UINVERSITY OF CALIFORNIA

Los Angeles

Small Molecule Probes of Mitochondrial Translocases Elucidate

PINK1 Trafficking in PINK1/Parkin Regulated Mitophagy

A dissertation submitted in partial satisfaction of the

requirements for the degree Doctor of Philosophy

in Biochemistry, Molecular, and Structural Biology

by

Michael Conti

2018

© Copyright by

Michael Conti

2018 ABSTRACT OF THE DISSERTATION

Small Molecule Probes of Mitochondrial Translocases Elucidate

PINK1 Trafficking in PINK1/Parkin Regulated Mitophagy

by

Michael Conti

Doctor in Philosophy in Biochemistry, Molecular, and Structural Biology

University of California, Los Angeles, 2018

Professor Carla Marie Koehler, Chair

Dysfunctional mitochondria result in many human diseases including some cancers and many metabolic diseases. More recently, momentum has been building to implicate defective mitochondria as a leading cause of neurodegenerative diseases. Because preventions and therapies to treat such diseases are extremely limited and non-efficacious, there remains a very large burden on human health. But before effective therapies can be developed, additional research is needed to understand the underlying mechanisms of disease progression on a molecular level. Approaches to investigate these diseases pathways, such as genetic manipulations and RNAi, are useful but limited in what they can achieve. To alleviate this, we have performed a small molecule screen designed to discover inhibitors of protein import into

ii mitochondria and have applied one of these inhibitors, named MitoBloCK-10, to further clarify a protein pathway involved in autosomal recessive Parkinson’s disease.

In this pathway, proteins PINK1 and Parkin work in concert to selectively remove dysfunctional mitochondria from the total cellular population. When either PINK1, which is partially imported into mitochondria, or Parkin, which is recruited to mitochondria by PINK1, is not functioning properly, damaged mitochondria accumulate and this leads to neuronal degradation and Parkinson’s disease. In this study, we use MitoBlock-10, one of the molecules discovered in the screen, to reveal an important role for the mitochondrial import protein

TIMM44 in PINK1 trafficking at mitochondria.

Two other molecules discovered in the screen are also of interest. DECA, an FDA- approved compound, shows an interesting interaction with PINK1 and offers promise as a therapy for Parkinson’s disease. MitoBloCK-11, a novel molecule, has a protein binding target,

Seo1, that was previously believed to exist exclusively at the plasma membrane. This collection of molecules not only increases the tools available for research, it has revealed important insights into PINK1 trafficking and Seo1 localization.

iii

The dissertation of Michael Conti is approved.

Catherine F. Clarke

Alexander M. van der Bliek

Carla Marie Koehler, Committee Chair

University of California, Los Angeles

2018

iv

DEDICATION

To my parents, for all of the encouragement. To Antoinette, for always understanding.

v

TABLE OF CONTENTS

List of Figures List of Tables Acknowledgements Vita

Chapter 1. Introduction– Discovering small molecules probes that attenuate mitochondrial protein trafficking ...... 1 Chapter 2. Adaptation of a Genetic Screen Reveals an Inhibitor for Mitochondrial Protein Import Component Tim44 ...... 32 Chapter 3. A small molecule probe elucidates the role of mitochondrial translocase TIMM44 in PINK1/Parkin regulated mitophagy...... 60 Chapter 4. DECA induces PINK1 accumulation and Parkin recruitment ...... 93 Chapter 5. A small molecule inhibitor of mitochondrial protein import possibly acts through transport protein Seo1 ...... 104

LIST OF FIGURES

Chapter 1. Figure 1.1 Overview of protein trafficking in mitochondria ...... 15 Figure 1.2 Pathological hallmarks of Parkinson’s disease ...... 16 Figure 1.3 A model of PINK1/Parkin-dependent mitophagy...... 17 Figure 1.4 Complex I may be an important link between mitophagy and Parkinson’s disease ...... 18 Figure 1.5 Growth phenotype strategy for small molecule screen ...... 19

Chapter 2. Figure 1 MB-10 is a potential attenuator of protein import into mitochondria ...... 34 Figure 2 MB-10 inhibits the import of substrates that use the TIM23 import pathway...... 35 Figure 3 MB-10 inhibition of import depends on specific chemical characteristics ...... 37 Figure 4 MB-10 targets Tim44 ...... 38 Figure 5 Mutations in the α4 helix of the C-terminal domain of Tim44 confer resistance to MB-10 ...... 38 Figure 6 MB-10 fits in a binding pocket in the C-terminal domain of Tim44 ...... 39 Figure 7 MB-10 inhibits Tim44 binding to the precursor and to Hsp70, but not other components of the PAM complex...... 40 Figure 8 MB-10 inhibits protein import into mammalian mitochondria ...... 41 Figure 9 MB-10 treatment impairs cardiac development and induces apoptosis in zebrafish ...... 42 Figure S1 MB-10 activity is influenced by specific chemical characteristics ...... 58 Figure S2 Mutations H292A, W316A, W316Y, and W417A in Tim44 confer a dominant negative phenotype ...... 59

vi

Chapter 3. Figure 3.1 MB-10 inhibits the accumulation of PINK1 on damaged mitochondria, but does not affect PINK1 import into healthy mitochondria ...... 74 Figure 3.2 MB-10 disrupts association of PINK1 with the TOM complex and inhibits CCCP- induced recruitment of Parkin to mitochondria in HeLa cells overexpressing stable EGFP-Parkin ...... 76 Figure 3.3 MB-10 slows down mitophagy but does not affect PINK1 transcription ...... 78 Figure 3.4 SAR compounds of MB-10 have no affect on PINNK1-dependent mitophagy ....80 Figure 3.5 TIMM44 knockdown inhibits PINK1 association with the TOM complex but does not interfere with normal PINK1 import ...... 82 Figure 3.S1 Human and yeast Tim44 are highly conserved ...... 84 Figure 3.S2 MB-10 inhibits the import of precursors that use the TIM23 import pathway ...... 86 Figure 3.S3 MB-10 inhibits CCCP-induced mitophagy ...... 87

Chapter 4. Figure 4.1 DECA strongly inhibits the import of precursors that use the TIM23 import pathway and modestly inhibits the import of precursors that use the redox- regulated pathway ...... 97 Figure 4.2 DECA causes accumulation of PINK1 on mitochondria ...... 98 Figure 4.3 DECA promotes Parkin recruitment to mitochondria that are already partially dysfunctional ...... 100 Figure 4.4 MB-10 enhances the activity of DECA ...... 102

Chapter 5. Figure 5.1 MB-11 inhibits precursor proteins that contain hydrophobic segments ...... 116 Figure 5.2 MB-11 targets a protein on the OM, but not Tom70 or Tom20 ...... 118 Figure 5.3 An approach to obtain spontaneous revertants that increase viability in the presence of MB-11 ...... 120 Figure 5.4 Growth assays for the ∆seo1 strain in the presence of various small molecules .122 Figure 5.5 MB-11 inhibits the import of some precursors into mammalian mitochondria...123 Figure 5.S1 MB-11 confers growth in media lacking uracil in a specific manner ...... 125 Figure 5.S2 MB-11 affects zebrafish development ...... 127

vii

LIST OF TABLES

Chapter 1. Table 1.1 Summary of the libraries from the small molecule screen ...... 20

Chapter 2. Table S1 Strains constructed in this study ...... 54 Table S2 Summary of chemical genetic screen ...... 56 Table S3 Results of the MM/PBA calculation for MB-10 bound to wild-type Tim44 and Tim44 mutants (F284L, T290S, I297V, and V298A) ...... 57

Chapter 5. Table 5.1 A summary of the results from MIC 50 assays with the spontaneous revertants ..128 Table 5.2 A complete list of all mutations located in any exon (non-synonymous mutations only), in any intron, that gain or lose a stop codon, or that remove a start codon from all the strains that were sequenced ...... 129

viii

ACKNOWLEDGEMENTS

This chapter of my life has truly challenged me. As the end approaches, I look back at all of the obstacles, all of the ups-and-downs, all of the successes, and I am in awe of just how far I have come. There have been many people that have helped me reach this end. I must extend my gratitude to my thesis advisor, Dr. Carla Koehler, for allowing me to fulfill my work in her lab. I think a skill that Carla has is being able to alter her management style to those in her lab. In contrast to some others in her lab, she has always allowed me to work independently and forge my own path in my research with little interference from her. For that I am sincerely thankful, for it helped me grow as both a scientist and a thinker. I would also like to thank my committee members, Dr. Catherine Clarke, Dr. James Gober, Dr. Michael Teitell, and Dr. Alex van der

Bliek for all of their guidance and support along the way.

I must also thank my mentor for my first year in the Koehler lab, Dr. Non Miyata. I really had no idea what I was getting myself into when I started graduate school, and Non helped me to settle into lab and provided excellent guidance. His presence kept me on track and I’m not sure how I would have fared without him. Although Non left the lab 4 years ago for the next step of his career in Japan, I still go to him with questions.

I should also acknowledge the other members of the lab whose paths I crossed while here, including post-docs Dr. Janos Steffen, Dr. Deepa Dabir, and Dr. Melania Abrahamian, graduate students Dr. Matt Maland, Dr. Meghan Johnson, Dr. Colin Douglas, Dr. Juwina Wijaya,

Eriko Shimada, Eric Torres, Vivian Zhang, and Shahriyar Jahanbakhsh, lab managers Tanya

Hioe, Jisoo Han, Jenny Ngo, Jerry Shmorgon, Kayla White, Sean Atamdede, and our many undergraduate students. Thank you all for the discussions about science and the discussions

ix about everything else, for the happy hours, for all being on this ride together, and for the memories. You’ve all affected my life in some way.

And finally, my most sincere appreciation and thanks to all of the people who have helped me outside of the lab. My parents for always believing in me, always being patient with me, and always encouraging me. Without you, I would not have had the perseverance to complete this journey. To my brother, I am grateful to have you in my life and call you family.

To Antoinette, thank you so much for always supporting me even through the lowest of times, I am forever thankful to have found you and would be lost without you. To the many people that I have met along the way- Danny, KidAdam, Boston Rob, Reid, Mike U, TJ, Sean, the rest of my cohort, all of the soccer guys, the members of the Chanfreau lab, most of the members of the

Gelbart lab, and so many others, thank you all for the friendships and the memories!

x

Chapter 1 is an introduction to protein trafficking at the mitochondria and how this process relates to the health of cells and whole organisms, with particular emphasis on

Parkinson’s disease. Additionally, a small molecule screen is described to discover inhibitors of protein import into mitochondria. Special thanks goes to Non Miyata for designing and performing the screen and to Robert Damoiseaux for assistance at the screening facility.

Chapter 2 is a reprint of Miyata, N.; Tang, Z.; Conti, M. A.; Johnson, M. E.; Douglas, C.

J.; Hasson, S. A.; Damoiseaux, R.; Chang, C. A.; Koehler, C. M., Adaptation of a Genetic Screen

Reveals an Inhibitor for Mitochondrial Protein Import Component Tim44. J Biol Chem 2017,

292 (13), 5429-5442. [DOI: 10.1074/jbc.M116.770131]. Non Miyata managed the research and

performed much of the lab work. My role on this project was to aid Non in planning and

performing the experiments; I also addressed all of the comments from the reviewers. Z. Tang

performed the molecular dynamics simulation work in the lab of C.A. Chang. Meghan Johnson

performed the zebrafish studies. Colin Douglas provided guidance on small molecule chemistry.

Sam Hasson aided in establishing the screening strain and provided guidance for the screen.

Robert Damoiseaux gave technical assistance at the screening facility. The Journal of Biological

Chemistry has a policy for reuse of its published material in a thesis or dissertation. I have

obtained this permission in writing from the journal editor, Sarah Crespi.

Chapter 3 is a version of a manuscript in preparation to be authored by myself and

Koehler, CM. The title is “A small molecule probe elucidates the role of mitochondrial

translocase TIMM44 in PINK1/Parkin regulated mitophagy” with plans to submit to J Cell Bio . I

would like to thank Dr. Richard Youle for the EGFP-Parkin expressing HeLa cells and the MFN-

1 antibody, and Dr. Janos Steffen and Dr. Juwina Wijaya for helpful discussions and technical

advice.

xi

Chapter 4 highlights the application of a molecule from the screen named DECA to induce PINK1 accumulation at mitochondria. Thank you to Non Miyata for performing the experiment and imaging of cells treated with just DMSO or DECA.

Chapter 5 characterizes another small molecule discovered in the screen, MB-11, and discussion of its potential protein binding target Seo1. Thank you to Non Miyata for performing some of the in vitro import reactions. Thank you to Yong Xue and Trent Su from the lab of

Siavash Kurdistani for guidance on preparing and submitting samples for the next generation sequencing and performing the bioinformatics on the subsequent data. Thank you to Meghan

Johnson for performing the zebrafish studies.

This work was partially funded by the Ruth L Kirschstein National Research Service

Award GM007185.

xii

VITA

2006-2010 B.S. Biochemistry Boston College, Chestnut Hill, MA

2009-2010 Research Assistant Microchips Biotech Inc., Burlington, MA

2010-2012 Laboratory Technician Addgene, Inc., Cambridge, MA

2012-2018 Graduate Student Researcher Department of Chemistry and Biochemistry University of California, Los Angeles

2012-2018 Teaching Assistant Department of Chemistry and Biochemistry University of California, Los Angeles

PUBLICATIONS

Miyata, N., Z. Tang, M.A. Conti , M.E. Johnson, C.J. Douglas, S.A. Hasson, R. Damoiseaux, C.A. Chang, and C.M. Koehler. 2017. Adaptation of a Genetic Screen Reveals an Inhibitor for Mitochondrial Protein Import Component Tim44. J Biol Chem . 292:5429- 5442

Filipuzzi, I., J. Steffen, M. Germain, L. Goepfert, M.A. Conti , C. Potting, R. Cerino, M. Pfeifer, P. Krastel, D. Hoepfner, J. Bastien, C.M. Koehler, and S.B. Helliwell. 2017. Stendomycin selectively inhibits TIM23-dependent mitochondrial protein import. Nat Chem Biol . 13:1239-1244.

INVITED SPEAKER PRESENTATIONS

2016 Small Molecule Inhbitors Display Distinct Phenotypes in a Parkinson’s Disease Mechanism. Presented at Gordon Research Conference. Presented at Gordon Research Conference: Protein Translocation Across Cell Membranes, Galveston, Texas.

2017 Elucidating the Mechanism of PINK1 Trafficking in Mitochondria Using a Small Molecule Probe. Presented at UCLA Mitochondria Symposium, Los Angeles, CA.

xiii

POSTER PRESENTATIONS

Michael A. Conti , Non Miyata, Robert Damoiseaux, and Carla M. Koehler (2014). A small molecule inhibitor of mitochondrial protein import and its application in Parkinson’s disease research. Presented at Gordon Research Conference: Protein Translocation Across Cell Membranes, Galveston, Texas.

Michael A. Conti , Non Miyata, Robert Damoiseaux, and Carla M. Koehler (2015). Chemical probes outline PINK1 import into mitochondria. Presented at EMBO Conference: Mechanisms and Regulation of Protein Translocation, Dubrovnik, Croatia and at UCLA MBI Retreat, Los Angeles, CA.

Michael A. Conti , Non Miyata, Robert Damoiseaux, and Carla M. Koehler (2016). Small molecule inhibitors display distinct phenotypes in a Parkinson’s disease mechanism. Presented at Gordon Research Conference: Protein Translocation Across Cell Membranes, Galveston, Texas and at UCLA MBI Retreat, Lake Arrowhead, CA.

Michael A. Conti , Non Miyata, and Carla M. Koehler (2017). Elucidating the mechanism of PINK1 trafficking in mitochondria using a small molecule probe. Presented at SoCal Biomedical Sciences Gradute Student Symposium, Los Angeles, CA and at UCLA Mitochondria Symposium, Los Angeles, CA.

HONORS AND AWARDS

2010 Cum Laude ; Dean’s List High Honors; Member of Golden Key International Honor Society Boston College, Chestnut Hill, MA

2012 Pauley Fellowship Physical Sciences Division University of California, Los Angeles

2013-2016 Ruth L. Kirschstein National Research Service Award GM007185 University of California, Los Angeles

2014 Excellence in 2 nd Year Academics and Research Department of Chemsitry and Biochemistry University of California, Los Angeles

2015 MBI Retreat Outstanding Poster Presentation Molecular Biology Institute, University of California, Los Angeles

2015 Certificate of completion: Proficiency in the Business of Science Business of Science Center University of California, Los Angeles xiv

Chapter 1: Introduction – Discovering small molecules probes that attenuate mitochondrial protein trafficking

The importance of mitochondria

Traditionally, research scientists were given free range to explore biological systems for the sole pursuit of knowledge. In the current research landscape, though, funding is very limited and there is more pressure to study scientific questions related to disease or that have a direct impact on improving the human condition 1. Not only is it a near requirement to have a functional application of your research to secure grant money, but concentrating studies in disease states also lends itself to developing partnerships with commercial partners, which can bring in additional large quantities of funding. Because of this, many labs that have traditionally stuck to basic research are now expanding into translational fields and focusing on topics that have established connections to human disease. With this considered, the mitochondrion has become of interest to many labs because of its importance to cellular and whole organism health.

The mitochondrion is a dual-membrane organelle in eukaryotic cells that originated by early eukaryotes engulfing mitochondria, presumably an exogenous prokaryote, forming an endosymbitoic relationship 2. Having two membranes makes the biogenesis of mitochondria

relatively complex and tightly controlled, with various import pathways for precursor proteins to

follow depending on their destination. There are many diseases associated with mutations in

components of these pathways, such as DNAJC19 in dilated cardiomyopathy with ataxia 3, DDP1 in deafness-dystonia syndrome 4, HSP60 in hereditary spastic paraplegia 5, amyloid-β

accumulation at import channels in Alzheimer’s disease 6, OPA1 in dominant optic atrophy 7, 1

OMA1 in some metabolic defects 8, XPNPEP3 in cystic kidney disease with ciliary dysfunction 9,

and PINK1 10 or α-synuclein 11 in Parkinson’s disease, among many others. Clearly there is a

significant need to study this organelle, particularly in relation these components of protein

trafficking.

There is also a need for new methods of study. Although a great deal of knowledge has

already been established, alternative approaches will drastically increase the breadth of studies

that we can perform and can often be developed into novel therapeutic options. Small molecule

probes are widely used as drugs to treat diseases, but there is much less adoption of them as tools

for research. Within the scope of mitochondrial protein trafficking, there are scant resources for

inhibiting or activating a specific component of the translocation machinery. To alleviate this

issue and expand our research capabilities, I have characterized molecular inhibitors of protein

trafficking at the mitochondria discovered in the Koehler lab and utilized them to elucidate the

role of certain proteins in a Parkinson’s disease protein pathway.

Protein trafficking at mitochondria

The mitochondrion contains its own genome located in the matrix that codes for

important membrane components of the respiration machinery. But this subset only represents

approximately 1% of total mitochondrial proteins, with the other 99% encoded in nuclear DNA,

synthesized by cytosolic ribosomes, and finally imported into mitochondria through various

transport machineries 12 . Due to its dual-membrane nature, there are four destinations for mitochondrial precursor proteins including the outer membrane (OM), the inter membrane space

(IMS), the inner membrane (IM), and the matrix, with various pathways to arrive at each 13 .

Precursor proteins typically contain targeting sequences that not only localize them to

2 mitochondria (MTS), but also direct their trafficking once there, ensuring they end up in the correct compartment (Figure 1.1).

Located on the OM is the General Import Pore (GIP) or TOM complex, which serves as the initial entry gate of nearly all precursor proteins 14, 15 . This dynamic complex exists in an oligomeric form consisting of the channel-forming protein Tom40, the essential protein Tom22, small subunits that function in assembly and disassembly Tom5, Tom6, and Tom7, and is associated with receptor proteins Tom20 and Tom70 16 . Whereas Tom20 acts as the initial

recognition site for soluble precursor proteins, typically by recognizing mitochondrial targeting

sequences 17 , Tom70 serves as the receptor for hydrophobic precursor proteins, which often contain internal targeting signals 18 . Both receptors transfer their cargo to the central receptor protein Tom22 19 , and each can partially compensate for the other if one becomes defective 20 .

Tom22 deposits the precursor proteins into the pore of Tom40, a β-barrel protein with an interior that contains both hydrophilic and hydrophobic regions to facilitate the import of both soluble and membrane-embedded precursor proteins 21 . Upon exiting the Tom40 channel, the precursor proteins bind to the IMS domain of Tom22 22 . In addition to providing binding sites on both the cytosolic and IMS sides of the OM, the membrane domain of Tom22 also provides a crucial role in the formation of the TOM complex 19 . The small subunits are also important for assembly and dynamics of the TOM complex, but are not essential 23 and it has been suggested that Tom5

participates in the precursor transfer from Tom22 to Tom40 24 .

The path for a precursor protein diverges once through the Tom40 channel depending on

its final destination, which is often determined by its targeting sequence. The most classical type

is the cleavable amphipathic presequence, typically located at the N-terminus of the precursor

protein and targets it to the mitochondrial matrix via the TIM23 complex 25 . These signals are

3 recognized by Tom20 and Tom22 for import initiation on the OM before being handed off 20, 26 .

The TIM23 complex is composed of proteins including Tim50, Tim23, Tim17, Tim21, and

Mgr2 27 . Tim50 acts as a receptor protein that coordinates with Tom22 to bind the precursor 28 .

The channel-forming protein Tim23 also interacts with Tim50 to keep the pore open when

precursors are entering and closed otherwise 29 . It is also believed that Tim17 acts in conjunction

with Tim23 to form the TIM23 complex pore 30 . Import of precursors through the TIM23

complex and into the matrix relies upon a membrane potential ( ∆ψ) across the IM to create an

electrophoretic effect on the positively charged targeting sequence 31 and to activate the channel 32 . Due to this ∆ψ dependence, the TIM23 complex is coupled to the respiratory chain

complexes III and IV by a transient interaction that is regulated by Tim21 33 .

Although many proteins containing amphipathic presequences are destined for the matrix, there is also a subset that are laterally sorted into the IM itself in a process known as the stop transfer mechanism 34 . These precursors contain a second targeting signal behind the presequence that is hydrophobic in nature 35 . An important component of this process is Mgr2, which is considered the lateral gatekeeper for laterally-sorted precursors by binding to the hydrophobic sorting signal to control its release from the TIM23 complex 36 . It is also believed that Tim17 may play an important role in this process 37 . For both matrix-targeted and laterally sorted precursors, the matrix protease MPP serves to cleave the presequence and form the mature protein 38 . For laterally inserted precursors, a second cleavage occurs by IMP 39 for removal of the

hydrophobic sorting sequence followed by release of the mature protein to the IMS or anchoring

in the membrane. The resulting peptides are degraded by PreP 40 .

Membrane potential is sufficient to transport the presequence of precursor proteins into the matrix, but for complete import additional energy is needed in the form of ATP. As matrix-

4 targeted proteins emerge from the Tim23 channel, the PAM complex identifies and pulls the protein fully through the pore into the membrane. Components of the PAM complex include

Tim44, mtHsp70, Mge1, Pam16, Pam17, and Pam18 41, 42 . The dual-domain protein Tim44 is a

peripherally associated membrane protein that recognizes the precursor as it comes through the

Tim23 channel, connecting it with the chaperone mtHsp70 43 . mtHsp70 drives the translocation of the precursor into the matrix in a reaction cycle that hydrolyzes ATP, a mechanism which is stimulated by the J-protein Pam18 and further regulated by the J-like protein Pam16 44, 45 . The

nucleotide exchange factor Mge1 promotes the release of ADP from mtHsp70, allowing for a

new round of ATP binding and hydrolysis 46 . Because the PAM complex is quite dynamic and

has many interacting sites with the TIM23 complex, it is a highly regulated machine. Much of

this organization is performed by Pam17, although its exact action remains elusive 47, 48 .

There are a large family of proteins that function as metabolite carriers, most of which are

embedded in the IM. These proteins often contain targeting sequences located internally 49 and

utilize multiple Tom70 molecules on the OM for recognition and initiation of import 50, 51 . As with other precursors, these proteins use the Tom40 channel. Upon entry into the IMS, small

TIM chaperones form heterohexameric complexes to guide the hydrophobic precursors. There are two known TIM chaperone complexes that form, the more prolific Tim9-Tim10 and the non- essential Tim8-Tim13 complexes 52, 53 . A fifth small Tim protein, Tim12, forms a complex with

Tim9-Tim10 to bind the TIM22 complex of the IM for precursor insertion into the IM 54 . This

complex contains proteins Tim54, Tim22, and Tim18-Sdh3. Tim54 acts as the receptor for the

complex and directly binds the Tim9-Tim10-Tim12 complex 55 . Tim22 serves as the channel- forming protein in the complex, allowing the insertion of the precursor into the IM in a ∆ψ- dependent process 56 . The role of the Tim18-Sdh3 module is not conclusively established, but it is

5 believed to be important for the assembly of the TIM22 complex 55 . In addition to the metabolite

carriers, the import of some other IM proteins containing multiple transmembrane segments,

such as members of the Tim23 family, are dependent upon the TIM22 complex.

Although some IMS precursors are imported by lateral sorting using the TIM23 complex,

there is also a redox-regulated import of small IMS proteins. These proteins have characteristic

Cysteine motifs to form intramolecular disulfide bonds which use the import machinery for

catalytic formation. The central component of this pathway is Mia40, an oxidoreductase that

catalyzes disulfide bond formation 57, 58 . Other factors are also involved in this pathway including

the sulfhydryl oxidase Erv1, cytochrome c, and the respiratory chain to form a disulfide relay 59 .

