104
Mitochondrial and Mitochondrial DNA Inheritance Checkpoints in the Budding Yeast, Saccharomyces cerevisiae David Garry Crider Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the Graduate School of Arts and Sciences COLUMBIA UNIVERSITY 2012

Mitochondrial and Mitochondrial DNA Inheritance

  • Upload
    others

  • View
    22

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Mitochondrial and Mitochondrial DNA Inheritance

Mitochondrial and Mitochondrial DNA Inheritance

Checkpoints in the Budding Yeast, Saccharomyces

cerevisiae

David Garry Crider

Submitted in partial fulfillment of the requirements for the

degree of Doctor of Philosophy in the Graduate School of

Arts and Sciences

COLUMBIA UNIVERSITY

2012

Page 2: Mitochondrial and Mitochondrial DNA Inheritance

© 2011

David Garry Crider

All rights reserved

Page 3: Mitochondrial and Mitochondrial DNA Inheritance

ABSTRACT

Mitochondrial and Mitochondrial DNA Inheritance

Checkpoints in the Budding Yeast, Saccharomyces cerevisiae

David Garry Crider

This dissertation analyzes the importance of mitochondria and mitochondrial DNA in

Saccharomyces cerevisiae during cell division. Movement and positional control of

mitochondria and other organelles are coordinated with cell cycle progression in the

budding yeast, Saccharomyces cerevisiae. Recent studies have revealed a checkpoint

that inhibits cytokinesis when there are severe defects in mitochondrial inheritance. An

established checkpoint signaling pathway, the mitotic exit network (MEN), participates in

this process. Here, we describe mitochondrial motility during inheritance in budding

yeast, emerging evidence for mitochondrial quality control during inheritance, and

organelle inheritance checkpoints for mitochondria and other organelles.

Page 4: Mitochondrial and Mitochondrial DNA Inheritance

i

TABLE OF CONTENTS

CHAPTER 1 ................................................................................................ 1

INTRODUCTION ........................................................................................ 1

MITOCHONDRIA .................................................................................................................... 2

CELL CYCLE .......................................................................................................................... 2

CHECKPOINTS ...................................................................................................................... 5

MITOCHONDRIAL INHERITANCE ....................................................................................... 10

QUALITY CONTROL DURING MITOCHONDRIAL INHERITANCE ...................................... 11

MITOTIC EXIT NETWORK FUNCTION IN THE MITOCHONDRIAL INHERITANCE

CHECKPOINT ...................................................................................................................... 16

CHAPTER 2 ........................................................................................................................ 20

MITOCHONDRIAL INHERITANCE IS REQUIRED FOR MEN-

REGULATED CYTOKINESIS IN BUDDING YEAST .......................................... 20

SUMMARY ........................................................................................................................... 21

RESULTS AND DISCUSSION .............................................................................................. 22

Mutations that inhibit mitochondrial inheritance produce multibudded cells in budding yeast. .......... 22

mdm10∆ cells exhibit defects in contractile ring closure. ................................................................... 26

Role for the MEN in regulation of cell cycle progression in mdm10∆ cells. ....................................... 32

Experimental Procedures ................................................................................................................... 38

Yeast strains, plasmids, and growth conditions: ................................................................................ 42

Yeast strains used for this study ........................................................................................................ 43

CHAPTER 3 .............................................................................................. 46

MtDNA INHERITANCE CHECKPOINT ..................................................... 46

BACKGROUND .................................................................................................................... 47

OTHER PROTEINS IMPLICATED IN mtDNA INHERITANCE .............................................. 48

mtDNA MUTATIONS: rho0 AND rho- CELLS ........................................................................ 49

Rad53 AND THE DNA DAMAGE CHECKPOINT .................................................................. 51

LINKS BETWEEN Rad53 AND mtDNA ................................................................................ 54

RESULTS ............................................................................................................................. 56

LOSS OF mtDNA INDUCES A G1 ARREST IN CELL CYCLE PROGRESSION ............................ 56

Page 5: Mitochondrial and Mitochondrial DNA Inheritance

ii

THE G1 TO S PROGRESSION DEFECT OBSERVED IN CELLS LACKING mtDNA IS NOT DUE

TO LOSS OF MITOCHONDRIAL RESPIRATORY ACTIVITY OR ENERGY PRODUCTION .......... 58

THE DEFECT IN CELL CYCLE PROGRESSION OBSERVED IN rho0 CELLS IS DUE TO LOSS OF

DNA IN MITOCHONDRIA AND NOT GENES ENCODED BY THAT DNA ....................................... 61

ROLE FOR A KNOWN CHECKPOINT PROTEIN (Rad53) IN REGULATION OF CELL CYCLE

PROGRESSION IN CELLS LACKING mtDNA .................................................................................. 62

DISCUSSION ....................................................................................................................... 65

EXPERIMENTAL PROCEDURES ........................................................................................ 68

Yeast strains, plasmids, and growth conditions: ................................................................................ 68

CHAPTER 4 .............................................................................................. 72

DISCUSSION ............................................................................................ 72

DISCUSSION ....................................................................................................................... 73

SUMMARY OF PROJECT 1: ................................................................................................ 74

FUTURE EXPERIMENTS: .................................................................................................... 75

SUMMARY OF PROJECT 2: ................................................................................................ 77

ROLE FOR DNA Pol γ AS A SENSOR FOR THE mtDNA INHERITANCE CHECKPOINT: ... 79

HOW IS THE SIGNAL THAT mtDNA LOSS TRANSMITTED FROM MITOCHONDRIA TO

THE NUCLEUS? .................................................................................................................. 83

POSSIBLE ROLE FOR PROHIBITINS IN THE mtDNA INHERITANCE CHECKPOINT ....... 85

HOW MY STUDIES MAY CONTRIBUTE TO OUR UNDERSTANDING OF

MITOCHONDRIAL DISEASES ............................................................................................. 87

BIBLIOGRAPHY ....................................................................................... 90

Page 6: Mitochondrial and Mitochondrial DNA Inheritance

iii

ACKNOWLEDGEMENTS

I would like to thank Liza Pon for her mentorship over the past 3 years. The lab

environment she created gives every graduate student the opportunity to

succeed. She set the perfect balance between scientific curiosity and reality; for

when to focus and finish up the story. I cannot thank her enough for her

dedication and guidance.

The Pon lab would not function as efficiently, or as well as it does, without Istvan

Boldogh. He is the backbone of the lab. He has such a wealth of knowledge, and

such a selfless personality that he will drop everything and help you whenever he

can. I am so grateful for all our conversations and the time we spent together. He

truly made the lab a fun place.

I would like to thank Tom Lipkin for taking the time to grab a cup of coffee and

encourage me to join the Pon lab. Theresa Swayne might not realize that she was

the first person that I talked to when I came to Columbia (during my orientation)

and I’m honored to be able to thank her on my last day at Columbia. To Jose

“Ricky” Ricardo McFaline Figueroa (pick a name man!) who is like an older, I

mean younger, brother that I never had, thank you. We discussed very little

science and for that I credit my current sanity. To Luis Garcia, who I have never

met, but would not have had such an exciting project if it wasn’t for his discovery

and previous work.

Page 7: Mitochondrial and Mitochondrial DNA Inheritance

iv

I would like to thank my first scientific mentor, Kenneth Boheler, who encouraged

out-of-the-box thinking and molded me into a scientist. To this day my most

exciting experiment was watching my stem cell derived cardiomyocytes beat on

the dish. Who would have thought that single experiment would be the reason I

met my wife. I’d like to thank those little cardiomyocytes for grabbing my wife’s

attention and giving me the chance to meet her, fall in love, and get married…my

next most exciting experiments are yet to come.

My family has provided unconditional support for me throughout the last decade

as I would finish one degree and start the next. They never questioned my

dedication and always understood and accepted the sacrifice that it took to reach

my goal. I can never get those lost holidays and time spent away from them back

but I hope the sacrifice is justified in the end.

To my pap, the best self-taught scientist that I know.

Everyone mentioned had a huge role into getting me to this day but no one

deserves more credit than my wife Cheryl. She is the reason I came back to

graduate school more focused and determined than ever. I cannot thank her

enough for her patience and understanding when I had to work every weekend

and stay in lab late but also for those days where she put her foot down and

physically dragged me out of lab. She knew when I was burning myself out and

when to tell me to get my ass off the couch and finish. I can’t describe how much

you mean to me and I couldn’t have completed this journey without you and I

cannot wait to start the next journey with you.

Page 8: Mitochondrial and Mitochondrial DNA Inheritance

1

CHAPTER 1

INTRODUCTION

Page 9: Mitochondrial and Mitochondrial DNA Inheritance

2

MITOCHONDRIA

Mitochondria are essential organelles that perform fundamental cellular functions

including aerobic metabolic fatty acid oxidation, amino acid metabolism, apoptosis and

biosynthesis of many cellular metabolites. Mitochondria contain their own DNA

(mtDNA), which encodes respiratory chain components and tRNAs necessary for

synthesis of mtDNA-encoded respiratory chain components, and cannot be made de

novo. Saccharomyces cerevisiae has become an ideal model system for studying

mitochondrial function and inheritance, processes that are required for cell survival. The

vital process of replicating and transferring ‘fit’ mitochondria to a new cell gives

researchers a glimpse at the complex mechanisms involved in maintaining quality

assurance for the next generation of cells. This thesis project will describe 2 major

findings on how mitochondria are regulated throughout the entire cell cycle and how the

presence and segregation of mitochondria and its genome in the daughter cell are

critical for normal cell division to occur. The first project investigates how the presence

of mitochondria in the daughter cell is required in order to complete cytokinesis, and the

second project explores how the presence of mitochondrial DNA (mtDNA) is monitored

and regulated by a conserved Rad53 checkpoint signaling pathway.

CELL CYCLE

Rudolph Virhow has been credited for the observation in 1855 that cells arise only from

pre-existing cells; which later lead embryologists to describe the cytology of cell division

in greater detail. However, the understanding of the underlying mechanisms of cell

Page 10: Mitochondrial and Mitochondrial DNA Inheritance

3

division took place over a century later in the 1970s and 1980s where molecular

biologists, cell biologists, biochemists, and geneticists joined forces to define and

dissect the cell cycle in molecular terms (Nurse et al., 1998). Their work revealed the

basic structure and components of cell division, leading to the understanding that this

event is highly regulated.

The cell cycle is a series of events and processes that lead to the replicating and

segregation of essential organelles to a new cell during cell cycle division. The cell cycle

is often composed of four phases, the time before DNA replication is known as (G1), the

DNA synthetic phase (S), the process after DNA replication (G2), and the mitotic phase

(M), which is where cells segregate duplicated chromosomes (Figure 1.1). The concept

of rate-limiting steps during known phases of the cell cycle lead to the discovery of

maturation promotion factor (MPF) in fission yeast mutants, (later found in starfish

oocytes, blastomeres of frog embryos and in human cells), that prematurely advanced

to mitosis and eventually led to the observation of cyclin dependent kinases (CDKs) as

regulators of the cell cycle (Nurse et al., 1998). Eukaryotic cells commit to each new

cycle in mid-G1. In yeast, this occurs before bud emergence. This commitment or

transition is referred to as the ‘restriction point’ or ‘START’.

Progressing beyond ‘START’ is the major target for growth factors and nutritional

signaling and occurs by activation of G1 phase cyclin-dependent kinases (CDKs). It also

marks the time in which the replication of the nuclear genome, an error prone process,

begins. Errors within the genome must be identified and corrected before the genome is

Page 11: Mitochondrial and Mitochondrial DNA Inheritance

4

inherited to the next generation of cells. Therefore, it is not surprising that cells have a

quality control mechanism that induces a transient arrest in G1 if DNA damage is

detected in order to repair mutagenic DNA damage before replication occurs in S-

phase.

DNA monitoring is one example of a regulatory surveillance and quality control

apparatus that is in place to limit and reduce erroneous DNA to be inherited by the

daughter cell. Eukaryotic cells also cannot divide until the genomes (DNA) are

replicated and transferred to the new dividing cells. This control during cell division is

not limited to just genomic material; the centrosome is a large macromolecular structure

that moves and assemblies and these events are also under surveillance during all

stages of the cell division. Indeed, cell division is under constant surveillance and the

initiation of late events is dependent on the completion of early events. Dependent

relationships seen in somatic cells are a key element in understanding the high fidelity

of organelle reproduction and distribution during cell division (Hartwell and Weinert,

1989). Saccharomyces cerevisiae, budding yeast, has been instrumental in the

identification and characterization of regulatory components during successful cell

divisions as well as identifying key proteins involved in this regulatory process.

Page 12: Mitochondrial and Mitochondrial DNA Inheritance

5

mpf.biol.vt.edu. Budding yeast homepage

CHECKPOINTS

Eukaryotic cells developed a complex network of signal pathways to ensure high fidelity

replication and transfer of macromolecules from mother cell to daughter cell. This

process is highly controlled: both intrinsic and external signals can activate a regulatory

signal known as a checkpoint. This regulation is a control mechanism called a

checkpoint. Checkpoints are in place to ‘check’ that critical events of the cell cycle, such

as genomic and macromolecular assembly and organization, are properly executed.

The concept of ‘checkpoints’ arose from the discovery of genes in budding yeast that

Figure 1.1 General

Outline of the Cell

Cycle in Budding

Yeast. The cell

commits to a round of

cell division once it

passes the ‘transition

or Start’ point in mid

G1. The replication of

the nuclear genome

takes place during S

Phase and nuclear

segregation shortly

follows the replication

during G2. Organelle

duplication and

migration is completed

in mitosis and

cytokinesis marks the

end of one division

and the beginning of

the next.

Page 13: Mitochondrial and Mitochondrial DNA Inheritance

6

regulate the process of cell cycle events and have the ability to delay or block the

continuation of cell division if errors are found (Hartwell and Weinert, 1989).

Checkpoints control the timing and coordination of events during cell division and were

originally identified in budding yeast (Hartwell and Weinert, 1989). Hartwell discovered

and generated a collection of cdc (cell-division cycle) mutants in which Hartwell and

others were able to isolate key cell cycle genes (Hartwell et al., 1970). This further

showed the ability and power of genetics to study and define cell cycle regulators. Yeast

provided an ideal organism to study the cdc mutants collection generated by Hartwell by

genetically and visually observing specific phases of the cell cycle in order to map out

the mechanisms involved. A remarkable finding was that most cdc genes are conserved

and have homologues in all eukaryotes and have similar key regulating functions in

species ranging from yeast to humans (Hartwell et al., 1970) (Nurse et al., 1998). An

important early finding by Hartwell was the identification of cdc28 as the gene that

controls START. This CDK gene was a major discovery that formed the basis of our

understanding of the cell cycle (Nurse et al., 1998).

Late cell cycle events are thought to be dependent on earlier events by either substrate-

product pathway or a control mechanism but cell cycle events that are not ‘hard wired’

together in the same manner as metabolic pathways and, therefore, they can be

mutated and further studied and dissected (Hartwell and Weinert, 1989) (Nurse et al.,

1998). These events and phases can be studied by mutants that specifically inhibit one

event of the cell cycle. Temperature sensitive mutants in S. cerevisiae exist for bud

Page 14: Mitochondrial and Mitochondrial DNA Inheritance

7

formation, spindle pole body enlargement, spindle elongation, initiation of DNA

(replication, elongation, and ligation), chromatin assembly, chromosome assembly and

segregation, nuclear division, and cytokinesis (Hartwell and Weinert, 1989). The

existence of control mechanisms are supported by conditions that permit late events to

occur even when an early normal prerequisite event is prevented, a term coined by

Hartwell as ‘relief of dependence’ (Hartwell and Weinert, 1989). Relief of dependency

experiments identified a number of checkpoints control mechanisms.

Hartwell et al. identified a regulatory checkpoint control mechanism for making mitosis

dependent on the completion of DNA replication (Hartwell and Weinert, 1989).

Coordinated events allow the stoppage of cell division during multiple stages if errors

are detected. The delays are mediated by genetically encoded checkpoint controls that

are not a consequence of the damage per se and are not essential for cell cycle events

(e.g. DNA replication or mitosis). Instead, checkpoints only initiate a delay after damage

is detected (Weinert, 1998). Perturbation of DNA replication by inhibitors or by

mutations, results in inactivation of replication enzymes, which in turn prevents passage

through mitosis in yeast and many other eukaryotic organisms. cdc9 temperature

sensitive mutants were blocked at mitosis and failed to bud when grown at the

restrictive temperature but cells could proceed past mitosis when they contained an

addition rad9 mutant. Suggesting RAD9 gene and its components control this system

(Hartwell and Weinert, 1989).

Further, a role for RAD9 in cell cycle progression in response to defects in DNA

Page 15: Mitochondrial and Mitochondrial DNA Inheritance

8

synthesis was determined using a collection of temperature sensitive mutants that were

initially identified by failure to arrest the cell after DNA damage was induced by x-

irradiation (Hartwell and Weinert, 1989). Mutations in the RAD9 gene allow cells with

DNA damage to proceed through cell division, whereas irradiated wild-type cells arrest

in G2 until the DNA damage is repaired. This was illustrated by relieving the

dependency of mitosis on the completion of DNA synthesis through the use of

temperature sensitive mutants involved in DNA replication and by knocking out RAD9

which enabled the cells to continue through mitosis into the next cell cycle (Hartwell and

Weinert, 1989). Cell cycle progression was accessed by visualization of bud size and

analysis of DNA content either through DAPI staining or flow cytometry. Temperature

sensitive mutants were used to confirm the control of cell cycle progression by RAD9

monitoring components of DNA replication. This illustrates that a control pathway,

RAD9, must be active in order to arrest DNA replication defective cells before mitosis

and is thought to be the main principle that applies to many checkpoints.

Checkpoints and other quality control mechanisms increase the fidelity of cell division

by triggering pathways to repair errors or to cause cell death if the error cannot be

repaired. DNA damage cascades are linked to at least 3 checkpoints: G1/S (G1)

checkpoint, intra-S phase checkpoint, and G2/M checkpoint. However, quality control is

not limited to just the integrity of DNA. There are checkpoints that monitor the presence

of specific structures such as spindle formation and localization. This checkpoint, known

as the spindle checkpoint, illustrates the importance of polarity throughout the entire cell

cycle. Upon activation, the spindle checkpoint arrests cell cycle at M phase until all

Page 16: Mitochondrial and Mitochondrial DNA Inheritance

9

chromosomes are aligned on the spindle. Another cytoskeletal checkpoint is referred to

as the morphogenesis checkpoint and has also been identified in yeast. This checkpoint

detects abnormality in the cytoskeleton and arrests the cell cycle at G2/M transition.

These checkpoints ensure that genomic integrity, replication, and segregation work in

harmony with mechanisms underlying movement and localization of segregating cellular

constituents. Saccharomyces cerevisiae has been invaluable in dissecting the

mechanisms involved in identifying the series of events from moving macromolecules

from one cell to the next during cell division.

The elimination of checkpoints may have both a subtle or catastrophic consequence

depending on prevailing conditions. The human homologues of several S. cerevisiae

checkpoint genes map to chromosomal regions implicated in the etiology of a wide

variety cancers. Specifically, Rad9 expression has been associated with prostate,

breast, lung, skin, thyroid and gastric cancers and high expression has been associated

with human prostate cancer growth (Broustas and Lieberman, 2011). Also, the deletion

of Rad9 in mouse models shows a higher incidence of skin cancer, therefore, it is

thought that Rad9 can act as an oncogene or tumor suppressor (Broustas and

Lieberman, 2011). Control of cell cycle phase progression and the link to cancer is not

limited to Rad9 only; many checkpoint proteins such as ATM, ATR, Chk1, and Chk2 are

major signaling molecules that are involved with both endogenous and exogenous

sources of DNA damage. Many cell cycle protein mutations are thought to have

fundamental roles in the pathogenesis of human cancers (Dai and Grant, 2011; Hartwell

and Weinert, 1989) (Hartwell and Kastan, 1994).

Page 17: Mitochondrial and Mitochondrial DNA Inheritance

10

The understanding of checkpoint proteins could shed light on the mechanism involved

in cancer development and provide therapeutic targets. Further, Cell cycle checkpoints

are important in the protection and ability to rescue the cells from DNA damage induced

by current chemotherapeutic agents and radiation therapy (Dai and Grant, 2011). Yeast

remains an ideal organism to study checkpoint proteins because of large percent of

homology between human and yeast checkpoint mechanisms and proteins.

