adhere to poly-

lysine coated slides to form a monolayer, and then treated with primary (1:25 anti-Pma1p

mouse monoclonal [40B7]; Abcam, ab4645) and secondary antibodies (1:250 Goat Anti-

Mouse Alexa Fluor 488; Molecular Probes, A11001) in PBS with 1% BSA. After final

aspiration, mounting solution containing 2.5% DABCO, 100 mM Tris HCl, pH8.8, 50%

Glycerol, 0.2 μg/ml DAPI was added and slides sealed. Images were taken on a Zeiss

Page 131: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

114

Observer.Z1 fluorescence microscope with an Apotome VH optical sectioning grid

(Zeiss, Gena, Germany) under 100X magnification.

3. Results

3.1. Pma1p Mutants and Sod1p Deficiency

We previously reported that Pma1p activity is low in S. cerevisiae sod1∆ mutants

based on assays of isolated membranes [18]. One possible mechanism might involve

mis-sorting of Pma1p in the secretory pathway, as has been shown for yeast mutants

defective in vacuolar function [33]. Since S. cerevisiae sod1∆ cells have abnormal

vacuoles [35], we tested whether loss in Pma1p activity might reflect Pma1p mis-sorting.

By indirect immunofluorescence microscopy, Pma1p in WT cells is predominantly

localized to the cell-surface as seen in Fig. 3-1A. This precise pattern of Pma1p

localization was mirrored in sod1∆ cells. Furthermore, the intensity of Pma1p staining

was comparable in WT and sod1∆ cells (Fig. 3-1A), consistent with normal Pma1p

polypeptide levels in sod1∆ cells [18]. Thus, sod1∆ mutations do not alter Pma1p protein

levels or localization. Pma1p ATPase activity at the cell surface must be affected.

Pma1p activity at the cell surface is normally controlled by glucose and we sought

to determine the effects of sod1∆ mutations when such glucose control of Pma1p has

been altered. To this end, we utilized mutant alleles of Pma1p known to affect glucose

control and activity of the ATPase. These alleles target the C-terminal inhibitory and

interacting ATPase domains (Fig. 3-1B) and include Ala and phospho-mimic Asp

substitutions at the phosphorylation sites S899, S911, and T912, and Cys scanning

substitutions in stalk segment 5 that increase or decrease Pma1p activity as summarized

Page 132: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

115

in Table 3-1 [36,37]. Strains expressing these various Pma1p alleles in combination with

sod1∆ were initially created under anaerobic conditions to minimize sod1∆ oxidative

damage. When tested for growth in atmospheric oxygen, we were surprised to find a

strong “synthetic” effect [25,26] when sod1∆ mutations were combined with a specific

allele of Pma1p. Specifically, the Pma1-T912D mutant was not viable when combined

with sod1∆ mutations (Figs. 3-1C and D). The results were reproducible in two distinct

strain backgrounds and the synthetic lethality was only seen with oxygen, as these same

strains grew well under anaerobic conditions (Figs. 3-1C and D). For the purposes of this

study, this synthetic lethal mutant will be referred to as T912D sod1∆.

The lethal effect of combining sod1∆ and Pma1p mutations appeared highly

specific to T912D sod1∆. As seen in Fig. 3-1C and Fig. 3-2, 13 out of 14 Pma1p mutants

in the C-terminal inhibitory and ATPase domain showed no evidence of a synthetic effect

with sod1∆, i.e., there was no obvious changes in cell growth of the double mutant

compared to the single Pma1p or sod1∆ mutant. Even Pma1-T912A showed no

inhibitory effect when combined with sod1∆; the lethality appeared unique to Pma1-

T912D (Fig. 3-1C). In the course of our studies, we noted that single T912D mutants of

Pma1p have a high tendency to accumulate rho- mutations, or losses of mitochondrial

DNA, particularly in the BY4741 strain background (see Materials and Methods).

However, T912A Pma1p mutants also have a high tendency to go rho- but are not lethal

with sod1∆, and more importantly, all of our T912D sod1∆ mutants are RHO+ (see

Materials and Methods). Thus the synthetic lethality of T912D sod1∆ cannot be

attributed to loss of mitochondrial DNA and other mechanisms must be involved.

3.2. Suppressors of the T912D sod1Δ Aerobic Lethality

Page 133: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

116

To help understand the basis for the T912D sod1Δ lethality, we used a genetic

suppressor approach. After prolonged incubation in air, viable colonies of T912D sod1Δ

could be identified at a frequency of ≈105. When isolated and retested, those colonies

retained full aerobic growth (Fig. 3-3A, representative sample), indicating stable genetic

suppression of either Sod1- or Pma1-deficiency. To decipher between these two, we

tested for aerobic lysine auxotrophy, a hallmark of Sod1p deficiency [38]. Mutants of

sod1∆ cannot grow aerobically without lysine due to oxidative damage to the lysine

biosynthetic pathway, and previously characterized genetic suppressors of sod1∆ (e.g.,

pmr1∆, bsd2∆) correct this lysine defect [39-44]. However, none of the suppressors of

T912D sod1∆ could grow aerobically without lysine (Fig. 3-3B), suggesting suppression

reversed the deficiency of Pma1p, not Sod1p. To test for intragenic suppressors in

Pma1p, the entire PMA1 gene was sequenced in four representative suppressors and each

was found to have a distinct additional missense mutation within the PMA1 coding region

(Fig. 3-1B). F666L lies within the aforementioned Pma1p stalk region and in fact F666C

was previously shown by Miranda and colleagues to result in constitutive activation of

Pma1p [36]. The T540A, L586V mutations lie within the ATPase domain and A906G is

within the C-terminal inhibitory domain (Fig. 3-1B). These findings are highly

reminiscent of intragenic suppressors previously reported for the Pma1-S911A/T912A

double mutant. Like T912D sod1∆, the double Pma1-S911A/T912A is a synthetic lethal

mutant, and this lethality is rescued by mutations in the Pma1p ATPase domain (P536L,

A565T, G587N, G648S), the stalk domain (P669L, G670S) and the C-terminus (M907I)

[16,17]. M907I is in close juxtaposition to the A906G suppressor identified here and was

Page 134: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

117

reported by Eraso and Portillo [17] to be a hyperactive Pma1p allele. We therefore chose

Pma1-A906G for further analysis.

To confirm that Pma1-A906G by itself suppresses T912D sod1∆, we introduced

this identical mutation into an independent strain background. As seen in Fig. 3-3C,

A906G was capable of restoring aerobic growth to the T912D sod1∆ strain in the

BY4741 background and as before (Fig. 3-3B), this reflects specific suppression of

Pma1p and not Sod1p deficiency, since the strain remains incapable of aerobic growth

without lysine (Fig. 3-3C). To explore how A906G suppresses T912D, we examined the

polypeptide levels, cellular localization and activity of Pma1-T912D with and without the

A906G suppressor. As seen in Fig. 3-4, the A906G mutation significantly enhances

Pma1p ATPase activity without changing Pma1p polypeptide levels or its localization at

the cell surface. In fact, activity with the Pma1-T912D/A906G double mutant is even

higher than that of WT Pma1p (Fig. 3-4C). A906G behaves as a hyperactive allele of

Pma1p, as has been reported for its close neighbor M907I [17]. Thus, the synthetic

lethality of T912D sod1∆ can be rescued by hyperactivation of Pma1p. These findings

strongly implicate profound loss of essential Pma1p activity as the cause of death in the

T912D sod1∆ strain.

3.3. Oxidative Damage, Not Loss of Yck1p Signaling Causes the Synthetic Lethality of

T912D sod1Δ

Our previous studies suggested that inhibition of Pma1p activity in sod1∆ cells

may reflect defective signaling by casein kinases. Sod1p stabilizes the yeast casein

kinase I homologue, Yck1p, which in turn regulates Pma1p [18,23,24]. Activity of

Yck1p at the plasma membrane can be abolished by deleting the palmitoyl transferase

Page 135: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

118

akr1∆ [45,46], and like sod1∆ mutations, akr1∆ mutants have low Pma1p activity [18].