Precursors are initially imported through the Tom40 channel in a reduced form 60 but utilize none of the Tom receptors for recognition 61 . Instead, an oxidized form of Mia40 recognizes a

hydrophobic signal with a Cysteine residue and forms intermolecular disulfide bonds with the

substrate 62-64 . The precursors are then released in an oxidized form with intact disulfide bonds,

representing their mature form. The now reduced form of Mia40 shuttles the electrons gained

from the substrate to Erv1, which subsequently transfers them to either cytochrome c and further to the respiratory chain or transfers them directly to molecular oxygen 65, 66 . Recent evidence has also demonstrated that the MIA pathway aids import for precursors destined to other mitochondrial compartments, including IM proteins Tim17 and Tim22 67, 68 and the matrix

protein Mrp10 69 .

Finally, precursors destined for the OM belong in one of two types, either β-barrel or α-

helical proteins, which are imported by distinct pathways. Although there are only a limited

number of β-barrel proteins in the OM, well-known proteins such as Tom40 and porin fall into this category. These precursors are initially targeted to the TOM complex by a signal that

6 contains a β-hairpin element 70 and translocated through the Tom40 channel, where they are bound by the small TIM chaperone complexes 71 , similar to the carrier proteins of the IM. The precursor then uses the SAM complex in the OM for insertion as a β-barrel, a process that has

been conserved from bacteria. This complex contains Sam50 and two peripheral membrane

proteins Sam35 and Sam37 72 . Although the mechanism of β-barrel insertion is not entirely established, it is known that precursors contain C-terminal signals that bind to Sam50-Sam35 to guide their membrane insertion 73 . SAM37 functions to transiently tether the SAM complex to the

Tom complex by an interaction with the cytosolic domain of Tom22 74 . This TOM-SAM

supercomplex improves the efficiency of β-barrel precursor transfer from TOM to SAM. For formation of the TOM complex specifically, an additional subunit of the SAM complex has been identified known as Mdm10, a protein that is also a part of ERMES 75 , suggesting a regulatory

network that links mitochondrial protein trafficking and mitochondria-ER junctions.

There are three classes of α-helical proteins at the OM including signal-anchored, tail- anchored, and polytopic which seem to be inserted through a variety of mechanisms. For the signal-anchored and polytopic classes, the MIM complex, consisting of Mim1 and Mim2, acts as an insertase 76, 77 . The mechanism of the MIM complex remains elusive, but it has been shown

that Tom70 acts as a receptor for transfer to MIM for insertion of polytopic precursors 78 and that proper lipid composition is required for efficient insertion79 . This requisite also applies to tail- anchored precursors, although little else is known about the mechanism of their insertion 80, 81 .

There are possibly many distinct insertion routes and more investigation is needed to tease out

components and mechanisms.

7

It is abundantly clear that mitochondrial biogenesis is a highly organized and regulated process. The many dynamic and oft-interacting complexes that form the translocation machinery are very important for mitochondrial, cellular, and organism health.

Mitochondrial health and neurodegeneration

Because mitochondria perform or take part in many essential processes such as energy production, iron-sulfur cluster formation, calcium homeostasis, and apoptosis, it follows that maintaining their health is vital for cell health. It has long been understood that improperly functioning mitochondria can play a role in metabolic diseases such as diabetes 82 , but recent

evidence now implicates mitochondrial dysfunction in neurodegenerative disease including

Alzheimer’s, Parkinson’s, ALS, and Huntingon’s 83 . Parkinson’s disease (PD), which affects

more than 4 million people worldwide 84 and has no efficacious therapy or prevention, is of

particular interest to those studying protein trafficking at the mitochondria.

As a progressive disease, PD has a mean age at onset of 55 with incidences increasingly

dramatically with age. It has numerous motor symptoms including bradykinesia, dyskinesia,

akinesia, and tremor all of which greatly impair the quality of life, often leading to psychological

symptoms such as depression, all of which is accompanied by memory impairment 85 . PD is also

a direct or indirect cause of death in many who suffer from it. Pathologically, Parkinson’s

disease develops by the gradual loss of nigrostriatal dopaminergic neurons in the substantia nigra

pars compacta (SNpc), a group of cells that normally project to the basal ganglia and synapse of

the striatum to deliver dopamine to dampen neuronal signals. The deficiency in dopamine results

in uncontrolled signaling through the striatum and the subsequent motor symptoms displayed in

disease patients. The presence of α-synuclein positive intraneuronal inclusion bodies, called

“Lewy Bodies,” is also a hallmark of the disease (Figure 1.2)86 .

8

Over 90% of PD cases are considered sporadic, lacking any apparent genetic linkage. But various genetic factors have been identified in autosomal recessive PD (ARPD) including Parkin,

PTEN-induced kinase 1 (PINK1), LRRK2, DJ-1, ATP13A2, FBXO7, DNAJC6, SYNJ1, and

PLA2G6 87 . PINK1 and Parkin, in particular, have been the focus of many studies in recent years.

Over two decades ago, the PARK2 , which codes for Parkin, was linked to ARPD in

Japanese families 88, 89 . It was later discovered that Parkin is an E3 ubiquitin ligase containing an

N-terminal ubiquitin-like domain and a C-terminal ubiquitin ligase domain 90 . Soon after, the

PARK6 gene, which codes for PINK1, was identified as having identical clinical symptoms to

those patients with ARPD when mutated 91 . Sequence analysis revealed that PINK1 contains a kinase domain at its C-terminus and a MTS at its N-terminus, confirming that it is present at mitochondria 10 . These proteins normally work together in a common pathway to govern mitochondrial quality control in a mitochondrial-specific form of autophagy known as mitophagy.

PINK1 is constitutively target to mitochondria, where it is partially imported via the

TOM complex, cleaved by proteases MPP and PARL, and subsequently degraded by the proteasome 92-95 . Upon insult, the mitochondria maintain PINK1 on the OM where its kinase activity actively recruits and activates Parkin 96, 97 . Parkin has a dual role at mitochondria where it ubiquitinates OM proteins, such as MFN-1 and TOMM20, for degradation and recruits autophagy machinery for degradation of the entire mitochondria 98, 99 (Figure 1.3). This pathway

is capable of removing small pieces of mitochondria100 which may indicate that damaged proteins are aggregated to one section of mitochondria and explain why mitochondrial fission is tightly connected to PINK1/Parkin-activated mitophagy 101 . When this process is functioning

normally, it maintains mitochondrial fidelity by selectively removing only the damaged

9 organelles from the total population. But loss of function mutations in PINK1 or Parkin lead to the accumulation of damaged mitochondria, a contributing factor in Parkinson’s disease 102 . It has

been difficult to connect PINK1/Parkin-dependent mitophagy to neuronal death in vivo , although

it has been demonstrated that Parkin protects dopaminergic neurons against mitochondrial

dysfunction and therefore protecting the neurons from degradation, the connection to mitophagy

is still unclear 103 . There is evidence to suggest that the failure to remove respiratory complex I when damaged may be a link between impaired mitophagy and PD 104 (Figure 1.4).

While many studies have teased out the mechanism of PINK1/Parkin-dependent

mitophagy, little is known about PINK1 trafficking at the mitochondria. As previously

mentioned, it is known that PINK1 is initially imported via the TOM complex and its

maintenance on the OM is dependent on TOMM7. It has also been established that the N-

terminal MTS is cleaved by MPP, followed by an additional cleavage by PARL in the IM to

release the protein for degradation. A recent study using a small molecule probe that inhibits the

TMI23 complex, presumably by binding to Tim17, has suggested that PINK1 does not use the

TIM23 complex for import into mitochondria, but does require it for accumulation on the OM 105 .

Further investigation is needed to understand the mechanism by which PINK1 translocates into mitochondria and the mechanism by which it is maintained at mitochondria.

A small molecule screen to discover mitochondrial protein transport inhibitors

The current research methods available to life scientists are quite vast. But because research reveals an ever-evolving wealth of knowledge, there is always the need for additional modes of investigation and tools to further that knowledge. Traditionally, yeast are studied by genetic modification, whether by knocking out , overexpressing genes, genetic fusions, tagging genes with markers, or by some other genetic alteration 106 . Once the genetic makeup of

10 an organism is altered, it can be tested for a detectable phenotype. But these approaches have limitations and are reaching the limit of what they are capable of revealing. For instance, knocking out one gene can alter the protein expression of a second in what are known as pleiotropic effects 107 , or knocking down a protein that exists in a complex may severely hinder complex formation 108 . It has also been quite challenging to utilize genetic approaches outside of simple organisms like yeast, such as in vertebrate cell models or whole organism studies 109 . New methods of study are needed to overcome these obstacles.

As research tools, small molecule probes have many advantages over genetic approaches.

Small molecules are able have an immediate effect with only brief incubation times as opposed to many other techniques such as RNA interference (RNAi), which typically takes at least 48 hours to have a useful effect 110 . Because of these shorter timelines, experiments are easier to plan and execute, and the cell does not have a chance to compensate for the experimental manipulations. Another advantage over RNAi and other genetic approaches is that small molecules do not remove the actual protein structure, but instead remove the functionality of the protein. This is important because it should allow protein complexes to still form normally and it will limit the pleiotropic effects often seen with genetic approaches. Small molecules expand what studies can be performed because they are often capable of being washed away, allowing the target protein to return to its normal function if one wanted to perform a rescue experiment.

Finally, small molecules can be developed into therapeutics for disease, so have a significant role outside of basic research.

There are also disadvantages to using small molecules as research tools. Off-target effects can be a concern, particularly as dosage increases. But perhaps the largest obstacle to using molecular probes is determining the protein target. This is particularly true because many

11 molecules are discovered in phenotypic-based screens 111 , in which molecules are discovered based on inducing a desired phenotype, with the target being irrelevant in screen design. The target must be determined after the screen in secondary assays, which can often be very challenging. A second method of screening is a target-based screen, in which the target of the molecule is known, but how effective it is as a research tool is unknown and its potential as a therapy is severely limited 112 . Even with these drawbacks, small molecules are increasingly

being utilized in research labs because of their many advantages, but mitochondria-directed

studies have yet to fully embrace the potential of these probes.

Taking this need for new approaches into account, we have performed a small molecule

screen to discover novel inhibitors for mitochondrial protein import. The screening design is

similar to a prior screen performed to discover components of mitochondrial protein transport by

identifying mutations that caused defective import and conferred a growth phenotype 113 . We

took advantage of the yeast uracil biosynthetic pathway which relies on the URA3 protein,

among others, to synthesize the nucleotide. URA3 is typically localized to the cytosol and

confers growth to yeast even in conditions where exogenous uracil is lacking. We made a

chimeric protein by fusing a MTS that uses the TOM and TIM23 complexes for import to the N-

terminus of the Ura3 gene, localizing it to the cytosol but preserving its functionality. This yeast

strain is no longer able to grow in media lacking uracil because URA3 must be in the cytosol to

take part in the biosynthetic pathway. By hindering mitochondrial protein import, though, the

URA3 protein will once again accumulate in the cytosol and restore growth in media lacking

uracil (Figure 1.5).

Using this strategy, we screened for small molecule inhibitors that conferred growth of

this yeast strain in media lacking uracil. In collaboration with the Molecular Screening Shared

12

Resource (MSSR) at UCLA’s California NanoSystems Institute (CNSI), we tested 68,000 compounds from various libraries using their robotics and sample handling systems. Because of screening limitations, we were only able to test each molecule at 10 M and test for one 24 hour

endpoint. From this screen, we obtained 64 hit compounds that reproducibly conferred growth in

media lacking uracil that have become a part of the MitoBloCK (MB) series being developed in

the Koehler lab. Relevant hit molecules include the FDA-approved dequalinium chloride

(DECA; MB-12) 114, 115 from the Prestwick library as well as the previously undiscovered MB-10

and MB-11 from the ChemBridge library (Table 1.1). Characterizing these compounds and

applying them in mitochondrial-associated disease research, particularly neurodegeneration,

became the topic of my dissertation.

Concluding remarks

Due to the relevance in many diseases, mitochondria should begin to gain appreciation as

an attractive therapeutic target. Additional research is needed to understand the underlying

mechanism of many of these conditions to progress potential remedies. The development of

small molecule probes to dissect mitochondrial function and biogenesis will yield fruitful

discoveries that can contribute human health. These new tools can also be combined with other

methods, such as genome engineering, for a multifaceted approach that should lend itself to new

opportunities in research. Furthermore, hit molecules from high throughput screens may be

developed as drugs to treat disease. Although the screen design needs to be optimized for this

type of end goal, it has been an adopted strategy for industry researchers for many years. In

short, small molecules offer possibilities as both research tools and as eventual therapies, and

therefore their discovery and characterization is vitally useful. Taking advantage of this, we have

performed a small molecule screen to identify inhibitors of protein import into mitochondria and

13 characterized hit compounds. From there, we have used these compounds to explore the trafficking of the PD-related protein PINK1 to find surprising roles for translocation components. We hope that these molecules will be adopted as tools by other investigators in their studies, and that our discoveries related to PINK1 import into mitochondria will help push the field forward.

14

Figures

Figure 1.1116 . Overview of protein trafficking at mitochondria. Most mitochondrial precursor proteins enter the mitochondria via the TOM complex. From there, the pathways are divergent depending on the final destination and type of targeting signal of the precursor. Pathways include

TIM23 for matrix- and some IM- or IMS-bound precursors, TIM22 for metabolite carries in the

IM, MIA for small soluble IMS proteins, SAM for β-barrel proteins in the OM, and MIM for many α-helical proteins in the OM.

15

Figure 1.286 . Pathological hallmarks of Parkinson’s disease. (A, B) The nigrostriatal pathway is composed of dopaminergic neurons within the SNpc that extend to the basal ganglia (putamen and caudate). Normal pigmentation is shown against a PD patient. (C) Immunohistochemical labeling of intraneuronal Lewy bodies in a SNpc dopaminergic neuron. The α-synuclein stain is

much more intense relative to the ubiquitin stain.

16

Figure 1.3102 . A model of PINK1/Parkin-dependent mitophagy. Damaged mitochondria are selectively removed from the total population by a pathway initiated by PINK1 accumulation,

Parkin recruitment and activation, autophagy initiation, and finally lysosomal degradation.

17

Figure 1.4117 . Complex I may be an important link between mitophagy and Parkinson’s disease. There are various causes of mitochondrial dysfunction, including loss of function mutations in PINK1 or Parkin, but the connection to neurodegeneration remains elusive.

Complex I deficiency may play an important role in linking the two.

18

Figure 1.5. Growth phenotype strategy for small molecule screen. Normal yeast have the

URA3 protein localized in the cytosol where it is active in the uracil biosynthetic pathway, conferring growth in media lacking uracil. When fused with a MTS (Su9), the URA3 is imported into the mitochondrion, and although the URA3 is still functional it is unable to perform its role in the uracil biosynthetic pathway, so the yeast can no longer grow with exogenous uracil. If the mitochondrial protein import is blocked, such as by a small molecule inhibitor, the URA3 is once again localized to the cytosol where it can rescue the growth in media lacking uracil. In this way, small molecules that inhibit mitochondrial protein import can be identified if they confer growth in media lacking uracil.

19

# of compounds # of verified Compounds Library Abbreviation (1,000s) hits of interest

Diverse EAM 20 11 -

Targeted TAR 8 4 -

ChemBridge ChemBridge 30 6 MB-10, MB-11 diverse set E

Lead-like DL 20 28 -

Smart SS 7 - 40 Diverse/Smart ES 5 -

Prestwick PW 6 DECA 2 Biomol Biomol 1 -

Total 120 68 -

Table 1.1. Summary of the libraries from the small molecule screen. In total, 68,000 molecules were tested from 8 different libraries. Each library conferred at least 1 hit, and 68 verified hits were identified across all libraries. Of particular interest to my dissertation are two compounds from the ChemBridge set and one compound from the Prestwick library.

20

References

1. Staff, S. N., A grim budget day for U.S. science:analysis and reaction to Trump's plan. Science: 2017.

2. Martin, W.; Mentel, M., The Origin of Mitochondria. Nature Education 2010; Vol. 3, p 58.

3. Davey, K. M.; Parboosingh, J. S.; McLeod, D. R.; Chan, A.; Casey, R.; Ferreira, P.; Snyder, F. F.; Bridge, P. J.; Bernier, F. P., Mutation of DNAJC19, a human homologue of yeast inner mitochondrial membrane co-chaperones, causes DCMA syndrome, a novel autosomal recessive Barth syndrome-like condition. J Med Genet 2006, 43 (5), 385-93.

4. Koehler, C. M.; Leuenberger, D.; Merchant, S.; Renold, A.; Junne, T.; Schatz, G., Human deafness dystonia syndrome is a mitochondrial disease. Proc Natl Acad Sci U S A 1999, 96 (5), 2141-6.

5. Hansen, J. J.; Dürr, A.; Cournu-Rebeix, I.; Georgopoulos, C.; Ang, D.; Nielsen, M. N.; Davoine, C. S.; Brice, A.; Fontaine, B.; Gregersen, N.; Bross, P., Hereditary spastic paraplegia SPG13 is associated with a mutation in the gene encoding the mitochondrial chaperonin Hsp60. Am J Hum Genet 2002, 70 (5), 1328-32.

6. Devi, L.; Prabhu, B. M.; Galati, D. F.; Avadhani, N. G.; Anandatheerthavarada, H. K., Accumulation of amyloid precursor protein in the mitochondrial import channels of human Alzheimer's disease brain is associated with mitochondrial dysfunction. J Neurosci 2006, 26 (35), 9057-68.

7. Zanna, C.; Ghelli, A.; Porcelli, A. M.; Karbowski, M.; Youle, R. J.; Schimpf, S.; Wissinger, B.; Pinti, M.; Cossarizza, A.; Vidoni, S.; Valentino, M. L.; Rugolo, M.; Carelli, V., OPA1 mutations associated with dominant optic atrophy impair oxidative phosphorylation and mitochondrial fusion. Brain 2008, 131 (Pt 2), 352-67.

8. Quirós, P. M.; Ramsay, A. J.; Sala, D.; Fernández-Vizarra, E.; Rodríguez, F.; Peinado, J. R.; Fernández-García, M. S.; Vega, J. A.; Enríquez, J. A.; Zorzano, A.; López-Otín, C., Loss of mitochondrial protease OMA1 alters processing of the GTPase OPA1 and causes obesity and defective thermogenesis in mice. EMBO J 2012, 31 (9), 2117-33.

9. O'Toole, J. F.; Liu, Y.; Davis, E. E.; Westlake, C. J.; Attanasio, M.; Otto, E. A.; Seelow, D.; Nurnberg, G.; Becker, C.; Nuutinen, M.; Kärppä, M.; Ignatius, J.; Uusimaa, J.; Pakanen, S.; Jaakkola, E.; van den Heuvel, L. P.; Fehrenbach, H.; Wiggins, R.; Goyal, M.; Zhou, W.; Wolf, M. T.; Wise, E.; Helou, J.; Allen, S. J.; Murga-Zamalloa, C. A.; Ashraf, S.; Chaki, M.; Heeringa, S.; Chernin, G.; Hoskins, B. E.; Chaib, H.; Gleeson, J.; Kusakabe, T.; Suzuki, T.; Isaac, R. E.; Quarmby, L. M.; Tennant, B.; Fujioka, H.; Tuominen, H.; Hassinen, I.; Lohi, H.; van Houten, J. L.; Rotig, A.; Sayer, J. A.; Rolinski, B.; Freisinger, P.; Madhavan, S. M.; Herzer, M.; Madignier, F.; Prokisch, H.; Nurnberg, P.; Jackson, P. K.; Jackson, P.; Khanna, H.; Katsanis, N.;

21

Hildebrandt, F., Individuals with mutations in XPNPEP3, which encodes a mitochondrial protein, develop a nephronophthisis-like nephropathy. J Clin Invest 2010, 120 (3), 791-802.

10. Valente, E. M.; Abou-Sleiman, P. M.; Caputo, V.; Muqit, M. M.; Harvey, K.; Gispert, S.; Ali, Z.; Del Turco, D.; Bentivoglio, A. R.; Healy, D. G.; Albanese, A.; Nussbaum, R.; González- Maldonado, R.; Deller, T.; Salvi, S.; Cortelli, P.; Gilks, W. P.; Latchman, D. S.; Harvey, R. J.; Dallapiccola, B.; Auburger, G.; Wood, N. W., Hereditary early-onset Parkinson's disease caused by mutations in PINK1. Science 2004, 304 (5674), 1158-60.

11. Di Maio, R.; Barrett, P. J.; Hoffman, E. K.; Barrett, C. W.; Zharikov, A.; Borah, A.; Hu, X.; McCoy, J.; Chu, C. T.; Burton, E. A.; Hastings, T. G.; Greenamyre, J. T., α-Synuclein binds to TOM20 and inhibits mitochondrial protein import in Parkinson's disease. Sci Transl Med 2016, 8 (342), 342ra78.

12. Chacinska, A.; Koehler, C. M.; Milenkovic, D.; Lithgow, T.; Pfanner, N., Importing mitochondrial proteins: machineries and mechanisms. Cell 2009, 138 (4), 628-44. 13. Neupert, W.; Herrmann, J. M., Translocation of proteins into mitochondria. Annu Rev Biochem 2007, 76 , 723-49.

14. Kiebler, M.; Pfaller, R.; Söllner, T.; Griffiths, G.; Horstmann, H.; Pfanner, N.; Neupert, W., Identification of a mitochondrial receptor complex required for recognition and membrane insertion of precursor proteins. Nature 1990, 348 (6302), 610-6.

15. Hill, K.; Model, K.; Ryan, M. T.; Dietmeier, K.; Martin, F.; Wagner, R.; Pfanner, N., Tom40 forms the hydrophilic channel of the mitochondrial import pore for preproteins [see comment]. Nature 1998, 395 (6701), 516-21.

16. Bausewein, T.; Mills, D. J.; Langer, J. D.; Nitschke, B.; Nussberger, S.; Kühlbrandt, W., Cryo-EM Structure of the TOM Core Complex from Neurospora crassa. Cell 2017, 170 (4), 693- 700.e7.

17. Saitoh, T.; Igura, M.; Obita, T.; Ose, T.; Kojima, R.; Maenaka, K.; Endo, T.; Kohda, D., Tom20 recognizes mitochondrial presequences through dynamic equilibrium among multiple bound states. EMBO J 2007, 26 (22), 4777-87.

18. Steger, H. F.; Söllner, T.; Kiebler, M.; Dietmeier, K. A.; Pfaller, R.; Trülzsch, K. S.; Tropschug, M.; Neupert, W.; Pfanner, N., Import of ADP/ATP carrier into mitochondria: two receptors act in parallel. J Cell Biol 1990, 111 (6 Pt 1), 2353-63.

19. van Wilpe, S.; Ryan, M. T.; Hill, K.; Maarse, A. C.; Meisinger, C.; Brix, J.; Dekker, P. J.; Moczko, M.; Wagner, R.; Meijer, M.; Guiard, B.; Hönlinger, A.; Pfanner, N., Tom22 is a multifunctional organizer of the mitochondrial preprotein translocase. Nature 1999, 401 (6752), 485-9.

22

20. Yamano, K.; Yatsukawa, Y.; Esaki, M.; Hobbs, A. E.; Jensen, R. E.; Endo, T., Tom20 and Tom22 share the common signal recognition pathway in mitochondrial protein import. J Biol Chem 2008, 283 (7), 3799-807.

21. Shiota, T.; Imai, K.; Qiu, J.; Hewitt, V. L.; Tan, K.; Shen, H. H.; Sakiyama, N.; Fukasawa, Y.; Hayat, S.; Kamiya, M.; Elofsson, A.; Tomii, K.; Horton, P.; Wiedemann, N.; Pfanner, N.; Lithgow, T.; Endo, T., Molecular architecture of the active mitochondrial protein gate. Science 2015, 349 (6255), 1544-8.

22. Moczko, M.; Bömer, U.; Kübrich, M.; Zufall, N.; Hönlinger, A.; Pfanner, N., The intermembrane space domain of mitochondrial Tom22 functions as a trans binding site for preproteins with N-terminal targeting sequences. Mol Cell Biol 1997, 17 (11), 6574-84.

23. Model, K.; Meisinger, C.; Prinz, T.; Wiedemann, N.; Truscott, K. N.; Pfanner, N.; Ryan, M. T., Multistep assembly of the protein import channel of the mitochondrial outer membrane. Nat Struct Biol 2001, 8 (4), 361-70.

24. Dietmeier, K.; Hönlinger, A.; Bömer, U.; Dekker, P. J.; Eckerskorn, C.; Lottspeich, F.; Kübrich, M.; Pfanner, N., Tom5 functionally links mitochondrial preprotein receptors to the general import pore. Nature 1997, 388 (6638), 195-200.

25. Schulz, C.; Schendzielorz, A.; Rehling, P., Unlocking the presequence import pathway. Trends Cell Biol 2015, 25 (5), 265-75.

26. Moberg, P.; Nilsson, S.; Ståhl, A.; Eriksson, A. C.; Glaser, E.; Mäler, L., NMR solution structure of the mitochondrial F1beta presequence from Nicotiana plumbaginifolia. J Mol Biol 2004, 336 (5), 1129-40.

27. van der Laan, M.; Schrempp, S. G.; Pfanner, N., Voltage-coupled conformational dynamics of mitochondrial protein-import channel. Nat Struct Mol Biol 2013, 20 (8), 915-7. 28. Rahman, B.; Kawano, S.; Yunoki-Esaki, K.; Anzai, T.; Endo, T., NMR analyses on the interactions of the yeast Tim50 C-terminal region with the presequence and Tim50 core domain. FEBS Lett 2014, 588 (5), 678-84.