MITOCHONDRIAL INHERITANCE

In early characterizations of mitochondrial morphology and distribution mutants, Sogo

and Yaffe (Sogo and Yaffe, 1994) noted a multibudded phenotype in yeast bearing a

mutation in MDM10, now known to encode a component of the mitochore. My thesis

work revealed that the multibudded phenotype observed in mdm10∆ occurs as a result

of a mitochondrial inheritance checkpoint: a mechanism that inhibits cell cycle

progression at cytokinesis when there are defects in mitochondrial inheritance (Garcia-

Rodriguez et al., 2009). I also found that a known cell cycle checkpoint signaling

pathway, the Mitotic Exit Network, regulates the mitochondrial inheritance checkpoint.

Below, I describe the events that occur during mitochondrial inheritance in yeast, and

the mechanisms underlying mitochondrial movement, immobilization and segregation

during yeast cell division.

Page 18: Mitochondrial and Mitochondrial DNA Inheritance

11

QUALITY CONTROL DURING MITOCHONDRIAL INHERITANCE

Quality control and the establishment of signaling cascades to ensure proper temporal

events is not limited to just the genome. Many diverse mechanisms have evolved in

eukaryotic organisms for the inheritance of their organelles upon cell division. These

mechanisms depend on the characteristics of a given organelle, such as its structure,

abundance and arrangement in the cell, and are coupled to the nature of the process by

which a cell divides.

Cell division in S. cerevisiae takes place by an asymmetric process. S. cerevisiae grow

and divide asymmetrically by budding and, therefore, the inheritance of their organelles

from mother to daughter cells relies on an active segregation process. This process is

highly regulated and coordinated with the progression of the cell cycle. Although each

organelle is transferred to the bud by a different mechanism, there are important

common strategies used by yeasts to inherit organelles (for review see (Fagarasanu

and Rachubinski, 2007).

As illustrated in the case of mitochondria (Fig. 1.2), a fraction of the organelles are

engaged in linear poleward movements towards the bud (anterograde movement) and

away from the bud (retrograde movement). This movement is initiated during S phase

of the cell cycle, when the bud emerges, and continues until the end of the cell division

cycle (Boldogh and Pon, 2007). In addition, some mitochondria are immobilized at the

distal tip of the mother cell, ensuring that not all organelles are transferred into the bud.

Page 19: Mitochondrial and Mitochondrial DNA Inheritance

12

Other mitochondria, which have been transported to the bud tip, are anchored there.

Anchorage of mitochondria in the bud tip ensures that these organelles are retained in

the bud. Poleward mitochondrial movement, together with anchorage of the organelle at

the poles, results in the equal partitioning of mitochondria between mother and daughter

cells during yeast cell division.

Figure 1.2. The mitochondrial inheritance cycle in budding yeast. Mitochondria align along the mother-bud axis and orient toward the site of bud

emergence at G1. During S, G2, and mitosis, mitochondria move linearly toward the tip

of the bud or mother cell. Mitochondria are immobilized at the bud tip or mother cell tip

until the end of the cycle, when they are released and redistributed throughout the

cytoplasm.

Page 20: Mitochondrial and Mitochondrial DNA Inheritance

13

Segregation of mitochondria is a complex process that relies on a large number of

proteins with diverse functions. The movement and segregation of mitochondria and

other organelles in S. cerevisiae depends on the actin cytoskeleton. Therefore, a

significant proportion of these proteins are involved in structuring and remodeling the

actin cytoskeleton.

In yeast, there are two F-actin-containing structures that persist throughout the cell

division cycle: actin patches and actin cables (reviewed in ref. (Moseley and Goode,

2006). Actin patches were named for their appearance in phalloidin-stained cells; but

they are actually endosomes, invested with an F-actin coat, that form in the bud and

bud tip (Fehrenbacher et al., 2004). Actin cables are bundles of F-actin that contain the

actin-bundling proteins fimbrin (Sac6p) and Abp140p, and two tropomyosin isoforms

(Tpm1p and Tpm2p). Formin proteins, Bni1p and Bnr1p, mediate nucleation and

elongation of F-actin filaments that are then bundled into actin cables. Bni1p and Bnr1p

localize to the bud tip and bud neck, which are the sites of actin cable assembly.

Actin cables extend from their assembly sites along the mother-bud axis of the cell and

are also dynamic structures that undergo retrograde movement from the bud toward the

mother cell (Fig. 1.3A) (Yang and Pon, 2002). Retrograde actin cable flow is driven in

part by actin cable assembly and elongation, which occurs continuously and provides a

pushing force for retrograde flow. A type II myosin at the bud neck, Myo1p, provides

pulling force during retrograde actin cable flow (Huckaba et al., 2004) (Huckaba et al.,

2006).

Page 21: Mitochondrial and Mitochondrial DNA Inheritance

14

Figure 1.3. Model for actin-driven

bidirectional movement of mitochondria in

budding yeast. Actin cables are bundles of

F-actin that align along the mother-bud axis.

Anterograde mitochondrial movement, toward

the bud tip, and retrograde movement of the

organelle, toward the mother cell tip, are both

dependent on actin cables. The mitochore

mediates reversible binding of mitochondria

and mitochondrial DNA nucleoids to actin

cables for anterograde and retrograde

movement. (A) Anterograde mitochondrial

movement is driven by the Arp2/3 complex,

which is recruited to the mitochondrial

surface by Jsn1p/Puf1p and stimulates force

generation for bud-directed movement along actin cables through actin polymerization.

Puf3p promotes anterograde mitochondrial movement by linking the mitochore and

mitochondria associated Arp2/3 complex. (B) Retrograde mitochondrial movement is

driven by the retrograde flow of actin cables, which is driven by the push of formin-

stimulated actin polymerization and assembly into actin cables and the pull of type II

myosins. Thus, mitochore-mediated binding of mitochondria to actin cables undergoing

retrograde flow links the organelle to its ‘conveyor belt’ for retrograde mother tip directed

movement.

Page 22: Mitochondrial and Mitochondrial DNA Inheritance

15

The essential function of actin cables in budding yeast is to drive intracellular

movements that are required for bud growth and organelle inheritance. Mitochondria

undergo reversible binding to F-actin in vitro, require actin cables for movement and

inheritance, and undergo anterograde and retrograde movement along actin cables in

vivo (Lazzarino et al., 1994) (Simon et al., 1997) (Fehrenbacher et al., 2004).

Association of mitochondria with actin cables for anterograde and retrograde movement

requires a mitochondrial outer membrane protein complex, the mitochore, which

consists of the proteins Mdm10p, Mdm12p, and Mmm1p (Boldogh et al., 1998)

(Boldogh et al., 2003). For retrograde movement, mitochondria undergo mitochore

dependent binding to actin cables, and use the forces of retrograde actin cable flow to

drive their movement toward the distal tip of the mother cell (Fig. 1.3B) (Fehrenbacher

et al., 2004).

For anterograde movement of many organelles, propulsion is provided by myosin motor

proteins of the class V family of myosins, which move and transport cargoes toward the

barbed ends of actin filaments within actin cables. In S. cerevisiae, there are two class

V myosins: Myo2p, which transports secretory vesicles, vacuoles, peroxisomes, and

late Golgi vesicles; and Myo4p, which transports the cortical ER (cER, see below) and

mRNA (Fagarasanu et al., 2007).

Myo2p can bind to mitochondria in vitro and is required for normal mitochondrial

distribution (Altmann et al., 2008) (Itoh et al., 2004) (Itoh et al., 2002). However, live-cell

imaging and biochemical evidence implicate an alternative propulsion mechanism in

Page 23: Mitochondrial and Mitochondrial DNA Inheritance

16

which the mitochore mediates association of mitochondria with actin cables, and the

Arp2/3 complex generates force through actin polymerization for movement of

mitochondria along the cables (Fig. 1.3A) (Boldogh et al., 2001) (McKane et al., 2005)

(Senning and Marcus, 2010). The Arp2/3 complex assembles F-actin by binding to the

side of a preexisting filament and stimulating nucleation of a new filament at a 70° angle

relative to the pre-existing filament (recently reviewed in ref. (Campellone and Welch,

2010). Jsn1p, a Pumilio family protein that localizes to the mitochondrial outer

membrane, is an Arp2/3 complex receptor on mitochondria (Fehrenbacher et al., 2005).

Puf3p, another mitochondria-associated Pumilio family protein, mediates association of

the Arp2/3 complex with the mitochore, which coordinates anterograde forces

generated on mitochondria with association of the organelle with actin cables (Garcia-

Rodriguez et al., 2007).

MITOTIC EXIT NETWORK FUNCTION IN THE MITOCHONDRIAL

INHERITANCE CHECKPOINT

When I began my thesis research, it was clear that mitochondria undergo cell cycle

linked changes in position and movement, which ensure equal segregation of the

organelle between mother and daughter cells. Indeed, segregation of mitochondria

during cell division in budding yeast exhibits features that resemble chromosome

segregation: both undergo poleward movement followed by anchorage at the poles.

Moreover, there are many known checkpoints that inhibit cell cycle progression and

activate cellular repair pathways when there are defects in chromosome duplication

Page 24: Mitochondrial and Mitochondrial DNA Inheritance

17

and/or segregation. However, there are no known checkpoints for inheritance of

organelles other than the nucleus.

My thesis research revealed a mitochondrial inheritance checkpoint that inhibits

cytokinesis when there are defects in mitochondrial inheritance in budding yeast, and a

novel role for the MEN in this process. My thesis research also revealed a mtDNA

inheritance checkpoint that inhibits G1-to-S progression in response to defects in

mtDNA inheritance, and a role for the DNA damage checkpoint in this process. Below, I

describe the MEN. In the introduction to Chapter 3, I described the DNA damage

checkpoint, and functional links between this pathway and mtDNA.

The mitotic exit network (MEN) is a GTPase-driven signal transduction cascade that

was originally identified for its role in coordinating chromosome segregation and exit

from mitosis, and ensuring proper segregation of genetic information (Krapp et al.,

2004; Seshan and Amon, 2004). MEN regulators are conserved in yeasts, C. elegans,

and mammalian cells (Fig. 1.4). The central players in the MEN are the protein

phosphatase Cdc14p, the small G protein Tem1p, and Lte1p and Bub2p/Bfa1p, the

activator and GAP/inhibitor, respectively, for Tem1p. Activation of Tem1p, which

ultimately leads to activation of Cdc14p and localization of active Cdc14p to its sites of

action, is required for mitotic exit and completion of cytokinesis.

Page 25: Mitochondrial and Mitochondrial DNA Inheritance

18

Several studies indicate that the MEN also has a role in regulating cytokinesis in yeast.

First, several MEN components localize to the site of contractile ring assembly (Luca

and Winey, 1998) (Frenz et al., 2000) (Song et al., 2000) (Xu et al., 2000) (Yoshida and

Toh-e, 2001) (Bembenek et al., 2005). Second, conditions that bypass MEN function in

mitotic exit (including mutations that weaken interactions of Cdc14p with its inhibitor

Cfi1p/Net1p, overexpression of the CDK inhibitor Sic1p, or mutations that inhibit export

of active Cdc14p from the nucleus to the cytosol) produce severe cytokinesis defects

(Song et al., 2000) (Bembenek et al., 2005) (Jimenez et al., 1998) (Shou et al., 1999)

(Lippincott et al., 2001) (Hwa Lim et al., 2003). Third, recent studies indicate that the

MEN may control cytokinesis by targeting the proteins implicated in septa formation to

the bud neck (Blondel et al., 2005) (Meitinger et al., 2010) (Nishihama et al., 2009).

Figure 1.4 Cdc14p is regulated by FEAR and MEN

signaling pathways. During early anaphase the

FEAR (for Cdc-fourteen early anaphase release)

signaling phosphatase, Cdc14p, is sequestered in the

nucleolus by an inhibitor complex Cfi1p/Net1p. When

correct spindle orientation occurs in early anaphase,

some Cdc14p is activated and released from the

nucleolus by the FEAR pathway (Dimmer et al.,

2005). Further activation of Cdc14 occurs in late

anaphase by the MEN. Tem1p, a small GTPase

protein in the MEN pathway, is negatively regulated

by a GTPase-activating protein (GAP) complex

Bub2p-Bfa1p, which is negatively regulated by Lte1p

and Cdc5p. Activated, GTP-bound Tem1p then

initiates a signaling cascade by activation of the

protein kinase Cdc15p. Cdc15p then activates the

Dbf2p/Mob1p protein kinase complex, which further

activates Cdc14p allowing for progression through

mitotic exit and cytokinesis.

Page 26: Mitochondrial and Mitochondrial DNA Inheritance

19

My thesis research supports a role for the MEN in inhibiting cytokinesis in response to

defects in mitochondrial inheritance. These studies provide additional evidence for a

role of the MEN in regulating cytokinesis, independent of its function in regulation of

mitotic exit. They also reveal a broader surveillance function for the MEN than

previously observed.

Page 27: Mitochondrial and Mitochondrial DNA Inheritance

20

CHAPTER 2

MITOCHONDRIAL INHERITANCE IS REQUIRED

FOR MEN-REGULATED CYTOKINESIS IN

BUDDING YEAST

Page 28: Mitochondrial and Mitochondrial DNA Inheritance

21

SUMMARY

Mitochondrial inheritance, the transfer of mitochondria from mother to daughter cell

during cell division, is essential for daughter cell viability. The mitochore, a mitochondrial

protein complex containing Mdm10p, Mdm12p and Mmm1p, is required for

mitochondrial motility, leading to inheritance in budding yeast. We observe a defect in

cytokinesis in mitochore mutants and another mutant (mmr1∆ gem1∆) with impaired

mitochondrial inheritance. This defect is not observed in yeast that have no

mitochondrial DNA or defects in mitochondrial protein import or assembly of β-barrel

proteins in the mitochondrial outer membrane. Deletion of MDM10 inhibits contractile

ring closure, but does not inhibit contractile ring assembly, localization of a

chromosomal passenger protein to the spindle during early anaphase, spindle

alignment, nucleolar segregation or nuclear migration during anaphase. Release of the

mitotic exit network (MEN) component, Cdc14p, from the nucleolus during anaphase is

delayed in mdm10∆ cells. Finally, hyperactivation of the MEN by deletion of BUB2

restores defects in cytokinesis in mdm10∆ and mmr1∆ gem1∆ cells, and reduces the

fidelity of mitochondrial segregation between mother and daughter cells in wild-type and

mdm10∆ cells. Our studies identify a novel MEN-linked regulatory system that inhibits

cytokinesis in response to defects in mitochondrial inheritance in budding yeast.

Page 29: Mitochondrial and Mitochondrial DNA Inheritance

22

RESULTS AND DISCUSSION

Mutations that inhibit mitochondrial inheritance produce multibudded

cells in budding yeast.

Equal segregation of mitochondria between mother and daughter cells during yeast cell

division occurs as a result of bidirectional movement of mitochondria to the bud tip and

mother cell tip and anchorage of the organelle at those sites (Fehrenbacher et al.,

2004). The mitochore, a mitochondrial membrane protein complex containing the

proteins Mmm1p, Mdm10p and Mdm12p, is required for binding of mitochondria to actin

filaments in vitro, actin cable-dependent bidirectional mitochondrial movement, and

mitochondrial inheritance (Fehrenbacher et al., 2004) (Boldogh et al., 1998) (Boldogh et

al., 2003). In early characterizations of mitochondrial morphology and distribution

mutants, Sogo and Yaffe (Sogo and Yaffe, 1994) noted the presence of a multibudded

phenotype in mdm10∆ cells. We find that multibudded clusters consisting of 3-5 buds

are present during mid-log phase and accumulate with growth time in mdm10∆ cells.

This multibudded phenotype is observed in mdm10∆ cells in three different genetic

backgrounds: S288C, W303 and A264A (data not shown).

In wild-type yeast, mitochondria constitute a dynamic and tubular reticulum (Fig. 2.1A-B)

(Fehrenbacher et al., 2004). In mdm10∆ cells, mitochondria are large spherical

structures that fail to move from mother cells to buds and undergo rapid loss of

mitochondrial DNA (mtDNA) (Boldogh et al., 1998; Boldogh et al., 2003). The large

spherical mitochondria typical of mdm10∆ cells are usually present in only one cell

Page 30: Mitochondrial and Mitochondrial DNA Inheritance

23

within a multibudded clump (Fig. 2.1E-F). Visualization of DNA confirmed that mdm10∆

cells have no mtDNA and revealed that each cell body in mdm10∆ clumps contains a

nucleus (Fig. 2.1G-H). The viability of wild-type and mdm10∆ cells during mid-log phase

growth, assessed using FUN-1 staining, is 93.5% and 76.5%, respectively. Thus, a

mutation in MDM10 that results in severe defects in mitochondrial morphology and

inheritance also produces defects in mother-daughter cell separation but does not

inhibit nuclear inheritance or compromise cell viability in SC medium.

Deletion of MDM10, MDM12 or MMM1 also results in defects in maintenance of mtDNA,

mitochondrial morphology and assembly of β-barrel proteins in the mitochondrial outer

membrane (OM) (Boldogh et al., 1998; Hobbs et al., 2001; Meisinger et al., 2004; Sogo

and Yaffe, 1994). Therefore, we tested whether the multibudded phenotype of mdm10∆

cells is due to defects in these mitochondrial inheritance-independent processes by

analysis of yeast bearing deletions in mtDNA, MAS37 or TOM7. rho0 cells have no

mtDNA and severe defects in mitochondrial respiration (Goldring et al., 1970). Mas37p

is a subunit of the SAM/TOB complex, which mediates assembly of β-barrel proteins

into the mitochondrial OM (Wiedemann et al., 2003). Tom7p is a subunit of the protein-

translocating pore in the mitochondrial OM (Honlinger et al., 1996). Deletion of TOM7

produces defects in mitochondrial morphology that are similar to those observed in

mdm10∆ cells as well as defects in mitochondrial protein import (Meisinger et al., 2004).

Tom7p also promotes the segregation of Mdm10p from the SAM/TOB complex

(Meisinger et al., 2006).

Page 31: Mitochondrial and Mitochondrial DNA Inheritance

24

rho0, mas37∆ and tom7∆ cells exhibit significantly lower defects in mitochondrial

inheritance and lower levels of multibudded cells compared to mitochore mutants (Fig.

2.1 I-J). Thus, the multibudded phenotype observed in mdm10∆ cells is not a

consequence of loss of mtDNA, or of defects in mitochondrial respiratory activity,

protein import, or OM β-barrel protein assembly. Moreover, we observed a link between

the extent of multibudded cells in late-log phase cultures and the severity of the

mitochondrial inheritance defect in yeast carrying mutations in mitochore subunits:

mdm10∆ = mmm1- 1 > > mdm12∆ (Fig. 2.1 I-J). Mdm12p coordinates mitochondrial

inheritance and biogenesis through its direct interactions with the PUF family protein

Puf3p (Garcia-Rodriguez et al., 2007). Thus, mdm12∆ cells may have less severe

multibudded and inheritance phenotypes compared to mdm10∆ or mmm1-1 mutants

because Mdm12p has regulatory effects on mitochondrial motility, while Mdm10p and

Mmm1p have predominant roles in mediating mitochondrial motility. Overall, the

multibudded phenotype observed in all mutants analyzed correlates with defects in

mitochondrial inheritance.

Page 32: Mitochondrial and Mitochondrial DNA Inheritance

25

Figure 2.1. Cell separation defects in mitochondrial inheritance mutants. Wild-type (A-D)

(BY4741) or mdm10∆ (E-H) (398) cells were grown in SC medium at 30°C to mid-log phase.

Cells were either stained for mitochondria using MitoTracker Red (MtRed), or fixed with

formaldehyde and stained using the DNA-binding dye DAPI. Images of MtRed or DAPI stained

cells are 2-D projections of the reconstructed 3-D volume that are superimposed on the

corresponding phase image. (A-B and E-F) Phase images and MtRed staining of wild-type and

mdm10∆ cells, respectively. The arrow points to the original mother cell in a multibudded

mdm10∆ cluster that contains a large spherical mitochondrion (F). Bar: 1 µm. (C-D and G-H)

Phase images and DAPI staining of wild-type and mdm10∆ cells, respectively. n: nuclear DNA.

m: mtDNA. (I-J) Quantification of mitochondria-free buds in cell bearing small buds (I) and

multibudded cells (J) in wild-type (ISY001), rho0 (ISY001-rho0), mas37∆ (ISY005), tom7∆

(ISY006), mdm12∆ (ISY003), mmm1-1 (ISY065) and mdm10∆ (ISY002) cells (n>800). Cells

were grown in SC medium at 30°C for 12-16 hrs to late-log phase (OD600 = 1.2 – 1.4). Error

bars are standard deviations.