We therefore predicted that akr1∆ mutations would phenocopy sod1∆ mutations and

show synthetic lethality with Pma1-T912D. However, in contrast to T912D sod1Δ, the

T912D akr1Δ double mutant was quite viable under all conditions tested (Fig. 3-5A).

This unexpected viability did not reflect rapid accumulation of genetic suppressors in this

strain. There were no second-site mutations in PMA1, as determined by complete

sequence analysis of the gene, and there was no obvious suppression of akr1∆, as the

large cell size and elongated bud morphology of akr1∆ remained intact (data not shown)

[46]. Since akr1∆ does not phenocopy sod1∆ when combined with Pma1-T912D, the

synthetic lethality of T912D sod1Δ may not be due to loss of Yck1p activation.

Aside from its role in regulating Yck1p, S. cerevisiae Sod1p is also important for

oxidative stress protection [47-49] and the oxidative damage of sod1∆ cells can be

reversed by growing cells under anaerobic conditions or by supplying high levels of non-

SOD antioxidants, such as reactive manganese complexes that can scavenge superoxide

[50-55]. Neither of these treatments will correct the Yck1p-defect of sod1∆ cells which

is independent of oxidative stress [18]. The synthetic lethality of T912D sod1∆ strains is

completely reversed by anaerobic conditions (Figs. 3-1C, 3-1D, and 3-3A), strongly

indicative of oxidative stress. The strain begins to grow when oxygen levels drop to 5%

(Fig. 3-5B). Furthermore, the synthetic lethality of T912D sod1 could be reversed by

treatment with high levels of manganese, which form non-proteinaceous antioxidants

(Fig. 3-5C) [41,50-55]. Together, these observations strongly indicate that oxidative

damage from sod1∆, rather than loss of Yck1p casein kinase signaling, contributes to the

synthetic lethality of T912D sod1∆. The combined effect of this damage together with

Page 136: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

119

Pma1-T912D must sufficiently inhibit essential Pma1p activity to halt S. cerevisiae

growth.

4. Discussion

The cytosolic Cu/Zn-SOD enzyme can broadly impact cell metabolism and

growth. On one hand, Sod1p eliminates superoxide that can cause oxidative damage to

biomolecules, while on the other hand, Sod1p generation of peroxide can act in signaling

pathways independent of oxidative stress protection [18,56]. It was previously shown

that in S. cerevisiae, Sod1p can signal glucose repression through a mechanism involving

peroxide stabilization of Yck1p, and one downstream target of this signal is the proton

ATPase Pma1p [18]. We now provide evidence that Sod1-control of Pma1p extends

beyond Yck1p. Specifically, we find that through certain oxidative damage effects,

sod1∆ mutations are lethal to cells expressing Pma1p with a T912D mutation in the C-

terminal inhibitory domain. This synthetic lethality must be independent of Yck1p, as

Pma1-T912D is without consequence when Yck1p is inactivated in SOD1+ strains (e.g.,

by akr1∆ mutations). Furthermore, the T912D sod1∆ lethality is reversed by anti-oxidant

treatments that do not rescue the sod1∆ defect in Yck1p. We propose that Sod1p helps to

maintain Pma1p in the active state in two distinct ways: one dependent on Yck1p-

glucose signaling and the other via protection against oxidative damage.

Why are T912D sod1∆ mutants inviable in air? Since lethality is reversed by

hyper-activating mutations in Pma1p (e.g., A906G), death likely results from profound

inhibition of the essential proton ATPase Pma1p. As an alternative, lethality might

reflect severe oxidative damage to the cell from the combined effects of sod1∆ and

Page 137: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

120

T912D PMA1 mutations. However, we have no evidence that Pma1-T912D mutations by

themselves cause oxidative damage. The single Pma1-T912D mutants are not more

sensitive to the environmental oxidant paraquat or to conditions of hyperoxia that are

lethal to sod1∆ strains (Fig. 3-6A,C). We did observe a slight sensitivity towards the

oxidant menadione compared to WT Pma1p strains, but this effect was orders of

magnitude diminished compared to the menadione sensitivity of sod1∆ strains (Fig. 3-

6B). We therefore favor a model in which profound inhibition of the essential Pma1p

ATPase is the cause of the aerobic lethality of T912D sod1∆ strains.

Of all the Pma1p mutants tested, only T912D showed synthetic lethality with

sod1∆; even T912A was without effect. In some studies, the T912D mutation has been

reported to inhibit the ATPase but the defect in Pma1p caused by T912A was more

severe [14,17] (Table 3-1). Thus the allele specificity for synthetic lethality cannot be

explained simply by T912D inhibition of the ATPase. Instead, effects on protein

conformation may be key. Based on trypsin digestion experiments by Lecchi and

colleagues [14], Pma1-T912D is predicted to adopt a more open conformation compared

to Pma1-WT or -T912A (summarized in Table 3-1). WT Pma1p has been reported to be

relatively resistant to oxidative damage compared to other P-type ATPases [57], but the

conformation of Pma1-T912D may render this transporter hypersensitive to such insults.

Pma1p has nine cysteine residues including five in the catalytic domain and modification

of these can inhibit the ATPase [58]. It will be of interest to study the effect of T912D

and sod1∆ mutations on the oxidation status of these cysteines as a potential target for

oxidative damage.

Page 138: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

121

While direct oxidative damage of Pma1p is one possibility, it is also possible that

sod1∆-oxidative stress can affect Pma1p indirectly by damaging one of the Pma1p

regulatory proteins. Pma1p regulation is complex and numerous factors have been

implicated in its regulation, including those for ubiquitination and protein degradation

(RSP5, UBC4, proteasome) [59], glucose sensing (SNF3, GPA2) [60], phosphorylation

(Yck1p, Ptk2, Hxk1, Hxk2, Glk1, etc.) [18,23,24,61] and calcium signaling through

calcineurin [62]. In this regard, Sod1p has been shown to bind to and protect calcineurin

from inactivation [63]. Regardless of whether Sod1p is protecting Pma1p directly or

indirectly, these studies underscore the multiple ways this antioxidant enzyme can affect

growth and metabolism of aerobic cells.

Page 139: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

122

TABLE 3-1. Effect of Pma1 mutations on activity and C-terminal conformation

The effect of mutations used in this study on Pma1p activity and C-terminal

conformation was compiled from the literature cited. For Pma1p activity, “↑” indicates

increased activity, “↓” decreased activity, and “↓↓” severely decreased activity relative to

Pma1-WT in the same study. The C-terminal conformation of Pma1p mutants is based

on the ability of trypsin to cleave off the C-terminus of Pma1p, as performed by Lecchi

and colleagues in the presence of glucose [14]). Here, “closed” and “open” refer to

constitutive states unaffected by glucose availability, whereas “≈ WT” indicates the

mutations retain glucose-responsive conformational changes. N.D. = not determined

Pma1 Mutant Activity* C-terminal Conformation References

3CHT ≈ WT N.D. [36] V665C ↓ N.D. [36] A668C ↑ N.D. [36] D676C ↓ N.D. [36] L678C ↓ N.D. [36]

I684C ↑ ATPase, ↓ H+ transport N.D. [36]

H686C ↓ N.D. [36] R687C ↓ N.D. [36] Y689C ↑ N.D. [36] S899A ≈ WT or ↓ ≈ WT [14,17] S899D ≈ WT or ↑ ≈ WT [14,17] S911A ≈ WT N.D. [16] S911D N.D. N.D. - T912A ↓↓ closed [14,16,17,64] T912D ≈ WT or ↓ open [14,17]

* Primarily refers to ATPase activity but may also reflect H+-pumping activity where

measured. Unless otherwise noted, the identified mutation affected ATPase and H+-

pumping activity in a similar manner.