29. Meinecke, M.; Wagner, R.; Kovermann, P.; Guiard, B.; Mick, D. U.; Hutu, D. P.; Voos, W.; Truscott, K. N.; Chacinska, A.; Pfanner, N.; Rehling, P., Tim50 maintains the permeability barrier of the mitochondrial inner membrane. Science 2006, 312 (5779), 1523-6.

30. Martinez-Caballero, S.; Grigoriev, S. M.; Herrmann, J. M.; Campo, M. L.; Kinnally, K. W., Tim17p regulates the twin pore structure and voltage gating of the mitochondrial protein import complex TIM23. J Biol Chem 2007, 282 (6), 3584-93.

31. Martin, J.; Mahlke, K.; Pfanner, N., Role of an energized inner membrane in mitochondrial protein import. Delta psi drives the movement of presequences. J Biol Chem 1991, 266 (27), 18051-7.

23

32. Truscott, K. N.; Kovermann, P.; Geissler, A.; Merlin, A.; Meijer, M.; Driessen, A. J.; Rassow, J.; Pfanner, N.; Wagner, R., A presequence- and voltage-sensitive channel of the mitochondrial preprotein translocase formed by Tim23. Nat Struct Biol 2001, 8 (12), 1074-82.

33. Wiedemann, N.; van der Laan, M.; Hutu, D. P.; Rehling, P.; Pfanner, N., Sorting switch of mitochondrial presequence translocase involves coupling of motor module to respiratory chain. J Cell Biol 2007, 179 (6), 1115-22.

34. Glick, B. S.; Brandt, A.; Cunningham, K.; Müller, S.; Hallberg, R. L.; Schatz, G., Cytochromes c1 and b2 are sorted to the intermembrane space of yeast mitochondria by a stop- transfer mechanism. Cell 1992, 69 (5), 809-22.

35. Botelho, S. C.; Osterberg, M.; Reichert, A. S.; Yamano, K.; Björkholm, P.; Endo, T.; von Heijne, G.; Kim, H., TIM23-mediated insertion of transmembrane α-helices into the mitochondrial inner membrane. EMBO J 2011, 30 (6), 1003-11.

36. Ieva, R.; Schrempp, S. G.; Opali ński, L.; Wollweber, F.; Höß, P.; Heißwolf, A. K.; Gebert, M.; Zhang, Y.; Guiard, B.; Rospert, S.; Becker, T.; Chacinska, A.; Pfanner, N.; van der Laan, M., Mgr2 functions as lateral gatekeeper for preprotein sorting in the mitochondrial inner membrane. Mol Cell 2014, 56 (5), 641-52.

37. Chacinska, A.; Lind, M.; Frazier, A. E.; Dudek, J.; Meisinger, C.; Geissler, A.; Sickmann, A.; Meyer, H. E.; Truscott, K. N.; Guiard, B.; Pfanner, N.; Rehling, P., Mitochondrial presequence translocase: switching between TOM tethering and motor recruitment involves Tim21 and Tim17. Cell 2005, 120 (6), 817-29.

38. Hawlitschek, G.; Schneider, H.; Schmidt, B.; Tropschug, M.; Hartl, F. U.; Neupert, W., Mitochondrial protein import: identification of processing peptidase and of PEP, a processing enhancing protein. Cell 1988, 53 (5), 795-806.

39. Mossmann, D.; Meisinger, C.; Vögtle, F. N., Processing of mitochondrial presequences. Biochim Biophys Acta 2012, 1819 (9-10), 1098-106.

40. Johnson, K. A.; Bhushan, S.; Ståhl, A.; Hallberg, B. M.; Frohn, A.; Glaser, E.; Eneqvist, T., The closed structure of presequence protease PreP forms a unique 10,000 Angstroms3 chamber for proteolysis. EMBO J 2006, 25 (9), 1977-86.

41. Krayl, M.; Lim, J. H.; Martin, F.; Guiard, B.; Voos, W., A cooperative action of the ATP- dependent import motor complex and the inner membrane potential drives mitochondrial preprotein import. Mol Cell Biol 2007, 27 (2), 411-25.

42. Slutsky-Leiderman, O.; Marom, M.; Iosefson, O.; Levy, R.; Maoz, S.; Azem, A., The interplay between components of the mitochondrial protein translocation motor studied using purified components. J Biol Chem 2007, 282 (47), 33935-42.

24

43. Banerjee, R.; Gladkova, C.; Mapa, K.; Witte, G.; Mokranjac, D., Protein translocation channel of mitochondrial inner membrane and matrix-exposed import motor communicate via two-domain coupling protein. Elife 2015, 4, e11897.

44. Kozany, C.; Mokranjac, D.; Sichting, M.; Neupert, W.; Hell, K., The J domain-related cochaperone Tim16 is a constituent of the mitochondrial TIM23 preprotein translocase. Nat Struct Mol Biol 2004, 11 (3), 234-41.

45. Mokranjac, D.; Bourenkov, G.; Hell, K.; Neupert, W.; Groll, M., Structure and function of Tim14 and Tim16, the J and J-like components of the mitochondrial protein import motor. EMBO J 2006, 25 (19), 4675-85.

46. Horst, M.; Oppliger, W.; Rospert, S.; Schönfeld, H. J.; Schatz, G.; Azem, A., Sequential action of two hsp70 complexes during protein import into mitochondria. EMBO J 1997, 16 (8), 1842-9.

47. Ting, S. Y.; Schilke, B. A.; Hayashi, M.; Craig, E. A., Architecture of the TIM23 inner mitochondrial translocon and interactions with the matrix import motor. J Biol Chem 2014, 289 (41), 28689-96.

48. Hutu, D. P.; Guiard, B.; Chacinska, A.; Becker, D.; Pfanner, N.; Rehling, P.; van der Laan, M., Mitochondrial protein import motor: differential role of Tim44 in the recruitment of Pam17 and J-complex to the presequence translocase. Mol Biol Cell 2008, 19 (6), 2642-9.

49. Brix, J.; Rüdiger, S.; Bukau, B.; Schneider-Mergener, J.; Pfanner, N., Distribution of binding sequences for the mitochondrial import receptors Tom20, Tom22, and Tom70 in a presequence-carrying preprotein and a non-cleavable preprotein. J Biol Chem 1999, 274 (23), 16522-30.

50. Wiedemann, N.; Pfanner, N.; Ryan, M. T., The three modules of ADP/ATP carrier cooperate in receptor recruitment and translocation into mitochondria. EMBO J 2001, 20 (5), 951-60.

51. Wu, Y.; Sha, B., Crystal structure of yeast mitochondrial outer membrane translocon member Tom70p. Nat Struct Mol Biol 2006, 13 (7), 589-93.

52. Curran, S. P.; Leuenberger, D.; Schmidt, E.; Koehler, C. M., The role of the Tim8p- Tim13p complex in a conserved import pathway for mitochondrial polytopic inner membrane proteins. J Cell Biol 2002, 158 (6), 1017-27.

53. Curran, S. P.; Leuenberger, D.; Oppliger, W.; Koehler, C. M., The Tim9p-Tim10p complex binds to the transmembrane domains of the ADP/ATP carrier. EMBO J 2002, 21 (5), 942-53.

25

54. Gebert, N.; Chacinska, A.; Wagner, K.; Guiard, B.; Koehler, C. M.; Rehling, P.; Pfanner, N.; Wiedemann, N., Assembly of the three small Tim proteins precedes docking to the mitochondrial carrier translocase. EMBO Rep 2008, 9 (6), 548-54.

55. Wagner, K.; Gebert, N.; Guiard, B.; Brandner, K.; Truscott, K. N.; Wiedemann, N.; Pfanner, N.; Rehling, P., The assembly pathway of the mitochondrial carrier translocase involves four preprotein translocases. Mol Cell Biol 2008, 28 (13), 4251-60.

56. Rehling, P.; Model, K.; Brandner, K.; Kovermann, P.; Sickmann, A.; Meyer, H. E.; Kühlbrandt, W.; Wagner, R.; Truscott, K. N.; Pfanner, N., Protein insertion into the mitochondrial inner membrane by a twin-pore translocase. Science 2003, 299 (5613), 1747-51.

57. Chacinska, A.; Pfannschmidt, S.; Wiedemann, N.; Kozjak, V.; Sanjuán Szklarz, L. K.; Schulze-Specking, A.; Truscott, K. N.; Guiard, B.; Meisinger, C.; Pfanner, N., Essential role of Mia40 in import and assembly of mitochondrial intermembrane space proteins. EMBO J 2004, 23 (19), 3735-46.

58. Banci, L.; Bertini, I.; Cefaro, C.; Ciofi-Baffoni, S.; Gallo, A.; Martinelli, M.; Sideris, D. P.; Katrakili, N.; Tokatlidis, K., MIA40 is an oxidoreductase that catalyzes oxidative protein folding in mitochondria. Nat Struct Mol Biol 2009, 16 (2), 198-206.

59. Mesecke, N.; Terziyska, N.; Kozany, C.; Baumann, F.; Neupert, W.; Hell, K.; Herrmann, J. M., A disulfide relay system in the intermembrane space of mitochondria that mediates protein import. Cell 2005, 121 (7), 1059-69.

60. Lu, H.; Allen, S.; Wardleworth, L.; Savory, P.; Tokatlidis, K., Functional TIM10 chaperone assembly is redox-regulated in vivo. J Biol Chem 2004, 279 (18), 18952-8.

61. Gornicka, A.; Bragoszewski, P.; Chroscicki, P.; Wenz, L. S.; Schulz, C.; Rehling, P.; Chacinska, A., A discrete pathway for the transfer of intermembrane space proteins across the outer membrane of mitochondria. Mol Biol Cell 2014, 25 (25), 3999-4009.

62. Milenkovic, D.; Gabriel, K.; Guiard, B.; Schulze-Specking, A.; Pfanner, N.; Chacinska, A., Biogenesis of the essential Tim9-Tim10 chaperone complex of mitochondria: site-specific recognition of cysteine residues by the intermembrane space receptor Mia40. J Biol Chem 2007, 282 (31), 22472-80.

63. Peleh, V.; Cordat, E.; Herrmann, J. M., Mia40 is a trans-site receptor that drives protein import into the mitochondrial intermembrane space by hydrophobic substrate binding. Elife 2016, 5.

64. Sideris, D. P.; Petrakis, N.; Katrakili, N.; Mikropoulou, D.; Gallo, A.; Ciofi-Baffoni, S.; Banci, L.; Bertini, I.; Tokatlidis, K., A novel intermembrane space-targeting signal docks cysteines onto Mia40 during mitochondrial oxidative folding. J Cell Biol 2009, 187 (7), 1007-22.

26

65. Dabir, D. V.; Leverich, E. P.; Kim, S. K.; Tsai, F. D.; Hirasawa, M.; Knaff, D. B.; Koehler, C. M., A role for cytochrome c and cytochrome c peroxidase in electron shuttling from Erv1. EMBO J 2007, 26 (23), 4801-11.

66. Bien, M.; Longen, S.; Wagener, N.; Chwalla, I.; Herrmann, J. M.; Riemer, J., Mitochondrial disulfide bond formation is driven by intersubunit electron transfer in Erv1 and proofread by glutathione. Mol Cell 2010, 37 (4), 516-28.

67. Wrobel, L.; Sokol, A. M.; Chojnacka, M.; Chacinska, A., The presence of disulfide bonds reveals an evolutionarily conserved mechanism involved in mitochondrial protein translocase assembly. Sci Rep 2016, 6, 27484.

68. Wrobel, L.; Trojanowska, A.; Sztolsztener, M. E.; Chacinska, A., Mitochondrial protein import: Mia40 facilitates Tim22 translocation into the inner membrane of mitochondria. Mol Biol Cell 2013, 24 (5), 543-54.

69. Longen, S.; Woellhaf, M. W.; Petrungaro, C.; Riemer, J.; Herrmann, J. M., The disulfide relay of the intermembrane space oxidizes the ribosomal subunit mrp10 on its transit into the mitochondrial matrix. Dev Cell 2014, 28 (1), 30-42.

70. Jores, T.; Klinger, A.; Groß, L. E.; Kawano, S.; Flinner, N.; Duchardt-Ferner, E.; Wöhnert, J.; Kalbacher, H.; Endo, T.; Schleiff, E.; Rapaport, D., Characterization of the targeting signal in mitochondrial β-barrel proteins. Nat Commun 2016, 7, 12036.

71. Wiedemann, N.; Truscott, K. N.; Pfannschmidt, S.; Guiard, B.; Meisinger, C.; Pfanner, N., Biogenesis of the protein import channel Tom40 of the mitochondrial outer membrane: intermembrane space components are involved in an early stage of the assembly pathway. J Biol Chem 2004, 279 (18), 18188-94.

72. Wiedemann, N.; Kozjak, V.; Chacinska, A.; Schönfisch, B.; Rospert, S.; Ryan, M. T.; Pfanner, N.; Meisinger, C., Machinery for protein sorting and assembly in the mitochondrial outer membrane. Nature 2003, 424 (6948), 565-71.

73. Kutik, S.; Stojanovski, D.; Becker, L.; Becker, T.; Meinecke, M.; Krüger, V.; Prinz, C.; Meisinger, C.; Guiard, B.; Wagner, R.; Pfanner, N.; Wiedemann, N., Dissecting membrane insertion of mitochondrial beta-barrel proteins. Cell 2008, 132 (6), 1011-24.

74. Wenz, L. S.; Ellenrieder, L.; Qiu, J.; Bohnert, M.; Zufall, N.; van der Laan, M.; Pfanner, N.; Wiedemann, N.; Becker, T., Sam37 is crucial for formation of the mitochondrial TOM-SAM supercomplex, thereby promoting β-barrel biogenesis. J Cell Biol 2015, 210 (7), 1047-54.

75. Meisinger, C.; Rissler, M.; Chacinska, A.; Szklarz, L. K.; Milenkovic, D.; Kozjak, V.; Schönfisch, B.; Lohaus, C.; Meyer, H. E.; Yaffe, M. P.; Guiard, B.; Wiedemann, N.; Pfanner, N., The mitochondrial morphology protein Mdm10 functions in assembly of the preprotein translocase of the outer membrane. Dev Cell 2004, 7 (1), 61-71.

27

76. Becker, T.; Pfannschmidt, S.; Guiard, B.; Stojanovski, D.; Milenkovic, D.; Kutik, S.; Pfanner, N.; Meisinger, C.; Wiedemann, N., Biogenesis of the mitochondrial TOM complex: Mim1 promotes insertion and assembly of signal-anchored receptors. J Biol Chem 2008, 283 (1), 120-7.

77. Dimmer, K. S.; Papi ć, D.; Schumann, B.; Sperl, D.; Krumpe, K.; Walther, D. M.; Rapaport, D., A crucial role for Mim2 in the biogenesis of mitochondrial outer membrane proteins. J Cell Sci 2012, 125 (Pt 14), 3464-73.

78. Becker, T.; Wenz, L. S.; Krüger, V.; Lehmann, W.; Müller, J. M.; Goroncy, L.; Zufall, N.; Lithgow, T.; Guiard, B.; Chacinska, A.; Wagner, R.; Meisinger, C.; Pfanner, N., The mitochondrial import protein Mim1 promotes biogenesis of multispanning outer membrane proteins. J Cell Biol 2011, 194 (3), 387-95.

79. Vögtle, F. N.; Keller, M.; Taskin, A. A.; Horvath, S. E.; Guan, X. L.; Prinz, C.; Opali ńska, M.; Zorzin, C.; van der Laan, M.; Wenk, M. R.; Schubert, R.; Wiedemann, N.; Holzer, M.; Meisinger, C., The fusogenic lipid phosphatidic acid promotes the biogenesis of mitochondrial outer membrane protein Ugo1. J Cell Biol 2015, 210 (6), 951-60.

80. Kemper, C.; Habib, S. J.; Engl, G.; Heckmeyer, P.; Dimmer, K. S.; Rapaport, D., Integration of tail-anchored proteins into the mitochondrial outer membrane does not require any known import components. J Cell Sci 2008, 121 (Pt 12), 1990-8.

81. Krumpe, K.; Frumkin, I.; Herzig, Y.; Rimon, N.; Özbalci, C.; Brügger, B.; Rapaport, D.; Schuldiner, M., Ergosterol content specifies targeting of tail-anchored proteins to mitochondrial outer membranes. Mol Biol Cell 2012, 23 (20), 3927-35.

82. Sivitz, W. I.; Yorek, M. A., Mitochondrial dysfunction in diabetes: from molecular mechanisms to functional significance and therapeutic opportunities. Antioxid Redox Signal 2010, 12 (4), 537-77.

83. Lezi, E.; Swerdlow, R. H., Mitochondria in neurodegeneration. Adv Exp Med Biol 2012, 942 , 269-86.

84. Calabrese, V. P., Projected number of people with Parkinson disease in the most populous nations, 2005 through 2030. Neurology 2007, 69 (2), 223-4; author reply 224.

85. Savitt, J. M.; Dawson, V. L.; Dawson, T. M., Diagnosis and treatment of Parkinson disease: molecules to medicine. J Clin Invest 2006, 116 (7), 1744-54.

86. Dauer, W.; Przedborski, S., Parkinson's disease: mechanisms and models. Neuron 2003, 39 (6), 889-909.

87. Lin, M. T.; Beal, M. F., Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature 2006, 443 (7113), 787-95.

28

88. Matsumine, H.; Saito, M.; Shimoda-Matsubayashi, S.; Tanaka, H.; Ishikawa, A.; Nakagawa-Hattori, Y.; Yokochi, M.; Kobayashi, T.; Igarashi, S.; Takano, H.; Sanpei, K.; Koike, R.; Mori, H.; Kondo, T.; Mizutani, Y.; Schäffer, A. A.; Yamamura, Y.; Nakamura, S.; Kuzuhara, S.; Tsuji, S.; Mizuno, Y., Localization of a gene for an autosomal recessive form of juvenile Parkinsonism to 6q25.2-27. Am J Hum Genet 1997, 60 (3), 588-96.

89. Kitada, T.; Asakawa, S.; Hattori, N.; Matsumine, H.; Yamamura, Y.; Minoshima, S.; Yokochi, M.; Mizuno, Y.; Shimizu, N., Mutations in the parkin gene cause autosomal recessive juvenile parkinsonism. Nature 1998, 392 (6676), 605-8.

90. Shimura, H.; Hattori, N.; Kubo, S.; Mizuno, Y.; Asakawa, S.; Minoshima, S.; Shimizu, N.; Iwai, K.; Chiba, T.; Tanaka, K.; Suzuki, T., Familial Parkinson disease gene product, parkin, is a ubiquitin-protein ligase. Nat Genet 2000, 25 (3), 302-5.

91. Valente, E. M.; Brancati, F.; Caputo, V.; Graham, E. A.; Davis, M. B.; Ferraris, A.; Breteler, M. M.; Gasser, T.; Bonifati, V.; Bentivoglio, A. R.; De Michele, G.; Dürr, A.; Cortelli, P.; Filla, A.; Meco, G.; Oostra, B. A.; Brice, A.; Albanese, A.; Dallapiccola, B.; Wood, N. W.; Disease, E. C. o. G. S. i. P. s., PARK6 is a common cause of familial parkinsonism. Neurol Sci 2002, 23 Suppl 2 , S117-8.

92. Greene, A. W.; Grenier, K.; Aguileta, M. A.; Muise, S.; Farazifard, R.; Haque, M. E.; McBride, H. M.; Park, D. S.; Fon, E. A., Mitochondrial processing peptidase regulates PINK1 processing, import and Parkin recruitment. EMBO Rep 2012, 13 (4), 378-85.

93. Meissner, C.; Lorenz, H.; Weihofen, A.; Selkoe, D. J.; Lemberg, M. K., The mitochondrial intramembrane protease PARL cleaves human Pink1 to regulate Pink1 trafficking. J Neurochem 2011, 117 (5), 856-67.

94. Yamano, K.; Youle, R. J., PINK1 is degraded through the N-end rule pathway. Autophagy 2013, 9 (11), 1758-69.

95. Hasson, S. A.; Kane, L. A.; Yamano, K.; Huang, C. H.; Sliter, D. A.; Buehler, E.; Wang, C.; Heman-Ackah, S. M.; Hessa, T.; Guha, R.; Martin, S. E.; Youle, R. J., High-content genome- wide RNAi screens identify regulators of parkin upstream of mitophagy. Nature 2013, 504 (7479), 291-5.

96. Narendra, D. P.; Jin, S. M.; Tanaka, A.; Suen, D. F.; Gautier, C. A.; Shen, J.; Cookson, M. R.; Youle, R. J., PINK1 is selectively stabilized on impaired mitochondria to activate Parkin. PLoS Biol 2010, 8 (1), e1000298.

97. Kondapalli, C.; Kazlauskaite, A.; Zhang, N.; Woodroof, H. I.; Campbell, D. G.; Gourlay, R.; Burchell, L.; Walden, H.; Macartney, T. J.; Deak, M.; Knebel, A.; Alessi, D. R.; Muqit, M. M., PINK1 is activated by mitochondrial membrane potential depolarization and stimulates Parkin E3 ligase activity by phosphorylating Serine 65. Open Biol 2012, 2 (5), 120080.

29

98. Narendra, D.; Walker, J. E.; Youle, R., Mitochondrial quality control mediated by PINK1 and Parkin: links to parkinsonism. Cold Spring Harb Perspect Biol 2012, 4 (11).

99. Gegg, M. E.; Cooper, J. M.; Chau, K. Y.; Rojo, M.; Schapira, A. H.; Taanman, J. W., Mitofusin 1 and mitofusin 2 are ubiquitinated in a PINK1/parkin-dependent manner upon induction of mitophagy. Hum Mol Genet 2010, 19 (24), 4861-70.

100. McLelland, G. L.; Soubannier, V.; Chen, C. X.; McBride, H. M.; Fon, E. A., Parkin and PINK1 function in a vesicular trafficking pathway regulating mitochondrial quality control. EMBO J 2014, 33 (4), 282-95.

101. Mao, K.; Wang, K.; Liu, X.; Klionsky, D. J., The scaffold protein Atg11 recruits fission machinery to drive selective mitochondria degradation by autophagy. Dev Cell 2013, 26 (1), 9- 18.

102. Pickrell, A. M.; Youle, R. J., The roles of PINK1, parkin, and mitochondrial fidelity in Parkinson's disease. Neuron 2015, 85 (2), 257-73.

103. Pickrell, A. M.; Huang, C. H.; Kennedy, S. R.; Ordureau, A.; Sideris, D. P.; Hoekstra, J. G.; Harper, J. W.; Youle, R. J., Endogenous Parkin Preserves Dopaminergic Substantia Nigral Neurons following Mitochondrial DNA Mutagenic Stress. Neuron 2015, 87 (2), 371-81.

104. Lopez-Fabuel, I.; Martin-Martin, L.; Resch-Beusher, M.; Azkona, G.; Sanchez-Pernaute, R.; Bolaños, J. P., Mitochondrial respiratory chain disorganization in Parkinson's disease- relevant PINK1 and DJ1 mutants. Neurochem Int 2017, 109 , 101-105.

105. Filipuzzi, I.; Steffen, J.; Germain, M.; Goepfert, L.; Conti, M. A.; Potting, C.; Cerino, R.; Pfeifer, M.; Krastel, P.; Hoepfner, D.; Bastien, J.; Koehler, C. M.; Helliwell, S. B., Stendomycin selectively inhibits TIM23-dependent mitochondrial protein import. Nat Chem Biol 2017, 13 (12), 1239-1244.

106. Boone, C.; Bussey, H.; Andrews, B. J., Exploring genetic interactions and networks with yeast. Nat Rev Genet 2007, 8 (6), 437-49.

107. Setoguchi, K.; Otera, H.; Mihara, K., Cytosolic factor- and TOM-independent import of C-tail-anchored mitochondrial outer membrane proteins. EMBO J 2006, 25 (24), 5635-47.

108. Waizenegger, T.; Habib, S. J.; Lech, M.; Mokranjac, D.; Paschen, S. A.; Hell, K.; Neupert, W.; Rapaport, D., Tob38, a novel essential component in the biogenesis of beta-barrel proteins of mitochondria. EMBO Rep 2004, 5 (7), 704-9.

109. Khurana, V.; Tardiff, D. F.; Chung, C. Y.; Lindquist, S., Toward stem cell-based phenotypic screens for neurodegenerative diseases. Nat Rev Neurol 2015, 11 (6), 339-50.

110. Weiss, W. A.; Taylor, S. S.; Shokat, K. M., Recognizing and exploiting differences between RNAi and small-molecule inhibitors. Nat Chem Biol 2007, 3 (12), 739-44. 30

111. Moffat, J. G.; Rudolph, J.; Bailey, D., Phenotypic screening in cancer drug discovery - past, present and future. Nat Rev Drug Discov 2014, 13 (8), 588-602.

112. Eder, J.; Herrling, P. L., Trends in Modern Drug Discovery. Handb Exp Pharmacol 2016, 232 , 3-22.

113. Maarse, A. C.; Blom, J.; Grivell, L. A.; Meijer, M., MPI1, an essential gene encoding a mitochondrial membrane protein, is possibly involved in protein import into yeast mitochondria. EMBO J 1992, 11 (10), 3619-28.

114. Weissenbacher, E. R.; Donders, G.; Unzeitig, V.; Martinez de Tejada, B.; Gerber, S.; Halaška, M.; Špa ček, J.; Group, F. S., A comparison of dequalinium chloride vaginal tablets (Fluomizin®) and clindamycin vaginal cream in the treatment of bacterial vaginosis: a single- blind, randomized clinical trial of efficacy and safety. Gynecol Obstet Invest 2012, 73 (1), 8-15.