Page 33: Mitochondrial and Mitochondrial DNA Inheritance

26

mdm10∆ cells exhibit defects in contractile ring closure.

mdm10∆ cells that enter the cell cycle have a significant delay in the ability to enter G2

phase compare to wild-type cells (Fig. 2.2). Spindle assembly and disassembly as well

as the appearance and disappearance of mitotic cyclin are delayed to a similar extent in

mdm10∆ compared to wild-type cells (Fig 2.3). Formation of the second bud (d2) in

multibudded mdm10∆ cells occurs 150 min after release from pheromone-induced G1

arrest, 25 min after the first bud (d1) undergoes Clb2p degradation and spindle

disassembly (Fig 2.3).

Page 34: Mitochondrial and Mitochondrial DNA Inheritance

27

Fig 2.2. mdm10∆ cells exhibit a delay in progression from G1 to G2. Wild-type (ISY001), mdm10∆(ISY002) or rho0 (ISY001-rho0) yeast were grown in SC medium at 300C to mid-log phase (OD600 = 0.5 – 0.8). Cells were incubated with a-factor (10 µM) for 2.5 hrs to induce arrest in G1 phase. Thereafter, they were washed with pheromone-free SC medium and incubated at 30°C. At various times min after release from pheromone-induced G1 arrest, cells were fixed, stained with propidium iodide and analyzed for DNA content by flow cytometry. The percentage of total dividing cells in G2

phase, as a function of time after release from pheromone-induced G1 arrest, is shown. The basis for the reduced rate of progression of mdm10∆ cells from G1 to G2 is not clear. However, rho0 cells, which have no mtDNA and severe defects in mitochondrial respiratory activity, can progress through the G2 phase at a similar rate as that of wild-type cells. Thus, the reduced rate of cell cycle progression observed in mdm10∆ cells does not appear to be linked to mitochondrial respiration or dependent upon mtDNA.

Page 35: Mitochondrial and Mitochondrial DNA Inheritance

28

Fig 2.3 mdm10∆ cells exhibit a delay anaphase entry and mitotic exit. A) A wild-type strain (BY4741) and an mdm10∆ mutant strain (ISY002) were synchronized as for Fig. 2.2 and grown at 30°C. Aliquots were removed from cultures at the time indicated. Levels of Clb2p in synchronized cell cultures were determined by Western blot analysis using a polyclonal anti-Clb2p antibody. Hxk1p was used as a loading control. B) Wild-type (LGY020; black line) and mdm10∆ (LGY021; grey line) cells expressing plasmid-borne mCherry-tagged tubulin were synchronized and aliquots were removed from cultures at the times indicated, fixed and visualized by fluorescence microscopy. The number of cells with spindles > 4 µm in length was determined as a function of time after release from pheromone-induced G1 arrest. Spindle assembly and disassembly are also delayed in mdm10∆ compared to wild type cells. We detect the maximum number of anaphase spindles, spindles that are 4-7 µm in length, within 60 min after release of wild-type cells from pheromone-induced G1 arrest, and spindle disassembly 100 min after release from G1 arrest. The kinetics of anaphase spindle assembly and disassembly in these synchronized wild-type cells are similar to those reported previously (Goldring et al., 1970). In contrast, deletion of MDM10 results in a 25 min delay in spindle assembly and a 35 min delay in spindle disassembly. Accounting for the delay in anaphase onset, mdm10∆ cells exhibit a 10 min delay in mitotic exit. The delay in early anaphase and mitotic exit observed in mdm10∆ is similar to the delay in G1 to G2 phase progression. C) Quantitation of the timecourse for formation of second buds in multibudded mdm10∆ cells.

Page 36: Mitochondrial and Mitochondrial DNA Inheritance

29

Interestingly, rho0 cells undergo a G1 delay in cell cycle progression similar to that

observed in mdm10∆; therefore, the decrease in cell cycle progression in mdm10∆ may

be due to loss of mtDNA. However, the multibudded phenotype in mdm10∆ cells is not

due to loss of mtDNA (Fig. 2.1 J), or to defects in septation (degradation of the cell wall

between mother and daughter cells) (data not shown). Spindle alignment was confirmed

to be normal by looking at tubulin staining and nucleolar segregation was visualized by

the visualization of a nucleolus protein, nop1 (data not shown). Rather, it is due to

defects in contractile ring closure. Actomyosin ring contraction was visualized in wild-

type and mdm10∆ cells using a fully-functional fusion protein consisting of the type II

myosin (Myo1p) fused to GFP (Lippincott and Li, 1998), mitochondria-targeted DsRed,

and 4-D imaging (time lapse imaging combined with 3-D reconstruction).

Deletion of MDM10 has no effect on contractile ring assembly: Myo1p-GFP localizes to

a ring at the mother-bud junction in both wild-type and mdm10∆ cells (Fig. 2.4 A-D).

Moreover, mdm10∆ cells have the capacity to undergo contractile ring closure (Fig. 2. 4

B), and to do so with kinetics (14.2 ± 3.5 min, n = 48) similar to that of wild-type cells

(10.4 ± 2.1 min, n = 43). There is some loss of synchrony in mdm10∆ cells at the time of

contractile ring closure. Nonetheless, mdm10∆ cells that undergo contractile ring

closure do so 20-40 min later in the cell cycle compared to wild-type cells (n = 48).

However, mdm10∆ cells exhibit defects in contractile ring closure, which correlates with

defects in mitochondrial inheritance (Fig. 2. 4 C). To quantitate the frequency of

contractile ring closure, Myo1p-GFP and DsRed-labeled mitochondria were visualized in

Page 37: Mitochondrial and Mitochondrial DNA Inheritance

30

cells that bore large buds at the onset of imaging for 2 hrs. During this time, contractile

ring closure occurred in 100% of the wild-type cells examined (n=19) and in only 29% of

the mdm10∆ cell examined (n=38). To evaluate mitochondrial inheritance as a function

of contractile ring closure, we measured the mitochondrial content in buds of mdm10∆

cells that undergo contractile ring closure (Fig. 2. 4 B) and in the first buds (d1) of

multibudded mdm10∆ that failed to undergo contractile ring closure at the mother cell:d1

junction (Fig. 2. 4 E). In wild-type and mdm10∆ cells that undergo contractile ring

closure 43±2% (n = 32), and 36.7±3% (n=37) of mitochondria are in the bud,

respectively. In contrast, there are no detectable mitochondria in 87% of d1 cells within

multibudded mdm10∆ cells (n = 100).

Page 38: Mitochondrial and Mitochondrial DNA Inheritance

31

Figure 2.4. Multibudded clusters of mdm10∆ cells are due to defects in contractile ring closure. A-D) Still frames from time-lapse imaging of Myo1p-GFP (green) and DsRed-labeled mitochondria (red) in synchronized wild-type (ISY008) (A) and mdm10∆ (ISY009) (B-D) cells. Unbudded cells were isolated from mid-log phase cultures by centrifugation through a 10-35% sorbitol gradient for 12 min at 56 x g and visualized by 4D time lapse imaging 1 hr after bud formation for a total of 1 hr. Images were acquired at 3 and 4 min intervals for wild-type and mdm10∆ cells, respectively. Images shown are 2D projections of 3D reconstructions. Arrows point to buds. Numbers indicate time of image acquisition from the onset of bud formation. Bar, 1 µm. A) Wild-type cell undergoing contractile ring closure. B) mdm10∆ cell that has mitochondria in the bud and undergoes contractile ring closure. C) mdm10∆ cells that does not undergo contractile ring closure and has no detectable mitochondria in the bud. D) Multibudded mdm10∆ cell in which the first bud (d1) has no detectable mitochondria, and a contractile ring has assembled at the site of growth of the second daughter cell (d2). E) Mitochondrial morphology and distribution in multibudded cells from synchronized mdm10∆ cells. Cells were grown in SC medium at 30

0C to mid-log

phase (OD600 = 0.5 – 0.8) and incubated with a-factor (10 µM) for 2.5 hrs. Cells were washed and resuspended in medium, fixed at various times after release from pheromone-induced G1 arrest and stained with Calcufluor white to stain bud scars on the mother cell (m) but not on the first or second daughter cell (d1 and d2, respectively) produced from that mother cell (middle panel). DsRed labeled mitochondria are present in the mother cell but not in daughter cells (left panel). Bar, 1 µm. F) Quantification of mitochondrial content in mother cells (m), their first (d1) and second (d2) daughter cells in multibudded mdm10∆ cells from synchronized cell cultures. n = 100 clumps with 3 cell bodies.

Page 39: Mitochondrial and Mitochondrial DNA Inheritance

32

Role for the MEN in regulation of cell cycle progression in mdm10∆

cells.

The MEN regulates cell cycle progression in response to spindle alignment and

elongation, and to the transfer of the nucleus from mother to daughter cell during the

anaphase-to-telophase transition. Cdc14p activation and localization of the active

protein to its sites of action are essential for degradation of a mitotic cyclin (Clb2p),

inactivation of a mitotic cyclin-dependent kinase (CDK; Cdc28p/Clb2p),

dephosphorylation of CDK substrates, and exit from mitosis (Stegmeier and Amon,

2004). However, several studies indicate that the MEN also has a direct role in

regulating contractile ring closure during cytokinesis in budding yeast (Bembenek et al.,

2005; Blondel et al., 2005; Clifford et al., 2008; Corbett et al., 2006; Luca et al., 2001;

Song and Lee, 2001).

mdm10∆ cells undergo mitotic exit, as assessed by degradation of Clb2p and spindle

disassembly (Fig 2.3). To evaluate the role of the MEN in the observed cytokinesis

defect, we studied the localization of Cdc14p-GFP in mdm10∆ and wild-type cells.

Cdc14p is released from its inhibitor Cfi1p/Net1p in the nucleolus during two stages in

the cell division cycle. In early anaphase, separase, as part of the Cdc fourteen early-

anaphase release (FEAR) pathway, promotes a transient and partial release of Cdc14p

from the nucleolus. In a second phase, signal transduction through the MEN releases

the remaining Cdc14p, which facilitates mitotic exit and cytokinesis (D'Amours and

Amon, 2004).

Page 40: Mitochondrial and Mitochondrial DNA Inheritance

33

We confirmed that Cdc14p-GFP in wild-type cells localizes to the nucleolus through

early stages of the cell division cycle, and is released from the nucleolus and localizes

to the spindle pole bodies and bud neck as the spindle apparatus elongates (Fig. 2.5A).

When the spindle is at its maximum length (6-8 µm), 100% of the Cdc14p-GFP is

released from the nucleolus (Fig. 2.5C). In mdm10∆ cells, some cytosolic Cdc14p

localizes to the spindle pole body in mdm10∆ cells bearing fully elongated spindles.

However, release of Cdc14p-GFP from the nucleolus is inhibited by 50% in mdm10∆

cells bearing 4-6 µm spindles, and to a lesser extent in cells with 6-8 µm spindles

compared to wild type cells (Fig. 2.5 B-C). Thus, deletion of MDM10 results in a delay in

release of Cdc14p from the nucleolus.

Page 41: Mitochondrial and Mitochondrial DNA Inheritance

34

Figure 2.5. Cdc14p is mislocalized in mdm10∆cells.

Wild-type (LGY020) and mdm10∆ (LG0Y21) cells

expressing Cdc14p-GFP and mCherry-tagged tubulin were

grown to mid-log phase, fixed and stained with DAPI as for

Fig. 2.1. The images shown are 2-D projections from

reconstructed 3-D volumes. An overlay of Cdc14p-GFP

(green) and tubulin in the mitotic spindle (red) are shown

(left). An overlay of Cdc14p-GFP (green) and DAPI (blue)

are shown (right). Cell outlines are shown in white. White

arrow: spindle pole body. White arrowhead: nucleus. Red

arrowhead: mother-bud neck. Bar, 1 µm. A) Cdc14p-GFP

localization in wild-type cells. Cdc14p-GFP localizes to the

nucleolus in cells bearing short, but detectable spindles

(upper panels), to the nucleus and spindle pole bodies in

early anaphase when spindles are 4-6 µm in length

(middle panels) and to spindle pole bodies and the mother-

bud neck during telophase when spindles have elongated

and reached their maximum length of 8-10 µm (lower

panels). B). Defects in localization of Cdc14p-GFP in

mdm10∆ cells. C) Quantitation of the release of Cdc14p

from the nucleolus in wild-type and mdm10∆ cells as a

function of spindle length. Error bars show standard error

of the mean (n>200).

Sli15p, a chromosomal passenger protein and substrate for Cdc14p that is present at

kinetochores during metaphase and transfers to the spindle midzone during early

anaphase (D'Amours and Amon, 2004), localizes to the spindle to the same extent in

mdm10∆ and in wild-type cells (data not shown).

Page 42: Mitochondrial and Mitochondrial DNA Inheritance

35

Thus, mislocalization of Cdc14p in mdm10∆ cells is due to an alteration in MEN-

mediated control of Cdc14p and not the FEAR pathway. In light of these findings and

our observation that release of Cdc14p from the nucleolus is partially inhibited in

mdm10∆ cells, it is possible that the level of MEN-mediated Cdc14p activation in

mdm10∆ cells is sufficient to support mitotic exit but insufficient to support cytokinesis.

Consistent with this, conditions that hyperactivate the MEN promote cytokinesis in

mdm10∆ cells. Deletion of BUB2 suppresses the subtle mitotic exit defect observed in

mdm10∆ cells, but has no effect on the time of entry of mdm10∆ cells into anaphase

Deletion of BUB2 or overexpression of CDC5 in mdm10∆ cells results in a 67%

decrease in the number of multibudded cells in late-log phase cell cultures compared to

mdm10∆ cells (Fig. 2.6 A-B). Thus, conditions that bypass MEN regulation bypass the

cytokinesis defects observed in mdm10∆cells. To determine whether other mutations

that inhibit mitochondrial inheritance also affect cytokinesis, we studied GEM1, a

member of the rho (Miro) family of GTPases and MMR1, a protein that localizes to

mitochondria, binds to the type V myosin Myo2p and is required for anchorage of

mitochondria in the bud tip (Itoh et al., 2002) (Frederick et al., 2008). mmr1∆ or gem1∆

mutants exhibit subtle defects in mitochondrial inheritance, and low but detectable

defects in cytokinesis. However, gem1∆ mmr1∆ double mutants exhibit mitochondrial

distribution and inheritance defects that are significantly greater than those observed in

either single mutant (Frederick et al., 2008) and a cytokinesis defect that is more

Page 43: Mitochondrial and Mitochondrial DNA Inheritance

36

severe than that observed in either single mutants and similar to that observed in the

mdm10∆ mutant. In addition, deletion of BUB2 suppresses the cytokinesis defect

observed in the gem1∆ mmr1∆ double mutant (Fig. 2.6C). These findings provide

additional evidence for the existence of a mechanism to inhibit cell cycle progression at

cytokinesis when there are severe defects in mitochondrial inheritance.

Finally, the primary function of a checkpoint is to ensure that critical cell division

processes occur with high fidelity and at the correct time as cells divide. Thus, if the

MEN regulates cell cycle progression in response to mitochondrial inheritance, then

hyperactivation of the MEN should reduce the fidelity of mitochondrial inheritance.

Indeed, we find that conditions that bypass MEN regulation, deletion of BUB2 or

overexpression of CDC5, result in defects in partitioning of mitochondria between

mother cells and buds (Fig. 2.6D). Deletion of BUB2 reduces the amount of

mitochondria in daughter cells. Deletion of MDM10 produces more severe defects in

the fidelity of mitochondrial inheritance. Finally, mdm10∆ mutants bearing a deletion in

BUB2 or overexpressing CDC5 exhibit defects in mitochondrial partitioning that are

more severe than that in mdm10∆ mutants.

Page 44: Mitochondrial and Mitochondrial DNA Inheritance

37

Figure 2.6. Hyperactivation of the MEN suppresses the defect in cytokinesis defect observed in mdm10∆ cells. A) mdm10∆ cells that expressed mitochondria-targeted DsRed and contained either no plasmid (ISY002) or plasmid-borne CDC5 under control of the GAL promoter (ISY048) incubated in galactose-based media for 5.5 hrs. Images are phase-contrast images superimposed on fluorescence images of DsRed-labeled mitochondria. Bar, 3 µm. B) Quantitation of multibudded cells in wild-type cells and mdm10∆ cells that overexpress CDC5 or carry a BUB2 deletion. Wild-type, CDC5 overexpression, mdm10∆ and mdm10∆ overexpressing CDC5 strains ISY001, ISY048, ISY002 and ISY013. Wild-type, bub2∆, mdm10∆, and bub2∆ mdm10∆ strains are BY4741, 6189, 398, and LGY025. Cell culture and quantitation were carried out as for Fig. 2.1 D. Error bars show standard deviations (n>800). C) Quantitation of multibudded cells in wild-type (BY4741), mmr1∆ (4139), gem1∆ (357), mmr1∆ gem1∆ (DCY001) and

mmr1∆ gem1∆ bub2∆ cells (DCY002) that were analyzed after sonication and treatment with zymolyase 20T, a protein mixture that catalyzes cell wall degradation, (0.1 mg/ml for 10 min at RT). Error bars show standard deviations (n>100). D) Hyperactivation of the mitotic exit network results in mitochondrial partitioning defects. Mid-log phase wild-type, bub2∆, mdm10∆, mdm10∆ bub2∆ and CDC5 overexpressing cells (ISY001, ISY028, ISY002, ISY029, and ISY013), which express mitochondria-targeted DsRed, were fixed, and images of yeast bearing large buds (buds >2/3 the length of their mother cells) were collected at 1-µm z-intervals. Mitochondrial area in the mother cell or bud was measured in each z-section using a user-defined threshold, and these areas were summed over the mother cell or bud to determine mitochondrial volume. Mitochondrial partitioning ratios are the mitochondrial volume in the bud/mother. Error bars show standard error of the mean (n>250).

Page 45: Mitochondrial and Mitochondrial DNA Inheritance

38

Overall, there are numerous cell cycle checkpoints to monitor events associated with

nuclear inheritance, including replication of nuclear DNA and segregation of

chromosomes and nuclei. Here, we provide evidence for a mitochondrial inheritance

checkpoint that inhibits cytokinesis when there are defects in mitochondrial inheritance

in budding yeast, and for a role for the MEN in this process. Since the mitochore has

been implicated in association of mitochondria with ER (Kornmann et al., 2009), it is

possible that these interactions could contribute to cytokinesis. Moreover, in Drosophila

melanogaster, mitochondrial second messengers, either ROS or ATP, can function as

two independent signals to enforce checkpoints at G1/S that are not due to metabolic

restriction (Owusu-Ansah et al., 2008). Our findings indicate that a checkpoint for

mitochondrial inheritance, that is also independent to metabolic restriction, exist in

budding yeast. Finally, since there are mechanisms to insure the inheritance of many

organelles and the MEN is a conserved pathway, our findings also raise the possibility

that there are similar checkpoints for organelle inheritance in yeast and other cell types.

Experimental Procedures

A summary of the materials and methods used for this study is included. Please refer to

Supplemental Information for more detailed description.

Yeast strains, plasmids, and growth conditions: Yeast strains used in this work are listed

in Table 2.1. Strain ISY065 is a derivative of W303. Strains MYY291 and DNY416 were

derived from A364A. All other strains were derived from BY4741. rho0 derivatives were

Page 46: Mitochondrial and Mitochondrial DNA Inheritance

39

generated from wild-type cells expressing plasmid-borne mitochondria-targeted DsRed

(ISY001), as described by Goldring et al. (Goldring et al., 1970). Other yeast methods

were performed according to Sherman (Sherman, 2002).