Page 140: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

123

TABLE 3-2. BY4741 strains used in this study

Strain Parent Strain Partial Genotype Plasmid Reference

BY4741 - MATa his3Δ1 leu2Δ0 met15Δ0 ura3Δ0 - [65] sod1Δ BY4741 sod1::kanMX - [66] akr1Δ BY4741 akr1::kanMX - [66] JC02 BY4741 PMA1-T912A (URA3), rho- pJC001 This study JC03 BY4741 PMA1-T912D (URA3), rho- pJC002 This study JC04 BY4741 PMA1-S911A (URA3) pJC003 This study JC05 BY4741 PMA1-S911D (URA3) pJC004 This study JC06 BY4741 PMA1-S899A (URA3) pJC005 This study JC07 BY4741 PMA1-S899D (URA3) pJC006 This study JC08 BY4741 PMA1-A906G (URA3) pJC007 This study JC09 sod1Δ sod1::kanMX, PMA1-T912A (URA3) pJC001 This study JC10 sod1Δ sod1::kanMX, PMA1-T912D (URA3) pJC002 This study JC11 sod1Δ sod1::kanMX, PMA1-S911A (URA3) pJC003 This study JC12 sod1Δ sod1::kanMX, PMA1-S911D (URA3) pJC004 This study JC13 sod1Δ sod1::kanMX, PMA1-S899A (URA3) pJC005 This study JC14 sod1Δ sod1::kanMX, PMA1-S899D (URA3) pJC006 This study JC15 sod1Δ s sod1::kanMX, PMA1-A906G (URA3) pJC007 This study JC20 BY4741 PMA1-T912D/A906G (URA3) pJAB008 This study JC61 BY4741 akr1::kanMX, PMA1-T912D (URA3) pJC002 This study JAB140 JC20 PMA1-T912D/A906G (URA3), sod1::LEU2 pKS1 This study

Page 141: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

124

TABLE 3-3. NY13 strains used in this study

Strain Parent Strain Partial Genotype Plasmid Reference PMA1-WT NY13 PMA1-WT (URA3) - [14] PMA1-T912D NY13 PMA1-T912D (URA3) - [14] PMA1-3CHT NY13 10His-PMA1-3C (lacks all Cys

residues except C376, C409, & C472) - [37]

PMA1-V665C NY13 PMA1-V665C (URA3) - [36] PMA1-A668C NY13 PMA1-A668C (URA3) - [36] PMA1-D676C NY13 PMA1-D676C (URA3) - [36] PMA1-L678C NY13 PMA1-L678C (URA3) - [36] PMA1-I684C NY13 PMA1-I684C (URA3) - [36] PMA1-H686C NY13 PMA1-H686C (URA3) - [36] PMA1-R687C NY13 PMA1-R687C (URA3) - [36] PMA1-Y689C NY13 PMA1-Y689C (URA3) - [36] JAB020-1 PMA1-T912D PMA1-T912D (URA3), sod1::kanMX pJAB002 This study JAB021 PMA1-WT PMA1-WT (URA3), sod1::kanMX pJAB002 This study JAB124 JAB020-1 PMA1-T912D/T540A (URA3),

sod1::kanMX - This study

JAB125 JAB020-1 PMA1-T912D/L586V (URA3), sod1::kanMX

- This study

JAB126 JAB020-1 PMA1-T912D/F666L (URA3), sod1::kanMX

- This study

JAB142 PMA1-T912D PMA1-T912D/A906G (URA3), sod1::kanMX

- This study

JC22 PMA1-3CHT PMA1-3CHT, sod1::kanMX pJAB002 This study JC23 PMA1-L678C PMA1-L678C, sod1::kanMX pJAB002 This study JC24 PMA1-Y689C PMA1-Y689C, sod1::kanMX pJAB002 This study JC25 PMA1-A668C PMA1-A668C, sod1::kanMX pJAB002 This study JC27 PMA1-D676C PMA1-D676C, sod1::kanMX pJAB002 This study JC28 PMA1-H686C PMA1-H686C, sod1::kanMX pJAB002 This study JC29 PMA1-I684C PMA1-I684C, sod1::kanMX pJAB002 This study JC30 PMA1-V665C PMA1-V665C, sod1::kanMX pJAB002 This study JC31 PMA1-R687C PMA1-R687C, sod1::kanMX pJAB002 This study

Page 142: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

125

TABLE 3-4. Plasmids used in this study

Plasmid Description Digest* Reference pGW201 Pma1-WT (URA3) HindIII [29] pJAB002 sod1::kanMX BamHI [27] pJAB008 Pma1-T912D/A906G (URA3) HindIII This Study pJC001 Pma1-T912A (URA3) HindIII This Study pJC002 Pma1-T912D (URA3) HindIII This Study pJC003 Pma1-S911A (URA3) HindIII This Study pJC004 Pma1-S911D (URA3) HindIII This Study pJC005 Pma1-S899A (URA3) HindIII This Study pJC006 Pma1-S899D (URA3) HindIII This Study pJC007 Pma1-A906G (URA3) HindIII This Study pKS1 sod1::LEU2 PstI [28]

*Digest indicates the restriction enzyme used to linearize the cassette prior to

transformation to ensure proper integration.

Page 143: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

126

FIGURE 3-1. Effect of sod1∆ on Pma1p localization and growth of Pma1p mutants

with altered glucose regulation

(A) Indirect immunofluorescence of endogenous Pma1p was performed on WT and

sod1∆ cells in the BY4741 background as described in Materials and Methods. (B)

Topology map of Pma1p based on a map previously published by Lecchi, et al. [14] with

circles representing individual amino acids. The N- and C-termini are designated by “N”

and “C”, respectively. Red circles indicate the location of mutants in stalk segment 5 and

the C-terminus that were tested for synthetic lethality with sod1∆ in parts C, D, and Fig.

3-2. The specific effects of these mutants on Pma1p activity can be found in Table 3-1.

Green circles and bold type denote the Pma1p residues in the central ATPase region and

C-terminus that were identified as secondary mutations by sequencing in strains with

suppressors of the T912D sod1∆ synthetic lethality (see Fig. 3-3). (C, D) Plate-based

growth tests were performed with 104, 103, and 102 cells spotted onto YPD (aerobic) or

YPDE (anaerobic) and incubated at 30°C for 5 days. Plates were maintained

anaerobically using the GasPak EZ Anaerobe Container System. (C) pma1 single and

pma1sod1∆ double mutants in the BY4741 strain background. (D) pma1 single and pma1

sod1∆ double mutants in the NY13 background.

Page 144: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

127

Page 145: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

128

FIGURE 3-2. Mutations in Pma1p stalk segment 5 that alter glucose regulation are

not synthetically lethal with sod1∆

sod1∆ was introduced into the Pma1p stalk segment 5 mutants indicated (NY13

background) by transformation of pJAB002 [27] and tested for growth aerobically and

anaerobically as described for Fig. 3-1C,D.

Page 146: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

129

FIGURE 3-3. T912D sod1∆ cells develop suppressors with secondary mutations in

PMA1

Individual colonies of T912D sod1∆ in the NY13 strain background that grew aerobically

were isolated (see Materials and Methods) and along with appropriate controls were (A)

retested for their ability to grow aerobically as described for Fig. 3-1C,D or (B) tested to

determine if they suppress the sod1∆ lysine auxotrophy by spotting onto SC media

(+Lys) or SC media lacking Lys (-Lys) and incubating at 30°C for 3 days. Four

representative suppressors are shown. Each had a unique secondary mutation in the

Pma1p coding region as indicated. (C) In the BY4741 background, the WT PMA1 gene

was replaced with PMA1-A906G or PMA1-T912D/A906G in both WT and sod1∆ strains

as described in Materials and Methods and tested for lysine auxotrophy alongside

appropriate controls as described for part B.

Page 147: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

130

Page 148: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

131

FIGURE 3-4. The PMA1-A906G mutation increases Pma1p ATPase activity but

does not alter Pma1p localization or protein level

The strains indicated were grown at 30°C shaking in YPD to mid-log phase and tested by

(A) indirect immunofluorescence to determine Pma1p localization and (B) western blot

to determine total cellular Pma1p protein level using a monoclonal antibody against

Pma1p as described in Materials and Methods. (C) Pma1p activity was determined in

membranes prepared from cells grown in the same manner by measuring the

orthovanadate-sensitive ATPase activity as described in Materials and Methods. Values

shown represent the average of two biological replicates normalized to WT; error bars

indicate standard error. All strains in this figure are in the BY4741 background.