115. Lyrawati, D.; Trounson, A.; Cram, D., Expression of GFP in the mitochondrial compartment using DQAsome-mediated delivery of an artificial mini-mitochondrial genome. Pharm Res 2011, 28 (11), 2848-62.

116. Wiedemann, N.; Pfanner, N., Mitochondrial Machineries for Protein Import and Assembly. Annu Rev Biochem 2017, 86 , 685-714.

117. Bose, A.; Beal, M. F., Mitochondrial dysfunction in Parkinson's disease. J Neurochem 2016, 139 Suppl 1 , 216-231.

31

Chapter 2

32

33

34

35

36

37

38

39

40

41

42

43

44

45

SUPPLEMENTAL DATA Adaptation of a genetic screen reveals an inhibitor for mitochondrial protein import component Tim44

Non Miyata, Zhiye Tang, Meghan E. Johnson, Colin J. Douglas, Samuel A. Hasson, Robert Damoiseaux, Chia-en A. Chang, and Carla M. Koehler

TABLE OF CONTENTS

Supplemental Experimental Procedures References Table S1. Strains constructed in this study Table S2. Summary of chemical genetic screen Table S3. Results of the MM/PBSA calculation for MB-10 bound to wild-type Tim44 and Tim44 mutants (F284L, T290S, I297V, and V298A). Figure S1. MB-10 activity is influenced by specific chemical characteristics. Figure S2. Mutations H292A, W316A, W316Y, and W417A in Tim44 confer a dominant negative phenotype.

46

SUPPLEMENTAL EXPERIMENTAL PROCEDURES Computational details for BD, MD simulations and molecular docking . Starting from the crystal structure of yeast Tim44 (PDB code: 2FXT, resolution 3.2 Å), we ran Brownian dynamics to search for potential MB-10 binding sites on the surface of Tim44 using the GeomBD2 package (1,2). The structure of MB-10 was generated using Vega ZZ (3). We used the GeomBD2 scripts to parameterize the ligand and generate parameterized grids for protein using AMBER14SB and GAFF force fields. The program diffused MB-10 around the protein and obtained 26886000 MB-10 conformations that reached the surface of Tim44 for analysis.

We used the AMBER12 package with GPU acceleration to run the MD simulations and perform setup for all MD runs. We used the same structure of the free Tim44 from the protein data bank (PDB) code 2FXT to study the motions of the free protein. We also constructed the MB-10 and Tim44 complex for the wild-type protein and four mutants, F284L, T290S, I297V, and V298A. The AMBER 99SB and GAFF force fields were used for the protein and MB-10, respectively (4,5). VCharge was used to build the partial charges for MB-10 (6). The protein or the complexes were solvated with a box of TIP3P water molecules at least 12 Å away from the solutes, and the systems had ~43400 atoms. The standard procedures for setting up and equilibrating each MD are detailed in a previous publication (7). All production runs were performed for 100 ns at 298 K in an isothermic-isobaric (NPT) ensemble, and the temperature was maintained at 298 K by Langevin dynamics (8). The particle mesh Ewald (PME) was used for the long-range electrostatic interactions. The time step of the simulations was set to 2.0 fs with a non-bonded cutoff of 8.0 Å, and we collected frames every 1 ps.

To evaluate the binding of MB-10 to the wildtype Tim44 and Tim44 mutants, we performed MM/PBSA calculation on the MD trajectories (9). MM/PBSA calculation computes the interaction energy between MB-10 and protein but does not calculate changes of entropy. The solvent effect is estimated by the Poisson Boltzmann implicit solvent model (10). The interaction energy is computed by < ∆E>=--, where , and are the average potential energies of the protein and MB-10 complex, the free protein and MB-10, respectively.

We used AutoDock, a molecular docking program, to place MB-10 in the selected pocket of Tim44 (11). In addition to docking MB-10 to a crystal structure (PDB code 2FTX), we also selected multiple conformations from the MD simulation for wild-type apo Tim44 for our docking processes. Residues 315, 316, 404, 415 and 417 were kept flexible during docking. The Lamarckian genetic algorithm was used to search for ligand binding poses. A total of 20 docking results was generated and analyzed, and the lowest energy conformation was reported. All 3D molecular structures were generated using VMD (12).

Plasmids and strains. In general, a standard set of genetic and molecular techniques was used to generate the strains in this study (Table S1). The screening strains were generated from the parental S. cerevisiae yeast strain GA74 (13). The tim23-2 mutant strain was described previously (14,15). SNQ1 and PDR5 were deleted using a PCR-mediated deletion approach (16,17). Amino acids 1-69 of the targeting sequence from subunit 9 F 1Fo ATPase targeting sequence from N. crassa were fused in frame to Ura3 and a myc tag was placed at the C- 47 terminus of Ura3 (18). The construct was assembled in pRS305 with the glyceraldehyde-3- phosphate dehydrogenase promoter and the phosphoglycerate kinase terminator and then integrated at the LEU2 (19). The TIM44-His10 strain was generated using PCR-based integration at the 3’ end of TIM44 and integration was selected using the HISMX6 selection cassette (20). Media and Reagents. Media reagents were purchased from EMD Biosciences and US Biological. Chemical reagents were from Chembridge and Sigma unless otherwise noted. YPEG medium is 1% Bacto-yeast extract, 2% Bacto-peptone, 3% glycerol, and 3% ethanol. Synthetic dextrose (SD) media contains 0.17% yeast nitrogen base, 0.5% ammonium sulfate, and the appropriate amino acid dropout mixture. HeLa and SH-SY5Y cells were cultured in DMEM and DMEM/F12 (1:1) medium, respectively, supplemented with 10% FBS.

Yeast MIC 50 assays. An early-log growth of the indicated yeast strains in YPEG media was diluted to an initial OD 600 of 0.01 in fresh YPEG media (16). This diluted stock was dispensed in 50 l aliquots into individual wells of a 96-well glass bottom plate. Subsequently, MB-10 in DMSO (or DMSO alone) in YPEG (50 l) was serially diluted by a factor of 2 to the indicated concentration. The final concentration of DMSO was 1%. Plates were then incubated at 25˚C in a humidified chamber for 24 hours. During this time, the OD 600 reached approximately 0.5. Each plate was shaken in a Beckman orbital shaker to resuspend settled cells, and the OD600 in each well was read by a Wallac Victor plate reader (Perkin Elmer). For studies in Figure S1E, the OD 600 of the WT strain in YPEG with 1% DMSO was set as 100% survival. For studies in Figure 4c, the OD 600 of the WT strain in YPEG with 1% DMSO was set as 100% survival and all strains grew at approximately the same rate.

Cell viability assays. Cell viability was measured with an MTT-based toxicology assay kit (Sigma). HeLa cells were grown in 24-well dishes to 80% confluency. Cells were treated with 1% DMSO or various concentrations of MB-10 for 24 hr. After treatment, cells were incubated with MTT solution for additional 4 hr as described in the manufacturer protocols. Purification of Mitochondria. Mitochondria were purified from yeast cells grown in YPEG as described in previous studies (21,22). Yeast cultures were kept at a constant 25°C with vigorous shaking during growth. After the concentration was measured by bicinchoninic acid (BCA) assay, mitochondria were stored in 25 mg/ml aliquots at -80°C. Mammalian mitochondria were isolated from HeLa cells (23). A 6 x 10 cm culture dish of HeLa cells with 100% confluency was harvested and homogenized in 1 ml of homogenization buffer (70 mM sucrose, 220 mM mannitol, 1% BSA, 20 mM Hepes-KOH, pH 7.4) using a glass dounce homogenizer with a Teflon pestle. The homogenate was centrifuged at 770 x g for 5 min to remove the nuclei. The post-nuclear supernatant was centrifuged at 10,000 x g for 10 min. The pellet was washed with the homogenization buffer without BSA and centrifuged again at 10,000 x g for 10 min. Mitochondria were finally resuspended in the homogenization buffer lacking BSA; the mitochondrial concentration was determined using the BCA assay. Mammalian mitochondria were freshly prepared before import assays. Import of radiolabeled proteins into mitochondria and crosslinking assays. Prior to import into purified mitochondria, [ 35 S]-methionine and cysteine labeled precursor proteins were generated with the TNT Quick Coupled Transcription/Translation kit (Promega). Import 48 reactions with yeast mitochondria were conducted according to established methods (22). Because the compounds are typically soluble in DMSO, the final DMSO concentration of 1% was used, unless noted. The compounds in DMSO or DMSO vehicle were added to the mitochondria (25-50 µg/ml) in import buffer and incubated for 15 min incubation at 25°C. Import reactions were then initiated by the addition of precursor. Aliquots were removed at intervals during the reaction time-course and import was terminated with 25 g/ml trypsin on ice. After 15 min, 200 µg/ml soybean trypsin inhibitor was subsequently added. Mitochondria were pelleted by centrifugation at 8,000 x g for 5 min. For import reactions with mammalian mitochondria, isolated mitochondria were added at a final concentration of 200 µg/ml to the import buffer (20 mM Hepes-KOH pH 7.4, 70 mM sucrose, 220 mM mannitol, 0.5 mM MgCl 2, 1 mM ATP, 20 mM sodium succinate, and 5 mM NADH), followed by DMSO or small molecule treatment for 15 min. Import reactions were likewise initiated by the addition of translation mix. Aliquots were removed at intervals during the reaction time-course and import was terminated with 12.5 µg/ml trypsin. After 15 min, 200 µg/ml soybean trypsin inhibitor was subsequently added. Mitochondria were pelleted by centrifugation at 12,000 x g for 5 min. Import of membrane proteins (AAC and Tom40) included a carbonate extraction step to remove proteins that had not inserted into the membrane (24). After a final recovery by centrifugation, mitochondria were disrupted in Laemmli sample buffer and analyzed by SDS-PAGE and autoradiography. Crosslinking assays were reported previously (16). Following import in the presence of 1 µM methotrexate (MTX), import-arrested precursor was subjected to crosslinking with 200 µM disuccinimidyl suberate (DSS) for 30 min on ice. After quenching the crosslinking reactions with 100 mM Tris, mitochondria were pelleted by centrifugation at 10,000 x g for 10 min and analyzed by SDS-PAGE and autoradiography. Oxygen consumption assays. Oxygen consumption of WT mitochondria was measured using a Clark-type oxygen electrode as described previously (25). Briefly, state II respiration was induced to a suspension of 100 µg/ml mitochondria in 0.23 M sucrose, 20 mM KCl, 20 mM Tris- HCl, 0.5 mM EDTA, 4 mM KH 2PO 4 and 3 mM MgCl 2, pH 7.2 by adding 2 mM NADH. The consumption rate was monitored for approximately 2 min. Small molecule or DMSO was added to a final vehicle concentration of 1% and respiration was measured for another approximately 2 min. Uncoupled respiration was achieved by adding 20 µM carbonyl cyanide 3- chlorophenylhydrazone (CCCP) to the chamber. Target identification using the protease protection assay. 40 µg of isolated yeast mitochondria were lysed with buffer containing 20 mM Hepes pH 7.4, 50 mM KCl, and 0.2 % Triton X-100, followed by centrifugation at 20,000 x g for 10 min. The lysate was treated with 1% DMSO or compound at 4°C for 15 min. and then incubated with 0.6 µg/ml of pronase (Sigma-Aldrich) at 25°C. One quarter of the reaction was removed at each time point, and pronase was immediately quenched by the addition of Laemmli sample buffer, followed by incubation at 95°C for 15 min. 1 µg of recombinant Tim44 was treated with 1 % DMSO or small molecule in buffer containing 20 mM Hepes-KOH pH 7.4, 50 mM KCl, and 0.01% BSA for 15 min at 4°C. The proteins were treated with pronase as indicated at 25°C for 15 min.

49

Isolation of MB-10 resistant TIM44 mutants. The TIM44 gene including 300 bp 5’ and 3’ of the open reading frame was cloned into pRS314 for mutagenesis. TIM44 was mutagenized using error-prone PCR as described previously (26), the buffer contained Mg 2+ and the concentration of the adenine nucleotide was decreased 5-fold. The mutagenized fragment was co-transformed with the linearized vector pRS315 into the ∆tim44 strain that expressed TIM44 from the centromeric URA3 vector pRS316 (designated the shuffling strain) and transformants were selected on media lacking tryptophan. The transformants were subsequently transferred to plates with 5-fluoroorotic acid (FOA) to remove the pRS316-TIM44 plasmid. Cell harboring tim44 mutants were selected for their ability to grow in the presence of 15 µM MB-10. The mutant tim44 plasmids were subsequently recovered and sequenced and retransformed to confirm that resistance to MB-10 was plasmid dependent. tim44 mutants were reconstructed with single mutations MB-10 resistant mutations (T290S and I297V) and mutations of predicted MB-10 coordinating points(H292A, W316A, W417A, W316Y and W417Y) were synthesized. Immunoprecipitation. 20 g of isolated yeast mitochondria were lysed with buffer containing 20 mM Hepes pH 7.4, 80 mM KCl, and 0.2 % Triton X-100, followed by centrifugation at 20,000 x g for 10 min. The lysate was treated with 1% DMSO or 100 µM MB-10 at 25°C for 15 min. After small molecule treatment, 5 mM MgCl 2 and 1 mM ATP were added to the lysate, followed by immunoprecipitation with anti-mHsp70 antibody and Protein A Sepharose beads (Amersham Biosciences). Tim44 was detected by immunoblot analysis. Pulldown assays. Mitochondria (400 g) were pretreated with 1% DMSO or the indicated concentration of MB-10 for 15 min and subsequently lysed in buffer A [1% digitonin, 20 mM K+HEPES (pH 7.4), 80 mM KCl, 10% glycerol, and 1mM PMSF], followed by centrifugation at 20,000 x g for 30 minutes to remove insoluble material. 50 g lysate was removed from the supernatant to use as a reference for the total sample (T). The remaining 350 g was diluted to 0.4 g/ l buffer A. Pulldowns were performed for 2 hr at 4˚C with rotation using 30 ul of HisPur Ni 2+ beads (Thermo Scientific). After binding, 50 g of lysate was removed as a reference for the material that did not bind (denoted flow-through, FT). The beads were washed 4 times with buffer. The samples were eluted in Laemmli sample buffer supplemented with 15% β-mercaptoethanol and 300 mM imidazole. The eluate (E), total, and flow through were separated by SDS-PAGE and analyzed by immunoblotting.

Cell manipulations. For microscopy experiments, HeLa cells treated with DMSO, MB-10, or CCCP for 24 hrs were fixed with 4% formaldehyde and permeabilized with ice-cold methanol. Immunostaining was performed with rabbit anti-cytochrome c antibody (Santa Cruz) and Alexa fluor 488 goat anti-rabbit IgG (Invitrogen). Cells were visualized with a microscope 9Axiovert- 200M (Carl Zeiss) using a Plan-Fluor 63x oil objective. Images were acquired at room temperature with a charge-coupled device camera (ORCA ER, Hamamatsu Photonics) controlled by Axiovision software (Carl Zeiss). Images were resized to 300 dpi without resampling using Photoshop software (Adobe). Zebrafish manipulations. Zebrafish displaying fluorescent hearts were derived from transgenic TL fish expressing a fusion of the CoxIV targeting sequence with DsRed regulated by a cmlc2 (cardiac myocyte light chain-2) promoter (22,27 ). Lines were maintained in a 14-hr light/10-hr dark cycle and mated for one hour to obtain synchronized embryonic development. Embryos were grown for 3 hpf in E3 buffer (5 mM NaCl, 0.17 mM KCl, 0.33 mM CaCl 2, 0.33 mM 50

Mg 2SO 4) and then incubated with E3 buffer supplemented with 1% DMSO, MB-10, MB-10.2, or Analog-4 for 3 days at 28.5°C. Following treatment, embryos were imaged using a Leica MZ16F fluorescent stereoscope (TexasRed filter set) at 5X magnification. Alternatively, 3-day embryos were stained with 10 µg/ml acridine orange (28) and incubated for 30 min. Embryos were then washed with E3 buffer to remove residual stain, and apoptotic cells were likewise imaged using a Leica MZ16F fluorescent stereoscope (FITC filter set). Images were resized to 300 dpi without resampling using Photoshop software (Adobe). Miscellaneous. Western blotting was performed using standard protocols. Proteins were transferred to nitrocellulose membranes and immune complexes were visualized with HRP labeled goat secondary antibody against rabbit or mouse IgG. Chemiluminescent and autoradiographic imaging was performed on film unless otherwise noted.

SUPPLEMENTAL REFERENCES 1. Roberts, C. C., and Chang, C. E. (2015) Modeling of enhanced catalysis in multienzyme nanostructures: effect of molecular scaffolds, spatial organization, and concentration. J. Chem. Theory Comput. 11 , 286-292

2. Roberts, C. C., and Chang, C. E. (2016) Analysis of Ligand-Receptor Association and Intermediate Transfer Rates in Multienzyme Nanostructures with All-Atom Brownian Dynamics Simulations. J. Phys. Chem. B 120 , 8518-8531

3. Pedretti, A., Villa, L., and Vistoli, G. (2004) VEGA--an open platform to develop chemo- bio-informatics applications, using plug-in architecture and script programming. J. Comput. Aided Mol. Des. 18 , 167-173

4. Hornak, V., Abel, R., Okur, A., Strockbine, B., Roitberg, A., and Simmerling, C. (2006) Comparison of multiple Amber force fields and development of improved protein backbone parameters. Proteins 65 , 712-725

5. DePaul, A. J., Thompson, E. J., Patel, S. S., Haldeman, K., and Sorin, E. J. (2010) Equilibrium conformational dynamics in an RNA tetraloop from massively parallel molecular dynamics. Nucleic Acids Res. 38 , 4856-4867

6. Gilson, M. K., Gilson, H. S., and Potter, M. J. (2003) Fast assignment of accurate partial atomic charges: an electronegativity equalization method that accounts for alternate resonance forms. J. Chem. Inf. Comput. Sci. 43 , 1982-1997

7. You, W., Huang, Y. M., Kizhake, S., Natarajan, A., and Chang, C. E. (2016) Characterization of Promiscuous Binding of Phosphor Ligands to Breast-Cancer-Gene 1 (BRCA1) C-Terminal (BRCT): Molecular Dynamics, Free Energy, Entropy and Inhibitor Design. PLoS Comput. Biol. 12 , e1005057

51

8. Loncharich, R. J., Brooks, B. R., and Pastor, R. W. (1992) Langevin dynamics of peptides: the frictional dependence of isomerization rates of N-acetylalanyl-N'- methylamide. Biopolymers 32 , 523-535

9. Miller, B. R., 3rd, McGee, T. D., Jr., Swails, J. M., Homeyer, N., Gohlke, H., and Roitberg, A. E. (2012) MMPBSA.py: An Efficient Program for End-State Free Energy Calculations. J. Chem. Theory Comput. 8, 3314-3321

10. Borukhov, I., Andelman, D., and Orland, H. (1997) Steric Effects in Electrolytes: A Modified Poisson-Boltzmann Equation. Phys. Rev. Lett. 79 , 435-438

11. Morris, G. M., Huey, R., Lindstrom, W., Sanner, M. F., Belew, R. K., Goodsell, D. S., and Olson, A. J. (2009) AutoDock4 and AutoDockTools4: Automated docking with selective receptor flexibility. J. Comput. Chem. 30 , 2785-2791

12. Humphrey, W., Dalke, A., and Schulten, K. (1996) VMD: visual molecular dynamics. J. Mol. Graph 14 , 33-38, 27-38

13. Jarosch, E., Tuller, G., Daum, G., Waldherr, M., Voskova, A., and Schweyen, R. J. (1996) Mrs5p, an essential protein of the mitochondrial intermembrane space, affects protein import into yeast mitochondria. J. Biol. Chem. 271 , 17219-17225

14. Dekker, P. J., Martin, F., Maarse, A. C., Bomer, U., Muller, H., Guiard, B., Meijer, M., Rassow, J., and Pfanner, N. (1997) The Tim core complex defines the number of mitochondrial translocation contact sites and can hold arrested preproteins in the absence of matrix Hsp70-Tim44. EMBO J.. 16 , 5408-5419

15. Hwang, D. K., Claypool, S. M., Leuenberger, D., Tienson, H. L., and Koehler, C. M. (2007) Tim54p connects inner membrane assembly and proteolytic pathways in the mitochondrion. J. Cell Biol. 178 , 1161-1175

16. Hasson, S. A., Damoiseaux, R., Glavin, J. D., Dabir, D. V., Walker, S. S., and Koehler, C. M. (2010) Substrate specificity of the TIM22 mitochondrial import pathway revealed with small molecule inhibitor of protein translocation. Proc. Natl. Acad. Sci. U S A 107 , 9578-9583

17. Guldener, U., Heck, S., Fielder, T., Beinhauer, J., and Hegemann, J. H. (1996) A new efficient gene disruption cassette for repeated use in budding yeast. Nucleic Acids Res. 24 , 2519-2524

18. Schmidt, B., Hennig, B., Kohler, H., and Neupert, W. (1983) Transport of the precursor to neurospora ATPase subunit 9 into yeast mitochondria. Implications on the diversity of the transport mechanism. J. Biol. Chem. 258 , 4687-4689

19. Schena, M., and Yamamoto, K. R. (1988) Mammalian glucocorticoid receptor derivatives enhance transcription in yeast. Science 241 , 965-967 52

20. Longtine, M. S., McKenzie, A., 3rd, Demarini, D. J., Shah, N. G., Wach, A., Brachat, A., Philippsen, P., and Pringle, J. R. (1998) Additional modules for versatile and economical PCR-based gene deletion and modification in Saccharomyces cerevisiae. Yeast 14 , 953- 961.

21. Glick, B. S., and Pon, L. A. (1995) Isolation of highly purified mitochondria from Saccharomyces cerevisiae. Methods Enzymol. 260 , 213-223

22. Dabir, D. V., Hasson, S. A., Setoguchi, K., Johnson, M. E., Wongkongkathep, P., Douglas, C. J., Zimmerman, J., Damoiseaux, R., Teitell, M. A., and Koehler, C. M. (2013) A small molecule inhibitor of redox-regulated protein translocation into mitochondria. Dev. Cell 25 , 81-92

23. Kato, H., and Mihara, K. (2008) Identification of Tom5 and Tom6 in the preprotein translocase complex of human mitochondrial outer membrane. Biochem. Biophys. Res. Commun. 369 , 958-963

24. Koehler, C. M., Jarosch, E., Tokatlidis, K., Schmid, K., Schweyen, R. J., and Schatz, G. (1998) Import of mitochondrial carriers mediated by essential proteins of the intermembrane space. Science 279 , 369-373

25. Claypool, S. M., Boontheung, P., McCaffery, J. M., Loo, J. A., and Koehler, C. M. (2008) The cardiolipin transacylase, tafazzin, associates with two distinct respiratory components providing insight into Barth syndrome. Mol. Biol. Cell 19 , 5143-5155

26. Koehler, C. M., Merchant, S., Oppliger, W., Schmid, K., Jarosch, E., Dolfini, L., Junne, T., Schatz, G., and Tokatlidis, K. (1998) Tim9p, an essential partner subunit of Tim10p for the import of mitochondrial carrier proteins. EMBO J. 17 , 6477-6486

27. Shu, X., Huang, J., Dong, Y., Choi, J., Langenbacher, A., and Chen, J. N. (2007) Na,K- ATPase alpha2 and Ncx4a regulate zebrafish left-right patterning. Development 134 , 1921-1930

28. van Ham, T. J., Mapes, J., Kokel, D., and Peterson, R. T. (2010) Live imaging of apoptotic cells in zebrafish. FASEB J. 24 , 4336-4342

53

Table S1. Strains constructed in this study

Strain Backgroun Genotype Source d WT GA74 his3 leu2 ade8 trp1 ura3 (13) tim23-2 GA74 his3 leu2 ade8 trp1 ura3 tim23-2:TRP1 (14)