The carboxy terminus of Myo1p and Cdc14p were tagged with GFP using PCR-based

insertion into the chromosomal copies of the MYO1 or CDC14 loci (Longtine et al.,

1998). Table 2.2 lists primers used to tag these genes. Standard molecular techniques

for cloning procedures were used. Mitochondria were visualized using a fusion protein

expressed from the plasmid pRS426ADH + PreFoATPase-(subunit 9)-DsRed or from

the plasmid pTDT104GAL1 + PreFoATPase-(subunit 9)-DsRed (gifts from Dr. J. Shaw,

University of Utah). Tubulin was visualized using fusion proteins expressed from the

plasmid [pAFS125 TUB1-GFP] or [pRS406 TUB1-mCherry] (gifts from Dr. K. Bloom,

University of North Carolina at Chapel Hill). GAL-CDC5 bub2∆ cells (A4453) and

plasmid pGal(myc)3CDC5-306 were gifts from Dr. A. Amon (MIT) and plasmid

[pGP195-2 (pRS305-SLI15-GFP KanMX6)] a gift from Dr. E Schiebel (University of

Heidelberg). Growth conditions for individual experiments are described in the figure

legends.

To construct the mmr1∆ gem1∆ double mutant (DCY001), MMR1 was deleted in a

gem1∆ strain (357) using an insertion cassette in which the selectable marker LEU2

was flanked with loxP recombination sites. LEU2 was later excised from DCY001 using

plasmid-borne CreLox under control of the galactose inducible promoter (pSH47). To

construct the mmr1∆ gem1∆ bub2∆ triple mutant (DCY002), pSH47, which carried a

Page 47: Mitochondrial and Mitochondrial DNA Inheritance

40

URA3 marker, was cured from DCY001 using FOA, and BUB2 was deleted using an

insertion cassette containing a LEU2 marker.

Yeast cell viability was measured using FUN-1, a halogenated unsymmetric cyanine

dye that was developed for assessing the viability and metabolic activity of yeast

(Millard et al., 1997). FUN-1 is membrane permeate, that binds to nucleic acids and is

biochemically processed in living yeast to produce a cylindrical intravacuolar structure

that is red shifted in fluorescence emission compared to the unprocessed form.

Incubation of yeast with FUN-1 was carried out according to manufacturer’s

recommendations (Invitrogen - Molecular Probes, Carlsbad, CA). The conversion of

FUN-1 by viable yeast cells was quantified using fluorescence microscopy. Cells with

prominent fluorescent intravacuolar structures were scored as live. Cells that lacked

these structures and had diffuse cytosolic fluorescence that was green or yellow were

scored as non-viable.

Fluorescence microscopy, image analysis and cytology: Cells were gently pelleted and

mounted directly in 2% low melting agarose on a coverslip. Fluorescence/phase

microscopic images were collected using an E600 microscope (Plan-Apo 100X/1.4 NA

objective) (Nikon, Melville, NY) equipped with a cooled CCD camera (Orca-ER,

Hamamatsu, Japan), and a Dual-View image splitter (Optical Insights, Tucson, AZ) for

simultaneous two-color imaging. Openlab 3.1.5 software (Improvision, Lexington, MA)

was used to acquire images. Z-stacks of 0.2-µm slices were obtained and the out-of-

Page 48: Mitochondrial and Mitochondrial DNA Inheritance

41

focus light was removed using an iterative deconvolution algorithm in Volocity 2.6

(Improvision, Lexington, MA). All z-sections were assembled and 3-D projections were

generated with comparable parameters and thresholds.

Each cluster of more than two attached cell bodies was counted as cell separation

failure and those cells were scored as multibudded cells. To stain bud scars,

formaldehyde-fixed cells were incubated in 10 µg/ml Calcofluor (Invitrogen Molecular

Probes, Carlsbad CA) for 30 min at RT. For cell wall digestion, cells were fixed by

incubation with formaldehyde (3.7%) for 1 hr at RT. After washes to remove the

fixative, cells were incubated in 0.1 mg/ml zymolyase 20T (Seikagaku Corp., Tokyo

Japan) for 10 min at RT. To stain DNA, formaldehyde-fixed cells were incubated with 1

µg/ml DAPI (Molecular Probes, Eugene, OR) for 5 min at RT. For cell cycle

synchronization, cells were incubated with α-factor (10 µM) for 2.5 hrs. Cells were

released from arrest by washing and were transferred to pheromone-free media.

Induction of CDC5 expression in cells carrying the pGal(myc)3CDC5-306 plasmid was

carried out by growth in SC-Raff (2%) medium and transfer to SC medium containing

raffinose (2%) and galactose (2%).

Flow cytometry: Analysis of DNA content in propidium iodide-stained, synchronized cell

cultures was determined according to Paulovich and Hartwell using a fluorescence-

activated cell analyzer (Becton Dickerson LSRII, Franklin Lakes, NJ). The percent of

Page 49: Mitochondrial and Mitochondrial DNA Inheritance

42

total cells in G2 phase was determined using the Flowjo program (TreeStar Inc.,

Ashland, OR).

Protein and immunological techniques: Protein extracts of mid-log phase yeast cells for

Western blot analysis were obtained as described (Boldogh et al., 1998). The

bicinchoninic acid (BCA) assay (Pierce Chemical, Rockford, IL) was used for protein

concentration determinations. Immunoblot analysis of the total amount of Clb2p and

Hxk1p was performed with antibodies specific for Clb2p (a gift from Dr. Doug Kellogg,

University of California) and Hxk1p (a gift from Dr. Gottfried Schatz, University of Basel).

HRP-conjugated secondary antibodies and Supersignal detection (Pierce Chemical,

Rockford, IL) were used to visualize bands.

Yeast strains, plasmids, and growth conditions:

Yeast strains used in this work are listed in Table 2.1. rho0 derivatives were generated

from wild-type cells expressing plasmid-borne mitochondria-targeted DsRed (ISY001),

as described by Goldring et al. (Goldring et al., 1970). Other yeast methods were

performed according to Sherman (Sherman, 2002). Yeast cell viability was measured

using FUN-1 (Millard et al., 1997).

Page 50: Mitochondrial and Mitochondrial DNA Inheritance

43

The carboxy terminus of Myo1p and Cdc14p were tagged with GFP using PCR-based

insertion into the chromosomal copies of the MYO1 or CDC14 loci (Longtine et al.,

1998). Standard molecular techniques for cloning procedures were used sambrook

1998 Coldspring harbor.

Fluorescence microscopy, image analysis and cytology: Mitochondria, tubulin and

Sli15p were visualized using plasmid borne GFP fusion proteins. Chitin in bud scars and

DNA were visualized using Calcofluor White and DAPI. Acquisition, manipulation and

analysis of fluorescence images was carried out as described previously (Sogo and

Yaffe, 1994).

Yeast strains used for this study

Table 2.1

Strains Genotype Source

357 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 gem1∆::KANMX Open

Biosystems

398 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 mdm10∆::KANMX Open

Biosystems

4139 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 mmr1∆::KANMX Open

Biosystems

BY4741 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 Open

Biosystems

DCY001 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 gem1∆::KANMX

mmr1∆::LEU2

This study

Page 51: Mitochondrial and Mitochondrial DNA Inheritance

44

DCY002 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 gem1∆::KANMX

mmr1∆ bub2∆::LEU2

This study

DNY416 mdm10::URA3 ura3 leu2 his3 Boldogh et

al., 2003

ISY001 MATa his3∆1 leu2∆0 met15∆0 ura3∆0

[pRS426ADH+PreF0ATPase-DsRed]

This study

ISY002 MATa mdm10∆::KANMX6 his3∆1 leu2∆0 met15∆0 ura3∆0

[pRS426ADH+PreF0ATPase-DsRed]

This study

ISY003 MATa mdm12∆::KANMX6 his3∆1 leu2∆0 met15∆0 ura3∆0

[pRS426ADH+PreF0ATPase-DsRed]

This study

ISY005 MATa mas37∆::KANMX6 his3∆1 leu2∆0 met15∆0 ura3∆0

[pRS426ADH+PreF0ATPase-DsRed]

This study

ISY006 MATa tom7∆::KANMX6 his3∆1 leu2∆0 met15∆0 ura3∆0

[pRS426ADH+PreF0ATPase-DsRed]

This study

ISY007 MATa cbk1∆::HIS3 his3∆1 leu2∆0 met15∆0 ura3∆0 This study

ISY008 MATa MYO1-GFP::HIS3 his3∆1 leu2∆0 met15∆0 ura3∆0

[pRS426ADH+PreF0ATPase-DsRed]

This study

ISY009 MATa mdm10∆::KANMX6 MYO1-GFP::HIS3 his3∆1 leu2∆0

met15∆0 ura3∆0 [pRS426ADH+PreF0ATPase-DsRed]

This study

ISY013 MATa mdm10∆::KANMX6 CDC14-GFP::HIS3 his3∆1

leu2∆0 met15∆0 ura3∆0 [pGal(myc)3CDC5-306]

[pTDT104GAL1+ PreF0ATPase-DsRed]

This study

ISY016 MATa CDC14-GFP::HIS3 his3∆1 leu2∆0 met15∆0 ura3∆0

[pRS426ADH+PreF0ATPase-DsRed]

This study

ISY018 MATa mdm10∆::KANMX6 CDC14-GFP::HIS3 his3∆1

leu2∆0 met15∆0 ura3∆0 [pRS426ADH+PreF0ATPase-

DsRed]

This study

ISY028 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 bub2∆::HIS3

[pRS426ADH+PreF0ATPase-DsRed]

This study

ISY029 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 bub2∆::HIS3 This study

Page 52: Mitochondrial and Mitochondrial DNA Inheritance

45

mdm10∆::KANMX [pRS426ADH+PreF0ATPase-DsRed]

ISY048 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 [pGal(myc)3CDC5-

306] [pRS426ADH+PreF0ATPase-DsRed]

This study

ISY065 MAT∆ mmm1-1, leu2-∆1 trp1-∆1 his3-∆200 ura3 ade2 his3

leu2 lys2 trp2 ura3 [pRS426ADH+PreF0ATPase-DsRed]

This study

LGY020 MATa CDC14-GFP::HIS3 his3∆1 leu2∆0 met15∆0 ura3∆0

[pRS406 TUB1-mCherry]

This study

LGY021 MATa mdm10 �::KANMX6 CDC14-GFP::HIS3 his3∆1 leu2∆0

met15∆0 ura3∆0 [pRS406 TUB1-mCherry]

This study

LGY022 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 [pRS406 TUB1-

mCherry] [pGP195-2 (pRS305-SLI15-GFP KanMX6)]

This study

LGY023 MATa mdm10∆::KANMX6 his3∆1 leu2∆0 met15∆0 ura3∆0

[pRS406 TUB1-mCherry] [pGP195-2 (pRS305-SLI15-GFP

KanMX6)]

This study

LGY024 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 bub2∆::HIS3 This study

LGY025 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 bub2∆::HIS3

mdm10∆::KANMX

This study

MYY291 ura3 leu2 his3 Yaffe and

Smith, 1991

Page 53: Mitochondrial and Mitochondrial DNA Inheritance

46

CHAPTER 3

MtDNA INHERITANCE CHECKPOINT

Page 54: Mitochondrial and Mitochondrial DNA Inheritance

47

BACKGROUND

Mitochondria contain >1,000 proteins. The majority of these proteins are encoded by

nuclear DNA, undergo Mendelian inheritance and are imported from the cytoplasm into

the mitochondria (Chen and Butow, 2005). In addition, mitochondria contain DNA

(mtDNA), which encodes respiratory chain components or RNAs that are essential for

mitochondrial protein synthesis. mtDNA varies in size among organisms. The budding

yeast Saccharomyces cerevisiae has a large mitochondrial genome of around 75-80 kb.

In contrast, the human mitochondrial genome is 16.5 kb (Chen and Butow, 2005).

Human mtDNA contains 37 genes, 13 encode proteins that participate in oxidative

phosphorylation, 22 encode tRNAs and 2 encode ribosomal RNAs (Wallace, 2010).

mtDNA is organized and packaged as a DNA-protein complex called the mtDNA

nucleoid. Haploid cells of the yeast Saccharomyces cerevisiae with wild-type

mitochondrial genomes (rho+ cells) contain 10–20 mtDNA nucleoids per cell (Williamson

and Fennell, 1979). This macromolecular complex is inherited from mother to daughter

cells during cell division (Williamson and Fennell, 1979) (Stevens, 1981/coldspring

harbor). Many proteins that are required for stabilizing and or packaging of mtDNA

nucleoids have been identified including DNA binding proteins (e.g. the abundant high

mobility group-box, mtDNA-binding nucleoid packing protein Abf2p), mtDNA replication

proteins (e.g. human DNA polymerase γ and Mip1 in yeast), and transcription (e.g.

human TFAM (transcription factor A and yeast Abf2) (Zelenaya-Troitskaya et al., 1998).

Another protein is Mgm101p, a DNA-binding protein that is essential for mtDNA

Page 55: Mitochondrial and Mitochondrial DNA Inheritance

48

maintenance and required for repairing oxidative damaged mtDNA and localizes to a

‘subset’ area around mtDNA nucleoids (Meeusen and Nunnari, 2003) (Chen et al.,

1993). Mgm101p also interacts with Mmm1p, a mitochore protein that is required for

mitochondrial inheritance and for maintenance of mtDNA (Meeusen and Nunnari, 2003).

One other protein that localizes to mtDNA nucleoids and is relevant to my thesis

research is Pif1. Pif1p is part of the superfamily 1 DNA helicases, proteins that unwind

DNA and are essential for DNA replication, recombination, and repair (Budd et al.,

2006) (Chang et al., 2009). There are two forms of Pif1. One localizes to the nucleus

and contains a nuclear localization sequence. Recent studies indicate that DNA

damage results in Rad53-dependent phosphorylation of Pif1 at nuclear DNA damage

breaks, which inhibits telomerase activity at those sites, and prevents telomerase-

dependent addition of telomeres at DNA breaks (Makovets and Blackburn, 2009). The

other form of Pif1 contains a mitochondrial targeting sequence, localizes to mtDNA

nucleoids and is required for mtDNA maintenance (Foury and Kolodynski, 1983). Cheng

et al. reported that pif1∆ cells have mtDNA breaks at specific sites and proposed that

Pif1p either prevents or repairs mtDNA dsDNA breaks (Cheng et al., 2007).

OTHER PROTEINS IMPLICATED IN mtDNA INHERITANCE

A number of proteins on the mitochondrial surface, including the mitochore and

Mdm34p, have also been implicated in maintenance of mtDNA. These proteins localize

to punctate structures on mitochondria that are adjacent to mtDNA nucleoids and are

Page 56: Mitochondrial and Mitochondrial DNA Inheritance

49

required for maintenance of mtDNA (Boldogh et al., 2003) (Hobbs et al., 2001)

(Youngman et al., 2004).There is also evidence that the mitochore may interact with the

mtDNA replication protein complexes. Mip1p is a subunit of mtDNA polymerase gamma

of yeast. The Mgm101p/Mip1p protein complex associates with mtDNA nucleoids and is

essential for mtDNA replication and mtDNA genome maintenance (Meeusen and

Nunnari, 2003). Mitochore-containing foci are adjacent to and move with a protein

complex consisting of Mgm101p and Mip1p. Consistent with this, Mgm101p has

physical and genetic interactions with the mitochore subunit Mmm1p. These data reveal

a physical interaction between the mitochore and mtDNA nucleoids, and provide

evidence that the mechanism for mitochondrial motility during inheritance is linked to the

process of mtDNA replication and inheritance.

mtDNA MUTATIONS: rho0 AND rho- CELLS

mtDNA has a very high mutation rate, resulting in clinical manifestations (Wallace,

2010). Epidemiological studies indicate that the frequency of human mtDNA mutations

is in the order of 1/5000 (Wallace, 2010). Further, gene expression microarray studies

revealed that loss of human mtDNA results in down-regulation of cell cycle regulatory

genes and a reduction of cell replication rates resulting in a longer S-phase (Mineri et

al., 2009).

Since yeast are facultative aerobes, they can withstand mtDNA mutations that are not

readily tolerated by mammalian cells. In yeast there are two broad classes of mtDNA

mutants: rho- and rho0 cells. rho0 cells contain no mtDNA, while rho- cells contain

Page 57: Mitochondrial and Mitochondrial DNA Inheritance

50

mtDNA that carries mutations that block mitochondrial respiratory activity. As a result,

both of these cell types have defects in respiration and cannot grow in media containing

non-fermentable carbon sources like glycerol or lactate. Moreover, upon growth on

fermentable carbon sources like glucose, rho0 and rho- cells produce small colonies

called petites.

Early studies on rho- cells that carry large deletions in mtDNA revealed that yeast cells

exercise tight control of the mtDNA genome size. Interestingly, the total size of the

mtDNA genome in these rho- cells is similar to that of rho+ cells (~80 kb) (Dujon, 1981-

coldspring harbor publication). Later studies revealed that rho- cells maintain the size of

their mtDNA genome by amplifying existing mtDNA and creating tandem repeats of the

amplified DNA until the critical size of ~80kb is reached (Figure 3.1). rho- cells contain

fewer mtDNA nucleoids per cell that stain more brightly with DNA binding dyes

compared to rho+ cells, but they still continue to contain the same mass of mtDNA

(Dujon, 1981-coldspring harbor publication).

The ability of yeast to amplify mtDNA to a critical mass supports the existence of

surveillance machinery to monitor mtDNA mass, and regulatory machinery to up-

regulate mtDNA synthesis until it reaches a critical genomic mass. We identified a

mtDNA mass surveillance system, which inhibits cell cycle progression from G1 to S

phase in response to loss of mtDNA. Moreover, we find that this mtDNA inheritance

checkpoint is regulated by a conserved G1 to S checkpoint pathway, the Rad53 DNA

damage checkpoint.

Page 58: Mitochondrial and Mitochondrial DNA Inheritance

51

FIGURE 3.1- mtDNA content is under tight cellular control. Yeast have the ability to

amply by tandem repeats small fragments of mtDNA until the total size of the mtDNA is

similar to that found in cells that contain wild type mtDNA (75 kb). This control suggests

that there is a regulatory system that monitors a critical mass of mtDNA. This cartoon

depicts a small fragment that is amplified by tandem repeats to the critical size of 75kb.

This new ‘amplified’ genome is a respiratory incompetent rho- cell.

Rad53 AND THE DNA DAMAGE CHECKPOINT

The yeast proteins kinases, Mec1 and Rad53 are components of a conserved

checkpoint that arrests the cell cycle at G1, S, and G2 phase and activates pathways to

repair damaged DNA in response to damage of nuclear DNA (Elledge, 1996). The

response to DNA damage is dependent on when the insult takes place and the type of

damage that occurs. Two major checkpoints occur in response to DNA damage or

replication errors. The DNA damage checkpoint functions throughout the entire cell

Page 59: Mitochondrial and Mitochondrial DNA Inheritance

52

cycle including S-phase whereas the DNA replication checkpoint pathway is restricted

to cells undergoing active replication (Schwartz et al., 2003).

Mec1 is a homolog of the human ATR (Ataxia- and Rad-related) gene, while Rad53 is a

homolog of human CHK2 (Chabes et al., 2003). In response to DNA damage, Mec1 and

Ddc2 complexes are recruited to the site of damage or stalled DNA replication fork,

which leads to phosphorylation and activation of Rad53 (Cordon-Preciado et al., 2006).

Activation of Rad53 leads to cell cycle arrest and activation of genes that contribute to

DNA repair. Rad53 triggers a G1 arrest by stimulating the phosphorylation of Swi6,

which leads to down-regulation of the transcription of the G1 cyclins, CLN1 and CLN2

(Sidorova and Breeden, 1997).

Rad53 has 16 potential phosphorylation sites, a catalytic domain and two forkhead

homology-associated (FHA) domains (FHA-1, FHA-2) (Lee et al., 2003). Both FHA

domains are required for the DNA damage dependent pathway; however, either FHA1

or FHA2 domain allows activation of Rad53 in response to a replication block (Schwartz

et al., 2003). FHA domains interact with both upstream and downstream effectors of the

DNA checkpoint signaling pathways (Schwartz et al., 2003). In response to DNA

damage, Rad9 is phosphorylated and binds to Rad53 via the FHA-1 and FHA-2

domains. Replicative stress activates Rad53 by a Rad9 independent mechanism that

may involve an intermediate kinase, Mrc1 (Alcasabas et al., 2001).