Page 149: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

132

FIGURE 3-5. T912D sod1∆ synthetic lethality is due to oxidative stress and not

Yck1p misregulation

(A) Pma1-WT was replaced with PMA1-T912D in WT and akr1∆ null cells in the

BY4741 background and tested for aerobic and anaerobic growth as described in Fig. 3-

1C,D. (B) 104, 103, and 102 cells of each strain indicated (NY13 background) were

spotted onto YPDE (0% oxygen) or YPD (1-20% oxygen) and incubated in the oxygen

concentration specified for 3 days at 30°C. (C) Cells were spotted and grown

anaerobically on YPDE or aerobically on YPD with or without 1 mM MnCl2 as described

in Fig. 3-1C,D.

Page 150: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

133

Page 151: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

134

FIGURE 3-6. Pma1-T912D mutant cells are not sensitive to environmental oxidants

or hyperoxia

(A, B) 106 cells/ml were inoculated into liquid YPD containing the concentration of

paraquat (A) or menadione (B) indicated and shaken at 30°C for either 18 hrs (A) or 15

hrs (B). (B) Growth in menadione is graphed on a log scale. Error bars indicate standard

error of duplicate biological measurements. (C) 104, 103, and 102 cells of the strains

indicated (NY13 background) were plated onto YPD and incubated in air or at 100% O2

or onto YPDE and incubated anaerobically using the GasPak system, as described in

Materials and Methods, at 30°C for 2 days.

Page 152: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

135

Page 153: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

136

5. Reference List

[1] F. Portillo, Regulation of plasma membrane H+-ATPase in fungi and plants,

Biochim. Biophys. Acta, Rev. Biomembr. 1469 (2000) 31-42.

[2] F. Portillo, R. Serrano, Growth control strength and active site of yeast plasma

membrane ATPase studied by site-directed mutagenesis, Eur. J. Biochem. 186 (1989)

501-507.

[3] R. Rao, D. Drummond-Barbosa, C.W. Slayman, Transcriptional regulation by glucose

of the yeast PMA1 gene encoding the plasma membrane H(+)-ATPase, Yeast. 9 (1993)

1075-1084.

[4] B. Panaretou, P.W. Piper, Plasma-membrane ATPase action affects several stress

tolerances of Saccharomyces cerevisiae and Schizosaccharomyces pombe as well as the

extent and duration of the heat shock response, J. Gen. Microbiol. 136 (1990) 1763-

1770.

[5] B. Benito, F. Portillo, R. Lagunas, In vivo activation of the yeast plasma membrane

ATPase during nitrogen starvation Identification of the regulatory domain that controls

activation, FEBS Lett. 300 (1992) 271-274.

[6] G.A. Monteiro, I. Sá-Correia, In vivo activation of yeast plasma membrane H+-

ATPase by ethanol: effect on the kinetic parameters and involvement of the carboxyl-

terminus regulatory domain, Biochim. Biophys. Acta, Biomembr. 1370 (1998) 310-316.

Page 154: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

137

[7] M. Stratford, G. Nebe-von-Caron, H. Steels, M. Novodvorska, J. Ueckert, D.B.

Archer, Weak-acid preservatives: pH and proton movements in the yeast Saccharomyces

cerevisiae, Int. J. Food Microbiol. 161 (2013) 164-171.

[8] R. Serrano, In vivo glucose activation of the yeast plasma membrane ATPase, FEBS

Lett. 156 (1983) 11-14.

[9] M. García-Arranz, A.M. Maldonado, M.J. Mazón, F. Portillo, Transcriptional control

of yeast plasma membrane H(+)-ATPase by glucose. Cloning and characterization of a

new gene involved in this regulation. J. Biol. Chem. 269 (1994) 18076-18082.

[10] T. Ferreira, A.B. Mason, C.W. Slayman, The yeast Pma1 proton pump: a model for

understanding the biogenesis of plasma membrane proteins, J. Biol. Chem. 276 (2001)

29613-29616.

[11] B. Benito, E. Moreno, R. Lagunas, Half-life of the plasma membrane ATPase and its

activating system in resting yeast cells, Biochim. Biophys. Acta, Biomembr. 1063

(1991) 265-268.

[12] F. Portillo, I.F. de Larrinoa, R. Serrano, Deletion analysis of yeast plasma membrane

H+ -ATPase and identification of a regulatory domain at the carboxyl-terminus, FEBS

Lett. 247 (1989) 381-385.

[13] R. Serrano, F. Portillo, Catalytic and regulatory sites of yeast plasma membrane H+-

ATPase studied by directed mutagenesis, Biochim. Biophys. Acta, Bioenerg. 1018

(1990) 195-199.

Page 155: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

138

[14] S. Lecchi, K.E. Allen, J.P. Pardo, A.B. Mason, C.W. Slayman, Conformational

changes of yeast plasma membrane H(+)-ATPase during activation by glucose: role of

threonine-912 in the carboxy-terminal tail, Biochemistry. 44 (2005) 16624-16632.

[15] S. Lecchi, C.J. Nelson, K.E. Allen, D.L. Swaney, K.L. Thompson, J.J. Coon, M.R.

Sussman, C.W. Slayman, Tandem phosphorylation of Ser-911 and Thr-912 at the C

terminus of yeast plasma membrane H+-ATPase leads to glucose-dependent activation,

J. Biol. Chem. 282 (2007) 35471-35481.

[16] F. Portillo, P. Eraso, R. Serrano, Analysis of the regulatory domain of yeast plasma

membrane H+-ATPase by directed mutagenesis and intragenic suppression, FEBS Lett.

287 (1991) 71-74.

[17] P. Eraso, F. Portillo, Molecular mechanism of regulation of yeast plasma membrane

H(+)-ATPase by glucose. Interaction between domains and identification of new

regulatory sites. J. Biol. Chem. 269 (1994) 10393-10399.

[18] A. Reddi, V. Culotta, SOD1 Integrates Signals from Oxygen and Glucose to Repress

Respiration, Cell. 152 (2013) 224-235.

[19] J.M. McCord, I. Fridovich, Superoxide

Dismutase: An enzymic function for erythrocuprein (hemocuprein). J. Biol. Chem. 244

(1969) 6049-6055.

[20] H.R. Warner, Superoxide dismutase, aging, and degenerative disease, Free Radical

Biol. Med. 17 (1994) 249-258.

Page 156: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

139

[21] V.C. Culotta, Superoxide dismutase, oxidative stress, and cell metabolism, Curr.

Top. Cell. Regul. 36 (2000) 117-132.

[22] J.M. McCord, Superoxide dismutase in aging and disease: An overview, in:

Anonymous Methods in Enzymology, Academic Press, 2002, pp. 331-341.

[23] J. Kolarov, J. Kulpa, M. Baijot, A. Goffeau, Characterization of a protein serine

kinase from yeast plasma membrane. J. Biol. Chem. 263 (1988) 10613-10619.

[24] E. Estrada, P. Agostinis, J.R. Vandenheede, J. Goris, W. Merlevede, J. François, A.

Goffeau, M. Ghislain, Phosphorylation of Yeast Plasma Membrane H+-ATPase by

Casein Kinase I, J. Biol. Chem. 271 (1996) 32064-32072.

[25] C. Boone, H. Bussey, B.J. Andrews, Exploring genetic interactions and networks

with yeast, Nat. Rev. Genet. 8 (2007) 437-449.

[26] J.L. Hartman 4th, B. Garvik, L. Hartwell, Principles for the buffering of genetic

variation, Science. 291 (2001) 1001-1004.

[27] J.M. Leitch, L.T. Jensen, S.D. Bouldin, C.E. Outten, P.J. Hart, V.C. Culotta,

Activation of Cu,Zn-superoxide dismutase in the absence of oxygen and the copper

chaperone CCS, J. Biol. Chem. 284 (2009) 21863-21871.