WT ∆pdr5 ∆snq2 GA74 his3 leu2 ade8 trp1 ura3 pdr5 ∆0:: HIS3 This study snq2 ∆0:: KANMX tim23-2∆pdr5 ∆snq2 GA74 his3 leu2 ade8 trp1 ura3 tim23-2:TRP1 This study pdr5 ∆0:: HIS3, snq2 ∆0:: KANMX tim10-1∆pdr5 ∆snq2 GA74 his3 leu2 ade8 trp1 ura3 tim10-1:LEU2 (16) ∆tim10 :: HIS3 pdr5 ∆0:: HGR snq2 ∆0:: KANMX WT ∆pdr5 ∆snq2 GA74 his3 leu2 ade8 trp1 ura3pdr5 ∆0:: HIS3 This study [Su9-URA3] snq2 ∆0:: KANMX Su9-URA3:LEU2 tim23-2∆pdr5 ∆snq2 GA74 his3 leu2 ade8 trp1 ura3 tim23-2:TRP1 This study [Su9-URA3] pdr5 ∆0:: HIS3 snq2 ∆0:: KANMX Su9- URA3:LEU2 WT ∆pdr5 ∆snq2 ∆tim GA74 his3 leu2 ade8 trp1 ura3 pdr5 ∆0 snq2 ∆0 This study 44 [pRS316-TIM44 ] tim44 ∆0:: HIS3 [pRS316-TIM44:URA3 ] WT ∆pdr5 ∆snq2 ∆tim GA74 his3 leu2 ade8 trp1 ura3 pdr5 ∆0 snq2 ∆0: This study 44 [pRS315-TIM44 ] tim44 ∆0:: HIS3 [pRS315-TIM44:LEU2 ] WT ∆pdr5 ∆snq2 ∆tim GA74 his3 leu2 ade8 trp1 ura3pdr5 ∆0 snq2 ∆0 This study 44 [pRS315- tim44 ∆0:: HIS3 [pRS315- TIM44T290S ] tim44T290S:LEU2 ] WT ∆pdr5 ∆snq2 ∆tim GA74 his3 leu2 ade8 trp1 ura3pdr5 ∆0 snq2 ∆0 This study 44 [pRS315- tim44 ∆0:: HIS3 [pRS315- TIM44H292Y ] tim44H292Y:LEU2 ] WT ∆pdr5 ∆snq2 ∆tim GA74 his3 leu2 ade8 trp1 ura3 pdr5 ∆0 snq2 ∆0 This study 44 [pRS315- tim44 ∆0:: HIS3 [pRS315- TIM44H292A ] tim44H292A:LEU2 ] WT ∆pdr5 ∆snq2 ∆tim GA74 his3 leu2 ade8 trp1 ura3pdr5 ∆0 snq2 ∆0 This study 44 [pRS315- tim44 ∆0:: HIS3 [pRS315- TIM44I297V] tim44I297V:LEU2 ] WT ∆pdr5 ∆snq2 ∆tim GA74 his3 leu2 ade8 trp1 ura3 pdr5 ∆0 snq2 ∆0 This study 44 [pRS315- tim44 ∆0:: HIS3 [pRS315-tim44 TIM44W316A] W316A:LEU2 ] WT ∆pdr5 ∆snq2 ∆tim GA74 his3 leu2 ade8 trp1 ura3 pdr5 ∆0 snq2 ∆0 This study 44 [pRS315- tim44 ∆0:: HIS3 [pRS315-tim44 TIM44W316Y] W316Y:LEU2 ] WT ∆pdr5 ∆snq2 ∆tim GA74 his3 leu2 ade8 trp1 ura3 pdr5 ∆0 snq2 ∆0 This study 44 [pRS315- tim44 ∆0:: HIS3 [pRS315- TIM44W417Y ] tim44W417Y:LEU2 ] WT ∆pdr5 ∆snq2 ∆tim GA74 his3 leu2 ade8 trp1 ura3 pdr5 ∆0 snq2 ∆0 This study 54

44 [pRS315- tim44 ∆0:: HIS3 [pRS315- TIM44W417A ] tim44W417A:LEU2 ] TIM44HIS10 GA74 his3 leu2 ade8 trp1 ura3 TIM44- This study HIS10:HISMX6

55

Table S2. Summary of chemical genetic screen Screening candidates Number Chembridge library 30,000 Candidates in primary screen that conferred growth 25 Confirmed hits 6 General uncouplers of respiration 0 Inhibitor of respiration 4 Membrane permeabilizers 0 Protein import modulators 2 (MB-10, MB-11)

56

Table S3. Results of the MM/PBSA calculation for MB-10 bound to wild-type Tim44 and Tim44 mutants (F284L, T290S, I297V, and V298A).*

Tim44 <∆ETotal > <∆EvdW > <∆EElec > <∆EPB > <∆ECavity > <∆EDisper > WT -14.89 ± 2.89 -46.44 -2.01 14.92 -28.81 47.45 F284L -10.30 ± 3.02 -43.78 -1.98 16.60 -27.80 46.66 T290S -9.66 ± 2.94 -42.64 0.47 13.92 -25.24 43.82 I297V -7.39 ± 3.19 -41.84 -8.83 23.74 -26.74 46.28 V298A -7.76 ± 3.50 -39.83 -2.78 14.94 -25.72 45.63 *Units are kcal/mol; < ∆ETotal >, < ∆EvdW >, < ∆EElec > < ∆EPB >, < ∆ECavity >and < ∆EDisper > are the average total interaction, van der Waals interaction, coulombic interaction, solvation energy change from Poisson Boltzmann model, interaction from cavity change and dispersion energy change, respectively. The more negative < ∆ETotal> value typically indicates stronger ligand binding and the less negative < ∆ETotal > value indicates weaker ligand binding.

57

Figure S1. MB-10 activity is influenced by specific chemical characteristics. Structures of MB-10 and the similar compounds from the screening library that did not confer growth of WT[Su9-Ura3] strain upon addition and were not identified as inhibitors.

58

Figure S2. Mutations H292A, W316A, W316Y, and W417A in Tim44 confer a dominant negative phenotype. Mutations in amino acids 292, 316, and 417 of Tim44 were generated because of potential contact sites with MB-10 from MD studies. Plasmids were transformed into the ∆tim44 [p TIM44:URA3 ] and plated on media lacking leucine, which was the selectable marker of the tim44 mutant plasmids. Despite several tries, viable colonies were not recovered for the ∆tim44 [p TIM44:URA3 ] strain with plasmids that harbored tim44 mutants H292A, W316A, W316Y, and W417A, indicating that these mutants conferred a dominant negative phenotype. Viable colonies were recovered for the strains harboring tim44 H292Y and tim44 W417Y; the strain with the TIM44 WT plasmid was included as a control.

59

Chapter 3: A small molecule probe elucidates the role of mitochondrial translocase TIMM44 in PINK1/Parkin regulated mitophagy

Abstract

Neurodegenerative diseases have been linked to dysfunctional mitochondrial quality control system that is partially maintained by proteins PINK1 and Parkin. Whereas mitophagy pathways are becoming well-characterized, little is known about the molecular mechanisms for

PINK1 trafficking. Accordingly, we have used a small molecule probe (MitoBloCK-10) that modulates the mitochondrial translocase TIMM44, an essential component of the protein translocation motor, to characterize PINK1 trafficking. MB-10 did not inhibit PINK1 import into, and subsequent degradation of, healthy mitochondria. However, MB-10 altered trafficking upon mitochondrial insult such that PINK1 accumulation and Parkin recruitment were impaired.

This data suggests that PINK1 may not follow a canonical import pathway and that the protein translocation motor has a central role in regulating PINK1-dependent mitophagy. Our studies also provide a probe for dissecting PINK1/Parkin events at mitochondria as well as studying

PINK1-dependent mitophagy in in vivo models.

Introduction

Deregulation of mitophagy, the process of clearing dysfunctional mitochondria, has been implicated in various human diseases, including neurodegeneration 1. Certain mutations in PTEN- induced putative kinase (PINK1), a mitochondrial serine kinase, and Parkin, a cytosolic E3- ubiquitin ligase, lead to early onset familial Parkinson’s disease (PD) 2. These proteins selectively degrade damaged mitochondria 3, which can be induced by the loss of the mitochondrial

60 membrane potential 4 or accumulation of unfolded proteins in the mitochondrial matrix 5, PINK1 and Parkin coordinate the ubiquitination of outer membrane (OM) proteins for proteasomal degradation 6, 7 and the recruitment of autophagic machinery for eventual mitophagy 8-10 .

The overall mechanism of this pathway is well-established. Under steady-state conditions, PINK1 is constitutively targeted to the mitochondria by its N-terminal mitochondrial targeting sequence (MTS), where it is partially imported through TOMM40 of the TOM complex into mitochondria and subsequently cleaved by MPP in the matrix 11 and PARL between Ala103

and Phe104 in the inner membrane (IM) 12 . This cleaved form of PINK1 undergoes proteasomal

degradation resulting in no accumulation of PINK1 on the OM 13 . Alternatively, upon mitochondrial insult, MTS of PINK1 is not cleaved, causing the accumulation of the precursor form at the OM with the kinase domain remaining in the cytosol 14 . PINK1 at the OM has the dual activity of phosphorylating Parkin at Ser65 and ubiquitin at Ser65 15-17 , both of which activate the ligase function of Parkin to maximize activity at the mitochondria 18 . This pathway not only stimulates ubiquitin-proteasome degradation of select OM proteins, but also recruits and activates the autophagy machinery at mitochondria 10 .

Conversely, the molecule mechanism of PINK1 translocation into mitochondria is

relatively unknown. It has been shown that TOMM40 is important for the import of precursor

PINK1 and the association of PINK1 with the OM 14, 19 . TOMM7, a homolog of the S. cerevisiae

protein Tom7 that modulates TOM complex dynamics and functions in protein import 20 , is

necessary for PINK1 accumulation on the OM but not for the PARL cleavage of PINK1 21 .

Recent evidence also suggests that the Tim23 complex has no role in PINK1 trafficking on healthy mitochondria, but is needed for stability upon insult 22 . Additionally, new insights pertaining to the import of PINK1 have revealed that PINK1 contains a second targeting

61 sequence, the outer mitochondrial membrane localization signal (OMS), between the canonical

MTS (amino acids 1-34) and the transmembrane domain at amino acids 94-110 23 . On healthy mitochondria, the MTS reaches the matrix to initiate the PINK1 degradation pathway, and the

OMS remains dormant. But upon depolarization, the MTS does not pass through to the IM and the OMS becomes functional to associate with the TOM complex and activate the mitophagy pathway.

Characterization of the PINK1 translocation pathway into healthy mitochondria and its retention on the OM in dysfunctional mitochondria may facilitate the development of therapeutic strategies 24 . Although methods to study the pathway exist, such as RNAi and molecular screens

to discover regulators of Parkin 22, 25 , there remains limited techniques to dissect the biochemistry

of PINK1 at the mitochondria or to recapitulate the pathway in a specific, controllable manner.

Currently, investigators commonly utilize the protonophore carbonyl cyanide m-chlorophenyl

hydrazone (CCCP) or the expression of misfolded proteins targeted to the matrix to activate

PINK1/Parkin-mediated mitophagy 5, 26 , but more precise methods are needed.

Here we report the characterization of a small molecule probe in PINK1/Parkin dynamics. This probe was identified in a chemical screen for mitochondrial protein import inhibitors in yeast. A Tim44 modulator, MitoBloCK-10 (MB-10) 27 , had no effect on PINK1

import into healthy mitochondria, but inhibited the accumulation of PINK1 on the mitochondrial

OM in the presence of CCCP or other insult, dramatically reducing the progression of

mitophagy. This molecule further elucidates the mechanism of PINK1 import, specifically the

role of TIMM44, and should provide a valuable tool in future research.

Results and discussion

MB-10 targets TIMM44 and inhibits PINK1 accumulation but not import

62

Because MB-10 (Figure 3.S1A) targets yeast Tim44, we first confirmed that human and yeast homologs were conserved. Sequence analysis shows that four of the seven Tim44 residues critical for MB-10 binding are identical in TIMM44, with a fifth residue sharing similarity

(Figure 3.S1B). Prior studies have also shown that the binding region of MB-10 shares structural similarities in yeast and human 28 . Furthermore, TIMM44 was localized to the matrix and is

loosely integrated in the inner membrane, most likely in a protein-rich environment, similar to

yeast Tim44 (Figure 3.S1C,D). As in yeast mitochondria, MB-10 only inhibited the import of

matrix-targeted precursors and not OM (MFN-1) or intermembrane space (IMS; CHCHD2)

precursors (Figure 3.S2A-F). Based on these and a previous study 27 , MB-10 inhibits TIMM44

activity, likely by holding the TIM23 precursor tightly in the translocon.

PINK1 is targeted to the mitochondria via its MTS and uses the TOM translocon for

import through the OM, but arrests translocation in the IM resulting in cytosolic presentation of

its kinase domain at the OM as demonstrated by trypsin sensitivity 14 (Figure 3.S1E). We investigated the role of the PAM complex in PINK1 import and stability by inhibiting TIMM44 with MB-10. HeLa cells (lacking Parkin expression) treated with uncoupler amass PINK1 on the

OM, but MB-10 treatment inhibited the uncoupler-dependent PINK1 accumulation. MB-10 alone did not alter the presence of PINK1 on mitochondria (Figure 3.1A,B). As controls,

MG132, a proteasome inhibitor, caused the accumulation of PARL-cleaved PINK1 (PINK1*) and valinomycin, a potassium ionophore, resulted in acute accumulation of full length (FL)

PINK1.

To confirm that this effect is not specific to CCCP, MB-10 similarly inhibited PINK1 accumulation upon mitochondrial impairment by antimycin A and oligomycin treatment or valinomycin treatment (Figure 3.1C,D). To determine if the reduction in PINK1 on mitochondria

63 was caused by inhibiting the initial import of PINK1, we treated cells with both MB-10 and

MG132. We show that MB-10 did not inhibit the formation of the PARL-cleaved band formed upon MG132 treatment, indicating that MB-10 does not alter normal PINK1 import (Figure

3.1E).

In vitro radiolabeled PINK1 import studies with MB-10 showed similar results to experiments with intact cells (Figure 3.1F-H). MB-10 treatment did not alter PARL cleavage of imported PINK1 (designated PINK1*) into isolated HeLa mitochondria. In contrast, uncoupler addition resulted in loss of the PARL-cleaved PINK1 intermediate (Figure 3.1F,G), providing additional evidence that TIMM44 does not play a direct role in normal PINK1 import. There is an accumulation of full-length PINK1 (designated FL-PINK1) in import assays that is uncoupler- independent that MB-10 again did not affect relative to vehicle control (Figure 3.1F,H).

MB-10 inhibits PINK1 association with the TOM complex and Parkin recruitment

Although import was not altered, radiolabeled PINK1 association with the TOM complex in the presence of CCCP was reduced by MB-10 treatment as shown by BN-PAGE (Figure

3.2A); however, MB-10 alone did not alter the amount of PINK1 at the TOM complex.

Additionally, intact HeLa cells overexpressing EGFP-Parkin, useful for assessing Parkin activity, treated with MB-10 prior to CCCP addition also displayed a marked reduction of PINK1 at the

TOM complex relative to CCCP alone (Figure 3.2B) and total Parkin recruitment to mitochondria was decreased, including the Parkin-VDAC complex 29 . To further explore the

decrease in Parkin recruitment, the mitochondrial fraction can also be run on SDS-PAGE.

Because Parkin expression in these cells is very high, a low level of Parkin was present at the

mitochondria even in untreated cells. However, high levels of phospho-ubiquitinated Parkin were

present at mitochondria following mitochondrial insult, coinciding with the degradation of the

64

OM protein MFN-1 (Figure 3.2C-E). MB-10 treatment reduced the amount of phospho- ubiquitinated Parkin and the subsequent OM events caused by CCCP treatment.

We also tracked the distribution of EGFP-Parkin within cells treated with MB-10 by fluorescent microscopy (Figure 3.2F,G). Under normal conditions (DMSO), Parkin localization was diffuse throughout the cytosol and the mitochondrial network, as marked by robust

MitoTracker staining, was distributed throughout the cell. With an increasing MB-10 concentration, Parkin localization remained diffuse and the membrane potential was maintained, but the mitochondrial network eventually fragmented, perhaps due to an alteration of dynamics proteins associated with the decrease in functional TIMM44 30 . Upon addition of CCCP, the

mitochondrial membrane potential was dissipated, causing strong Parkin recruitment to

mitochondria. The addition of MB-10, even at 20 M, inhibited Parkin recruitment to

mitochondria even though membrane depolarization was induced by CCCP addition.

To study the kinetics of mitophagy with and without MB-10, we used the tandem sensor

RFP-GFP-LC3. The LC3 causes the fusion protein to localize at autophagic structures, and the

RFP and GFP fluoresce to monitor mitophagic progression (Figure 3.3A). In HeLa cells treated

with CCCP alone, there is an initial rise in the number of autophagosomed form, followed by a

reciprocal rise in autolysosome formation, with the total number of autophagic events remaining

relatively constant through the timecourse. Treatment with MB-10 slows mitophagy progression

with no discernible rise in the number of autophagosomes even after 3 hours (Figure 3.3B,C).

Because PINK1/Parkin-mediated mitophagy can clear the entire population of mitochondria in

24 hrs 8, 10 , we tested if MB-10 treatment blocked mitophagy over longer periods by assessing the loss of mtDNA (Figure 3.S3). CCCP treatment the EGFP-Parkin overexpression cell line resulted in nearly complete removal of the mitochondria after 24 hrs as evidenced by a lack of

65 mtDNA staining (in nucleoid puncta surrounding the cell nucleus) with an anti-DNA antibody, whereas MB-10 treatment inhibited the degradation of mtDNA and mitochondria caused by the uncoupler (Figure 3.S3).

Reduction in PINK1 accumulation by MB-10 is specifically due to TIMM44 inhibition

Because specificity of MB-10 in cells is of some concern, we tested this using a variety of methods. We did not note any difference in PINK1 expression in HeLa cells using quantitiative PCR (qPCR) in cells that had been treated with 40 M MB-10 and vehicle control, with or without the addition of 10 M CCCP (Figure 3.3D). This indicates that the addition of

the molecules is not inducing cellular stress that slows protein expression, which would reduce

PINK1 levels.

We also used previously identified SAR compounds 27 that do not inhibit yeast Tim44 but

are structurally similar to MB-10 to rule out off-target effects of MB-10 (Figure 3.4A). These

Analogs do not inhibit the import of Su9-DHFR into isolated HeLa mitochondria, demonstrating

that they do not inhibit TIMM44 (Figure 3.4B,C). Furthermore, the compounds do not disrupt

the association of PINK1 with the TOM complex (Figure 3.4D). Analysis of downstream effects

of PINK1 recruitment, such as Parkin activation at mitochondria and MFN-1 degradation, shows

that these Analogs do not inhibit mitophagy (Figure 3.4E-G). Taken together, this data suggests

that the inhibition of PINK1 accumulation caused by MB-10 is not due to off-target effects of the

molecule.

Finally, we knocked down TIMM44 using RNAi and tested protein import. With less

TIMM44 available, Su9-DHFR import into isolated mitochondria was partially impaired (Figure

3.5A). There was only a slight reduction in formation of the PARL-cleaved form of PINK1

during import upon TIMM44 knockdown, but the association of PINK1 with TOM complex was

66 disrupted (Figure 3.5B) The ratio of band reduction to TIMM44 knockdown efficiency was similar between Su9-DHFR import and PINK1/TOM formation, but much lower for formation of the PARL-cleaved form of PINK1 (Figure 3.5C). This data would suggest that TIMM44 does not alter PINK1 import and release on healthy mitochondria but is necessary for PINK1 stability with the TOM complex.

Discussion

Taken together, we provide initial characterization of PINK1/Parkin trafficking on mitochondria using a small molecule probe that attenuates protein translocation (Figure 3.5D-F).

The data demonstrates that inhibiting TIMM44 with MB-10 treatment has no consequence on normal PINK1 import, cleavage, and ultimately degradation (Figure 3.5D). Upon mitochondrial insult, inhibiting TIMM44 activity alters the trafficking of PINK1 at the mitochondrial OM and subsequently reduces Parkin recruitment (Figure 3.5F). This is unexpected because PINK1 is predicted to bypass the TIM23 translocon and TIMM44 in the absence of a membrane potential

(Figure 3.5E). However, MB-10 studies support an important role for TIMM44 and the TIM23 translocon for PINK1 presentation on the OM for subsequent Parkin recruitment. Previous studies showed that TOMM7 also lacked a direct role in PINK1 import, but is necessary for the association of PINK1 with the TOM complex 21 , indicating that there may be coordination

between the TOM and TIM23 complexes. Recent evidence also supports the notion the TIM23

and PAM complex may not play a role in normal PINK1 import but are needed for PINK1

maintenance 22 . Because PINK1 contains two targeting sequences 23 , it is possible that TIMM44

interacts exclusively with the second targeting sequence, which is critical for the interaction

between PINK1 and the TOM complex. Our research provides new insights into the molecular

mechanism of PINK1 trafficking and offers a new tool for further studies.

67

Materials and Methods

Cell culture and transfection

HeLa cells including those with stable expression of EGFP-Parkin 8 were cultured in

Dulbecco’s modified eagle medium (DMEM), 25 mM glucose, 1 mM pyruvate (Life

Technologies 11995-065), supplemented with heat-inactivated 10% FBS (v/v; Atlanta

Biologicals S11150H) and 1% penicillin/streptomycin (Life Technologies 15140122) in a

humidified incubator at 37˚C under 5% CO 2 (v/v). Tranfections were performed using jetPRIME transfection reagent (Polyplus 114-07) according to the manufacturer’s instructions.

Small molecule treatment with cultured cells

Small molecules are dissolved in 100% DMSO in brown glass vials (Wheaton 224750) and stored in a dessicator at room temperature. The MB-10 stock was 20 mM; fresh stocks were made monthly. Cells in culture were grown to 100% confluency followed by seeding to 25% confluency. After 48 hrs, cells were treated with indicated small molecules dissolved in fresh media; the final DMSO for all conditions was 1%, unless otherwise noted. For experiments using

MB-10 on intact cells, MB-10 was added 2 hrs prior to the addition of other small molecules. For all experiments analyzed by BN-PAGE, 10 cm dishes (CELLSTAR 664-160) were used with 10 mL of media; all other experiments used 6 cm dishes (CELLSTAR 628-160) with 2.5 mL of media. At the time points indicated in each figure, media was aspirated and the cells were washed with PBS.

Mitochondrial purification for immunoblotting analysis

Cells were removed from the culture dishes by incubating in Trypsin-EDTA (0.05%, Life

Technologies 25300-054) for 5 min at 37˚C. Cells were collected in media and centrifuged at

770 x g for 5 min. at 4˚C. Cells were washed with ice-cold PBS (Life Technologies, 14190-144)

68 and centrifuged again. The pellet was resuspended in 1 ml homogenizer buffer (20 mM HEPES-

KOH pH 7.4, 220 mM mannitol, 70 mM sucrose) supplemented with 2 mg/mL BSA and 0.5 mM

PMSF. Cells were lysed on ice by 20 strokes through a 25-gauge x 5/8 needle (Becton Dickson

305122) and 1 mL syringe followed by centrifugation at 770 x g for 5 min. at 4˚C to remove unbroken cells. The supernatant was centrifuged at 10,000 x g for 10 min at 4˚C, washed with homogenizer buffer (lacking BSA or PMSF), and re-centrifuged at 10,000 x g for 10 min at 4˚C to obtain mitochondrial pellets. The pellets were resuspended in 100 µl of homogenizer buffer, and the protein concentration was determined using the Pierce BCA Protein Assay Kit

(Thermofisher). For each condition, 20 g of mitochondria was aliquoted into tubes and centrifuged at 10,000 x g for 10 min at 4˚C. The supernatant was removed and mitochondria were disrupted with Laemmli-SDS sample buffer; the proteins were separated by SDS-PAGE followed by transfer to PVDF membranes, and analyzed by immunoblotting for indicated proteins.

The antibodies used in this study include anti-Mortalin (UC Davis/NIH Neuromab

Facility clone N52A/42, 75-127), anti-Parkin (Santa Cruz Biotechnology, sc-32283), anti-AIF

(Santa Cruz Biotechnology, sc-13116), anti-TIMM23 (BD Biosciences, 611223), anti-PINK1

(Novus Biologicals clone 8E10.1D6, NBP2-36488), anti-TOMM20 (Santa Cruz Biotechnology, sc-11415), anti-TIMM44 (Proteintech, 13859-1-AP), anti-TOMM70 (Proteintech, 14528-1-AP), and anti-LRP-130 (Santa Cruz Biotechnology, sc-66844). The GST-MFN-1 (gift from Richard

Youle), anti-PreP, anti-YME1L, anti-GFP, and anti-TOMM40 were polyclonal antibodies generated by vendor Pacific Immunology from recombinant proteins.

69

Mitochondrial purification and in vitro protein import assays

Mitochondria isolation and imports were based on established methods 21 . Samples were prepared as described for lysis with a syringe but a Teflon pestle (Bellco) was used.

[35 S]-labeled proteins were synthesized in vitro via a coupled transcription/translation rabbit reticulocyte kit (Promega) incubated at 30˚C for 90 min with [ 35 S]Met/Cys metabolic

labeling reagent (MP Biomedicals, 015100907) and plasmid DNA expressing human PINK1,

CHCHD2, or Su9-DHFR via an Sp6 promoter or MFN1-myc via a T7 promoter.

Isolated mitochondria were diluted to 0.4 g/ L in homogenizer buffer (70 g per

reaction) supplemented with 17.5 L of 10X energy generating solution (10 mM ATP, 5 mM magnesium chloride, 200 mM sodium succinate, 50 mM NADH in homogenizer buffer) and

0.1% BSA adjusted to a volume of 140 L. DMSO vehicle or compounds dissolved in DMSO

(1% final) were added to the mitochondria and incubated at 25˚C for 15 min. Reactions were then initiated by addition of 35 uL of precursor and aliquots were withdawn and stopped at specified time points.

The import of Su9-DHFR was terminated by the addition of ice-cold 10 g/mL trypsin

(Sigma) in homogenizer buffer. After 15 min on ice, soybean trypsin inhibitor (VWR Amresco) was added at 50 g/mL. Mitochondria are pelleted by centrifugation at 10,000 x g for 5 min at

4˚C. Import of OM proteins (PINK1 and MFN1-myc) was terminated by the addition of ice-cold homogenizer buffer and subsequent centrifugation at 10,000 x g for 5 min. Samples were then

31 subject to carbonate extraction in 100 mM Na 2CO 3 at pH 11 . For all precursors, after final

recovery by centrifugation, mitochondria were disrupted in Laemmli SDS-sample buffer and

analyzed by SDS-PAGE and autoradiography.