Page 60: Mitochondrial and Mitochondrial DNA Inheritance

53

Figure 3.2 Rad53 activation during DNA damage. Checkpoints govern many responses to DNA damage and blocks to DNA replication. A. Rad9 or POL2 class genes encode proteins that recognize either damage (* denotes damage by single-strand gaps, stalled replication fork) and then activate Mec1, Rad53). The signal transducers mediate multiple responses and are dependent on the type of insult or where in the cell cycle the damage took place. B. Dun1 can also be activated from double stranded breaks or other DNA insults. C. Replication fork recognize DNA lesions due to damage or dNTP deprivation. Helicases and polymerases uncouple exposing the DNA where RPA can bind to and activate the checkpoint response. Mec1 is recruited which initiates the signaling cascade involving Rad53 and other downstream kinases.

Ribonucleotide reductase (RNR) is an important target that is activated by Rad53 and

serves to promote repair of DNA damage. RNR catalyzes the rate-limiting step in dNTP

synthesis (Zhao et al., 1998) (Reichard, 1988) and affects both the fidelity of DNA

replication and cell viability by regulating the dNTP pools (Zhao and Rothstein, 2002).

Rad53 increases dNTP levels by activating the downstream kinase Dun1, which leads

to the transcriptional induction of the RNR genes. In conjunction, Mec1/Rad53 increase

Page 61: Mitochondrial and Mitochondrial DNA Inheritance

54

RNR levels by regulating the degradation of the negative inhibitor Sml1 (Zhao and

Rothstein, 2002). Interesting, regulation of dNTP levels is the essential function of

Rad53: the Mec1 or Rad53 lethality can be suppressed by overexpression of RNR

genes or by inhibition of Sml1 (Suppressor of Mec1 Lethality) (Koc and Merrill, 2007)

(Zhao et al., 1998) (Chabes et al., 2003).

After successful repair of DNA damage, Rad53 is downregulated by phosphatase-

mediated de-phosphorylation and not degradation (O'Neill et al., 2007), which allows

cells to reset cell cycle progression (Cordon-Preciado et al., 2006; Pellicioli et al., 2001).

Dephosphorylation of Rad53 by PTC2 and PTC3 phosphatases can also allow cells to

be released from the checkpoint when DNA repair attempts fail. During this process,

which is referred to as recovery from the checkpoint, cells reenter the cell cycle with

damaged DNA.

LINKS BETWEEN Rad53 AND mtDNA

Cells contain multiple copies of mtDNA, ranging from 20–50 in yeast to as many as

10,000 in mammalian cells. Nuclear genes have direct interactions with genes that

control mtDNA replication, maintenance and copy number in mammalian cells and

yeast (Taylor et al., 2005) (Tyynismaa et al., 2004) (Matsushima et al., 2004) (Schultz et

al., 1998) (Zelenaya-Troitskaya et al., 1998). While conserved genes have been

implicated in this process, signaling pathways involved in modulating cellular mtDNA

content have not been fully elucidated in any organism.

Page 62: Mitochondrial and Mitochondrial DNA Inheritance

55

The Shadel laboratory obtained the first evidence that mtDNA copy number is regulated

by DNA damage checkpoint pathways in yeast and mammalian cells. They found that

the Ataxia-Telangiectasia mutated (ATM) gene, a kinase that regulates cell cycle

progression in response to DNA damage, and its target, ribonucleotide reductase (RNR)

regulates mtDNA copy number in human cells (Eaton et al., 2007). Specifically, they

found that mtDNA copy number is reduced in Ataxia-telangiectasia (A-T) patient

fibroblasts, wild-type fibroblasts treated with an ATM inhibitor, and cells in which RR is

inhibited by drug treatment or RNA interference (RNAi).

Their studies in yeast revealed that deletion of the RNR inhibitor SML1 increases

mtDNA copy number and that deletion of RAD53 and SML1 results in an increase in

mtDNA copy number that is greater than that observed in sml1∆ cells (Lebedeva and

Shadel, 2007). Thus, mtDNA copy number in yeast is regulated by Rad53, in part

through effects on RNR activity and dNTP pool size. Finally, other studies indicate that

Pif1 and RRM3, helicases that are associated with nuclear DNA and mtDNA have

genetic interactions with Rad53: 1) overexpression of the RAD53 target, RNR,

suppresses the loss of mtDNA associated with deletion of PIF1, and 2) deletion of PIF1

and RRM3 results in activation of Rad53 (Taylor et al., 2005).

Together, these studies revealed a functional link between the DNA damage

checkpoints and mtDNA. Moreover, they may explain some of the symptoms of A-T that

are not readily explained by damage to nuclear DNA. Here, I present evidence for a

mtDNA inheritance checkpoint, a pathway that inhibits progression from G1 to S in

Page 63: Mitochondrial and Mitochondrial DNA Inheritance

56

response to loss of mtDNA. Furthermore, we find that the checkpoint is regulated by

Rad53, an established checkpoint signaling molecule that has known effect on nuclear

DNA and mtDNA.

RESULTS

LOSS OF mtDNA INDUCES A G1 ARREST IN CELL CYCLE PROGRESSION

Treatment with ethidium bromide (EtBr) results in the conversion of respiratory

competent yeast (grandes) to respiratory deficient mutants (petites) at efficiencies close

to 100%. Early studies revealed that EtBr treatment inhibits mtDNA synthesis and

increases mtDNA degradation, which results in production of rho0 cells that have no

mtDNA (Goldring et al., 1970). We used EtBr to generate rho0 cells, and confirmed the

loss of mtDNA, by staining with the DNA binding dye 4',6-diamidino-2-phenylindole

(DAPI) which is a fluorescent stain that binds strongly to A-T rich regions in DNA.

Punctate mtDNA nucleoids are present in rho+ cells that have intact mtDNA but not in

rho0 cells that have no mtDNA (Figure 3.3A). We also confirmed that loss of mtDNA

results in loss of mitochondrial respiratory activity by assessing growth on a non-

fermentable carbon source, such as glycerol (SFig 3.1).

We used flow cytometry to assess the effect of the loss of mtDNA on cell cycle

progression. Wild-type rho+ cells exhibit normal cell cycle progression (Fig. 3.3). They

undergo transition from G1 to S phase 40-60 min, and transition from S to G2 phase

Page 64: Mitochondrial and Mitochondrial DNA Inheritance

57

100-120 min after release from pheromone-induced G1 arrest. In contrast, we find that

30-60% of rho0 cells fail to progress from G1 to S phase. Rho0 cells that do progress

through the cell cycle do so at a lower rate. They undergo transition from G1 to S

phase 60-100 min, and transition from S to G2 phase 150 min after release from

pheromone-induced G1 arrest.

Figure 3.3. Loss of mtDNA results in defects in progression from G1 to S phase. Wild-type (BY4741) or cells lacking mtDNA (rho0, mgm101∆ cells) were grown to mid-log phase (OD600 = 0.5–0.8) and arrested G1 phase by treatment with pheromone (α-factor; 10-100µm) for 2.5-3 hrs. Samples were released from pheromone-induced G1 arrest by washing, resuspending in synthetic complete (SC) media, and propagated at 30°C. Aliquots were removed at the time shown. Cells were fixed, stained with propidium iodine (PI) and their DNA content as a function of time after release from G1 arrest was analyzed by flow cytometry. (A) Maximum projections of deconvolved images of WT, rho0 and mgm101∆ rho0 cells stained with the DNA binding dye DAPI. n: nuclear DNA, m: mtDNA. Cell outlines are shown in white Bar, 1 µm. (B) Wild-type cells with mtDNA (BY4741 rho+) exhibit normal cell cycle progression. Cells lacking mtDNA either by EtBr treatment (BY4741 rho0) or by deletion of a gene required for mtDNA maintenance (mgm101∆ rho0) fail to progress normally and show a G1 defect. (C) Quantitation of progression through G1 phase was assessed as the ratio of the percent of cells in G2 phase at the time of release from pheromone-induced G1 arrest to the percent of cells in G2 at the time specified.

Page 65: Mitochondrial and Mitochondrial DNA Inheritance

58

There are significantly fewer mechanisms for repair of DNA damage in mitochondria

compared to nuclei. As a result, mtDNA is 10-times more sensitive to damage by

chemical induction compared to nuclei (Cheng et al., 2007). Therefore, it is likely that

EtBr is affecting mtDNA and not nuclear DNA. However, it is formally possible that the

EtBr that was used to generate rho0 cells is damaging nuclear DNA and triggering a cell

cycle progression defect through checkpoints that detect nuclear DNA damage. To

address this issue, we generated rho0 cells by deletion of MGM101, which encodes a

DNA-binding protein that is essential for mtDNA maintenance and required for repair of

oxidatively damaged mtDNA (Meeusen and Nunnari, 2003) (Chen et al., 1993). We

observed the rho0 mgm101∆ cells exhibit a stronger defect in the ability to exit G1

compared to rho0 cells produced by EtBr treatment (Fig. 3.3). The stronger defect in cell

cycle progression in rho0 mgm101∆ cells may be a result of the rapid loss of mtDNA that

occurs upon deletion of MGM101. Together, this data indicates that loss of mtDNA

results in defects in cell cycle progression from G1 to S phase.

THE G1 TO S PROGRESSION DEFECT OBSERVED IN CELLS LACKING mtDNA IS

NOT DUE TO LOSS OF MITOCHONDRIAL RESPIRATORY ACTIVITY OR ENERGY

PRODUCTION

Previous studies revealed the loss of mitochondrial respiratory activity results in a defect

in progression from G1 to S phase in Drosophila (Mandal et al., 2005) (Owusu-Ansah et

al., 2008). These studies revealed that defects in mitochondrial function produced by

either deletion of cytochrome oxidase complex Va or disruption of complex 1 results in

Page 66: Mitochondrial and Mitochondrial DNA Inheritance

59

two retrograde signals that lead to down-regulation of cyclin E and ROS signaling. To

determine whether the G1 to S phase transition defect observed in rho0 cells is due to

loss of mitochondrial energy production, we assessed cell cycle progression in cells

treated with oligomycin, an agent that binds to the F1F0 ATPase proton pump and

inhibits ATP production and respiration-driven yeast cell growth on non-fermentable

carbon sources (Fig. 3.4). Flow cytometry revealed that treatment with oligomyocin had

no effect on cell cycle progression. Thus, defects in the G1 to S phase progression in

rho0 cells is not due to loss of mitochondrial ATP production.

FIGURE 3.4- Inhibition of ATP production does not cause the defect in passage through G1 in cells. (A) Effect of oligomycin treatment on growth of wild type rho+ cells on fermentable and non-fermentable carbon sources. Cells were propagated in rich media containing glucose, glycerol or glycerol and ethanol in the presence or absence of oligomycin (1 µg/mL) shown for 48 hours, and cell density was assessed by determining the optical density of the culture at 600 nm. (B) Cell cycle progression of wild-type mtDNA-containing cells in the absence or presence of oligomycin. Cell synchronization, propagation in glucose based media and analysis by flow cytometry was carried out as for Fig 3.3. (C) Quantitation of progression into G2 phase was carried out as for Fig. 3.3.

Page 67: Mitochondrial and Mitochondrial DNA Inheritance

60

Next, we assessed the effect of deletion of subunit 5a of cytochrome c oxidase

(COX5A) on cell cycle progression. First, we confirmed that deletion of COX5A results

in defects in respiration driven yeast cell growth (data not shown). Moreover, we found

that cell cycle progression of cox5a∆ rho+ cells is similar to that observed in wild-type

rho+ cells (Fig. 3.5). Finally, we found that deletion of mtDNA in cox5a∆ cells (cox5a∆

rho0) results in defects in G1 to S phase transition that is similar that observed in rho0

cells (Fig. 3.5). These results confirm the finding that loss of mtDNA results in G1 to S

phase transition defects, and that the observed defects are not due to loss of

mitochondrial respiratory activity.

Figure 3.5- Loss of mitochondrial respiratory activity does not cause the defect in passage through G1. (A) Maximum projections of deconvolved images of cox5a∆ rho+ and cox5a∆ rho0 cells stained with the DNA binding dye DAPI. n: nuclear DNA, m: mtDNA. Cell outlines are shown in white Bar, 1 µm. B) Cell cycle progression of cox5A∆ cells with or without mtDNA (cox5a∆ rho+, cox5a∆ rho0 respectively) were determined as for Fig. 3.1. Images of DAPI stained cells are 2D projections of the reconstructed 3D volume. (C) Quantitation of progression through G2 phase was carried out as for Fig. 3.3.

Page 68: Mitochondrial and Mitochondrial DNA Inheritance

61

THE DEFECT IN CELL CYCLE PROGRESSION OBSERVED IN rho0 CELLS IS DUE

TO LOSS OF DNA IN MITOCHONDRIA AND NOT GENES ENCODED BY THAT DNA

To determine whether the observed cell cycle delay is due to loss of DNA and not

genes encoded by mtDNA, we studied cell cycle progression in a cell that has mtDNA

but does not have any genes encoded by that mtDNA. The mtDNA of the cell used

contains a tandem repeats of a 71 kb fragment of the cytochrome b gene (N24 rho-)

(Tzagoloff et al., 1979). This rho- cell contains mtDNA nucleoids (Fig. 3.6A); however,

since the mtDNA has no genes, the cell is respiratory incompetent and cannot grow

using a non-fermentable carbon source (Data not shown).

Figure 3.6 Loss of mtDNA function does not activate a cellular delay. Rho+, rho- and rho0 cells were grown to mid-log phase in glucose-based media (OD600 = 0.5–0.8). Unbudded cells were isolated by centrifugation through a 10%–35% sorbitol gradient for 12 min at 56 x g. Unbudded cells were propagated in fresh SC media, and cell cycle progression was analyzed as for Fig. 3.3. (A) Images of DAPI stained cells are 2D projections of the reconstructed 3D volume. (B) Cells with wild-type mtDNA (rho+) or mutant DNA with no genes (N24 rho-) have normal cell cycle progression. Cells lacking mtDNA (rho0) exhibit defects in G1 to S progression. (C) Quantitation of cell cycle progression was carried out as for Fig. 3.3.

Page 69: Mitochondrial and Mitochondrial DNA Inheritance

62

We assessed cell cycle progression in the N24 rho- cell and in rho+ and rho0 cells in the

same genetic background as the N24 rho- cell (D273-10b) (Fig. 3.6). Loss of mtDNA in

the D273-10b genetic background also results in a defect in G1 to S phase transition.

Thus, the observed cell cycle progression defect is not a consequence of genetic

background used. Equally important, we found that the N24 rho- cell has a similar cell

cycle progression profile to that seen in the wild-type rho+ cell. Thus, the defect in G1 to

S phase transition observed in rho0 cells is due to loss of DNA within mitochondria and

not due to loss of genes encoded by mtDNA.

ROLE FOR A KNOWN CHECKPOINT PROTEIN (Rad53) IN REGULATION OF CELL

CYCLE PROGRESSION IN CELLS LACKING mtDNA

Rad53 is a conserved cell cycle regulatory protein that can arrest the cell cycle at the

G1 to S phase transition, and has functional interactions with mtDNA in mammalian

cells and yeast. This raises the possibility that the cell cycle defects observed upon loss

of mtDNA may be regulated by Rad53. To test this hypothesis, we tested whether

deletion of RAD53 will suppress the cell cycle progression defects observed in rho0 cells

Page 70: Mitochondrial and Mitochondrial DNA Inheritance

63

Figure 3.7 Deletion of SML1, an inhibitor of dNTP synthesis, does not suppress the cell cycle progression defect observed in upon loss of mtDNA Cell cycle progression of wild-type (W303), sm1∆ rho+ or EtBr derived sml1∆ rho0 was carried out as described in Fig. 3.3. (A) Images of DAPI stained cells. n: nuclear DNA, m: mtDNA. Bar: 1 µm. (B-C) Cell cycle progression was assessed as for Fig. 3.3.

Since RAD53 is an essential gene and deletion of SML1 suppresses the lethality of

rad53∆, we assessed the effect of deletion of mtDNA in rad53∆ sml1∆ and sml1∆ cells.

Deletion of SML1 does not suppress the defect in G1 to S transition observed upon loss

of mtDNA (Fig. 3.7). In contrast, cell cycle progression of rad53∆/sml∆ rho0 cells is

similar to that observed in rad53∆/sml∆ rho+ cells (Fig. 3.8). The finding that deletion of

Rad53 allows for cell cycle progression in a rho0 cell provides additional support for the

finding that rho0 exhibit a defect in progression from G1 to S phase. It also supports a

role for Rad53 in regulation of cell cycle progression in response to loss of mtDNA.

Page 71: Mitochondrial and Mitochondrial DNA Inheritance

64

Figure 3.8 Role for Rad53 in regulating cell cycle progression in response to loss of mtDNA (A-C) Deletion of RAD53 suppresses the cell cycle defect observed in rho0 cells. Cell cycle progression of rho+ and rho0 rad53∆/sml1∆ cells was assessed as for Fig. 3.3. (A) Images of DAPI stained cells. n: nuclear DNA, m: mtDNA. Bar: 1 µm. (B-C) Cell cycle progression was assessed as for Fig. 3.3. (D) Pif1, a Rad53 target is elevated in rho0 cells compared to rho+ cells. Rho+ and rho0 cells containing Pif1p that is tagged at its chromosomal locus with 13 copies of the myc epitope were grown to mid-log phase in glucose based rich media. Cells were concentrated by centrifugation, and extracted with SDS sample buffer. Proteins in the whole cell extracts were analyzed using Western blots and antibodies raised against the myc epitope and hexokinase. The levels of hexokinase revealed that equal amounts of protein were analyzed in rho+ and rho0 cells. Using the myc antibody, it is clear that myc-tagged Pif1p is present at higher levels in rho0 compared to rho+ cells.

If Rad53 is regulating cell cycle progression at the G1 to S phase transition in response

to loss of mtDNA then loss of mtDNA should result in activation of Rad53p. To test this,

we compared the steady state level of Pif1p, a DNA helicase that is required for mtDNA

maintenance and has genetic interactions with Rad53, in rho+ and rho0 cells. We found

Page 72: Mitochondrial and Mitochondrial DNA Inheritance

65

that the steady state levels of Pif1p are elevated by 45% in rho0 cells compared to rho+

(Fig. 3.8D). This result indicates that deletion of mtDNA results in activation of Rad53p.

DISCUSSION

Surveillance of nuclear DNA to ensure genomic integrity before the replication and

segregation to a new daughter cell has been studied for several decades. The

involvement of the ATR/Mec1 and downstream kinases such as Rad53 has been well

accepted throughout the field. Here, we report that the existence of a surveillance

system that inhibits cell cycle progression from G1 to S phase when there is a loss of

mtDNA.

In Drosophila, defects in mitochondrial respiratory activity triggers a checkpoint that

inhibits G1 to S progression (Mandal et al., 2005; Owusu-Ansah et al., 2008). However,

we find that treatment with an agent that inhibits mitochondrial ATP production or

deletion of a mitochondrial respiratory chain component, which results in loss of

respiratory activity, does not affect transition from G1 to S phase in yeast. Instead, we

find that the observed cell cycle defect is triggered by loss of DNA within mitochondria

and not by loss of the genes encoded by mtDNA. Specifically, we find that rho- yeast,

which contain mtDNA that is similar in size to wild type mtDNA but does not bear any

genes, exhibits wild type progression from G1 to S phase.

Furthermore, we have implicated a known checkpoint signaling pathway, the Rad53

DNA Damage checkpoint, in control of progression through G1 in response to loss of

Page 73: Mitochondrial and Mitochondrial DNA Inheritance

66

mtDNA. Previous studies indicate that deletion of RAD53 results in an increase in

mtDNA copy number in yeast and mammalian cells, and that Pif1p, a DNA helicase that

localizes to nuclei and mitochondria, has genetic interactions with Rad53 and is a target

for Rad53. Thus, there is an established functional link between the Rad53 pathway and

mtDNA. We find that Pif1p is up-regulated in response to loss of mtDNA. Consistent

with this, we find that deletion of RAD53 by-passes the G1 to S transition defect in cells

with no mtDNA.