[28] V.C. Culotta, H.D. Joh, S.J. Lin, K.H. Slekar, J. Strain, A physiological role for

Saccharomyces cerevisiae copper/zinc superoxide dismutase in copper buffering,

J. Biol. Chem. 270 (1995) 29991-29997.

Page 157: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

140

[29] G. Wang, M.J. Tamás, M.J. Hall, A. Pascual-Ahuir, D.S. Perlin, Probing Conserved

Regions of the Cytoplasmic LOOP1 Segment Linking Transmembrane Segments 2 and 3

of the Saccharomyces cerevisiae Plasma Membrane H+-ATPase, J. Biol. Chem. 271

(1996) 25438-25445.

[30] D.S. Perlin, S.L. Harris, D. Seto-Young, J.E. Haber, Defective H(+)-ATPase of

hygromycin B-resistant pma1 mutants from Saccharomyces cerevisiae. J. Biol. Chem.

264 (1989) 21857-21864.

[31] J.A. Baron, K.M. Laws, J.S. Chen, V.C. Culotta, Superoxide triggers an acid burst in

Saccharomyces cerevisiae to condition the environment of glucose-starved cells,

J. Biol. Chem. 288 (2013) 4557-4566.

[32] B.C. Monk, M.B. Kurtz, J.A. Marrinan, D.S. Perlin, Cloning and characterization of

the plasma membrane H(+)-ATPase from Candida albicans. J. Bacteriol. 173 (1991)

6826-6836.

[33] G.A. Martínez-Muñoz, P. Kane, Vacuolar and Plasma Membrane Proton Pumps

Collaborate to Achieve Cytosolic pH Homeostasis in Yeast, J. Biol. Chem. 283 (2008)

20309-20319.

[34] C.J. Roberts, C.K. Raymond, C.T. Yamashiro, T.H. Stevens, Methods for studying

the yeast vacuole, in: Anonymous Methods in Enzymology, Academic Press, 1991, pp.

644-661.

Page 158: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

141

[35] L.B. Corson, J. Folmer, J.J. Strain, V.C. Culotta, D.W. Cleveland, Oxidative stress

and iron are implicated in fragmenting vacuoles of Saccharomyces cerevisiae lacking

Cu,Zn-superoxide dismutase, J. Biol. Chem. 274 (1999) 27590-27596.

[36] M. Miranda, K.E. Allen, J.P. Pardo, C.W. Slayman, Stalk segment 5 of the yeast

plasma membrane H+-ATPase: mutational evidence for a role in glucose regulation,

J. Biol. Chem. 276 (2001) 22485-22490.

[37] M. Miranda, J.P. Pardo, K.E. Allen, C.W. Slayman, Stalk segment 5 of the yeast

plasma membrane H(+)-ATPase. Labeling with a fluorescent maleimide reveals a

conformational change during glucose activation, J. Biol. Chem. 277 (2002) 40981-

40988.

[38] M.A. Wallace, L. Liou, J. Martins, M.H.S. Clement, S. Bailey, V.D. Longo, J.S.

Valentine, E.B. Gralla, Superoxide Inhibits 4Fe-4S Cluster Enzymes Involved in Amino

Acid Biosynthesis. Cross-compartment protection by CuZn-superoxide dismutase, J.

Biol. Chem. 279 (2004) 32055-32062.

[39] X.F. Liu, I. Elashvili, E.B. Gralla, J.S. Valentine, P. Lapinskas, V.C. Culotta, Yeast

lacking superoxide dismutase. Isolation of genetic suppressors, J. Biol. Chem. 267

(1992) 18298-18302.

[40] X.F. Liu, V.C. Culotta, The requirement for yeast superoxide dismutase is bypassed

through mutations in BSD2, a novel metal homeostasis gene. Mol. Cell. Biol. 14 (1994)

7037-7045.

Page 159: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

142

[41] P. Lapinskas, K. Cunningham, X. Liu, G. Fink, V. Culotta, Mutations in PMR1

suppress oxidative damage in yeast cells lacking superoxide dismutase, Mol. Cell. Biol.

15 (1995) 1382-1388.

[42] S.J. Lin, V.C. Culotta, Suppression of oxidative damage by Saccharomyces

cerevisiae ATX2, which encodes a manganese-trafficking protein that localizes to Golgi-

like vesicles, Mol. Cell. Biol. 16 (1996) 6303-6312.

[43] J. Strain, C.R. Lorenz, J. Bode, S. Garland, G.A. Smolen, D.T. Ta, L.E. Vickery,

V.C. Culotta, Suppressors of Superoxide Dismutase (SOD1) Deficiency in

Saccharomyces cerevisiae. IDENTIFICATION OF PROTEINS PREDICTED TO

MEDIATE IRON-SULFUR CLUSTER ASSEMBLY, J. Biol. Chem. 273 (1998)

31138-31144.

[44] L.T. Jensen, R.J. Sanchez, C. Srinivasan, J.S. Valentine, V.C. Culotta, Mutations in

Saccharomyces cerevisiae Iron-Sulfur Cluster Assembly Genes and Oxidative Stress

Relevant to Cu,Zn Superoxide Dismutase, J. Biol. Chem. 279 (2004) 29938-29943.

[45] Y. Feng, N.G. Davis, Akr1p and the Type I Casein Kinases Act prior to the

Ubiquitination Step of Yeast Endocytosis: Akr1p Is Required for Kinase Localization to

the Plasma Membrane, Mol. Cell. Biol. 20 (2000) 5350-5359.

[46] A.F. Roth, Y. Feng, L. Chen, N.G. Davis, The yeast DHHC cysteine-rich domain

protein Akr1p is a palmitoyl transferase, J. Cell Biol. 159 (2002) 23-28.

Page 160: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

143

[47] S. Elchuri, T.D. Oberley, W. Qi, R.S. Eisenstein, L. Jackson Roberts, H. Van

Remmen, C.J. Epstein, T.T. Huang, CuZnSOD deficiency leads to persistent and

widespread oxidative damage and hepatocarcinogenesis later in life, Oncogene. 24 (2005)

367-380.

[48] S. Sehati, M.H.S. Clement, J. Martins, L. Xu, V.D. Longo, J.S. Valentine, E.B.

Gralla, Metabolic alterations in yeast lacking copper–zinc superoxide dismutase, Free

Radical Biol. Med. 50 (2011) 1591-1598.

[49] I. Fridovich, Superoxide Radical - an Endogenous Toxicant,

Annu. Rev. Pharmacol. Toxicol. 23 (1983) 239-257.

[50] J.D. Aguirre, V.C. Culotta, Battles with iron: manganese in oxidative stress

protection, J. Biol. Chem. 287 (2012) 13541-13548.

[51] A.R. Reddi, V.C. Culotta, Regulation of manganese antioxidants by nutrient sensing

pathways in Saccharomyces cerevisiae, Genetics. 189 (2011) 1261-1270.

[52] A.R. Reddi, L.T. Jensen, A. Naranuntarat, L. Rosenfeld, E. Leung, R. Shah, V.C.

Culotta, The overlapping roles of manganese and Cu/Zn SOD in oxidative stress

protection, Free Radical Biol. Med. 46 (2009) 154-162.

[53] R.J. Sanchez, C. Srinivasan, W.H. Munroe, M.A. Wallace, J. Martins, T.Y. Kao, K.

Le, E.B. Gralla, J.S. Valentine, Exogenous manganous ion at millimolar levels rescues all

known dioxygen-sensitive phenotypes of yeast lacking CuZnSOD, J. Biol. Inorg. Chem.

10 (2005) 913-923.

Page 161: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

144

[54] K. Barnese, E.B. Gralla, D.E. Cabelli, J. Selverstone Valentine, Manganous

Phosphate Acts as a Superoxide Dismutase, J. Am. Chem. Soc. 130 (2008) 4604-4606.

[55] K. Barnese, E.B. Gralla, J.S. Valentine, D.E. Cabelli, Biologically relevant

mechanism for catalytic superoxide removal by simple manganese compounds, Proc.

Natl. Acad. Sci. U. S. A. 109 (2012) 6892-6897.

[56] J.C. Juarez, M. Manuia, M.E. Burnett, O. Betancourt, B. Boivin, D.E. Shaw, N.K.