70

BN-PAGE

After import of radiolabeled proteins into mitochondria (40 g per reaction), reactions

were stopped by the addition of ice-cold homogenizer buffer. Samples were centrifuged at

10,000 x g for 5 min at 4˚C and gently washed with homogenizer buffer. The mitochondrial pellets were dissolved in ice-cold lysis buffer [20 mM HEPES-KOH pH 7.4, 50 mM NaCl, 1% digitonin (Gold Biotechnology D-180-2.5), 10% glycerol, 2.5 mM MgCl 2, 0.5 mM PMSF, 0.5

mM EDTA] to give 2.5 g of mitochondria per % digitonin and kept on ice for 15 min. Samples

were subsequently centrifuged at 20,800 x g for 5 min at 4˚C. Supernatants were then separated

by BN-PAGE. Once complete, the gels were cut into two sections. One section was analyzed by

immunoblotting and the second section was analyzed by a phospho-screen and imaged with a

Bio-Rad Molecular Imager FX Pro Plus.

Sub-mitochondrial localization

Mitochondria were isolated from cells using Teflon dounce. Mitochondria (50 g) were subsequently incubated in homogenizer buffer (Mitos), 20 mM HEPES-KOH pH7.4 (to generate mitoplasts), or 1% Triton X-100 (TX-100) supplemented with or without 3 g/mL proteinase K.

Samples were incubated on ice for 30 min, followed by inhibition of proteinase K with 1 mM

PMSF. The samples were centrifuged at 20,800 x g for 10 min at 4˚C. The supernatant was precipitated in 10% trichloroacetic acid (TCA, w/v) followed by centrifugation to recover the precipitate. The samples were disrupted in Laemmli-SDS sample buffer, separated by SDS-

PAGE, and analyzed by immunoblotting.

Carbonate extraction

Isolated mitochondria (50 g) were incubated in homogenizer buffer or 100 mM Na 2CO 3 at increasing pH for 30 min on ice. Samples were centrifuged at 20,800 x g for 30 min at 4˚C.

71

The supernatant was TCA-precipitated and pellet and supernatant samples were disrupted in

Laemmli-SDS sample buffer, separated by SDS-PAGE, and analyzed by immunoblotting.

Immunofluorescence microscopy

Thirty min prior to fixing, MitoTracker Red CMXRos (Invitrogen, M7512) was added to each well to 20 nM final concentration. Cells on D-poly-Lysine coated coverslips were then fixed with pre-warmed 3.7% formaldehyde (v/v; VWR, VW3408-1), permeabilized with ice- cold methanol, and blocked with 1% fatty acid free BSA in PBS (v/v, PBS-BSA).

Where indicated, cells were subsequently incubated overnight at 4˚C with indicated primary antibodies diluted in PBS-BSA. Following thorough washing in PBS, cells were incubated for 1 hr at room temperature with Alexa Fluor 350-conjugated secondary antibodies

(Life Technologies) diluted in PBS-BSA. The following antibodies were used: anti-Mortalin primary with Alexa Fluor 350 goat-anti-mouse IgG (A11045), anti-TOMM20 primary with

Alexa Fluor 350 goat-anti-rabbit IgG (A11046), and anti-DNA (PROGEN Biotechnik clone AC-

30-10, 61014) mouse monoclonal IgM with Alexa Fluor 350 goat-anti-mouse IgM (A31552).

Cells were then set on slides and images were obtained using a Leica TCS SPE DMI 4000B inverted confocal microscope. qPCR

RNA was isolated from cells using TRIzol reagent, residual genomic DNA was removed with TURBO DNA-free kit, and cDNA was synthesized from the RNA with the Superscript III

First-Strand Synthesis System (ThermoFisher 15596026, AM1907, and 18080051) according toi manufacturer’s instructions. The qPCR was performed with 2x SYBR Green qPCR Master Mix

(Bimake B21202) on the Biorad CFX Connect Real-Time PCR detection system. PINK1 and

HRPT1 primers were obtained from Integrated DNA Technologies. PINK1 primer sequences

72 used were 5’-GGA CGC TGT TCC TCG TTA-3’ and 5’-ATC TGC GAT CAC CAG CCA-3’, as previously described 32 . siRNA knockdown

TIMM44 and negative control siRNA (QIAGEN SI03125969 and 1022076) were transfected at 20 nM using jetPRIME reagent (Polyplus 114-07) according to manufacturer’s instructions.

Statistical analyses

Data was plotted using GraphPad Prism version 5.02. Statistical significance of observations was determined using unpaired t tests (two-tailed), unless otherwise noted.

73

Figures

Figure 3.1. MB-10 inhibits the accumulation of PINK1 on damaged mitochondria, but does not affect PINK1 import into healthy mitochondria. (A) Intact HeLa cells were treated with indicated concentrations of MB-10 for 2 hrs, followed by the addition of 1 g/mL valinomycin,

1% DMSO, 10 M MG132, or 10 M CCCP for an additional 3 hrs as indicated. The

mitochondrial fraction was separated by SDS-PAGE, and the indicated proteins were detected by

immunoblot. FL-PINK1, full-length PINK1; PINK1*, PARL-cleaved PINK1. (B) Quantification

of the FL-PINK1 band. The band at 0 M MB-10 with CCCP represents 100% for each trial.

74

The data represent the average ± SD of n = 4 trials (* P < 0.05). (C) Intact HeLa cells were treated as in ‘A’ with 10 M antimycin A and 1 M oligomycin (A + O) instead of CCCP. (D)

Intact HeLa cells were treated as in ‘A’ with 1 g/mL valinomycin for 1 hour instead of CCCP

for 3 hours. (E) Intact HeLa cells were treated with DMSO or 40 M MB-10 for 2 hrs, followed

by the addition of DMSO, 1 g/mL valinomycin, or 10 M MG132 for an additional 3 hrs.

Samples were processed as in ‘A’. (F) Upper—a representative image of a time course assay for

radiolabeled PINK1 import into isolated HeLa mitochondria after a 15 min incubation in MB-10

or DMSO; -∆ψ = 50 M CCCP. The longer exposure shows the PARL-cleaved form of PINK1, designated PINK*. Lower—a shorter exposure showing full-length PINK1, designated FL-

PINK1. (G) Quantification of the PARL-cleaved band. The band at 0 M MB-10 at 20 min

represents 100% for each trial. The data represent the average ± SD of n = 3 trials. (H) As in ‘G’ for FL-PINK1; the -∆ψ band represents 100% for each trial.

75

Figure 3.2. MB-10 disrupts association of PINK1 with the TOM complex and inhibits

CCCP-induced recruitment of Parkin to mitochondria in HeLa cells overexpressing stable 76

EGFP-Parkin. (A) Radiolabeled PINK1 was imported into isolated HeLa cell mitochondria in

the presence of MB-10 or CCCP as indicated. Mitochondria were lysed with 1% digitonin and

lysates were fractionated on BN-PAGE. Lanes 1-7 display [ 35 S]PINK1; lanes 8-9 display a

TOMM40 immunoblot from the same gel. (B) Intact HeLa cells overexpressing stable EGFP-

Parkin were treated with MB-10 for 2 hrs followed by the addition of CCCP for another 3 hrs.

Samples were analyzed by BN-PAGE as in ‘A’ followed by immunoblot. PreP is included as a

loading control from the same samples run on SDS-PAGE. (C) Intact HeLa cells with stable

overexpression of EGFP-Parkin were treated as in ‘1A’. Ph-Ubi-Parkin marks the phoshpo-

ubiquitinated form and Ubi-MFN-1 marks the ubiquitinated form (D) Quantification of the Ph-

Ubi-Parkin band. The band at 0 M MB-10 with CCCP represents 100% for each trial. The data represent the average ± SD of n = 3 trials (* P < 0.05). (E) Quantification of the Ubi-MFN-1 band. The band at DMSO only represents 100% for each trial. The data represent the average ±

SD of n = 3 trials (* P < 0.05). (F) Representative fluorescent images after treatment with indicated concentrations of MB-10 for 2 hrs, followed by the addition of either DMSO of 10 M

CCCP for an additional 3 hrs. Scale bar = 15 m. (G) Quantification of Parkin translocation to mitochondria in ‘F’ by cell counting. 100 cells were counted for each condition; data represent the average ± SD of n = 4 trials (* P < 0.05).

77

Figure 3.3. MB-10 slows down mitophagy but does not affect PINK1 transcription. (A)

Cartoon depicting the processing of RFP-GFP-LC3. Upon mitophagy induction by CCCP, RFP-

GFP-LC3 colocalizes with mitochondria at the autophagosome, giving a green/yellow signal. As mitophagy progresses, the autophagosome fuses with the lysosome. The high acidity quenches the GFP, leaving on the red RFP signal. (B) Representative fluorescent images after cells are transfected with a plasmid the expresses RFP-GFP-LC3. After passaging, cells are treated with

78

DMSO or 40 M MB-10 for 2 hours, and finally 10 M CCCP. Scale bar = 10 m. (C) The number of green/yellow and red dots were counted for various timepoints post-CCCP treatment.

Data represent average ± SEM for n = 5 cells. (D) Analysis of PINK1 expression in HeLa cells

by qPCR. Data represent average ± SEM of n = 3.

79

Figure 3.4. SAR compounds of MB-10 have no affect on PINK1-dependent mitophagy. (A)

Two previously identified compounds that have similar structures to MB-10 but do not bind

80

TIMM44. Red circled indicate differences with the MB-10 structure. (B) Representative image of an import of radiolabeled Su9-DHFR into isolated HeLa mitochondria after 15 minute incubation with Analog-3 and -4. p = precursor; m = mature form. (C) Quantification of the mature band; the DMSO band represents 100%. Data represent average ± SD for n = 3 trials. (D)

HeLa mitochondria are treated as in ‘2A’ after treatment with Analog- or -4. (E) HeLa cells overexpressing EGFP-Parkin were treated as in ‘2B’ after treatment with Analog-3 or -4. (F)

Representative fluorescent images of HeLa cells overexpressing EGFP-Parkin after treatement with MB-10, Analog-3, or -4 similar to ‘2F’. Scale bar = 15 m. (G) Quantification of Parkin translocation to mitochondria in ‘F’ by cell counting. 100 cells were counted for each condition; data represent the average ± SD of n = 3 trials.

81

Figure 3.5. TIMM44 knockdown inhibits PINK1 association with the TOM complex but does not interfere with normal PINK1 import. (A) HeLa cells are treated with either scramble or TIMM44 siRNA. The mitochondria are isolated from these cells, followed by import of radiolabeled Su9-DHFR. An immunoblot of the mitochondria is also included to demonstrate

TIMM44 knockdown. (B) HeLa cells are treated with siRNA as in ‘A’, followed by radiolabeled

PINK1 import into the isolated mitochondria. Import reactions are run on both SDS-PAGE and

82

BN-PAGE as in ‘2A’. An immunoblot of the mitochondria is included to demonstrate TIMM44 knockdown. (C) Quantification of the ratio of % import inhibition to percent TIMM44 knockdown for each experiment in ‘A’ and ‘B’. The ratio for Su9-DHFR import is set to 1.0. (D-

F) Schematic depicting the activities of both TIMM44 and MB-10 (D) PINK1 is partially imported into healthy mitochondria, cleaved by PARL, and degraded by the proteasome.

TIMM44 plays no role. (E) PINK1 accumulates on the OM on dysfunctional mitochondria

(typically by uncoupler treatment) followed by Parkin recruitment and mitophagy. TIMM44 is a necessary component for PINK1 stability and therefore Parkin recruitment. (F) Treatment with

MB-10, a TIMM44 inhibitor, blocks the accumulation of PINK1 on damaged mitochondria in a form that precludes Parkin recruitment.

83

Figure 3.S1. Human and yeast Tim44 are highly conserved. (A) MB-10 structure. (B) A

sequence comparison between yeast Tim44 and human TIMM44 (lacking the MTS). 84

Information about the conservation score is provided above the sequences; single dot = dissimilar residues, two dots = similar residues, straight line = identical residues. The 7 residues that were identified as important for MB-10 binding26 are highlighted in red. (C) Isolated HeLa

mitochondria were incubated in normal buffer (Mitos), hypotonic buffer to rupture the OM and

generate mitoplasts (Plasts), or Triton X-100 (TX-100) for lysis in the presence and absence of

proteinase K. The pellet (P) was separated from the supernatant (S) and fractionated by SDS-

PAGE. The indicated proteins were detected by immunoblotting. Controls include presequence

peptidase (PreP) for the matrix, YME1L for the IMS, and TOMM70 for the OM. (D) Isolated

HeLa mitochondria were incubated in normal buffer or 100 mM Na 2CO 3 at the indicated pH.

Samples were processed as in ‘B’; markers include PreP (soluble), TIMM23 (integral), and AIF

(peripheral). (E) Radiolabeled PINK1 or Su9-DHFR were imported into isolated HeLa mitochondria. Post-import, the samples were treated with indicated concentrations of trypsin on ice for 15 min before fractionation by SDS-PAGE. Abbreviations, p, precursor and m, mature forms of Su9-DHFR; FL-PINK1 for full-length PINK1 and PINK* marking the PARL-cleaved form.

85

Figure 3.S2. MB-10 inhibits the import of precursors that use the TIM23 import pathway.

(A, C, E) Representative images of time course imports for radiolabeled precursors into isolated

HeLa mitochondria after a 15 min incubation with MB-10 or DMSO; -∆ψ = 50 M CCCP.

Precursors included Su9-DHFR (matrix), MFN1-myc (OM), and CHCHD2 (IMS). (B, D, F)

Quantification of the imports from ‘A’, ‘C’, and ‘E’; data represent the average ± SD of n = 3

trials. In ‘A’, the ‘m’ band was quantified. The longest time point at 0 M MB-10 band represents 100% import for ‘B’ and ‘D’, and the -∆ψ band represents 100% for

‘F’.Abbreviations, p, precursor and m, mature forms of Su9-DHFR.

86

Figure 3.S3. MB-10 inhibits CCCP-induced mitophagy. Representative images of HeLa cells with stable overexpression of EGFP-Parkin after treatment with indicated concentrations of MB-

10 for 2 hrs, followed by the addition of either DMSO of 10 M CCCP for an additional 24 hrs. 87

Both nuclear and mitochondrial DNA (as cytosolic puncta representative of nucleoids) were stained with an anti-DNA antibody. MitoTracker red marks mitochondria with a membrane potential. Scale bar = 7 m.

88

Acknowledgments

We thank Dr. Janos Steffen and Dr. Juwina Wijaya for technical assistance as well as Dr.

Richard Youle for the EGFP-Parkin HeLa cell line and the GST-MFN-1 antibody.

This work was supported by National Institutes of Health Grants DK101780 and GM61721,

California Institute of Regenerative Medicine Grants RT307678, and USAF-AFRL FA9550-15-

1-0406 (to C. M. K.); U.S. Public Health Service NRSA Grant T32GM07185 (to M.A.C.).

The authors declare no competing financial interests.

89

References

1. Springer, M. Z.; Macleod, K. F., In Brief: Mitophagy: mechanisms and role in human disease. J Pathol 2016, 240 (3), 253-255.

2. Pryde, K. R.; Smith, H. L.; Chau, K. Y.; Schapira, A. H., PINK1 disables the anti-fission machinery to segregate damaged mitochondria for mitophagy. J Cell Biol 2016, 213 (2), 163-71. 3. Pickrell, A. M.; Youle, R. J., The roles of PINK1, parkin, and mitochondrial fidelity in Parkinson's disease. Neuron 2015, 85 (2), 257-73.

4. Matsuda, N.; Sato, S.; Shiba, K.; Okatsu, K.; Saisho, K.; Gautier, C. A.; Sou, Y. S.; Saiki, S.; Kawajiri, S.; Sato, F.; Kimura, M.; Komatsu, M.; Hattori, N.; Tanaka, K., PINK1 stabilized by mitochondrial depolarization recruits Parkin to damaged mitochondria and activates latent Parkin for mitophagy. J Cell Biol 2010, 189 (2), 211-21.

5. Jin, S. M.; Youle, R. J., The accumulation of misfolded proteins in the mitochondrial matrix is sensed by PINK1 to induce PARK2/Parkin-mediated mitophagy of polarized mitochondria. Autophagy 2013, 9 (11), 1750-7.

6. Sarraf, S. A.; Raman, M.; Guarani-Pereira, V.; Sowa, M. E.; Huttlin, E. L.; Gygi, S. P.; Harper, J. W., Landscape of the PARKIN-dependent ubiquitylome in response to mitochondrial depolarization. Nature 2013, 496 (7445), 372-6.

7. Chan, N. C.; Salazar, A. M.; Pham, A. H.; Sweredoski, M. J.; Kolawa, N. J.; Graham, R. L.; Hess, S.; Chan, D. C., Broad activation of the ubiquitin-proteasome system by Parkin is critical for mitophagy. Hum Mol Genet 2011, 20 (9), 1726-37.

8. Narendra, D.; Tanaka, A.; Suen, D. F.; Youle, R. J., Parkin is recruited selectively to impaired mitochondria and promotes their autophagy. J Cell Biol 2008, 183 (5), 795-803.

9. Yamano, K.; Matsuda, N.; Tanaka, K., The ubiquitin signal and autophagy: an orchestrated dance leading to mitochondrial degradation. EMBO Rep 2016, 17 (3), 300-16.

10. Lazarou, M.; Sliter, D. A.; Kane, L. A.; Sarraf, S. A.; Wang, C.; Burman, J. L.; Sideris, D. P.; Fogel, A. I.; Youle, R. J., The ubiquitin kinase PINK1 recruits autophagy receptors to induce mitophagy. Nature 2015, 524 (7565), 309-14.

11. Greene, A. W.; Grenier, K.; Aguileta, M. A.; Muise, S.; Farazifard, R.; Haque, M. E.; McBride, H. M.; Park, D. S.; Fon, E. A., Mitochondrial processing peptidase regulates PINK1 processing, import and Parkin recruitment. EMBO Rep 2012, 13 (4), 378-85.

12. Deas, E.; Plun-Favreau, H.; Gandhi, S.; Desmond, H.; Kjaer, S.; Loh, S. H.; Renton, A. E.; Harvey, R. J.; Whitworth, A. J.; Martins, L. M.; Abramov, A. Y.; Wood, N. W., PINK1 cleavage at position A103 by the mitochondrial protease PARL. Hum Mol Genet 2011, 20 (5), 867-79.

90

13. Yamano, K.; Youle, R. J., PINK1 is degraded through the N-end rule pathway. Autophagy 2013, 9 (11), 1758-69.

14. Lazarou, M.; Jin, S. M.; Kane, L. A.; Youle, R. J., Role of PINK1 Binding to the TOM Complex and Alternate Intracellular Membranes in Recruitment and Activation of the E3 Ligase Parkin. Dev Cell 2012, 22 (2), 320-33.

15. Kondapalli, C.; Kazlauskaite, A.; Zhang, N.; Woodroof, H. I.; Campbell, D. G.; Gourlay, R.; Burchell, L.; Walden, H.; Macartney, T. J.; Deak, M.; Knebel, A.; Alessi, D. R.; Muqit, M. M., PINK1 is activated by mitochondrial membrane potential depolarization and stimulates Parkin E3 ligase activity by phosphorylating Serine 65. Open Biol 2012, 2 (5), 120080.

16. Kazlauskaite, A.; Kondapalli, C.; Gourlay, R.; Campbell, D. G.; Ritorto, M. S.; Hofmann, K.; Alessi, D. R.; Knebel, A.; Trost, M.; Muqit, M. M., Parkin is activated by PINK1-dependent phosphorylation of ubiquitin at Ser65. Biochem J 2014, 460 (1), 127-39.

17. Koyano, F.; Okatsu, K.; Kosako, H.; Tamura, Y.; Go, E.; Kimura, M.; Kimura, Y.; Tsuchiya, H.; Yoshihara, H.; Hirokawa, T.; Endo, T.; Fon, E. A.; Trempe, J. F.; Saeki, Y.; Tanaka, K.; Matsuda, N., Ubiquitin is phosphorylated by PINK1 to activate parkin. Nature 2014, 510 (7503), 162-6.

18. Ordureau, A.; Sarraf, S. A.; Duda, D. M.; Heo, J. M.; Jedrychowski, M. P.; Sviderskiy, V. O.; Olszewski, J. L.; Koerber, J. T.; Xie, T.; Beausoleil, S. A.; Wells, J. A.; Gygi, S. P.; Schulman, B. A.; Harper, J. W., Quantitative proteomics reveal a feedforward mechanism for mitochondrial PARKIN translocation and ubiquitin chain synthesis. Mol Cell 2014, 56 (3), 360- 75.

19. Gottschalk, W. K.; Lutz, M. W.; He, Y. T.; Saunders, A. M.; Burns, D. K.; Roses, A. D.; Chiba-Falek, O., The Broad Impact of TOM40 on Neurodegenerative Diseases in Aging. J Parkinsons Dis Alzheimers Dis 2014, 1 (1).

20. Honlinger, A.; Bomer, U.; Alconada, A.; Eckerskorn, C.; Lottspeich, F.; Dietmeier, K.; Pfanner, N., Tom7 modulates the dynamics of the mitochondrial outer membrane translocase and plays a pathway-related role in protein import. Embo J 1996, 15 (9), 2125-37.

21. Hasson, S. A.; Kane, L. A.; Yamano, K.; Huang, C. H.; Sliter, D. A.; Buehler, E.; Wang, C.; Heman-Ackah, S. M.; Hessa, T.; Guha, R.; Martin, S. E.; Youle, R. J., High-content genome- wide RNAi screens identify regulators of parkin upstream of mitophagy. Nature 2013, 504 (7479), 291-5.

22. Filipuzzi, I.; Steffen, J.; Germain, M.; Goepfert, L.; Conti, M. A.; Potting, C.; Cerino, R.; Pfeifer, M.; Krastel, P.; Hoepfner, D.; Bastien, J.; Koehler, C. M.; Helliwell, S. B., Stendomycin selectively inhibits TIM23-dependent mitochondrial protein import. Nat Chem Biol 2017, 13 (12), 1239-1244.

91

23. Okatsu, K.; Kimura, M.; Oka, T.; Tanaka, K.; Matsuda, N., Unconventional PINK1 localization to the outer membrane of depolarized mitochondria drives Parkin recruitment. J Cell Sci 2015, 128 (5), 964-78.

24. Schapira, A. H.; Olanow, C. W.; Greenamyre, J. T.; Bezard, E., Slowing of neurodegeneration in Parkinson's disease and Huntington's disease: future therapeutic perspectives. Lancet 2014, 384 (9942), 545-55.

25. Hasson, S. A.; Fogel, A. I.; Wang, C.; MacArthur, R.; Guha, R.; Heman-Ackah, S.; Martin, S.; Youle, R. J.; Inglese, J., Chemogenomic profiling of endogenous PARK2 expression using a genome-edited coincidence reporter. ACS Chem Biol 2015, 10 (5), 1188-97.

26. Jin, S. M.; Lazarou, M.; Wang, C.; Kane, L. A.; Narendra, D. P.; Youle, R. J., Mitochondrial membrane potential regulates PINK1 import and proteolytic destabilization by PARL. J Cell Biol 2010, 191 (5), 933-42.

27. Miyata, N.; Tang, Z.; Conti, M. A.; Johnson, M. E.; Douglas, C. J.; Hasson, S. A.; Damoiseaux, R.; Chang, C. A.; Koehler, C. M., Adaptation of a Genetic Screen Reveals an Inhibitor for Mitochondrial Protein Import Component Tim44. J Biol Chem 2017, 292 (13), 5429-5442.

28. Josyula, R.; Jin, Z.; Fu, Z.; Sha, B., Crystal structure of yeast mitochondrial peripheral membrane protein Tim44p C-terminal domain. J Mol Biol 2006, 359 (3), 798-804.

29. Callegari, S.; Oeljeklaus, S.; Warscheid, B.; Dennerlein, S.; Thumm, M.; Rehling, P.; Dudek, J., Phospho-ubiquitin-PARK2 complex as a marker for mitophagy defects. Autophagy 2017, 13 (1), 201-211.

30. Wang, Y.; Katayama, A.; Terami, T.; Han, X.; Nunoue, T.; Zhang, D.; Teshigawara, S.; Eguchi, J.; Nakatsuka, A.; Murakami, K.; Ogawa, D.; Furuta, Y.; Makino, H.; Wada, J., Translocase of inner mitochondrial membrane 44 alters the mitochondrial fusion and fission dynamics and protects from type 2 diabetes. Metabolism 2015, 64 (6), 677-88.

31. Koehler, C. M.; Jarosch, E.; Tokatlidis, K.; Schmid, K.; Schweyen, R. J.; Schatz, G., Import of mitochondrial carriers mediated by essential proteins of the intermembrane space. Science 1998, 279 (5349), 369-373. 1. Filipuzzi, I.; Steffen, J.; Germain, M.; Goepfert, L.; Conti, M. A.; Potting, C.; Cerino, R.; Pfeifer, M.; Krastel, P.; Hoepfner, D.; Bastien, J.; Koehler, C. M.; Helliwell, S. B., Stendomycin selectively inhibits TIM23-dependent mitochondrial protein import. Nat Chem Biol 2017, 13 (12), 1239-1244.

32. Gegg, M. E.; Cooper, J. M.; Schapira, A. H.; Taanman, J. W., Silencing of PINK1 expression affects mitochondrial DNA and oxidative phosphorylation in dopaminergic cells. PLoS One 2009, 4 (3), e4756.

92

Chapter 4: DECA induces PINK1 accumulation and Parkin recruitment

Introduction

In addition to MB-10, which inhibits the accumulation of PINK1 on mitochondria, we have also discovered the FDA-approved DECA 1, a small molecule that induces the accumulation of PINK1. Whereas the binding target and mechanism of action of MB-10 are established 2, the target of DECA remains elusive. The molecule is still very interesting because ARPD is associated with PINK1 mutations that drastically reduce its function, so initiating further recruitment of the kinase to mitochondria may serve as a potential therapy for ARPD. Here we demonstrate that DECA causes PINK1 accumulation, Parkin recruitment, and associated downstream events in a similar manner to CCCP, but DECA does not dissipate the mitochondrial membrane potential as CCCP does.