Although the essential function of Rad53 is in regulation of dNTP pools, our studies

indicate the regulation of dNTP pool size alone is not responsible for the defects in G1

to S progression in cells without mtDNA. Specifically, we find that deletion of SML1, a

negative regulator of the enzyme that catalyzes the rate-limiting step in dNTP synthesis

(Rnr1p), does not suppress the G1 to S transition defect observed in cells without

mtDNA. Thus, our evidence supports the model that Rad53 functions in regulating cell

cycle progression in response to loss of mtDNA as a checkpoint signaling molecule and

not as a regulator of dNTP levels.

Our previous studies revealed a checkpoint that inhibits cytokinesis in response to

defects in inheritance of mitochondria and is regulated by the Mitotic Exit Network

(Chapter 2). This checkpoint is triggered by loss of mitochondria and not by loss of

mtDNA. Here, we report the identification of a second checkpoint that monitors

mitochondrial inheritance. In this case, the trigger for cell cycle delay is loss of DNA in

mitochondria. Accordingly, we refer to this checkpoint as the mtDNA checkpoint. Our

Page 74: Mitochondrial and Mitochondrial DNA Inheritance

67

data also support a role for Rad53 in regulation of the mtDNA inheritance checkpoint.

Thus, we find that Rad53 checkpoint function is broader than previously appreciated: it

monitors DNA defects in both the nucleus and mitochondria. Finally, since Rad53 is

conserved, and affect mtDNA in yeast and mammalian cells, it is possible that a mtDNA

checkpoint exists in cells other than yeast.

Supplemental Fig 3.1 (SFig 3.1) Growth of all strains used in this study on fermentable

and non-fermentable carbon sources. A titration series of strain on glucose-based

media (YPD) and glycerol-based media (YPG). Rho0 and cells bearing a deletion in the

COX5 gene have defects in mitochondrial respiratory activity and can grow on a

fermentable carbon source (glucose) but not on a non-fermentable carbon source

(glycerol).

Page 75: Mitochondrial and Mitochondrial DNA Inheritance

68

Defects in respiration can be readily detected by assaying for growth on a non-

fermentable carbon source. All viable yeast strain exhibit growth on media containing

fermentable carbon sources. However, growth on non-fermentable carbon sources

requires mitochondrial respiratory activity. As a result, rho0 cells grow on glucose but not

on glycerol based media. All rho0 had the predicted phenotype: Lack of DAPI nucleoid

staining and glycerol growth (supplemental Fig 3.1).

EXPERIMENTAL PROCEDURES

Yeast strains, plasmids, and growth conditions:

Yeast strains used in this work are listed in Table 3.2. Strain ISY065 is a derivative of

W303. Strains MYY291 and DNY416 were derived from A364A. rho0 derivatives were

generated from wild-type by two consecutive two day treatments of 25 µM EtBr

(Goldring et al., 1970). Standard molecular techniques for cloning procedures were

used sambrook 1998 Coldspring harbor(28). Other yeast methods were performed

according to Sherman (Sherman, 2002).

Tagging of Pif1: The carboxy terminus of PIF1 was tagged with 13-myc using PCR-

based insertion into the chromosomal copies of the pif1 loci (27). Primers used to tag

these genes were: forward primer

CGAACCTCGTGGTCAGGATACCGAAGACCACATCTTAGAACGGATCCCCGGGTTA

ATTAA and reverse primer

GCAGTTTGTATTCTATATAACTATGTGTATTAATATGTACGAATTCGAGCTCGTTTAAAC.

Page 76: Mitochondrial and Mitochondrial DNA Inheritance

69

Standard molecular techniques for cloning procedures were used. Growth conditions for

individual experiments are described in the figure legends.

Synchronization: For cell cycle synchronization, cells were incubated with α-factor (10-

100 µM) for 2.5 hrs. Cells were released from arrest by washing and were transferred to

pheromone-free media.

Flow cytometry: Analysis of DNA content in propidium iodide-stained, synchronized

cell cultures was determined according to Paulovich and Hartwell (Paulovich and

Hartwell, 1995) using a fluorescence-activated cell analyzer (Becton Dickerson LSRII,

Franklin Lakes, NJ). The percent of total cells in G1 phase was determined using the

Flowjo program (TreeStar Inc., Ashland, OR).

Fluorescence microscopy, image analysis and cytology: Cells were gently pelleted,

resuspended in fresh media and mounted directly onto a pad consisting of 2% low

melting agarose and SC media. The sample was covered with a coverslip and

fluorescence/phase microscopic images were collected using an E600 microscope

(Plan-Apo 100X/1.4 NA objective) (Nikon, Melville, NY) equipped with a cooled CCD

camera (Orca-ER, Hamamatsu, Japan), and a Dual-View image splitter (Optical

Insights, Tucson, AZ) for simultaneous two-color imaging. Openlab 3.1.5 software

(Improvision, Lexington, MA) was used to acquire images. Z-stacks of 0.2-µm slices

were obtained and the out-of-focus light was removed using an iterative deconvolution

Page 77: Mitochondrial and Mitochondrial DNA Inheritance

70

algorithm in Volocity 5.5 (Improvision, Lexington, MA). All z-sections were assembled

and 3-D projections were generated with comparable parameters and thresholds.

Protein and immunological techniques: Protein extracts of mid-log phase yeast cells

for Western blot analysis were obtained as described (Boldogh et al., 1998). The

bicinchoninic acid (BCA) assay (Pierce Chemical, Rockford, IL) was used for protein

concentration determinations. Immunoblot analysis of the total amount of (Pif1) was

performed with antibodies specific for myc. HRP-conjugated secondary antibodies and

Supersignal detection (Pierce Chemical, Rockford, IL) were used to visualize bands.

Page 78: Mitochondrial and Mitochondrial DNA Inheritance

71

Table 3.1.

Strains Genotype Source

BY4741 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 Open Biosystems

DCY023 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 rho0 This study

6937 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 mgm101∆ Open Biosystems

7210 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 cox5a∆ Open Biosystems

DCY033 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 cox5a∆ rho0 This study

W1588-4C MATa leu2-3,112 trp1-1 can1-100 ura3-1 ade2-1

his3-11,15

R. Rothstein

DCY025 MATa leu2-3,112 trp1-1 can1-100 ura3-1 ade2-1

his3-11,15 rho0

This study

U952-3B MATa leu2-3,112 trp1-1 can1-100 ura3-1 ade2-1

kan-his3-11,15 sml1∆::HIS3

R. Rothstein

DCY027 MATa leu2-3,112 trp1-1 can1-100 ura3-1 ade2-1

kan-his3-11,15 sml1∆ rho0

This study

U960-5C MATa leu2-3,112 trp1-1 can1-100 ura3-1 ade2-1

his3-11,15 sml1-1 rad53∆::HIS3

R. Rothstein

DCY029 MATa leu2-3,112 trp1-1 can1-100 ura3-1 ade2-1

his3-11,15 sml1-1 rad53∆::HIS3 rho0

This study

D273-10B MATα mal G. Schatz

DCY015 MATα mal rho0 This study

N24 MATα mal N24 rho- Nobrega and Tzagoloff,

1980

PDY001 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 PIF1-13-

myc

This study

PDY002 MATa his3∆1 leu2∆0 met15∆0 ura3∆0 PIF1-13-

myc rho0

This study

Page 79: Mitochondrial and Mitochondrial DNA Inheritance

72

CHAPTER 4

DISCUSSION

Page 80: Mitochondrial and Mitochondrial DNA Inheritance

73

DISCUSSION

There are many surveillance mechanisms that control the progression of the mitotic cell

cycle in response to nuclear inheritance. Recent findings reveal that these surveillance

mechanisms are not limited to monitoring the inheritance of nuclei, but that they can

also sense and respond to the distribution and inheritance of other organelles including

mitochondria, cER and Golgi. The finding that checkpoint surveillance occurs for

organelles that can and cannot be produced de novo underscores the importance of

organelle inheritance to cell division and provides additional evidence that organelle

inheritance is an integral component of the cell division cycle. It also raises the

possibility that there are organelle inheritance checkpoints for other organelles in the

cell including lysosomes and peroxisomes. Indeed, emerging evidence supports the

existence of checkpoints for inheritance of the lysosome (vacuole) in yeast (Barelle et

al., 2003; Veses et al., 2009).

Recent findings also indicate that the checkpoints for inheritance of mitochondria, cER

and Golgi are regulated by conserved cell cycle regulatory mechanisms. This reveals

novel functions for known pathways and suggests that mechanisms underlying

organelle inheritance checkpoints are also conserved. Yet to be determined is the

mechanism(s) for communication between checkpoint machineries and their target

organelle. What is the mechanism for detecting defects in inheritance of mitochondria

and other organelles? How is information from inheritance checkpoint sensors

Page 81: Mitochondrial and Mitochondrial DNA Inheritance

74

transmitted to the machinery that mediates cell cycle arrest? Do organelle checkpoint

pathways activate proteins that promote or repair organelle inheritance when there are

defects in these processes? If so, how? Future studies will provide answers to each of

these fundamental questions on the checkpoint machineries of mitochondria and other

organelles.

SUMMARY OF PROJECT 1:

My first project revealed that the inheritance of mitochondria in the bud is required for

cell cycle progression at cytokinesis. We published that mutations that result in severe

defects in mitochondrial inheritance also results in defects in cell cycle progression

through cytokinesis. The mutants used in these studies are yeast bearing a deletion of

MDM10, which results in defects in mitochondrial movement during inheritance, or

deletion of MMR1, a protein that mediates anchorage of mitochondria in the bud tip,

combined with deletion of GEM1, a calcium binding GTPase that is required for normal

mitochondrial distribution. Deletion of MDM10 inhibits contractile ring closure, but does

not inhibit contractile ring assembly, localization of a chromosomal passenger protein to

the spindle during early anaphase, spindle alignment, nucleolar segregation or nuclear

migration during anaphase. Release of the mitotic exit network (MEN) component,

Cdc14p, from the nucleolus during anaphase is delayed in mdm10∆ cells.

We also showed that this mitochondrial inheritance checkpoint is regulated by the

mitotic exit network (MEN). I found that the mitochondrial inheritance checkpoint could

be bypassed in both the mdm10∆ and mmr1∆ gem1∆ cells by hyper-activating the MEN

Page 82: Mitochondrial and Mitochondrial DNA Inheritance

75

network by either knocking out the negative inhibitor BUB2 or by overexpressing the

polo kinase Cdc5p. Finally, since checkpoints are designed to increase the fidelity of

processes that are critical for the cell division cycle, we tested whether inhibition of the

MEN compromised mitochondrial inheritance. Here, we found that deletion of BUB2

results in a decrease in the fidelity of mitochondrial segregation between mother and

daughter cells in wild-type and mdm10∆ cells. Overall, our studies identify a novel MEN-

linked regulatory system that inhibits cytokinesis in response to defects in mitochondrial

inheritance in budding yeast.

FUTURE EXPERIMENTS:

An important goal for future studies is to identify the sensor that detects defects in

mitochondria inheritance, and how this sensor communicates that information to the

MEN. Preliminary data obtained from my studies and studies from other members of our

laboratory serve as a foundation for these studies.

I found that the passage of only a small fragment of mitochondria across the bud neck

can initiate contractile ring closure and commit to cytokinesis. Preliminary experiments

support a role for Dbf2 and Mob1 in this process. Dbf2 and Mob1 are part of a ser/thr

kinase complex that is involved in transcription and stress responses. The complex

functions as part of the MEN pathway and is activated by the MEN kinase Cdc15p.

Dbf2p has kinase activity and Mob1p binds to Dbf2, which results in conformational

changes in Dbf2p that allows an upstream kinase, Cdc15p, to phosphorylate Mob1p

Page 83: Mitochondrial and Mitochondrial DNA Inheritance

76

and Dbf2p. Phosphorylation of Dbf2p activates its kinase activity, which is required for

cytokinesis and cell separation.

During early stages in the cell division cycle in wild-type cells, Mob1p and Dbf2p localize

to the cytosol. During anaphase, Mob1p and Dbf2p relocalize to the spindle pole body

and mother bud neck, where they activate proteins that contribute to mitotic exit and

cytokinesis. Ongoing studies indicate that Mob1p and Dbf2p are mislocalized in two

strains (mdm10∆ and mmr1∆ gem1∆) with severe defects in mitochondrial inheritance

and cytokinesis. Both proteins are detected in the spindle pole body and at the bud

neck in mitochondrial inheritance mutants. However, Mob1p and Dbf2p also localize to

mitochondria in mdm10∆ and mmr1∆ gem1∆ mutants. This raises the possibility that

Mob1p and Dbf2p are part of the sensor for detecting mitochondria inheritance during

the yeast cell division cycle and that they do so through direct interactions with

mitochondria. Moreover, since Mob1p and Dbf2 are conserved in all cells studied, it is

possible that they may serve as sensors for mitochondrial inheritance in other cell types.

These findings raise additional questions: what proteins do Mob1p and Dbf2 interact

with on mitochondria? How is binding of Mob1p and Dbf2p to mitochondria regulated?

Does mislocalization of Dbf2p and Mob1p to mitochondria in mitochondrial inheritance

mutants prevent the active complex from localizing to its site of action at the mother-bud

neck?

Page 84: Mitochondrial and Mitochondrial DNA Inheritance

77

SUMMARY OF PROJECT 2:

My second project sheds light on the surveillance mechanism for mtDNA and how the

presence of mtDNA is required for normal G1 to S transition during yeast cell division.

Many other laboratories have previously published the repair mechanism needed when

there is damage, defects in replication, or defects in the segregation of nuclear DNA;

however, we are the first to show that there is a surveillance mechanism for defects in

mtDNA. For DNA damage checkpoints, the damage detected is not associated with a

specific gene or set of genes. Rather, these checkpoints monitor broader processes,

including stalled DNA replication fork, single or double stranded DNA breaks. We have

observed similar findings as in the mtDNA checkpoint. The mtDNA checkpoint observed

in yeast also does not monitor specific genes or function of genes encoded by mtDNA.

Instead, it monitors mtDNA content.

My results also indicate that the loss of mtDNA activates a DNA damage signaling

pathway, Rad53, which results in a defect in progression from G1 to S phase.

Specifically, I found that the loss of mtDNA results in an increase in the steady state

levels of Pif1p, a protein that is activated by the DNA damage checkpoint and thought to

be a substrate of Rad53. This finding indicates that Rad53 is activated by loss of

mtDNA. Moreover, since Pif1p is a DNA helicase that localizes both to the nucleus and

the mitochondrion, this finding indicates that activated Rad53p also promotes mtDNA

repair in response to loss of mtDNA. Consistent with this, I found that deletion of Rad53

Page 85: Mitochondrial and Mitochondrial DNA Inheritance

78

by-passes the cell cycle progression defect observed in cells that do not contain

mtDNA.

Although regulation of dNTP pools is an essential function of Rad53, my studies

suggest that the arrest found in a cell lacking mtDNA (rho0) is not due to the decrease in

deoxynucleotides (dNTP). Specifically, I found that deletion of the ribonucleotide

inhibitor (RNR) SML1, which increases dNTP pools, did not restore normal cell cycle

progression in a cell that has no mtDNA. Instead, my data supports the model that

Rad53 functions in the mtDNA inheritance checkpoint as a regulator of G1 to S

transition, and not as a regulator of dNTP levels.

Our findings are supported by reports in the literature in which cell cycle arrest was

observed in cells that have lost their mtDNA. One of these studies focused on DNA2, an

essential gene that encodes a endonuclease, and 5' to 3' DNA helicase that contributes

to Okazaki Fragment processing and repair of double strand DNA breaks. Mutation of

DNA2 results in a cell cycle arrest that is mediated by the Mec1/Rad53 pathway. (Budd

et al., 2006). Interestingly, mutations of DNA2 results in a high rate of mtDNA loss, and

cells that have lost their mtDNA undergo cell cycle arrest (Fiorentino and Crabtree,

1997). Thus, other laboratories have observed a link between mtDNA loss, cell cycle

progression defects and the Rad53 DNA damage pathway.

Page 86: Mitochondrial and Mitochondrial DNA Inheritance

79

ROLE FOR DNA Pol γγγγ AS A SENSOR FOR THE mtDNA INHERITANCE

CHECKPOINT:

The only DNA polymerase found solely in animal cell mitochondria is DNA polymerase

gamma (POLG). This polymerase bears the sole responsibility for DNA synthesis in all

replication, recombination, and repair transactions involving mitochondrial DNA

(mtDNA). Human polymerase γ is composed of a 140 kDa catalytic subunit A (POLγ A)

and a 55 kDa accessory subunit B, which increases the processivity of mtDNA

synthesis. POLγ A comprises a polymerase domain and an exonuclease proofreading

domain, separated by a linker region of 482 amino acids (Figure 4.1).

Over 60 PEO-associated mutations have been found in POLγγγγ (Lee et al., 2009) (Viikov

et al., 2011). Most of the dominant POLγγγγ mutations are in the polymerase domain,

whereas most of the recessive mutations are in the exonuclease or in the spacer

domains. Mutations in Pol γγγγ are associated with a spectrum of disease phenotypes

including autosomal dominant and recessive forms of progressive external

ophthalmoplegia, (PEO), spino-cerebellar ataxia and epilepsy, and Alpers-Huttenlocher

hepatocerebral poliodystrophy (Lee et al., 2009). Multiple deletions, or depletion of

mtDNA in affected tissues, are the molecular hallmarks of pol γγγγ mutations.

Page 87: Mitochondrial and Mitochondrial DNA Inheritance

80

Figure 4.1 Taken from (Lee et al., 2009) to illustrate the various domains of human

polymerase gamma.

The yeast orthologue of POLγγγγ is MIP1, which encodes for the catalytic subunit of DNA

polymerase γγγγ. Proteomic and genetic screens have revealed that Mip1, like Abf2 and

Mgm101, is associated with mtDNA nucleoids (Meeusen and Nunnari, 2003). Mip1

directly binds to mtDNA, synthesizes DNA stretches of up to several thousand

nucleotides without dissociation from the template, carries out DNA synthesis on

Page 88: Mitochondrial and Mitochondrial DNA Inheritance

81

double-stranded templates utilizing a strand displacement mechanism, and is active

even in cells that have no mtDNA (Viikov et al., 2011).

DNA repair complexes that associate with DNA at site of damage influence the

checkpoint response by facilitating the recruitment of checkpoint factors or by

generating the intermediate DNA structures that function as signals to the checkpoint

machinery. Although there are limited repair pathways for mtDNA, the only protein that

functions in the two known repair mechanisms, single nucleotide base excision repair

and long patch base excision repair, is Mip1p. Moreover, unlike other mtDNA binding

proteins like Pif1p and Abf2p that are present in reduced levels in mitochondria with no

mtDNA, Mip1 localizes to mitochondria and is fully functional in rho0 cells (Viikov et al.,

2011). This raises the possibility that Mip1p may serve as a sensor for the mtDNA

inheritance checkpoint.

If Mip1p is a mtDNA inheritance checkpoint sensor, then deletion of MIP1 should by-

pass the G1 to S transition defect observed in rho0 cells. Indeed, I found that deletion of

MIP1 bypasses the cell cycle delay associated with loss of mtDNA in rho0 cells that are

made through the deletion of MGM101 (Figure 4.2A). Thus, preliminary data indicate

that MIP1 is part of the sensor that detects mtDNA and activates Rad53 in response to

loss of mtDNA. Interestingly, Mip1p may function as a sensor for defects in mtDNA

replication because it is located at the replication fork. However, it would be impossible

Page 89: Mitochondrial and Mitochondrial DNA Inheritance

82

to distinguish between a sensory role versus a signal transduction role and it is possible

that it is the activity of an entire complex and not a single polymerase that must be intact

to properly sense replication (Elledge, 1996).

Figure 4.2 Role for DNA polymerase gamma, Mip1, in the mtDNA inheritance checkpoint- Cell progression is not affected by knocking out mgm101 in mip1∆ cells. (A) Cells cycle progression of mip1∆ and mip1∆ mgm101∆ rho0 cells. (B) Quantitation of cell cycle progression in mip1∆ and mip1∆ mgm101∆ rho0 with controls (WT Rho+ and mgm101∆ cells) was determined as described in Fig 3.2.

To further elucidate the possibility that Mip1 is a mtDNA inheritance sensor, I would take

advantage of previous work by Baruffini, E. Lodi, T. et al. and others who studied MIP1

mutations associated with mitochondrial diseases in humans (Baruffini et al., 2006).