Tonks, A.P. Mazar, F. Doñate, Superoxide dismutase 1 (SOD1) is essential for H2O2-

mediated oxidation and inactivation of phosphatases in growth factor signaling,

Proc. Natl. Acad. Sci. U. S. A. 105 (2008) 7147-7152.

[57] N. Stadler, M. Höfer, K. Sigler, Mechanisms of Saccharomyces Cerevisiae PMA1

H+-ATPase inactivation by Fe2+, H2O2 and Fenton reagents, Free Radical Res. 35

(2001) 643-653.

[58] V.V. Petrov, J.P. Pardo, C.W. Slayman, Reactive cysteines of the yeast plasma-

membrane H+-ATPase (PMA1). Mapping the sites of inactivation by N-ethylmaleimide,

J. Biol. Chem. 272 (1997) 1688-1693.

[59] N. de la Fuente, A.M. Maldonado, F. Portillo, Glucose activation of the yeast plasma

membrane H+-ATPase requires the ubiquitin–proteasome proteolytic pathway, FEBS

Lett. 411 (1997) 308-312.

Page 162: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

145

[60] M.A.A. Souza, M.J. Trópia, R.L. Brandão, New aspects of the glucose activation of

the H+-ATPase in the yeast Saccharomyces cerevisiae, Microbiology. 147 (2001) 2849-

2855.

[61] A. Goossens, N. de la Fuente, J. Forment, R. Serrano, F. Portillo, Regulation of

Yeast H+-ATPase by Protein Kinases Belonging to a Family Dedicated to Activation of

Plasma Membrane Transporters, Mol. Cell. Biol. 20 (2000) 7654-7661.

[62] J.L. Withee, R. Sen, M.S. Cyert, Ion Tolerance of Saccharomyces cerevisiae

Lacking the Ca2+/CaM-Dependent Phosphatase (Calcineurin) Is Improved by Mutations

in URE2 or PMA1, Genetics. 149 (1998) 865-878.

[63] X. Wang, V.C. Culotta, C.B. Klee, Superoxide dismutase protects calcineurin from

inactivation, Nature. 383 (1996) 434-437.

[64] I. Romero, A.M. Maldonado, P. Eraso, Glucose-independent inhibition of yeast

plasma-membrane H+-ATPase by calmodulin antagonists, Biochem. J. 322 ( Pt 3)

(1997) 823-828.

[65] C.B. Brachmann, A. Davies, G.J. Cost, E. Caputo, J. Li, P. Hieter, J.D. Boeke,

Designer deletion strains derived from Saccharomyces cerevisiae S288C: a useful set of

strains and plasmids for PCR-mediated gene disruption and other applications, Yeast. 14

(1998) 115-132.

[66] E.A. Winzeler, D.D. Shoemaker, A. Astromoff, H. Liang, K. Anderson, B. Andre, R.

Bangham, R. Benito, J.D. Boeke, H. Bussey, A.M. Chu, C. Connelly, K. Davis, F.

Page 163: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

146

Dietrich, S.W. Dow, M. El Bakkoury, F. Foury, S.H. Friend, E. Gentalen, G. Giaever,

J.H. Hegemann, T. Jones, M. Laub, H. Liao, N. Liebundguth, D.J. Lockhart, A. Lucau-

Danila, M. Lussier, N. M'Rabet, P. Menard, M. Mittmann, C. Pai, C. Rebischung, J.L.

Revuelta, L. Riles, C.J. Roberts, P. Ross-MacDonald, B. Scherens, M. Snyder, S.

Sookhai-Mahadeo, R.K. Storms, S. Veronneau, M. Voet, G. Volckaert, T.R. Ward, R.

Wysocki, G.S. Yen, K. Yu, K. Zimmermann, P. Philippsen, M. Johnston, R.W. Davis,

Functional characterization of the S. cerevisiae genome by gene deletion and parallel

analysis, Science. 285 (1999) 901-906.

Page 164: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

147

CHAPTER 4

Review and Future Directions

Page 165: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

148

1. Review of Thesis Research

Cellular pH homeostasis is regulated in a manner highly dependent on the

metabolic state of cells. This is particularly evident in yeast which readily shift from

respiratory to fermentative metabolism depending on carbon source availability. In the

presence of glucose, H+-ATPases rapidly extrude protons to tightly maintain pH

consuming significant amounts of cellular ATP in the process. In the absence of glucose,

H+-ATPase activities are diminished to conserve energy and environmental factors and

buffering become much more important.

Similarly, metabolism can regulate the ROS balance in cells. In fermenting yeast

growing with abundant glucose, the NADPH oxidase, Yno1p, along with other sources

produce a significant amount of ROS in the cytoplasm [1]. During respiration in the

absence of glucose, on the other hand, ROS in the mitochondrial matrix may be high.

The research in this thesis focuses on the role of ROS in regulating pH at these

two extremes of glucose abundance and glucose starvation. In abundant glucose, the

largely cytosolic superoxide dismutase, Sod1p, promotes activation of the major plasma

membrane H+-ATPase, Pma1p (see Fig. 4-1). Although earlier studies suggested Sod1p

works through a mechanism involving the Pma1p regulatory kinase, Yck1p [2], the

studies in this thesis indicate that the Sod1p-Pma1p connection extends beyond this

signaling pathway and involves oxidative stress protection as well (Chapter 3). While

Sod1p is important for pH homeostasis during glucose abundance, we find that the

mitochondrial matrix Sod2p and mitochondrial ROS become important during conditions

of glucose starvation. ROS in the mitochondrial matrix can actually be beneficial to yeast

during periods of long term glucose starvation. By damaging the TCA enzymes

Page 166: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

149

aconitase and SDH (Fig. 4-1B; Chapter 2), an acid burst occurs that is marked by large

increases in Ald4p-derived acetate, providing a new carbon source for cell survival.

Page 167: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

150

FIGURE 4-1. Models depicting ROS regulation of pH homeostasis at the metabolic

extremes of glucose abundance and glucose starvation.

(A) A model depicting Sod1p regulation of Pma1p during growth in abundant glucose

(Chapter 3). The plasma membrane H+-ATPase, Pma1p, is highly active in the presence

of glucose and is the primary protein responsible for ensuring pH homeostasis,

maintaining pHi ≈ 7.2 [3] while acidifying the surrounding media. Organic acids are also

secreted [4-6]. Sod1p acts through two pathways to ensure maximal Pma1p activation in

abundant glucose. Pathway 1 (top): Sod1p prevents degradation of the yeast casein

kinase, Yck1p, by producing H2O2 in a manner requiring oxygen and glucose. Yck1p

then activates Pma1p [2]. Pathway 2 (bottom): Sod1p scavenges ROS to prevent

oxidative damage, which inhibits Pma1p either directly or indirectly. (B) A model

depicting how mitochondrial ROS produced during glucose starvation can lead to the

acid burst and acidify the extracellular environment (Chapter 2). In the absence of

glucose, activity of the plasma membrane H+-ATPase, Pma1p, is low and yeast initially

alkalinize the surrounding media, presumably by producing ammonia (not shown) [7].

Later, yeast shift to a phase of acid production termed the acid burst. This period is

initiated by increased ROS in the mitochondrial matrix that inhibits the TCA cycle

enzymes aconitase (Aco1p) and SDH. TCA cycle inhibition prevents Ald4p-produced

acetate from entering the cycle (as acetyl-CoA) which subsequently builds up and is

secreted into the media. Acetate secretion alone causes acidification of the media, as

Pma1p activity remains low, and promotes yeast growth during glucose starvation.