Results

In precursor protein import assays, DECA (Figure 4.1A) strongly inhibited the import of the matrix-targeted precursor Su9-DHFR, partially inhibited the import of the IMS-targeted precursor CHCHD2, and failed to impair the import of the OM-targeted precursor MFN-1

(Figure 4.1B-G). This inhibition profile indicates that DECA targets a component of the import machinery that plays an important role in TIM23-dependent import and a minor role in redox- regulated import, but no role in import to the OM. This suggests that the target of DECA is a protein that is located in either the IMS or the IM, perhaps an IMS domain of one of the proteins of the TIM23 complex, and the target is very unlikely to be a part of the TOM complex.

93

Treatment of intact HeLa cells with DECA resulted in stabilization of PINK1 on the mitochondrial OM (Figure 4.2A). In HeLa cells expressing EGFP-Parkin, phospho- ubiquitination of Parkin was detected after an 8 hr treatment with 20 µM DECA, which was identical to treatment with CCCP or valinomycin (Figure 4.2B). Activation of Parkin by phospho-ubiquitination resulted in degradation of proteins such as MFN-1, TOMM20,

TOMM40, and YME1L, indicating that DECA induced mitophagy in a manner similar to treatment with uncouplers (Figure 4.2B). DECA also inhibited the formation of the PARL- cleaved form of PINK1 in import assays (Figure 4.2C,D), but did not increase the amount of FL-

PINK1 (Figure 4.2E,F), similar to CCCP treatment. We further confirmed that DECA initiated

Parkin recruitment to mitochondria in the EGFP-Parkin HeLa cells using fluorescent microscopy

(Figure 4.2G). At 2-4 M DECA, Parkin was not markedly recruited to mitochondria, failing to

co-localize with TOMM20. However, 10 M DECA treatment resulted in robust Parkin

recruitment to mitochondria, similar to CCCP treatment. As illustrated by MitoTracker staining,

the mitochondrial network was fragmented in both DECA and CCCP treated cells, correlating

with mitophagy induction, but the membrane potential was only maintained with DECA

treatment, in contrast to uncoupler CCCP.

We extended this study by determining if DECA treatment preferentially recruited Parkin

to impaired mitochondria rather than healthy mitochondria. The EGFP-Parkin HeLa cells were

treated with a titration of DECA in the presence and absence of 5 M CCCP, a dose that does not

uncouple mitochondria or result in Parkin recruitment to mitochondria (Figure 4.3A,B). In the

absence of CCCP, 7 M DECA was required to initiate Parkin recruitment to mitochondria.

Conversely, with the addition of 5 M CCCP, Parkin recruitment was initiated at a reduced

DECA concentration of 4 M, with no significant change in Parkin localization as the DECA

94 concentration was increased. In HeLa cells lacking Parkin, PINK1 stabilization on the mitochondrial OM mirrored that of the HeLa cells expressing EGFP-Parkin (Figure 4.3C), further suggesting that DECA seems to selectively trigger mitophagy in already-compromised mitochondria.

We also tested the consequences of a combination of DECA and MB-10 treatment in

HeLa cells (Figure 4.4). The PINK1 stabilization on mitochondria induced by treatment with 10

M DECA was enhanced in the presence of 20 M MB-10. This result was unexpectedly the

opposite of that with an MB-10/CCCP combination even though both DECA and CCCP cause

PINK1 stabilization, implying that events with DECA and its target may be dominant to that of

MB-10 and TIMM44, or DECA may act upstream of the interaction between PINK1 and

TIMM44.

Discussion

In this study, we characterize DECA as a useful probe for inducing the accumulation of

PINK1 at mitochondria, the subsequent recruitment of Parkin, and the induction of mitophagy

that is well tolerated in cell models. To this point, uncouplers and respiratory inhibtors such as

CCCP, valinomycin, or antimycin and oligomycin have been used to induce the accumulation of

PINK1, but these molecules dramatically alter cellular metabolism. DECA can serve as a useful

research tool to work around the negative side effects of such mitochondrial toxins to study

PINK1-dependent mitophagy. Additionally, because DECA is an already FDA-approved

compound, there remains the possibility that it could be develop into a therapeutic for some cases

of ARPD. An important future goal is to determine the binding target of DECA and elucidate its

mechanism of action. This would constitute a large step in furthering it as both a research tool

and drug.

95

Materials and Methods

Cell culture and transfection

HeLa cells including those with stable expression of EGFP-Parkin 3 were cultured in

Dulbecco’s modified eagle medium (DMEM), 25 mM glucose, 1 mM pyruvate (Life

Technologies 11995-065), supplemented with heat-inactivated 10% FBS (v/v; Atlanta

Biologicals S11150H) and 1% penicillin/streptomycin (Life Technologies 15140122) in a

humidified incubator at 37˚C under 5% CO 2 (v/v). For all experiments, 10 cm dishes

(CELLSTAR 664-160) were used with 10 mL of media or 6 cm dishes (CELLSTAR 628-160) with 2.5 mL of media.

Small molecule treatment with cultured cells

Small molecules are dissolved in 100% DMSO in brown glass vials (Wheaton 224750) and stored in a dessicator at room temperature. The DECA stock was 4 mM; fresh stocks were made monthly. Cells in culture were grown to 100% confluency followed by seeding to 25% confluency. After 48 hrs, cells were treated with indicated small molecules dissolved in fresh media; the final DMSO for all conditions was 1%, unless otherwise noted. At the time points indicated in each figure, media was aspirated and the cells were washed with PBS.

Mitochondrial purification for immunoblotting analysis

Cells were removed from the culture dishes by incubating in Trypsin-EDTA (0.05%, Life

Technologies 25300-054) for 5 min at 37˚C. Cells were collected in media and centrifuged at

770 x g for 5 min. at 4˚C. Cells were washed with ice-cold PBS (Life Technologies, 14190-144) and centrifuged again. The pellet was resuspended in 1 ml homogenizer buffer (20 mM HEPES-

KOH pH 7.4, 220 mM mannitol, 70 mM sucrose) supplemented with 2 mg/mL BSA and 0.5 mM

PMSF. Cells were lysed on ice by 20 strokes through a 25-gauge x 5/8 needle (Becton Dickson

96

305122) and 1 mL syringe followed by centrifugation at 770 x g for 5 min. at 4˚C to remove unbroken cells. The supernatant was centrifuged at 10,000 x g for 10 min at 4˚C, washed with homogenizer buffer (lacking BSA or PMSF), and re-centrifuged at 10,000 x g for 10 min at 4˚C to obtain mitochondrial pellets. The pellets were resuspended in 100 µl of homogenizer buffer, and the protein concentration was determined using the Pierce BCA Protein Assay Kit

(Thermofisher). For each condition, 20 g of mitochondria was aliquoted into tubes and centrifuged at 10,000 x g for 10 min at 4˚C. The supernatant was removed and mitochondria were disrupted with Laemmli-SDS sample buffer; the proteins were separated by SDS-PAGE followed by transfer to PVDF membranes, and analyzed by immunoblotting for indicated proteins.

The antibodies used in this study include anti-Mortalin (UC Davis/NIH Neuromab

Facility clone N52A/42, 75-127), anti-Parkin (Santa Cruz Biotechnology, sc-32283), anti-PINK1

(Novus Biologicals clone 8E10.1D6, NBP2-36488), anti-TOMM20 (Santa Cruz Biotechnology, sc-11415), and anti-LRP-130 (Santa Cruz Biotechnology, sc-66844). The GST-MFN-1 (gift from Richard Youle), anti-PreP, anti-YME1L, and anti-TOMM40 were polyclonal antibodies generated by vendor Pacific Immunology from recombinant proteins.

Mitochondrial purification and in vitro protein import assays

Mitochondria isolation and imports were based on established methods 4. Samples were prepared as described for lysis with a syringe but a Teflon pestle (Bellco) was used.

[35 S]-labeled proteins were synthesized in vitro via a coupled transcription/translation rabbit reticulocyte kit (Promega) incubated at 30˚C for 90 min with [ 35 S]Met/Cys metabolic

labeling reagent (MP Biomedicals, 015100907) and plasmid DNA expressing human PINK1,

CHCHD2, or Su9-DHFR via an Sp6 promoter or MFN1-myc via a T7 promoter.

97

Isolated mitochondria were diluted to 0.4 g/ L in homogenizer buffer (70 g per

reaction) supplemented with 17.5 L of 10X energy generating solution (10 mM ATP, 5 mM magnesium chloride, 200 mM sodium succinate, 50 mM NADH in homogenizer buffer) and

0.1% BSA adjusted to a volume of 140 L. DMSO vehicle or compounds dissolved in DMSO

(1% final) were added to the mitochondria and incubated at 25˚C for 15 min. Reactions were then initiated by addition of 35 uL of precursor and aliquots were withdawn and stopped at specified time points.

The import of Su9-DHFR was terminated by the addition of ice-cold 10 g/mL trypsin

(Sigma) in homogenizer buffer. After 15 min on ice, soybean trypsin inhibitor (VWR Amresco) was added at 50 g/mL. Mitochondria are pelleted by centrifugation at 10,000 x g for 5 min at

4˚C. Import of OM proteins (PINK1 and MFN1-myc) was terminated by the addition of ice-cold homogenizer buffer and subsequent centrifugation at 10,000 x g for 5 min. Samples were then

5 subject to carbonate extraction in 100 mM Na 2CO 3 at pH 11 . For all precursors, after final

recovery by centrifugation, mitochondria were disrupted in Laemmli SDS-sample buffer and

analyzed by SDS-PAGE and autoradiography.

Immunofluorescence microscopy

Thirty min prior to fixing, MitoTracker Red CMXRos (Invitrogen, M7512) was added to

each well to 20 nM final concentration. Cells on D-poly-Lysine coated coverslips were then

fixed with pre-warmed 3.7% formaldehyde (v/v; VWR, VW3408-1), permeabilized with ice-

cold methanol, and blocked with 1% fatty acid free BSA in PBS (v/v, PBS-BSA).

Where indicated, cells were subsequently incubated overnight at 4˚C with indicated

primary antibodies diluted in PBS-BSA. Following thorough washing in PBS, cells were

incubated for 1 hr at room temperature with Alexa Fluor 350-conjugated secondary antibodies

98

(Life Technologies) diluted in PBS-BSA. The anti-TOMM20 primary antibody with Alexa Fluor

350 goat-anti-rabbit IgG (A11046) was used. Cells were then set on slides and images were obtained using a Leica TCS SPE DMI 4000B inverted confocal microscope.

Statistical analyses

Data was plotted using GraphPad Prism version 5.02. Statistical significance of observations was determined using unpaired t tests (two-tailed), unless otherwise noted.

99

Figures

Figure 4.1. DECA strongly inhibits the import of precursors that use the TIM23 import pathway and modestly inhibits the import of precursors that use redox-regulated pathway.

(A) The chemical structure of DECA. (B, D, F) Representative images of time course imports for radiolabeled precursors into isolated HeLa mitochondria after a 15 min incubation with DECA or

DMSO; -∆ψ = 50 M CCCP. Precursors included Su9-DHFR (matrix), MFN1-myc (OM), and

CHCHD2 (IMS). (C, E, G) Quantification of the imports from ‘B’, ‘D’, and ‘F’; data represent

the average ± SD of n = 3 trials. In ‘B’, the ‘m’ band was quantified. The longest time point at 0

M DECA band represents 100% import for ‘C’ and ‘E’, and the -∆ψ band represents 100% for

‘G’. Abbreviations, p, precursor and m, mature forms of Su9-DHFR. 100

Figure 4.2. DECA causes accumulation of PINK1 on mitochondria. (A) Intact HeLa cells were treated with DECA, 10 M MG132, 10 M CCCP, or 1 g/ L valinomycin. The

mitochondrial fraction was separated by SDS-PAGE followed by immunoblotting. Additional

markers include LRP-130 and PreP (matrix); YME1L (IMS); and TOMM40 (OM). Panels of

short and long exposure were included to mark FL-PINK. (B) Intact HeLa cells with stable

overexpression of EGFP-Parkin were treated as in ‘A’. Ph-Ubi-Parkin marks the phospho- 101 ubiquitinated form and Ubi-MFN-1 marks the ubiquitinated form. Additional marker protein includes TOMM20 (OM). (C) A representative image of radiolabeled PINK1 imported into isolated HeLa mitochondria after a 15 min incubation in DECA or DMSO. (D) Quantification of the PARL-cleaved band. The band at 0 M DECA at 20 min represents 100% for each trial. Data

represent the average ± SD of n = 3 trials (* P<0.05). (E) A lighter exposure of the image

displayed in ‘C’. (F) Quantification as described in ‘C’ of the FL-PINK1 band from ‘E’; the -∆ψ band represents 100% for each trial. (G) Representative images of HeLa cells with stable overexpression of EGFP-Parkin after a 4 hr treatment with indicated concentrations of DECA or

CCCP; scale bar = 20 m.

102

Figure 4.3. DECA promotes Parkin recruitment to mitochondria that are already partially dysfunctional. (A) Representative images of HeLa cells with stable overexpression of EGFP- 103

Parkin after an 8 hr treatment with indicated concentrations of DECA, CCCP, or combination.

Scale bar = 7 m. (B) Quantification of Parkin translocation to mitochondria in ‘A’ by cell counting. 100 cells were counted for each condition; data represent the average ± SD of n = 4

trials (* P<0.05 and *** P<0.001). (C) Intact HeLa cells were treated as in ‘A’. Cells were

collected, mitochondria isolated on ice and fractionated by SDS-PAGE, and indicated proteins

detected by immunoblot. FL-PINK1, full-length PINK1; PINK*, PARL-cleaved form.

104

Figure 4.4. MB-10 enhances the activity of DECA. (A) Intact HeLa cells were treated with

DMSO or 20 M MB-10 for 2 hrs, followed by the addition of DMSO, 10 M MG132, or 10 M

DECA for an additional 8 hrs. Cells were then collected, mitochondria isolated on ice and fractionated by SDS-PAGE, and indicated proteins detected by immunoblot.

105

References

1. Miyata, N.; Steffen, J.; Johnson, M. E.; Fargue, S.; Danpure, C. J.; Koehler, C. M., Pharmacologic rescue of an -trafficking defect in primary hyperoxaluria 1. Proc Natl Acad Sci U S A 2014, 111 (40), 14406-11.

2. Miyata, N.; Tang, Z.; Conti, M. A.; Johnson, M. E.; Douglas, C. J.; Hasson, S. A.; Damoiseaux, R.; Chang, C. A.; Koehler, C. M., Adaptation of a Genetic Screen Reveals an Inhibitor for Mitochondrial Protein Import Component Tim44. J Biol Chem 2017, 292 (13), 5429-5442.

3. Narendra, D.; Tanaka, A.; Suen, D. F.; Youle, R. J., Parkin is recruited selectively to impaired mitochondria and promotes their autophagy. J Cell Biol 2008, 183 (5), 795-803.

4. Hasson, S. A.; Kane, L. A.; Yamano, K.; Huang, C. H.; Sliter, D. A.; Buehler, E.; Wang, C.; Heman- Ackah, S. M.; Hessa, T.; Guha, R.; Martin, S. E.; Youle, R. J., High-content genome-wide RNAi screens identify regulators of parkin upstream of mitophagy. Nature 2013, 504 (7479), 291-5.

5. Koehler, C. M.; Jarosch, E.; Tokatlidis, K.; Schmid, K.; Schweyen, R. J.; Schatz, G., Import of mitochondrial carriers mediated by essential proteins of the intermembrane space. Science 1998, 279 (5349), 369-73.

106

Chapter 5: A small molecule inhibitor of mitochondrial protein import possibly acts through transport protein Seo1

Abstract

Protein translocation pathways into mitochondria have been well established, but new methods for study are needed to further the field. There are a growing number of small molecule probes that inhibit protein translocation that were discovered in the laboratory of Dr. Carla

Koehler. This series of small molecules, known as the Mitochondrial import Blockers from the

Carla Koehler laboratory (MitoBloCK ; MB), includes probes for the PAM complex and redox- regulated import. We now describe the newest addition to this series, MB-11, that seems to act in a unique mechanism. Through a series of in vitro assays and a next generation sequencing approach, we have evidence that the molecule binds to Seo1, a transport protein that has been previously reported to be at the plasma membrane. We believe that Seo1 could have dual- localization to both mitochondria and the plasma membrane.

Introduction

Of the approximately 1,000 mitochondrial proteins in yeast 1, only about 1% are encoded in the mtDNA. The rest of the mitochondrial proteins are encoded in nuclear DNA and synthesized on cytosolic ribosomes to be imported into the mitochondria 2. Because mitochondria have two membranes and two aqueous compartments, the final destination of precursor proteins varies between the outer membrane (OM), inter membrane space (IMS), inner membrane (IM), and the matrix, necessitating multiple pathways of import. These protein import and sorting

107 machineries are vital for mitochondrial biosynthesis and for the synthesis of iron-sulfur clusters 3,

making these pathways attractive topics of research.

One of the more established import pathways involves proteins destined for the

mitochondrial matrix. Proteins that follow this pathway contain an N-terminal amphipathic helix

that serves as a mitochondrial targeting sequence (MTS) 4. This pathway utilizes the translocase

of the outer membrane (TOM) and the presequence translocase of the inner membrane (TIM23)

and its associated motor (PAM) complexes to reach the matrix. The TOM complex consists of

the pore Tom40, receptor proteins Tom22, Tom20, an associated receptor Tom70, and small

proteins that are important for dynamics and complex formation Tom5, Tom6, and Tom7 5, 6 .

Tom20 is traditionally considered the receptor for precursor proteins with an amphipathic helix 7, whereas Tom70 serves as the receptor for proteins that have transmembrane domains 8, 9 .

Precursor proteins are then transferred to the cytosolic domain of Tom22 10 prior to translocation through the Tom40 pore. Tom22 also has an IMS domain that receives the precursor protein as it exits the Tom40 channel 11 that coordinates with the TIM23 complex.

The TIM23 complex contains the channel-forming protein Tim23 12 , a protein important

for PAM recruitment and lateral sorting to the IM named Tim17 13 , a protein important for

interacting with the TOM complex named Tim21 14 , and a receptor protein Tim50 15 , among

others. Transport through the TIM23 complex is coupled to the membrane potential ( ∆ψ) between the matrix and IMS 16 . Finally, the PAM complex pulls the precursor into the matrix as it emerges from the TIM23 complex. mtHsp70 drives the movement of the precursor protein by hydrolyzing ATP with the aid of the nucleotide exchange factor Mge1, and J protein Pam18, and

J-like protein Pam16 17 . Additionally, Tim44 of the PAM complex serves as a recognition site for

mtHsp70 to link the chaperone to the TIM23 channel, and also binds the precursor as it emerges

108 from the channel 18, 19 . Once in the matrix, the presequences are cleaved by the protease MPP 20

and degraded to small fragments.

Although much is known about the presequence import pathway, additional methods are

needed to further our knowledge. To this end, we have performed a screen involving the

localization of fusion protein Su9-Ura3 to identify small molecule inhibitors of the TOM and

TIM23 complexes. The screening strategy was designed to identify import inhibitors by

identifying those molecules that conferred yeast growth in media lacking uracil, because the Su9-

Ura3 would be blocked from reaching the mitochondrial matrix. From this screen, we recovered

68 hit molecules, including MB-10, which we have previously characterized as binding to

Tim44 21 , and DECA, an FDA-approved drug 22 . We have now characterized another molecule from the screen, MB-11, that we believe targets Seo1, a transport protein that was previously believed to localize only to the yeast plasma membrane 23, 24 . We propose that Seo1 may also be

localized to the mitochondria and our molecule MB-11 inhibits its function.

Results

MB-11 confers growth in media lacking uracil in a specific manner

The small molecule MB-11 (Figure 5.1A) conferred growth for the screening strain in

media lacking uracil starting at around 1 M, with growth peaking at around 2.5 M (Figure

5.S1A). After this concentration, the positive growth phenotype began to diminish, most likely due to too much import inhibition or off-target effects. In the parental yeast strain, the MIC 50 of

MB-11 is around 0.67 M (Figure 5.S1B), indicating that the mechanism of inhibition is moderately deleterious for the yeast. We also need to ensure that the molecule is not acting on the mitochondria in a general way, such as by dissipating the membrane potential or solubilizing the membranes. Because the membrane potential is used to synthesize ATP by converting O 2 to

109

H2O, we can measure oxygen consumption to determine if the potential is still intact. When comparing the rate at which O 2 is utilized by isolated mitochondria, it is clear that MB-11 has no

effect relative to the vehicle control DMSO (Figure 5.S2C). Finally, we incubated isolated

mitochondria from both yeast and HeLa cells in either MB-11 or vehicle control before

separating the pellet from the supernatant. There was no affect of proteins leaking into the

supernatant upon MB-11 treatment by either Coomassie staining or immunoblot (Figure 5.S1D-

G). This data indicates that MB-11 does inhibit the import of Su9-Ura3 by conferring growth in

the screen and does so by targeting a specific protein rather than via general mitochondrial

alterations.

MB-11 targets a protein on the OM and only inhibits precursor proteins with hydrophobic

segments

To determine the protein complex that MB-11 targets, we performed a series of import

assays into isolated yeast mitochondria using S 35 -labeled precursor proteins synthesized in vitro

with a transcription/translation mix (Figure 5.1). The import of proteins that are inhibited by

MB-11 include Su9-Ura3, AAC, Tom40, and cyt b 2(167)A63P-DHFR, all of which contain

hydrophobic domains but are targeted to different compartment. In contrast, Su9-DHFR and

Cmc1 import is not inhibited, and they are also targeted to different compartment. Although Su9-

Ura3 is targeted to the matrix and uses the TIM23 pathway, AAC is targeted to the IM and uses

the TIM22 pathway 25, 26 , and TOM40 is targeted to the OM and uses the SAM complex 27 , they all use the TOM complex for initial entry into the mitochondria. Therefore, we conclude that

MB-11 must be binding to a component of the TOM complex, or at least a protein on the OM.

To further confirm this, we imported the screening precursor, Su9-Ura3, into mitoplasts, which are mitochondria that have undergone a hypotonic treatment to make them swell and

110 effectively ablate the OM (Figure 5.2A). This allows precursor proteins to bypass import through the OM and get directly imported through the IM. In this assay, MB-11 no longer inhibits Su9-

Ura3 import, further indicating the inhibition is at the OM. Because Tom70 is the receptor protein that initiates import for hydrophobic precursors, we decided to make a ∆tom70 yeast

strain. Using this strain, we tested the various stages of complex formation during Tom40

import 28 by BN-PAGE and compared it to Tom40 import into parent yeast with and without MB-

11 (Figure 5.2B). In the parental yeast strain, assembly intermediate I forms immediately and slowly diminishes over time. Assembly intermediate II begins forming between 10 and 30 minutes and tends to hold steady after that point. The TOM complex itself really only begins to show up at about 60 minutes and continues to develop through 90 minutes. MB-11 does not inhibit the formation of assembly intermediate I, but does seem to inhibit the formation of assembly intermediate and does not allow the Tom complex to form at all. This is very similar to the import profile that is seen with mitochondria from the ∆tom70 strain. This data led us to

believe that Tom70 was the binding target of MB-11.

MB-11 does not target Tom70 or Tom20

In an attempt to confirm that MB-11 binds Tom70 as its target, we used the ∆tom70 yeast

strain and made a yeast strain that over expresses Tom70. We determined the MIC 50 of each of

these strains relative to the parent strain, and saw no difference (Figure 5.2C,D). In theory,

knocking out the target of MB-11 should increase the cell survival relative to the parent because

the inhibition that is causing death will no longer be present. Alternatively, over expressing the

target should also increase the cell survival relative to wild type because there would be more

functional molecules to perform the proteins duty. These results led us to believe that Tom70 is

not, in fact, the target of MB-11, but a different protein on the OM is.

111

To further rule out Tom70 and to test Tom20, we performed an import of Su9-Ura3 into yeast mitochondria that had been exposed to trypsin prior to the import reaction (Figure 5.2E).

As expected, MB-11 inhibited the import of Su9-Ura3 into whole mitochondria. Although import into the trypsin-shaved mitochondria was partially impaired, MB-11 still showed an inhibition.

As can be seen by immunoblot, both Tom70 and Tom20 were completely degraded. These results indicate that neither of the OM receptor proteins are the target of MB-11.

An approach to find mutations that increase yeast growth in MB-11

We decided to take a more comprehensive approach in identifying the binding target of

MB-11 by finding spontaneous revertants that had increased viability in the presence of MB-11

(Figure 5.3A). Yeast were grown overnight in liquid media and diluted into fresh media containing MB-11. These yeast cultures were again grown overnight, plated the following day, and incubated until colonies began to appear. These colonies were validated by determining if they were more viable when grown in MB-11 relative to the parent by performing MIC 50 assays.

Whole genome sequencing was performed on the clones that did increase growth in MB-11 to search for common mutations that may have conferred the viable phenotype.

We performed three rounds of growth in MB-11, an initial screen using YPEG as the growth medium, and optimized second and third screens, one using YPD and the other YPEG again for the growth medium. The initial screen yielded 2 validated revertants (revertants 7.3 and

7.9), the YPD screen yielded 15 revertants (YPD.1-YPD.15), and the second YPEG screen yielded 16 initial hits, but only 7 showed an increase in viability in the MIC 50 assay (YPEG.9-

YPEG.15) (Figure 5.3B-D). A summary of all of the MIC 50 assays performed for each revertant

can be found in table 5.1. Whole genome sequencing was performed on each validated revertant,

112 on the parental strain, the parent of the parental strain, and two strains from the parent that had been made rho null using ethidium bromide, giving 28 total strains.