Specifically, I would be interested in studying the mutation within the exonuclease

domain (G303R) that has a severe phenotype and complete depletion of mtDNA and

Page 90: Mitochondrial and Mitochondrial DNA Inheritance

83

compare it to DNA binding mutations (R574W) and (P625R) that results in in less

severe phenotypes (Baruffini et al., 2006; Lee et al., 2009). I would predict that the cell

cycle regulation in response to loss of mtDNA may be dependent upon Mip1 DNA

binding but not necessarily dependent upon DNA polymerase activity.

This approach may also be used to identify other proteins that regulate the mtDNA

inheritance checkpoint. Here, I would screen for MIP1 mutations that inhibit its function

in the mtDNA inheritance checkpoint, but do not affect its function in mtDNA replication.

Thereafter, I would perform a 2-hybrid screen to identify proteins that interact with WT

MIP1 but do not interact with MIP that is defective in its mtDNA sensor function.

Finally, I would test if these proteins interact with MIP1 in vivo, and whether deletion of

candidate Mip1-interacting mtDNA sensor proteins suppress the cell cycle progression

defects observed in rho0 cells.

HOW IS THE SIGNAL THAT mtDNA LOSS TRANSMITTED FROM

MITOCHONDRIA TO THE NUCLEUS?

The mitochondrial retrograde (RTG) pathway is a pathway for communication between

the mitochondria and the nucleus. This pathway is active in both normal and

pathological conditions and involves multiple factors that sense and transmit signals to

effect changes in nuclear gene expression. These changes lead to a reconfiguration of

metabolism to accommodate cells with mitochondrial defects (Liu and Butow, 2006).

Retrograde signaling has been connected to nutrient sensing, TOR signaling, growth

Page 91: Mitochondrial and Mitochondrial DNA Inheritance

84

control, and other signaling processes that function in metabolic and organelle

homeostasis, as well as aging (Liu and Butow, 2006).

Interestingly, the RTG pathway affects mtDNA maintenance through regulation of the

RTG target gene, ACO1 (Liu and Butow, 2006). ACO1 is one of many bifunctional

genes in mitochondria. Aco1p, aconitase, function as a metabolic enzyme in the TCA

cycle and as a signaling molecule in the RTG pathway. Earlier studies linked Aco1 to

mtDNA inheritance through the identification of Aco1p being among proteins cross-

linked to mtDNA. Equally important, the loss of mtDNA observed upon deletion of the

mtDNA packaging protein Abf2p can be suppressed by activating ACO1 expression and

the RTG pathway (Liu and Butow, 2006). Moreover, Aco1p localizes to mtDNA, and

deletion of ACO1 results in severe mtDNA instability (Liu and Butow, 2006). Since loss

of mtDNA is not observed upon deletion of other TCA cycle genes, mtDNA instability

observed in ACO1 mutants is attribute to ACO1 function in the RTG pathway and not

central metabolism. (McCammon et al., 2003). These studies lead to the conclusion that

aconitase has two distinct function and activities: 1) enzymatic activity and 2) mtDNA

maintenance activity.

In light of this, it would be important to test whether the RTG pathway serves as a

signaling pathway to activate Rad53 in response to loss of mtDNA. To do so, I would

test whether mutation of ACO1 or other RTG signaling molecules by-passes the mtDNA

Page 92: Mitochondrial and Mitochondrial DNA Inheritance

85

inheritance checkpoint. Specifically, I would study G1 to S cell cycle progression,

Rad53 phosphorylation and Pif1p levels in rho0 cells and rho0 cells bearing mutations in

ACO1 and other RTG signaling molecules. Given evidence for a role of the RTG

pathway in the mtDNA inheritance checkpoint, I would begin the search for RTG targets

that play a role in this process by analysis of genes that are up-regulated in response to

activation of RTG that have genetic or physical interactions with Rad53.

POSSIBLE ROLE FOR PROHIBITINS IN THE mtDNA INHERITANCE

CHECKPOINT

Prohibitin was originally identified as a gene that inhibits cell cycle progression when

microinjected into mouse embryonic fibroblasts (MEFs) (Osman et al., 2009a). In yeast

there are two known prohibitin genes, prohibitin-1 (PHB1) and prohibitin-2 (PHB2),

which share more than 50% identical amino acid residues, are conserved and are

present in all eukaryotes sequenced to date (Van Aken et al., 2010). Moreover,

prohibitins localize and function predominantly to the mitochondria and the

mitochondrial inner membrane and are involved in mitochondria fusion through the

regulation of the dynamin-like GTPase OPA1 (Osman et al., 2009b) (Merkwirth and

Langer, 2009).

Interestingly, prohibitins localize to different cellular compartments, such as, the

nucleus, the plasma membrane, and mitochondria. In fact PHB2 has been shown to

Page 93: Mitochondrial and Mitochondrial DNA Inheritance

86

translocate from the mitochondria to the nucleus following the binding of estradiol by

ERα (Osman et al., 2009b). This raises the possibility that prohibitins function in

signaling from mitochondria to the nucleus, in addition to their function in regulation of

mitochondrial fusion. Consistent with this, mouse embryonic fibroblasts (MEFs) lacking

PHB2 were partially rescued by overexpressing the non-cleavable L-OPA1 isoform

(Merkwirth et al., 2008).

Equally important, prohibitins function in the maintenance and stability of mtDNA.

Crosslinking studies revealed that PHB1 and PHB2 are peripheral components of

mtDNA nucleoids (Wang and Bogenhagen, 2006) (Bogenhagen et al., 2008).

Furthermore, Kasashima et al. (2008) observed a link between prohibitins and steady

state of nucleoids. They found that downregulating PHB1 expression in HeLa cells

resulted in a decrease in the levels of TFAM and mtDNA. The absence of prohibitin also

leads to an increased generation of reactive oxygen species, disorganized mtDNA

nucleoids, abnormal cristae morphology, inhibition of complex I of the mitochondrial

electron transport chain and an increased sensitivity towards stimuli-elicited apoptosis

(Osman et al., 2009b) (Artal-Sanz and Tavernarakis, 2009). These findings suggest that

PHB1 maintains the organization and copy number of mtDNA by regulating TFAM

stability (Osman et al., 2009b) (Kasashima et al., 2008). An important goal of future

studies is to determine whether prohibitins serve as signaling molecules in the mtDNA

inheritance checkpoint.

Page 94: Mitochondrial and Mitochondrial DNA Inheritance

87

HOW MY STUDIES MAY CONTRIBUTE TO OUR UNDERSTANDING OF

MITOCHONDRIAL DISEASES

There has been some controversy on the interpretation between human samples and in

vitro cellular studies in regards to mitochondrial diseases. Human patients with

mitochondrial disease have a deterioration of the respiratory chain function and through

random segregation events may increase or decrease the mutated mtDNA. Impairment

in mitochondrial respiratory function not only reduces the supply of energy, which may

prevent energy-dependent apoptosis, but also enhances ROS production that may

induce mutation and oxidative damage to mitochondrial DNA (mtDNA) (Lee et al.,

2005). The accumulation of mtDNA mutations and alteration leading to increase

apoptosis contributes to the onset and progression of various neurodegenerative

diseases (Lee et al., 2005). Recently, it was reported that the expression level of the β

subunit of F1-ATPase required for mitochondrial ATP synthesis is decreased in human

cancers, including liver, gastric, lung, and colorectal cancers (Lee et al., 2005).

These types of results calls into question the proposed quality control mechanisms

making it clear that different cellular model needs to be constructed. A mutator mouse

has been made by Trifunovic et al. that shows a strong phenotype, interestingly not at

embryonic or early development but at 25 weeks (Trifunovic et al., 2004). This

premature aging phenotype and model is caused by the accumulation of point

mutations and the presence of linear deleted mtDNA molecules (Trifunovic et al., 2004)

(Park and Larsson, 2011).

Page 95: Mitochondrial and Mitochondrial DNA Inheritance

88

Decreased mitochondrial oxidative phosphorylation (OXPHOS) is one of the hallmarks

of cancer. Mutations in POLG are known to cause mtDNA depletion and decreased

OXPHOS, resulting in mtDNA depletion syndrome in humans. A study by Singh et al.

revealed that mtDNA was depleted in breast tumors. Consistently, mutant POLG, when

expressed in breast cancer cells, induced a depletion of mtDNA, decreased

mitochondrial activity, decreased mitochondrial membrane potential, and increased

levels of reactive oxygen species. He further proposed that decreased OXPHOS led to

an increased rate of aerobic glycolysis in most cancers. This phenomenon is described

as the Warburg effect (Singh et al., 2009).

Laboratories have measured mtDNA content directly in tumors and report a decrease in

mtDNA content in breast, renal, hepatocellular, gastric and prostate tumors (Singh et

al., 2009). It is also noteworthy that drugs used for treating human immunodeficiency

virus (HIV) inhibit POLG, which in turn induces mtDNA depletion (Kakuda, 2000).

Tamoxifen, a commonly used drug for the treatment of breast cancer, also depletes

mtDNA (Larosche et al., 2007). A recent study also demonstrates that depletion of

mtDNA correlates with tumor progression and prognosis in breast cancer patients (Yu et

al., 2007).

Many diseases have been linked to a mutation within a checkpoint pathway. The most

understood pathway studied in mamamalian cells is the p53 DNA damage pathway.

Page 96: Mitochondrial and Mitochondrial DNA Inheritance

89

The p53 gene is a tumor suppressor that is mutated in many human cancer cells and

encodes a transcription factor that is activated in response to DNA damage and leads to

increased nucleotide pools. Cells that are defective in the p53 gene cannot induce a G1

arrest and shows a reduced ability to induce an apoptotic response. Checkpoints such

as p53 are the targets for chemotherapeutic treatment.

Therefore, it is possible that a mtDNA inheritance checkpoint exists in mammalian cells,

and that defects in this process may contribute to cancer. My project identified fast and

simple readouts to identify proteins and the signaling pathways involved in the

regulation during segregation of both mitochondria and its genome, and to use budding

yeast as a translational tool to study human disease. It is also important to note that

there is no, as of yet, transfection system for mammalian mitochondria, and mouse

models either introduce a naturally occurring mtDNA mutation into mouse embryos or

manipulate nuclear genes that control maintenance and expression of mtDNA. These

limitations and the fact that yeast can propagate without the need to respire or use

oxidative phosphorylation to survive makes it an ideal model system for these human

disease studies. Thus, my research can be used to identify and further understand

signaling pathways and function between the mitochondria, its genome, and how it is

regulated during cell division in yeast and in mammalian cells including those that

contribute to human diseases.

Page 97: Mitochondrial and Mitochondrial DNA Inheritance

90

BIBLIOGRAPHY

Page 98: Mitochondrial and Mitochondrial DNA Inheritance

91

Alcasabas, A.A., A.J. Osborn, J. Bachant, F.H. Hu, P.J.H. Werler, K. Bousset, K. Furuya, J.F.X. Diffley, A.M.

Carr, and S.J. Elledge. 2001. Mrc1 transduces signals of DNA replication stress to activate Rad53. Nat Cell

Biol. 3:958-965.

Altmann, K., M. Frank, D. Neumann, S. Jakobs, and B. Westermann. 2008. The class V myosin motor

protein, Myo2, plays a major role in mitochondrial motility in Saccharomyces cerevisiae. J Cell

Biol. 181:119-130.

Artal-Sanz, M., and N. Tavernarakis. 2009. Prohibitin and mitochondrial biology. Trends Endocrinol

Metab. 20:394-401.

Barelle, C.J., E.A. Bohula, S.J. Kron, D. Wessels, D.R. Soll, A. Schafer, A.J. Brown, and N.A. Gow. 2003.

Asynchronous cell cycle and asymmetric vacuolar inheritance in true hyphae of Candida

albicans. Eukaryot Cell. 2:398-410.

Baruffini, E., T. Lodi, C. Dallabona, A. Puglisi, M. Zeviani, and I. Ferrero. 2006. Genetic and chemical

rescue of the Saccharomyces cerevisiae phenotype induced by mitochondrial DNA polymerase

mutations associated with progressive external ophthalmoplegia in humans. Hum Mol Genet.

15:2846-2855.

Bembenek, J., J. Kang, C. Kurischko, B. Li, J.R. Raab, K.D. Belanger, F.C. Luca, and H. Yu. 2005. Crm1-

mediated nuclear export of Cdc14 is required for the completion of cytokinesis in budding yeast.

Cell Cycle. 4:961-971.

Blondel, M., S. Bach, S. Bamps, J. Dobbelaere, P. Wiget, C. Longaretti, Y. Barral, L. Meijer, and M. Peter.

2005. Degradation of Hof1 by SCF(Grr1) is important for actomyosin contraction during

cytokinesis in yeast. EMBO J. 24:1440-1452.

Bogenhagen, D.F., D. Rousseau, and S. Burke. 2008. The layered structure of human mitochondrial DNA

nucleoids. J Biol Chem. 283:3665-3675.

Boldogh, I., N. Vojtov, S. Karmon, and L.A. Pon. 1998. Interaction between mitochondria and the actin

cytoskeleton in budding yeast requires two integral mitochondrial outer membrane proteins,

Mmm1p and Mdm10p. J Cell Biol. 141:1371-1381.

Boldogh, I.R., D.W. Nowakowski, H.C. Yang, H. Chung, S. Karmon, P. Royes, and L.A. Pon. 2003. A protein

complex containing Mdm10p, Mdm12p, and Mmm1p links mitochondrial membranes and DNA

to the cytoskeleton-based segregation machinery. Mol Biol Cell. 14:4618-4627.

Boldogh, I.R., and L.A. Pon. 2007. Mitochondria on the move. Trends Cell Biol. 17:502-510.

Boldogh, I.R., H.C. Yang, W.D. Nowakowski, S.L. Karmon, L.G. Hays, J.R. Yates, 3rd, and L.A. Pon. 2001.

Arp2/3 complex and actin dynamics are required for actin-based mitochondrial motility in yeast.

Proc Natl Acad Sci U S A. 98:3162-3167.

Broustas, C.G., and H.B. Lieberman. 2011. Contributions of RAD9 to tumorigenesis. J Cell Biochem.

Budd, M.E., C.C. Reis, S. Smith, K. Myung, and J.L. Campbell. 2006. Evidence suggesting that Pif1 helicase

functions in DNA replication with the Dna2 helicase/nuclease and DNA polymerase delta. Mol

Cell Biol. 26:2490-2500.

Campellone, K.G., and M.D. Welch. 2010. A nucleator arms race: cellular control of actin assembly. Nat

Rev Mol Cell Biol. 11:237-251.

Chabes, A., B. Georgieva, V. Domkin, X. Zhao, R. Rothstein, and L. Thelander. 2003. Survival of DNA

damage in yeast directly depends on increased dNTP levels allowed by relaxed feedback

inhibition of ribonucleotide reductase. Cell. 112:391-401.

Chang, M., B. Luke, C. Kraft, Z. Li, M. Peter, J. Lingner, and R. Rothstein. 2009. Telomerase is essential to

alleviate pif1-induced replication stress at telomeres. Genetics. 183:779-791.

Page 99: Mitochondrial and Mitochondrial DNA Inheritance

92

Chen, X.J., and R.A. Butow. 2005. The organization and inheritance of the mitochondrial genome. Nat

Rev Genet. 6:815-825.

Chen, X.J., M.X. Guan, and G.D. Clark-Walker. 1993. MGM101, a nuclear gene involved in maintenance

of the mitochondrial genome in Saccharomyces cerevisiae. Nucleic Acids Res. 21:3473-3477.

Cheng, X., S. Dunaway, and A.S. Ivessa. 2007. The role of Pif1p, a DNA helicase in Saccharomyces

cerevisiae, in maintaining mitochondrial DNA. Mitochondrion. 7:211-222.

Clifford, D.M., B.A. Wolfe, R.H. Roberts-Galbraith, W.H. McDonald, J.R. Yates, 3rd, and K.L. Gould. 2008.

The Clp1/Cdc14 phosphatase contributes to the robustness of cytokinesis by association with

anillin-related Mid1. J Cell Biol. 181:79-88.

Corbett, M., Y. Xiong, J.R. Boyne, D.J. Wright, E. Munro, and C. Price. 2006. IQGAP and mitotic exit

network (MEN) proteins are required for cytokinesis and re-polarization of the actin

cytoskeleton in the budding yeast, Saccharomyces cerevisiae. Eur J Cell Biol. 85:1201-1215.

Cordon-Preciado, V., S. Ufano, and A. Bueno. 2006. Limiting amounts of budding yeast Rad53 S-phase

checkpoint activity results in increased resistance to DNA alkylation damage. Nucleic Acids Res.

34:5852-5862.

D'Amours, D., and A. Amon. 2004. At the interface between signaling and executing anaphase--Cdc14

and the FEAR network. Genes Dev. 18:2581-2595.

Dai, Y., and S. Grant. 2011. Methods to study cancer therapeutic drugs that target cell cycle checkpoints.

Methods Mol Biol. 782:257-304.

Dimmer, K.S., S. Jakobs, F. Vogel, K. Altmann, and B. Westermann. 2005. Mdm31 and Mdm32 are inner

membrane proteins required for maintenance of mitochondrial shape and stability of

mitochondrial DNA nucleoids in yeast. J Cell Biol. 168:103-115.

Eaton, J.S., Z.P. Lin, A.C. Sartorelli, N.D. Bonawitz, and G.S. Shadel. 2007. Ataxia-telangiectasia mutated

kinase regulates ribonucleotide reductase and mitochondrial homeostasis. J Clin Invest.

117:2723-2734.

Elledge, S.J. 1996. Cell cycle checkpoints: preventing an identity crisis. Science. 274:1664-1672.

Fagarasanu, A., M. Fagarasanu, and R.A. Rachubinski. 2007. Maintaining peroxisome populations: a story

of division and inheritance. Annu Rev Cell Dev Biol. 23:321-344.

Fagarasanu, A., and R.A. Rachubinski. 2007. Orchestrating organelle inheritance in Saccharomyces

cerevisiae. Curr Opin Microbiol. 10:528-538.

Fehrenbacher, K.L., I.R. Boldogh, and L.A. Pon. 2005. A role for Jsn1p in recruiting the Arp2/3 complex to

mitochondria in budding yeast. Mol Biol Cell. 16:5094-5102.

Fehrenbacher, K.L., H.C. Yang, A.C. Gay, T.M. Huckaba, and L.A. Pon. 2004. Live cell imaging of

mitochondrial movement along actin cables in budding yeast. Curr Biol. 14:1996-2004.

Fiorentino, D.F., and G.R. Crabtree. 1997. Characterization of Saccharomyces cerevisiae dna2 mutants

suggests a role for the helicase late in S phase. Mol Biol Cell. 8:2519-2537.

Foury, F., and J. Kolodynski. 1983. pif mutation blocks recombination between mitochondrial rho+ and

rho- genomes having tandemly arrayed repeat units in Saccharomyces cerevisiae. Proc Natl

Acad Sci U S A. 80:5345-5349.

Frederick, R.L., K. Okamoto, and J.M. Shaw. 2008. Multiple pathways influence mitochondrial

inheritance in budding yeast. Genetics. 178:825-837.

Frenz, L.M., S.E. Lee, D. Fesquet, and L.H. Johnston. 2000. The budding yeast Dbf2 protein kinase

localises to the centrosome and moves to the bud neck in late mitosis. J Cell Sci. 113 Pt 19:3399-

3408.

Garcia-Rodriguez, L.J., D.G. Crider, A.C. Gay, I.J. Salanueva, I.R. Boldogh, and L.A. Pon. 2009.

Mitochondrial inheritance is required for MEN-regulated cytokinesis in budding yeast. Curr Biol.

19:1730-1735.

Page 100: Mitochondrial and Mitochondrial DNA Inheritance

93

Garcia-Rodriguez, L.J., A.C. Gay, and L.A. Pon. 2007. Puf3p, a Pumilio family RNA binding protein,

localizes to mitochondria and regulates mitochondrial biogenesis and motility in budding yeast. J

Cell Biol. 176:197-207.

Goldring, E.S., L.I. Grossman, D. Krupnick, D.R. Cryer, and J. Marmur. 1970. The petite mutation in yeast.

Loss of mitochondrial deoxyribonucleic acid during induction of petites with ethidium bromide. J

Mol Biol. 52:323-335.

Hartwell, L.H., J. Culotti, and B. Reid. 1970. Genetic control of the cell-division cycle in yeast. I. Detection

of mutants. Proc Natl Acad Sci U S A. 66:352-359.