Page 168: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

151

Page 169: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

152

2. Future Directions

2.1. Examination of the acid burst during glucose starvation (Chapter 2)

Under glucose starvation conditions, ROS stimulate the acid burst to cause media

acidification and enhance growth (Fig. 2-9C). This occurs quickly in sod2Δ null cells but

is also seen in WT yeast, where it occurs more gradually (Fig. 2-2). Čáp and colleagues

[8] previously examined ROS levels in the mitochondrial matrix during long-term

glucose starvation and demonstrated that matrix ROS increase in a subset of WT cells

and in a greater number of sod2Δ null cells after entrance into the alkali phase, which

agrees well with our proposal that ROS stimulate the shift from alkali-to-acid. One

important question that remains unanswered is why ROS increase in the mitochondrial

matrix. Is this through loss of antioxidant activities or through an increase in production

of mitochondrial ROS? Though Čáp and colleagues examined the activity of cytoplasmic

Sod1p and the catalase Ctt1p they did not assess the all important mitochondrial matrix

Sod2p [8]. As decreased activity of Sod2p is the most likely explanation for the

increased ROS, further work could evaluate the activity and protein levels of Sod2p

during the transition from alkaline to acidic phases. Enhanced ROS production might

also be the culprit and could be examined by comparing ROS production in mitochondria

or mitochondrial particles isolated at various times during growth using available probes

such as MitoSOX for superoxide [8] and Amplex Red for hydrogen peroxide [9,10].

The major mitochondrial aldehyde dehydrogenase, Ald4p, is required for

production of the acetate secreted during the acid burst (Fig. 2-7F and 2-8). In the

simplest model we proposed, Ald4p produces normal levels of acetate which build up due

to ROS-disruption of the TCA cycle and are eventually secreted. Another possibility is

Page 170: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

153

that Ald4p-acetate production is enhanced concomitantly with TCA cycle inhibition to

produce the acid burst. Future work could address the role of Ald4p in the acid burst

more closely by examining its expression, protein levels, and activity under glucose

starvation conditions.

Our work is the first to examine the mechanism by which yeast switch from

periods of media alkalinization to acidification during glucose starvation. It would be

interesting to examine cells of other organisms to determine if a similar mechanism

exists. At a minimum this is likely to be a general feature of yeasts, just as the ammonia

signaling during the period of alkalinization was shown by Palkova and colleagues to

occur in a variety of yeast strains [7]. However, it would also be interesting to examine

the intersection of nutrient deprivation, mitochondrial ROS, and growth in mammalian

cells. A number of similarities between cancer development, which is essentially an

overgrowth phenotype, and the acid burst of yeast suggest a similar pathway may exist in

mammalian cells. For one, defects in the TCA cycle initiate the acid burst and are also

known drivers of cancer formation [11]. Furthermore, when SDH mutations that are

known to cause cancer in humans are introduced into yeast grown on low glucose, the

resulting acid produced is very similar to the acid burst in both degree and composition

(i.e. high organic acid secretion, with particularly high levels of acetate) [12,13].

Secondly, increased mitochondrial ROS are associated with cancer formation [14,15].

Third, at least some studies show that glucose deprivation contributes to increased

mitochondrial ROS production in mammalian cells [14,16-19]. Few if any studies have

focused on the intersection between these factors and the potential role of extracellular

acidification in promoting growth.

Page 171: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

154

2.2. Glucose-dependent Sod1p regulation of Pma1p (Chapter 3)

Rescue of T912D sod1Δ aerobic lethality by Mn anti-oxidants and anaerobic

growth, along with the viability of T912D akr1Δ strongly suggests that the T912D sod1Δ

synthetic lethality is a result of oxidative damage and not misregulation of Yck1p (Fig. 3-

5)[2]. However, whether Sod1p protects Pma1p directly or indirectly remains unclear.

To determine if Pma1p is directly damaged by oxidative stress, Pma1p WT and T912D

could be immunoprecipitated from sod1Δ and SOD1+ cells and examined in western blots

to identify carbonyl formation (Oxyblot). Significantly more carbonyls in sod1Δ cells

would be indicative of oxidative damage to Pma1p. Comparison of Pma1-T912D and

WT carbonylation in sod1Δ cells might also indicate whether the Pma1-T912D mutant is

more sensitive to oxidative damage than WT. In a similar manner, the oxidation state of

the five cysteines located in the catalytic domain of Pma1p could be compared in sod1Δ

versus SOD1+ cells. To do this, WT and T912D Pma1p would be pretreated with

iodoacetamide (protects reduced cysteines) and run on SDS-PAGE gels under reducing

and non-reducing conditions. sod1Δ-induced changes in cysteine oxidation would appear

as a gel shift in the reduced versus non-reduced samples. As before, this may also allow

the comparison of cysteine oxidation in Pma1-T912D relative to WT to determine if the

mutant is more sensitive. Finally, oxidative modifications to Pma1p in sod1Δ versus

SOD1+ cells could also be examined more generally using mass spectrometry.

Our work suggests the Pma1-T912D mutant is sufficiently sensitive to oxidative

damage that it can be totally inactivated in sod1∆ null cells. However, the effect of

oxidative damage on WT Pma1p activity was not explored. Previously, the activity of

WT Pma1p was shown to be reduced ≈ 80% in sod1∆ null cells [2]. This reduction was

Page 172: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

155

attributed to destabilization of the Pma1p regulatory kinase, Yck1p, but no direct

evidence for this was provided and at least a portion is likely due to oxidative damage.

Future work could determine the extent that Sod1p regulation of Yck1p and Sod1p

oxidative stress protection play in promoting glucose-induced Pma1p activation. The

effect of oxidative damage could be explored by measuring Pma1p activity in sod1∆ null

cells that are grown in the presence of Mn to prevent oxidative damage, as this will not

rescue Yck1p destabilization. Conversely, Pma1p activity in sod1∆ null cells expressing

a mutant of Yck1p that is not degraded (YCK1-K383R/K386R/K390R) [2] would

indicate how important Sod1p regulation of Yck1p is to Pma1p activation.

The importance of Sod1p in regulating Pma1p, particularly by preventing

oxidative damage, was previously unexplored. This role of Sod1p may not be limited to

Pma1p and Sod1p is likely to play a role in regulating other transporters in both yeast and

mammalian cells. To identify transporters protected by Sod1p from oxidative damage, a

western blot for protein carbonyls could be used to identify membrane proteins modified

specifically in sod1∆, but not WT, yeast cells after crude membrane extraction and

protein separation by 2-dimensional gel. A similar approach could also be used in

mammalian cells. Interestingly, at least one study suggests that the plasma membrane

Na+/H+ exchanger, NHE1, that maintains pHi in mammalian cells (analogous to the role

of Pma1p in yeast) is sensitive to oxidative insult [20]. This suggests the possibility that

mammalian SOD1 also regulates NHE1 which to our knowledge has not been explored in

detail.

Page 173: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

156

3. Reference List

[1] M. Rinnerthaler, S. Buttner, P. Laun, G. Heeren, T.K. Felder, H. Klinger, M.

Weinberger, K. Stolze, T. Grousl, J. Hasek, O. Benada, I. Frydlova, A. Klocker, B.

Simon-Nobbe, B. Jansko, H. Breitenbach-Koller, T. Eisenberg, C.W. Gourlay, F. Madeo,

W.C. Burhans, M. Breitenbach, Yno1p/Aim14p, a NADPH-oxidase ortholog, controls

extramitochondrial reactive oxygen species generation, apoptosis, and actin cable

formation in yeast, Proc. Natl. Acad. Sci. U. S. A. 109 (2012) 8658-8663.

[2] A. Reddi, V. Culotta, SOD1 Integrates Signals from Oxygen and Glucose to Repress

Respiration, Cell. 152 (2013) 224-235.

[3] R. Orij, J. Postmus, A. Ter Beek, S. Brul, G.J. Smits, In vivo measurement of

cytosolic and mitochondrial pH using a pH-sensitive GFP derivative in Saccharomyces

cerevisiae reveals a relation between intracellular pH and growth, Microbiology. 155

(2009) 268-278.

[4] K. Sigler, A. Kotyk, A. Knotková, M. Opekarová, Processes involved in the creation

of buffering capacity and in substrate-induced proton extrusion in the yeast

Saccharomyces cerevisiae, Biochim. Biophys. Acta, Biomembr. 643 (1981) 583-592.

[5] C.R. Burtner, C.J. Murakami, B.K. Kennedy, M. Kaeberlein, A molecular mechanism

of chronological aging in yeast, Cell Cycle. 8 (2009) 1256-1270.