YAL067C was the most consistently mutated genetic locus

The results indicated that mutations in the gene locus YAL067C occurred the most consistently. This gene encodes for the putative permease Seo1, a member of the allantoate transporter subfamily of the major facilitator superfamily, and prior studies suggest that it is localized to the plasma membrane. The mutations that were conferred included Q26H in four revertants, V185I in one revertant, I430V in eleven revertants, Q26H/P579A in one revertant, and V185I/P579A in one revertant; oddly, the parent of the parent strain also had the I430V mutation. We believe that this set of mutations may form a pocket in which MB-11 is able to bind to inhibit function. A complete list of all mutations located in any exon (non-synonymous mutations only), in any intron, that gain or lose a stop codon, or that remove a start codon from all of the strains that were sequenced can be found in Table 5.2.

We made Seo1 knockout strains to confirm that it is the target of MB-11. We used these strains to perform growth assays in the presence of MB-11 and other small molecules. When comparing the growth of either parent or ∆seo1 to its own vehicle control, the ∆seo1 strain is more viable in MB-11 (Figure 5.4A). This is in contrast to other molecules, where the growth of the ∆seo1 strain is further reduced in MB-10, CCCP, and DECA (Figure 5.4B-D). Because all of these small molecules inhibit mitochondrial protein import following the TIM23 pathway, the data may indicate that Seo1 plays a role in import via a separate mechanism. By removing the

Seo1 pathway, the yeast become more susceptible to inhibition of the TIM23 pathway.

Additionally, the fact that the ∆seo1 strain can grow better in MB-11 suggests that Seo1 is the

binding target because removing it no longer confers inhibition.

113

MB-11 also affects systems beyond yeast

MB-11 also inhibits precursor protein import into mitochondria isolated from HeLa cells

(Figure 5.5). Interestingly, both Su9-Ura3 and Su9-DHFR import is inhibited by MB-11, and the full-length precursor form is protected in each case (Figure5.5A-B). In contrast to yeast mitochondria, the import of the hydrophobic IM carrier protein AAC is not inhibited by MB-11

(Figure 5.5C). Similarly to yeast mitochondria, cyt b 2(167)A63P-DHFR import was inhibited

and Cmc1 import was not inhibited by MB-11 (Figure 5.5D-E). This data suggests that the role

in mitochondrial protein import of the target of MB-11 differs between yeast and mammalian

systems.

When we treat zebrafish embryos with MB-11, there are noticeable phenotypes (Figure

5.S2). The most dramatic difference is the curvature of the organism at 5mM MB-11. In the

presence of vehicle control, the fish have straight body types with a small number of black spots

along the spine. At 2.5mM MB-11, the spine is still straight but there is an increase in the

number of black spots. By 5mM, the fish show a very strong curvature and are smaller in size.

The hearts of the fish also do not develop properly at 5mM MB-11. These results indicate that

inhibition of the target of MB-11 has a drastic effect on organism development, particularly in

the heart where energy dependence is very high.

Discussion

In this study, we have discovered a novel small molecule that inhibits protein import into

mitochondria. Our data suggests that the target of this molecule is Seo1, a putative transport

protein that has been previously believed to be located only at the plasma membrane in yeast. It

is possible that this protein is actually dual-localized, maintaining a population at both the

114 plasma membrane and mitochondrial membrane, and the localization at any given time may be based on the status of the cell.

Interestingly, the profile of protein import inhibition differs between yeast and mammalian mitochondria. In yeast, it seems that proteins that contain hydrophobic segments are exclusively inhibited by MB-11 and proteins that are completely soluble are not hindered. In

HeLa mitochondria, only proteins that have at least a domain targeted to the matrix are inhibited, independent of hydrophobicity. This difference suggests that the yeast and mammalian mitochondrial biogenesis mechanisms are not as conserved as previously thought, although many of the components may still play a role.

It will be interesting to further study the role of Seo1 at mitochondria. Because characterization of the transport protein has been severely limited, very little is established about its function or localization. MB-11 should serve as a useful tool to elucidate its function both in yeast and in mammals.

Materials and Methods

Purification of yeast mitochondria

Mitochondria were purified from yeast cells as previously described 29 . Yeast cultures were grown at a constant 30˚C in YPEG with vigorous shaking to an OD 600 of 3.0. After isolation, the

mitochondria concentration was measured using a BCA assay, diluted to 25 mg/mL, and stored

at -80˚C.

Purification of HeLa mitochondria

Isolation of mammalian mitochondria from HeLa cells was based on established methods 30 . A

Teflon pestle (Bellco) was used in place of a syringe for cell lysis. After isolation, mitochondria

concentration was measured using a BCA assay and used immediately, never frozen.

115

Import of radiolabel precursors

Import of radiolabel precursors was performed according to established methods for both yeast 31 and mammalian 30 mitochondria. The final concentration of DMSO in each reaction was 1%, unless otherwise noted. After final recovery of mitochondria by centrifugation, pellets were disrupted in Laemmli sample buffer, fractionated by SDS-PAGE, and analyzed by autoradiography.

Mitoplasting

Isolated mitochondria are pelleted and rapidly diluted in 20 mM HEPES(7.4) and immediately vortexed at high speed. The solution is placed on ice for 30 minutes with occasional vortexing.

The mitochondria are then pelleted by centrifugation at 10,000 x g. The supernatant is removed

leaving only the pelleted mitoplasts.

BN-PAGE

Pelleted mitochondria are dissolved in ice-cold lysis buffer (20 mM HEPES-KOH(7.4), 50 mM

NaCl, 1% digitonin, 10% glycerol, 2.5 mM MgCl 2, 0.5 mM PMSF, and 0.5 mM EDTA) and kept on ice for 15 minutes. Samples were then centrifuged at 20,800 x g and 4˚C for 5 minutes.

Supernatants were then fractionated by BN-PAGE and analyzed by autoradiography to detect

protein complex formation.

MIC 50 assay

21 The yeast MIC 50 assays were performed according to established methods . The yeast were grown for 24 hours at 30˚C. The dilution across the plate and initial small molecule concentration varied depending on how the given yeast strain tolerated the given molecule. For each strain in each assay, the well containing 1% DMSO with no molecule was set as 100% growth for its respective row.

116

Next generation sequencing

Yeast strains were grown overnight in YPD cultures at 30˚C. The genomic DNA was extracted from each culture using the QIAGEN Puregene Yeast/Bacteria kit. Genomic DNA samples were then sheared by sonication to achieve fragments of exclusively 200-500 basepairs in length. The fragments were appended with one of sixteen unique barcodes for sequencing using the NuGen

Encore Rapid DR Multiplex System kit and the efficiency of the aptamer binding was measured by the KAPA Biosystems Library Quantification Kit for Illumina Platforms. Sequencing was performed at the UCLA Broad Stem Cell Research Center (BSCRC) on an Illumina HiSeq machine using 50 basepair single-end sequencing with 14 samples used per lane. Sequences fragments were then compiled for each gDNA sample and mutations identified relative to the parental strain.

Oxygen consumption assays with electrode

Oxygen consumption of yeast mitochondria was done using a Clark-type electrode as previously described 21 . NADH was used to initiate respiration and CCCP was used to uncouple the

mitochondria to cause a rapid increase in oxygen consumption. DMSO was added to a final

concentration of 1% in each condition.

Permeability assay

Isolated mitochondria were resuspended in isotonic buffer and treated with indicated small

molecule for 30 minutes at 25˚C. Samples were then centrifuged to pellet the mitochondria. The

supernatant fraction was separated into a fresh tube for TCA precipitation of proteins. Both the

mitochondrial and supernatant fractions were disrupted in Laemmli sample buffer, fractionated

by SDS-PAGE, and analyzed by both Coomassie Brilliant Blue staining or immunoblot for

desired proteins.

117

Zebrafish studies

Zebrafish work was performed following established methods 21 . Zebrafish embryos were treated with the indicated concentration of small molecule and images were taken 3 days post fertilization.

118

Figures

Figure 5.1. MB-11 inhibits precursor proteins that contain hydrophobic segments. (A)

Molecular structure of MB-11. (B) Radiolabeled Su9-Ura3 is imported into isolated yeast mitochondria in the presence of indicated concentrations of MB-11. Total minutes incubating is indicated for each sample. Reaction is stopped with protease on ice. Mitochondrial fraction is fractionated by SDS-PAGE and detected on X-ray film. -∆ψ = 50 M CCCP. p = precursor form. m = mature form. (C) Same as in (B), but with Su9-DHFR as the precursor. (D) Same as in

119

(B), but a co-import of Su9-DHFR and Su9-Ura3 into the same mitochondria. (E-F) Same as in

(B), but using the indicated precursors. For AAC and Tom40, a 30 minute incubation in 100 mM

Na 2CO 3 on ice is also performed after protease treatment to remove and precursor protein that has not integrated into the membrane.

120

Figure 5.2. MB-11 targets a protein on the OM, but not Tom70 or Tom20. (A) As in Figure

5.1B, but the yeast mitochondria have undergone hypotonic treatment to ablate the OM prior to

Su9-Ura3 import. [MB-11] = 100 M. (B) As in Figure 5.1B, but the mitochondria are lysed in

1% digitonin and run on BN-PAGE to detect complexes. [MB-11] = 20 M. (C) Indicated yeast strains are grown in YPD and the indicated concentration of MB-11 at 30˚C for 24 hours and the 121 growth is determined by measuring the OD600 . The % growth for each strain is in reference to

vehicle control for that row. The curve for each is made on GraphPad Prism and n = 3 +/- SD.

(D) The same as in (C), but with the indicated strains. (E) Same as in 5.1B, but one subset of

mitochondria are trypsin-shaved prior to initiation of import. An immunoblot for Tom70,

Tom20, and loading control Tim44 is also shown.

122

Figure 5.3. An approach to obtain spontaneous revertants that increase viability in the presence of MB-11. (A) Illustration of the approach that was used to obtain spontaneous revertants that had increased viability in MB-11. Yeast are grown overnight, diluted to fresh media containing MB-11, grown overnight again, plated, allowed to grow colonies, colonies are validated, and finally whole genome sequencing was performed on colonies that were hits. (B-D)

Indicated revertant strains are grown in YPD and the indicated concentration of MB-11 at 30˚C 123 for 24 hours and the growth is determined by measuring the OD 600 . The % growth for each strain is in reference to vehicle control for that row. The curve for each is made on GraphPad Prism and n = 3 +/- SD.

124

Figure 5.4. Growth assays for the ∆seo1 strain in the presence of various small molecules.

The ∆seo1 and parental strains are grown in YPD and the indicated concentrations of the indicated molecule, including MB-11 (A), MB-10 (B), CCCP (C), and DECA (D), at 30˚C for 24 hours and the growth is determined by measuring the OD 600 . The % growth for each strain is in reference to vehicle control for that row. The curve for each is made on GraphPad Prism and n =

3 +/- SD.

125

Figure 5.5. MB-11 inhibits the import of some precursors into mammalian mitochondria.

(A) Radiolabeled Su9-Ura3 is imported into isolated HeLa mitochondria in the presence of indicated concentrations of MB-11. Total minutes incubating is indicated for each sample.

Reaction is stopped with protease on ice. Mitochondrial fraction is fractionated by SDS-PAGE and detected on X-ray film. -∆ψ = 50 M CCCP. p = precursor form. m = mature form. (B-E)

126

Same as in (A), but using the indicated precursors. For AAC, a 30 minute incubation in 100 mM

Na 2CO 3 on ice is also performed after protease treatment to remove and precursor protein that has not integrated into the membrane.

127

Figure 5.S1. MB-11 confers growth in media lacking uracil in a specific manner. (A) The screening strain expressing Su9-Ura3 increases growth in media lacking uracil as MB-11 is titrated in. Cultures are grown for 24 hours at 30˚C in SC-Ura and the indicated concentrations of

MB-11. (B) The parent strain of yeast, with Ura3 localized normally, is grown in YPEG and the indicated concentration of MB-11 for 24 hours at 30˚C. The % growth of yeast at each concentration is in reference to the growth with only vehicle notrol. The curve for each strain is 128 created using GraphPad Prism. (C) The O 2 consumption rate is measured in isolated yeast

mitochondria. NADH is added to begin respiration and CCCP is used as a control to uncouple

the mitochondria. DMSO represents the vehicle control for MB-11. (D) Isolated yeast

mitochondria are indicated in the indicated concentration of MB-11 for 30 minutes at 25˚C. The

pellet is separated from the supernatant by centrifugation and samples are fractionated by SDS-

PAGE. Proteins are detected by Coomassie Brilliant Blue staining. (E) The same as (D), but

proteins are detected by immunoblot. Markers include α-KDH (matrix), Cyt b 2 (IMS), Tim44

(IM), and Tim23 (IM). (F) Same as in (D), but with HeLa cells mitochondria in plave of yeast mitochondria. (G) Same as in (F), but proteins are detected by immunoblot. Markers include hPreP (matrix), YME1L (IM), hTOM40 (OM), and hMia40 (IMS).

129

Figure 5.S2. MB-11 affects zebrafish development. Zebrafish are treated with either DMSO or

MB-11 at the embryonic stage and images are taken multiple days later. Images are shown that were taken with a light microscope (“buffer”) or after treatment with various dyes.

130

Tables

Average Average ratio to P value to Strain MIC 50 1 MIC 50 2 MIC 50 3 MIC 50 parent parent Parent 1.407 1.664 1.549 1.540 - - 7.3 2.495 2.586 2.894 2.658 1.73 0.012 7.9 2.948 2.537 2.691 2.725 1.77 0.026

Parent 1.092 0.965 1.037 1.031 - - YPD.1 2.290 2.048 2.299 2.212 2.11 0.002 YPD.2 2.316 2.394 2.130 2.280 2.15 0.006 YPD.3 2.355 2.452 1.830 2.212 2.16 0.029 YPD.4 2.179 2.147 1.861 2.062 2.00 0.011 YPD.5 2.545 2.286 2.152 2.328 2.32 0.006 YPD.6 2.323 2.050 2.182 2.185 2.13 0.001 YPD.7 2.687 2.871 2.559 2.706 2.50 0.005 YPD.8 2.492 2.046 2.769 2.436 2.29 0.017 YPD.9 2.513 2.315 2.789 2.539 2.33 0.007 YPD.10 2.670 2.617 2.801 2.696 2.48 0.001 YPD.11 2.589 2.244 2.524 2.452 2.37 0.003 YPD.12 2.784 2.499 3.143 2.809 2.58 0.009 YPD.13 2.212 2.212 1.942 2.122 2.04 0.008 YPD.14 2.437 1.984 2.854 2.425 2.25 0.027 YPD.15 2.604 2.107 2.622 2.444 2.38 0.009

Parent 1.220 1.177 1.037 1.145 - - YPEG.1 0.937 1.106 - 1.021 0.85 - YPEG.2 1.119 1.011 - 1.065 0.89 - YPEG.3 1.014 1.091 - 1.053 0.88 - YPEG.4 1.093 1.014 - 1.054 0.88 - YPEG.5 1.314 1.051 - 1.183 0.98 - YPEG.6 1.116 1.000 - 1.058 0.88 - YPEG.7 1.236 0.950 - 1.093 0.91 - YPEG.8 0.968 0.963 - 0.966 0.81 - YPEG.9 2.966 2.070 2.813 2.616 2.30 0.037 YPEG.10 2.848 2.323 3.098 2.756 2.43 0.026 YPEG.11 2.968 2.082 2.705 2.585 2.27 0.033 YPEG.12 2.903 2.240 2.214 2.452 2.14 0.021 YPEG.13 2.785 2.130 2.688 2.534 2.23 0.024 YPEG.14 2.885 2.340 2.240 2.488 2.17 0.014 YPEG.15 2.830 1.981 2.875 2.562 2.26 0.046 YPEG.16 0.940 0.901 - 0.921 0.77 -

Table 5.1. A summary of the results from MIC 50 assays with the spontaneous revertants.

131

Table 5.2. A complete list of all mutations located in any exon (non-synonymous mutations only), in any intron, that gain or lose a stop codon, or that remove a start codon from all of the strains that were sequenced. 132

Table 5.2 continued. 133

Table 5.2 continued. 134

Table 5.2 continued.

135

Table 5.2 continued.

136

Table 5.2 continued.

137

References

1. Sickmann, A.; Reinders, J.; Wagner, Y.; Joppich, C.; Zahedi, R.; Meyer, H. E.; Schönfisch, B.; Perschil, I.; Chacinska, A.; Guiard, B.; Rehling, P.; Pfanner, N.; Meisinger, C., The proteome of Saccharomyces cerevisiae mitochondria. Proc Natl Acad Sci U S A 2003, 100 (23), 13207-12.

2. Chacinska, A.; Koehler, C. M.; Milenkovic, D.; Lithgow, T.; Pfanner, N., Importing mitochondrial proteins: machineries and mechanisms. Cell 2009, 138 (4), 628-44.

3. Lill, R.; Mühlenhoff, U., Maturation of iron-sulfur proteins in eukaryotes: mechanisms, connected processes, and diseases. Annu Rev Biochem 2008, 77 , 669-700.

4. Neupert, W.; Herrmann, J. M., Translocation of proteins into mitochondria. Annu Rev Biochem 2007, 76 , 723-49.

5. Ahting, U.; Thieffry, M.; Engelhardt, H.; Hegerl, R.; Neupert, W.; Nussberger, S., Tom40, the pore-forming component of the protein-conducting TOM channel in the outer membrane of mitochondria. J Cell Biol 2001, 153 (6), 1151-60.

6. Bausewein, T.; Mills, D. J.; Langer, J. D.; Nitschke, B.; Nussberger, S.; Kühlbrandt, W., Cryo-EM Structure of the TOM Core Complex from Neurospora crassa. Cell 2017, 170 (4), 693- 700.e7.

7. Saitoh, T.; Igura, M.; Obita, T.; Ose, T.; Kojima, R.; Maenaka, K.; Endo, T.; Kohda, D., Tom20 recognizes mitochondrial presequences through dynamic equilibrium among multiple bound states. EMBO J 2007, 26 (22), 4777-87.

8. Young, J. C.; Hoogenraad, N. J.; Hartl, F. U., Molecular chaperones Hsp90 and Hsp70 deliver preproteins to the mitochondrial import receptor Tom70. Cell 2003, 112 (1), 41-50.

9. Wiedemann, N.; Pfanner, N.; Ryan, M. T., The three modules of ADP/ATP carrier cooperate in receptor recruitment and translocation into mitochondria. EMBO J 2001, 20 (5), 951-60.

10. van Wilpe, S.; Ryan, M. T.; Hill, K.; Maarse, A. C.; Meisinger, C.; Brix, J.; Dekker, P. J.; Moczko, M.; Wagner, R.; Meijer, M.; Guiard, B.; Hönlinger, A.; Pfanner, N., Tom22 is a multifunctional organizer of the mitochondrial preprotein translocase. Nature 1999, 401 (6752), 485-9.

11. Kanamori, T.; Nishikawa, S.; Nakai, M.; Shin, I.; Schultz, P. G.; Endo, T., Uncoupling of transfer of the presequence and unfolding of the mature domain in precursor translocation across the mitochondrial outer membrane. Proc Natl Acad Sci U S A 1999, 96 (7), 3634-9.

138

12. Tamura, Y.; Harada, Y.; Shiota, T.; Yamano, K.; Watanabe, K.; Yokota, M.; Yamamoto, H.; Sesaki, H.; Endo, T., Tim23-Tim50 pair coordinates functions of translocators and motor proteins in mitochondrial protein import. J Cell Biol 2009, 184 (1), 129-41.

13. Martinez-Caballero, S.; Grigoriev, S. M.; Herrmann, J. M.; Campo, M. L.; Kinnally, K. W., Tim17p regulates the twin pore structure and voltage gating of the mitochondrial protein import complex TIM23. J Biol Chem 2007, 282 (6), 3584-93.

14. Chacinska, A.; Lind, M.; Frazier, A. E.; Dudek, J.; Meisinger, C.; Geissler, A.; Sickmann, A.; Meyer, H. E.; Truscott, K. N.; Guiard, B.; Pfanner, N.; Rehling, P., Mitochondrial presequence translocase: switching between TOM tethering and motor recruitment involves Tim21 and Tim17. Cell 2005, 120 (6), 817-29.

15. Mokranjac, D.; Sichting, M.; Popov-Celeketi ć, D.; Mapa, K.; Gevorkyan-Airapetov, L.; Zohary, K.; Hell, K.; Azem, A.; Neupert, W., Role of Tim50 in the transfer of precursor proteins from the outer to the inner membrane of mitochondria. Mol Biol Cell 2009, 20 (5), 1400-7.

16. Wiedemann, N.; van der Laan, M.; Hutu, D. P.; Rehling, P.; Pfanner, N., Sorting switch of mitochondrial presequence translocase involves coupling of motor module to respiratory chain. J Cell Biol 2007, 179 (6), 1115-22.

17. Li, Y.; Dudek, J.; Guiard, B.; Pfanner, N.; Rehling, P.; Voos, W., The presequence translocase-associated protein import motor of mitochondria. Pam16 functions in an antagonistic manner to Pam18. J Biol Chem 2004, 279 (36), 38047-54.

18. Slutsky-Leiderman, O.; Marom, M.; Iosefson, O.; Levy, R.; Maoz, S.; Azem, A., The interplay between components of the mitochondrial protein translocation motor studied using purified components. J Biol Chem 2007, 282 (47), 33935-42.

19. Krayl, M.; Lim, J. H.; Martin, F.; Guiard, B.; Voos, W., A cooperative action of the ATP- dependent import motor complex and the inner membrane potential drives mitochondrial preprotein import. Mol Cell Biol 2007, 27 (2), 411-25.

20. Taylor, A. B.; Smith, B. S.; Kitada, S.; Kojima, K.; Miyaura, H.; Otwinowski, Z.; Ito, A.; Deisenhofer, J., Crystal structures of mitochondrial processing peptidase reveal the mode for specific cleavage of import signal sequences. Structure 2001, 9 (7), 615-25.

21. Miyata, N.; Tang, Z.; Conti, M. A.; Johnson, M. E.; Douglas, C. J.; Hasson, S. A.; Damoiseaux, R.; Chang, C. A.; Koehler, C. M., Adaptation of a Genetic Screen Reveals an Inhibitor for Mitochondrial Protein Import Component Tim44. J Biol Chem 2017, 292 (13), 5429-5442.

22. Miyata, N.; Steffen, J.; Johnson, M. E.; Fargue, S.; Danpure, C. J.; Koehler, C. M., Pharmacologic rescue of an enzyme-trafficking defect in primary hyperoxaluria 1. Proc Natl Acad Sci U S A 2014, 111 (40), 14406-11.

139

23. Isnard, A. D.; Thomas, D.; Surdin-Kerjan, Y., The study of methionine uptake in Saccharomyces cerevisiae reveals a new family of amino acid permeases. J Mol Biol 1996, 262 (4), 473-84.

24. Nelissen, B.; De Wachter, R.; Goffeau, A., Classification of all putative permeases and other membrane plurispanners of the major facilitator superfamily encoded by the complete genome of Saccharomyces cerevisiae. FEMS Microbiol Rev 1997, 21 (2), 113-34.

25. Rehling, P.; Model, K.; Brandner, K.; Kovermann, P.; Sickmann, A.; Meyer, H. E.; Kühlbrandt, W.; Wagner, R.; Truscott, K. N.; Pfanner, N., Protein insertion into the mitochondrial inner membrane by a twin-pore translocase. Science 2003, 299 (5613), 1747-51.

26. Peixoto, P. M.; Graña, F.; Roy, T. J.; Dunn, C. D.; Flores, M.; Jensen, R. E.; Campo, M. L., Awaking TIM22, a dynamic ligand-gated channel for protein insertion in the mitochondrial inner membrane. J Biol Chem 2007, 282 (26), 18694-701.

27. Wiedemann, N.; Truscott, K. N.; Pfannschmidt, S.; Guiard, B.; Meisinger, C.; Pfanner, N., Biogenesis of the protein import channel Tom40 of the mitochondrial outer membrane: intermembrane space components are involved in an early stage of the assembly pathway. J Biol Chem 2004, 279 (18), 18188-94.

28. Becker, T.; Guiard, B.; Thornton, N.; Zufall, N.; Stroud, D. A.; Wiedemann, N.; Pfanner, N., Assembly of the mitochondrial protein import channel: role of Tom5 in two-stage interaction of Tom40 with the SAM complex. Mol Biol Cell 2010, 21 (18), 3106-13.

29. Glick, B. S.; Pon, L. A., Isolation of highly purified mitochondria from Saccharomyces cerevisiae. Methods Enzymol 1995, 260 , 213-23.

30. Hasson, S. A.; Kane, L. A.; Yamano, K.; Huang, C. H.; Sliter, D. A.; Buehler, E.; Wang, C.; Heman-Ackah, S. M.; Hessa, T.; Guha, R.; Martin, S. E.; Youle, R. J., High-content genome- wide RNAi screens identify regulators of parkin upstream of mitophagy. Nature 2013, 504 (7479), 291-5.

31. Dabir, D. V.; Hasson, S. A.; Setoguchi, K.; Johnson, M. E.; Wongkongkathep, P.; Douglas, C. J.; Zimmerman, J.; Damoiseaux, R.; Teitell, M. A.; Koehler, C. M., A small molecule inhibitor of redox-regulated protein translocation into mitochondria. Dev Cell 2013, 25 (1), 81-92.

140