Hartwell, L.H., and M.B. Kastan. 1994. Cell cycle control and cancer. Science. 266:1821-1828.

Hartwell, L.H., and T.A. Weinert. 1989. Checkpoints: controls that ensure the order of cell cycle events.

Science. 246:629-634.

Hobbs, A.E., M. Srinivasan, J.M. McCaffery, and R.E. Jensen. 2001. Mmm1p, a mitochondrial outer

membrane protein, is connected to mitochondrial DNA (mtDNA) nucleoids and required for

mtDNA stability. J Cell Biol. 152:401-410.

Honlinger, A., U. Bomer, A. Alconada, C. Eckerskorn, F. Lottspeich, K. Dietmeier, and N. Pfanner. 1996.

Tom7 modulates the dynamics of the mitochondrial outer membrane translocase and plays a

pathway-related role in protein import. EMBO J. 15:2125-2137.

Huckaba, T.M., A.C. Gay, L.F. Pantalena, H.C. Yang, and L.A. Pon. 2004. Live cell imaging of the assembly,

disassembly, and actin cable-dependent movement of endosomes and actin patches in the

budding yeast, Saccharomyces cerevisiae. J Cell Biol. 167:519-530.

Huckaba, T.M., T. Lipkin, and L.A. Pon. 2006. Roles of type II myosin and a tropomyosin isoform in

retrograde actin flow in budding yeast. J Cell Biol. 175:957-969.

Hwa Lim, H., F.M. Yeong, and U. Surana. 2003. Inactivation of mitotic kinase triggers translocation of

MEN components to mother-daughter neck in yeast. Mol Biol Cell. 14:4734-4743.

Itoh, T., E.A. Toh, and Y. Matsui. 2004. Mmr1p is a mitochondrial factor for Myo2p-dependent

inheritance of mitochondria in the budding yeast. EMBO J. 23:2520-2530.

Itoh, T., A. Watabe, E.A. Toh, and Y. Matsui. 2002. Complex formation with Ypt11p, a rab-type small

GTPase, is essential to facilitate the function of Myo2p, a class V myosin, in mitochondrial

distribution in Saccharomyces cerevisiae. Mol Cell Biol. 22:7744-7757.

Jimenez, J., V.J. Cid, R. Cenamor, M. Yuste, G. Molero, C. Nombela, and M. Sanchez. 1998.

Morphogenesis beyond cytokinetic arrest in Saccharomyces cerevisiae. J Cell Biol. 143:1617-

1634.

Kakuda, T.N. 2000. Pharmacology of nucleoside and nucleotide reverse transcriptase inhibitor-induced

mitochondrial toxicity. Clin Ther. 22:685-708.

Kasashima, K., M. Sumitani, M. Satoh, and H. Endo. 2008. Human prohibitin 1 maintains the organization

and stability of the mitochondrial nucleoids. Exp Cell Res. 314:988-996.

Koc, A., and G.F. Merrill. 2007. Checkpoint deficient rad53-11 yeast cannot accumulate dNTPs in

response to DNA damage. Biochem Biophys Res Commun. 353:527-530.

Kornmann, B., E. Currie, S.R. Collins, M. Schuldiner, J. Nunnari, J.S. Weissman, and P. Walter. 2009. An

ER-Mitochondria Tethering Complex Revealed by a Synthetic Biology Screen. Science. 325:477-

481.

Krapp, A., M.P. Gulli, and V. Simanis. 2004. SIN and the art of splitting the fission yeast cell. Curr Biol.

14:R722-730.

Larosche, I., P. Letteron, B. Fromenty, N. Vadrot, A. Abbey-Toby, G. Feldmann, D. Pessayre, and A.

Mansouri. 2007. Tamoxifen inhibits topoisomerases, depletes mitochondrial DNA, and triggers

steatosis in mouse liver. J Pharmacol Exp Ther. 321:526-535.

Lazzarino, D.A., I. Boldogh, M.G. Smith, J. Rosand, and L.A. Pon. 1994. Yeast mitochondria contain ATP-

sensitive, reversible actin-binding activity. Mol Biol Cell. 5:807-818.

Page 101: Mitochondrial and Mitochondrial DNA Inheritance

94

Lebedeva, M.A., and G.S. Shadel. 2007. Cell cycle- and ribonucleotide reductase-driven changes in

mtDNA copy number influence mtDNA Inheritance without compromising mitochondrial gene

expression. Cell Cycle. 6:2048-2057.

Lee, H.C., P.H. Yin, J.C. Lin, C.C. Wu, C.Y. Chen, C.W. Wu, C.W. Chi, T.N. Tam, and Y.H. Wei. 2005.

Mitochondrial genome instability and mtDNA depletion in human cancers. Ann N Y Acad Sci.

1042:109-122.

Lee, S.J., M.F. Schwartz, J.K. Duong, and D.F. Stern. 2003. Rad53 phosphorylation site clusters are

important for Rad53 regulation and signaling. Mol Cell Biol. 23:6300-6314.

Lee, Y.S., W.D. Kennedy, and Y.W. Yin. 2009. Structural insight into processive human mitochondrial DNA

synthesis and disease-related polymerase mutations. Cell. 139:312-324.

Lippincott, J., and R. Li. 1998. Dual function of Cyk2, a cdc15/PSTPIP family protein, in regulating

actomyosin ring dynamics and septin distribution. J Cell Biol. 143:1947-1960.

Lippincott, J., K.B. Shannon, W. Shou, R.J. Deshaies, and R. Li. 2001. The Tem1 small GTPase controls

actomyosin and septin dynamics during cytokinesis. J Cell Sci. 114:1379-1386.

Liu, Z., and R.A. Butow. 2006. Mitochondrial retrograde signaling. Annu Rev Genet. 40:159-185.

Longtine, M.S., A. McKenzie, 3rd, D.J. Demarini, N.G. Shah, A. Wach, A. Brachat, P. Philippsen, and J.R.

Pringle. 1998. Additional modules for versatile and economical PCR-based gene deletion and

modification in Saccharomyces cerevisiae. Yeast. 14:953-961.

Luca, F.C., M. Mody, C. Kurischko, D.M. Roof, T.H. Giddings, and M. Winey. 2001. Saccharomyces

cerevisiae Mob1p is required for cytokinesis and mitotic exit. Mol Cell Biol. 21:6972-6983.

Luca, F.C., and M. Winey. 1998. MOB1, an essential yeast gene required for completion of mitosis and

maintenance of ploidy. Mol Biol Cell. 9:29-46.

Makovets, S., and E.H. Blackburn. 2009. DNA damage signalling prevents deleterious telomere addition

at DNA breaks. Nat Cell Biol. 11:1383-1386.

Mandal, S., P. Guptan, E. Owusu-Ansah, and U. Banerjee. 2005. Mitochondrial regulation of cell cycle

progression during development as revealed by the tenured mutation in Drosophila. Dev Cell.

9:843-854.

Matsushima, Y., R. Garesse, and L.S. Kaguni. 2004. Drosophila mitochondrial transcription factor B2

regulates mitochondrial DNA copy number and transcription in schneider cells. J Biol Chem.

279:26900-26905.

McCammon, M.T., C.B. Epstein, B. Przybyla-Zawislak, L. McAlister-Henn, and R.A. Butow. 2003. Global

transcription analysis of Krebs tricarboxylic acid cycle mutants reveals an alternating pattern of

gene expression and effects on hypoxic and oxidative genes. Mol Biol Cell. 14:958-972.

McKane, M., K.K. Wen, I.R. Boldogh, S. Ramcharan, L.A. Pon, and P.A. Rubenstein. 2005. A mammalian

actin substitution in yeast actin (H372R) causes a suppressible mitochondria/vacuole phenotype.

J Biol Chem. 280:36494-36501.

Meeusen, S., and J. Nunnari. 2003. Evidence for a two membrane-spanning autonomous mitochondrial

DNA replisome. J Cell Biol. 163:503-510.

Meisinger, C., M. Rissler, A. Chacinska, L.K. Szklarz, D. Milenkovic, V. Kozjak, B. Schonfisch, C. Lohaus,

H.E. Meyer, M.P. Yaffe, B. Guiard, N. Wiedemann, and N. Pfanner. 2004. The mitochondrial

morphology protein Mdm10 functions in assembly of the preprotein translocase of the outer

membrane. Dev Cell. 7:61-71.

Meisinger, C., N. Wiedemann, M. Rissler, A. Strub, D. Milenkovic, B. Schonfisch, H. Muller, V. Kozjak, and

N. Pfanner. 2006. Mitochondrial protein sorting: differentiation of beta-barrel assembly by

Tom7-mediated segregation of Mdm10. J Biol Chem. 281:22819-22826.

Meitinger, F., B. Petrova, I.M. Lombardi, D.T. Bertazzi, B. Hub, H. Zentgraf, and G. Pereira. 2010.

Targeted localization of Inn1, Cyk3 and Chs2 by the mitotic-exit network regulates cytokinesis in

budding yeast. J Cell Sci. 123:1851-1861.

Page 102: Mitochondrial and Mitochondrial DNA Inheritance

95

Merkwirth, C., S. Dargazanli, T. Tatsuta, S. Geimer, B. Lower, F.T. Wunderlich, J.C. von Kleist-Retzow, A.

Waisman, B. Westermann, and T. Langer. 2008. Prohibitins control cell proliferation and

apoptosis by regulating OPA1-dependent cristae morphogenesis in mitochondria. Genes Dev.

22:476-488.

Merkwirth, C., and T. Langer. 2009. Prohibitin function within mitochondria: essential roles for cell

proliferation and cristae morphogenesis. Biochim Biophys Acta. 1793:27-32.

Millard, P.J., B.L. Roth, H.P. Thi, S.T. Yue, and R.P. Haugland. 1997. Development of the FUN-1 family of

fluorescent probes for vacuole labeling and viability testing of yeasts. Appl Environ Microbiol.

63:2897-2905.

Mineri, R., N. Pavelka, E. Fernandez-Vizarra, P. Ricciardi-Castagnoli, M. Zeviani, and V. Tiranti. 2009. How

do human cells react to the absence of mitochondrial DNA? PLoS One. 4:e5713.

Moseley, J.B., and B.L. Goode. 2006. The yeast actin cytoskeleton: from cellular function to biochemical

mechanism. Microbiol Mol Biol Rev. 70:605-645.

Nishihama, R., J.H. Schreiter, M. Onishi, E.A. Vallen, J. Hanna, K. Moravcevic, M.F. Lippincott, H. Han,

M.A. Lemmon, J.R. Pringle, and E. Bi. 2009. Role of Inn1 and its interactions with Hof1 and Cyk3

in promoting cleavage furrow and septum formation in S. cerevisiae. J Cell Biol. 185:995-1012.

Nurse, P., Y. Masui, and L. Hartwell. 1998. Understanding the cell cycle. Nat Med. 4:1103-1106.

O'Neill, B.M., S.J. Szyjka, E.T. Lis, A.O. Bailey, J.R. Yates, 3rd, O.M. Aparicio, and F.E. Romesberg. 2007.

Pph3-Psy2 is a phosphatase complex required for Rad53 dephosphorylation and replication fork

restart during recovery from DNA damage. Proc Natl Acad Sci U S A. 104:9290-9295.

Osman, C., M. Haag, C. Potting, J. Rodenfels, P.V. Dip, F.T. Wieland, B. Brugger, B. Westermann, and T.

Langer. 2009a. The genetic interactome of prohibitins: coordinated control of cardiolipin and

phosphatidylethanolamine by conserved regulators in mitochondria. J Cell Biol. 184:583-596.

Osman, C., C. Merkwirth, and T. Langer. 2009b. Prohibitins and the functional compartmentalization of

mitochondrial membranes. J Cell Sci. 122:3823-3830.

Owusu-Ansah, E., A. Yavari, S. Mandal, and U. Banerjee. 2008. Distinct mitochondrial retrograde signals

control the G1-S cell cycle checkpoint. Nat Genet. 40:356-361.

Park, C.B., and N.G. Larsson. 2011. Mitochondrial DNA mutations in disease and aging. J Cell Biol.

193:809-818.

Paulovich, A.G., and L.H. Hartwell. 1995. A checkpoint regulates the rate of progression through S phase

in S. cerevisiae in response to DNA damage. Cell. 82:841-847.

Pellicioli, A., S.E. Lee, C. Lucca, M. Foiani, and J.E. Haber. 2001. Regulation of Saccharomyces Rad53

checkpoint kinase during adaptation from DNA damage-induced G2/M arrest. Mol Cell. 7:293-

300.

Reichard, P. 1988. Interactions between deoxyribonucleotide and DNA synthesis. Annu Rev Biochem.

57:349-374.

Schultz, R.A., S.J. Swoap, L.D. McDaniel, B. Zhang, E.C. Koon, D.J. Garry, K. Li, and R.S. Williams. 1998.

Differential expression of mitochondrial DNA replication factors in mammalian tissues. J Biol

Chem. 273:3447-3451.

Schwartz, M.F., S.J. Lee, J.K. Duong, S. Eminaga, and D.F. Stern. 2003. FHA domain-mediated DNA

checkpoint regulation of Rad53. Cell Cycle. 2:384-396.

Senning, E.N., and A.H. Marcus. 2010. Actin polymerization driven mitochondrial transport in mating S.

cerevisiae. Proc Natl Acad Sci U S A. 107:721-725.

Seshan, A., and A. Amon. 2004. Linked for life: temporal and spatial coordination of late mitotic events.

Curr Opin Cell Biol. 16:41-48.

Sherman, F. 2002. Getting started with yeast. Methods Enzymol. 350:3-41.

Page 103: Mitochondrial and Mitochondrial DNA Inheritance

96

Shou, W., J.H. Seol, A. Shevchenko, C. Baskerville, D. Moazed, Z.W. Chen, J. Jang, H. Charbonneau, and

R.J. Deshaies. 1999. Exit from mitosis is triggered by Tem1-dependent release of the protein

phosphatase Cdc14 from nucleolar RENT complex. Cell. 97:233-244.

Sidorova, J.M., and L.L. Breeden. 1997. Rad53-dependent phosphorylation of Swi6 and down-regulation

of CLN1 and CLN2 transcription occur in response to DNA damage in Saccharomyces cerevisiae.

Genes Dev. 11:3032-3045.

Simon, V.R., S.L. Karmon, and L.A. Pon. 1997. Mitochondrial inheritance: cell cycle and actin cable

dependence of polarized mitochondrial movements in Saccharomyces cerevisiae. Cell Motil

Cytoskeleton. 37:199-210.

Singh, K.K., V. Ayyasamy, K.M. Owens, M.S. Koul, and M. Vujcic. 2009. Mutations in mitochondrial DNA

polymerase-gamma promote breast tumorigenesis. J Hum Genet. 54:516-524.

Sogo, L.F., and M.P. Yaffe. 1994. Regulation of mitochondrial morphology and inheritance by Mdm10p, a

protein of the mitochondrial outer membrane. J Cell Biol. 126:1361-1373.

Song, S., T.Z. Grenfell, S. Garfield, R.L. Erikson, and K.S. Lee. 2000. Essential function of the polo box of

Cdc5 in subcellular localization and induction of cytokinetic structures. Mol Cell Biol. 20:286-298.

Song, S., and K.S. Lee. 2001. A novel function of Saccharomyces cerevisiae CDC5 in cytokinesis. J Cell Biol.

152:451-469.

Stegmeier, F., and A. Amon. 2004. Closing mitosis: the functions of the Cdc14 phosphatase and its

regulation. Annu Rev Genet. 38:203-232.

Taylor, S.D., H. Zhang, J.S. Eaton, M.S. Rodeheffer, M.A. Lebedeva, W. O'Rourke T, W. Siede, and G.S.

Shadel. 2005. The conserved Mec1/Rad53 nuclear checkpoint pathway regulates mitochondrial

DNA copy number in Saccharomyces cerevisiae. Mol Biol Cell. 16:3010-3018.

Trifunovic, A., A. Wredenberg, M. Falkenberg, J.N. Spelbrink, A.T. Rovio, C.E. Bruder, Y.M. Bohlooly, S.

Gidlof, A. Oldfors, R. Wibom, J. Tornell, H.T. Jacobs, and N.G. Larsson. 2004. Premature ageing in

mice expressing defective mitochondrial DNA polymerase. Nature. 429:417-423.

Tyynismaa, H., H. Sembongi, M. Bokori-Brown, C. Granycome, N. Ashley, J. Poulton, A. Jalanko, J.N.

Spelbrink, I.J. Holt, and A. Suomalainen. 2004. Twinkle helicase is essential for mtDNA

maintenance and regulates mtDNA copy number. Hum Mol Genet. 13:3219-3227.

Tzagoloff, A., G. Macino, and W. Sebald. 1979. Mitochondrial genes and translation products. Annu Rev

Biochem. 48:419-441.

Van Aken, O., J. Whelan, and F. Van Breusegem. 2010. Prohibitins: mitochondrial partners in

development and stress response. Trends Plant Sci. 15:275-282.

Veses, V., A. Richards, and N.A. Gow. 2009. Vacuole inheritance regulates cell size and branching

frequency of Candida albicans hyphae. Mol Microbiol. 71:505-519.

Viikov, K., P. Valjamae, and J. Sedman. 2011. Yeast mitochondrial DNA polymerase is a highly processive

single-subunit enzyme. Mitochondrion. 11:119-126.

Wallace, D.C. 2010. Mitochondrial DNA mutations in disease and aging. Environ Mol Mutagen. 51:440-

450.

Wang, Y., and D.F. Bogenhagen. 2006. Human mitochondrial DNA nucleoids are linked to protein folding

machinery and metabolic enzymes at the mitochondrial inner membrane. J Biol Chem.

281:25791-25802.

Weinert, T. 1998. DNA damage checkpoints update: getting molecular. Curr Opin Genet Dev. 8:185-193.

Wiedemann, N., V. Kozjak, A. Chacinska, B. Schonfisch, S. Rospert, M.T. Ryan, N. Pfanner, and C.

Meisinger. 2003. Machinery for protein sorting and assembly in the mitochondrial outer

membrane. Nature. 424:565-571.

Williamson, D.H., and D.J. Fennell. 1979. Visualization of yeast mitochondrial DNA with the fluorescent

stain "DAPI". Methods Enzymol. 56:728-733.

Page 104: Mitochondrial and Mitochondrial DNA Inheritance

97

Xu, S., H.K. Huang, P. Kaiser, M. Latterich, and T. Hunter. 2000. Phosphorylation and spindle pole body

localization of the Cdc15p mitotic regulatory protein kinase in budding yeast. Curr Biol. 10:329-

332.

Yang, H.C., and L.A. Pon. 2002. Actin cable dynamics in budding yeast. Proc Natl Acad Sci U S A. 99:751-

756.

Yoshida, S., and A. Toh-e. 2001. Regulation of the localization of Dbf2 and mob1 during cell division of

saccharomyces cerevisiae. Genes Genet Syst. 76:141-147.

Youngman, M.J., A.E. Hobbs, S.M. Burgess, M. Srinivasan, and R.E. Jensen. 2004. Mmm2p, a

mitochondrial outer membrane protein required for yeast mitochondrial shape and

maintenance of mtDNA nucleoids. J Cell Biol. 164:677-688.

Yu, M., Y. Zhou, Y. Shi, L. Ning, Y. Yang, X. Wei, N. Zhang, X. Hao, and R. Niu. 2007. Reduced

mitochondrial DNA copy number is correlated with tumor progression and prognosis in Chinese

breast cancer patients. IUBMB Life. 59:450-457.

Zelenaya-Troitskaya, O., S.M. Newman, K. Okamoto, P.S. Perlman, and R.A. Butow. 1998. Functions of

the high mobility group protein, Abf2p, in mitochondrial DNA segregation, recombination and

copy number in Saccharomyces cerevisiae. Genetics. 148:1763-1776.

Zhao, X., E.G. Muller, and R. Rothstein. 1998. A suppressor of two essential checkpoint genes identifies a

novel protein that negatively affects dNTP pools. Mol Cell. 2:329-340.

Zhao, X., and R. Rothstein. 2002. The Dun1 checkpoint kinase phosphorylates and regulates the

ribonucleotide reductase inhibitor Sml1. Proc Natl Acad Sci U S A. 99:3746-3751.