[6] S. Yoshida, A. Yokoyama, Identification and characterization of genes related to the

production of organic acids in yeast, Journal of Bioscience and Bioengineering.

Page 174: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

157

[7] Z. Palkova, B. Janderova, J. Gabriel, B. Zikanova, M. Pospisek, J. Forstova,

Ammonia mediates communication between yeast colonies, Nature. 390 (1997) 532-536.

[8] M. Čáp, L. Váchová, Z. Palková, Yeast Colony Survival Depends on Metabolic

Adaptation and Cell Differentiation Rather Than on Stress Defense, J. Biol. Chem. 284

(2009) 32572-32581.

[9] K.J. Morten, B.A.C. Ackrell, S. Melov, Mitochondrial Reactive Oxygen Species in

Mice Lacking Superoxide Dismutase 2: ATTENUATION VIA ANTIOXIDANT

TREATMENT, J. Biol. Chem. 281 (2006) 3354-3359.

[10] A.A. Starkov, G. Fiskum, Regulation of brain mitochondrial H2O2 production by

membrane potential and NAD(P)H redox state, J. Neurochem. 86 (2003) 1101-1107.

[11] E.K. Oermann, J. Wu, K. Guan, Y. Xiong, Alterations of metabolic genes and

metabolites in cancer, Semin. Cell Dev. Biol. 23 (2012) 370-380.

[12] S.S.W. Szeto, S.N. Reinke, B.D. Sykes, B.D. Lemire, Ubiquinone-binding Site

Mutations in the Saccharomyces cerevisiae Succinate Dehydrogenase Generate

Superoxide and Lead to the Accumulation of Succinate, J. Biol. Chem. 282 (2007)

27518-27526.

[13] S.S.W. Szeto, S.N. Reinke, B.D. Sykes, B.D. Lemire, Mutations in the

Saccharomyces cerevisiae Succinate Dehydrogenase Result in Distinct Metabolic

Phenotypes Revealed Through 1H NMR-Based Metabolic Footprinting, J. Proteome

Res. 9 (2010) 6729-6739.

Page 175: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

158

[14] S.J. Ralph, S. Rodríguez-Enríquez, J. Neuzil, E. Saavedra, R. Moreno-Sánchez, The

causes of cancer revisited: “Mitochondrial malignancy” and ROS-induced oncogenic

transformation – Why mitochondria are targets for cancer therapy, Mol. Aspects Med. 31

(2010) 145-170.

[15] D.C. Wallace, Mitochondria and cancer, Nat. Rev. Cancer. 12 (2012) 685-698.

[16] N.K. Isaev, E.V. Stelmashook, U. Dirnagl, E.Y. Plotnikov, E.A. Kuvshinova, D.B.

Zorov, Mitochondrial free radical production induced by glucose deprivation in

cerebellar granule neurons, Biochemistry (Moscow). 73 (2008) 149-155.

[17] Qun Chen, Shadi Moghaddas, Charles L. Hoppel, Edward J. Lesnefsky, Ischemic

defects in the electron transport chain increase the production of reactive oxygen species

from isolated rat heart mitochondria, American Journal of Physiology - Cell Physiolog.

294 (2008) C460-C466.

[18] H. Shi, K.J. Liu, Effects of glucose concentration on redox status in rat primary

cortical neurons under hypoxia, Neurosci. Lett. 410 (2006) 57-61.

[19] G.A. Martens, Y. Cai, S. Hinke, G. Stange, M. Van de Casteele, D. Pipeleers,

Glucose suppresses superoxide generation in metabolically responsive pancreatic beta

cells, J. Biol. Chem. 280 (2005) 20389-20396.

[20] C. ELSING, A. VOSS, T. HERRMANN, I. KAISER, C.A. HUEBNER, T.

SCHLENKER, Oxidative Stress Reduces Na+/H+ Exchange (NHE) Activity in a Biliary

Epithelial Cancer Cell Line (Mz-Cha-1), Anticancer Research. 31 (2011) 459-465.

Page 176: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

159

J. ALLEN BARON 634 Melvin Dr. • Baltimore, MD 21230

Phone: 410-807-2955 • Email: [email protected]

EDUCATION Sept 2013, expected Ph.D. in Biochemistry and Molecular Biology Johns Hopkins University–Bloomberg School of Public Health Baltimore, MD

Dec 2005 B.S. in Biology (summa cum laude) Brigham Young University – Idaho Rexburg, ID RESEARCH EXPERIENCE 2007 – present Graduate research in the lab of Dr. Valeria C. Culotta. Cellular Redox and pH Signaling

• Biochemical characterization of cellular metabolites and enzyme activities central to energy metabolism and pH homeostasis affected by reactive oxygen species signaling

• Fractionation and isolation of subcellular components • Extensive experience in cloning and yeast molecular genetics • Fluorescence microscopy of protein localization using

immunofluorescence

2007 – 2010 Graduate research in the lab of Dr. Eric Grote. Calcium Regulation of Cell-Cell Fusion

• Used steady-state and time-lapse fluorescence microscopy to monitor cell-cell fusion events and dynamics of calcium signaling

• Developed a facile and high-throughput genetic screen in yeast to identify factors required for membrane fusion

2006 Five 7-week rotations during first year of Ph.D. introduced me to: • Mammalian cell culture • PCR-based genomic screening and genetic screens in plants • Rotations were in the labs of Drs. Terry Brown, Roger McMacken,

Judith Bender, Eric Grote, and William Wright

2005, summer IDeA Network of Biomedical Research Excellence (INBRE) Summer Fellow in the lab of Douglas G. Cole, Department of Biological Sciences, University of Idaho. Expression and Protein Interaction Analysis of Two Chlamydomonas Intraflagellar Transport Proteins

TEACHING EXPERIENCE 2012 Teaching Assistant, Principles of Cell Biology (Master’s level), Johns

Hopkins University

Page 177: THE DUAL NATURE OF REACTIVE OXYGEN SPECIES: …

160

2011 – 2013 Research mentor for Janice Chen (undergraduate), Johns Hopkins University

• Working with Pamela Silver, Ph.D. (Harvard) and applying to MD/PhD programs

2008 Research mentor for Preston Kramer (undergraduate), Johns Hopkins University

2007 Tutor, Molecular Biology and Genomics (Ph.D. level), Johns Hopkins University

2005 Tutor, Genetics, Brigham Young University – Idaho • College Reading & Learning Association Certified - Advanced Tutor

2004 Lab Assistant, General Microbiology, Brigham Young University – Idaho

PUBLICATIONS 2. Baron, J.A., Laws, K.M., Chen, J.S., and Culotta, V.C. (2013) Superoxide triggers

an acid burst in Saccharomyces cerevisiae to condition the environment of glucose-starved cells. J. Biol. Chem. 288, 4557-4566.

1. Leitch, J.M., Li, C.X., Baron, J.A., Matthews, L.M., Cao, X., Hart, P.J., and Culotta, V.C. (2012) Post-translational modification of Cu/Zn superoxide dismutase under anaerobic conditions. Biochemistry. 51, 677-685.

CONFERENCE PRESENTATIONS 3. Mitochondrial superoxide mimics the cellular response to nutrients in the yeast

Saccharomyces cerevisiae. Baron, J. Allen and Culotta, Valeria C. Society for Free Radical Biology and Medicine, 18th Annual Meeting (Nov 2011). Atlanta, GA, United States.

2. Ca2+ regulation of fusion/lysis in S. cerevisiae during plasma membrane fusion. Baron, J. Allen and Grote, Eric. Gordon Research Conference: Cell-Cell Fusion (Aug 2009). University of New England, Biddeford, ME, United States.

1. Expression and Protein Interaction Analysis of Two Chlamydomonas Intraflagellar Transport Proteins. Baron, J. Allen; Behal, Robert; Qin, Hongmin; Rosenbaum, Joel L.; Cole. Douglas G. IDeA Network of Biomedical Research Excellence (INBRE) Summer Research Conference (Aug 2005). Northwest Nazarene University, Nampa, ID, United States.