<<

UCLA UCLA Electronic Theses and Dissertations

Title Matrix Processing Peptidase and Putative Roles in Mitochondrial Biogenesis. Nuc1 and Porin influence L-A Killer Virus Loads Co-dependently in .

Permalink https://escholarship.org/uc/item/3600s4k4

Author Torres, Eric

Publication Date 2019

Peer reviewed|Thesis/dissertation

eScholarship.org Powered by the California Digital Library University of California

UNIVERSITY OF CALIFORNIA

Los Angeles

Matrix Processing Peptidase and Putative Roles in Mitochondrial Biogenesis

and

Nuc1 and Porin influence L-A Killer Virus Loads Co-dependently in Saccharomyces Cerevisiae

A dissertation submitted in partial satisfaction of the

requirements for the degree Doctor of Philosophy

in Biochemistry and Molecular Biology

by

Eric Rommel Torres

2019

© Copyright by

Eric Rommel Torres

2019 ABSTRACT OF THE DISSERTATION

Matrix Processing Peptidase and Putative Roles in Mitochondrial Biogenesis

and

Nuc1 and Porin Influence Killer Virus Loads Co-dependently in Saccharomyces Cerevisiae

by

Eric Rommel Torres

Doctor of Philosophy in Biochemistry and Molecular Biology

University of California, Los Angeles, 2019

Professor Carla Marie Koehler, Chair

Matrix Processing Peptidase and Putative Roles in Mitochondrial Biogenesis

The mitochondrial import and proteolytic system are important for homeostasis of the . Typically, precursor are targeted to the mitochondria by an N- terminal targeting sequence and are imported via TOM and TIM translocons of the outer and inner membrane. A series of proteases is required for the import and processing of precursors as well assembly and degradation of mitochondrial proteins. The matrix processing peptidase

(MPP) coordinates the removal of presequences from precursors and is important in maintaining mitochondrial homeostasis. Furthermore, MPP is associated with neurodegenerative diseases. A in PMPCA causes a form of non-progressive cerebellar ataxia, because the Friedreich’s

ii ataxia protein is not processed correctly. The Parkinson’s associated protein PINK1 is processed by MPP and attenuation of β-MPP has been shown to induce mitophagy. To characterize the function of MPP in mammalian systems and develop probes to attenuate MPP function, we developed a small molecule screening strategy to develop MPP-specific chemical probes from a library of over 100,000 drug-like small molecules. The screen yielded two structurally distinct and specific MPP inhibitors deemed MitoBloCK-50 and MitoBloCK-51. Our results suggest in addition to playing a role in the processing of precursors, MPP also plays a role in the import of precursors. Here I present data that characterize the utility of MitoBloCK-50 and MitoBloCK-51 for MPP attenuation in various model systems. In addition, an in vivo tagging system known as

BioID identified an interaction between MPP and core subunit Qcr2 of Complex III, suggesting an evolutionary conserved interaction with the respiratory chain.

Nuc1 and Porin Influence L-A Viral loads Co-dependently in Saccharomyces Cerevisiae

The double-stranded RNA (dsRNA) killer viruses of the Totiviridae family are well-characterized and one of the most prolific in laboratory yeast strains. Infected yeast are able to tolerate high levels of the killer virus without displaying any growth defects. The complex symbiotic relationship between host yeast and dsRNA viruses are still being elucidated. Notably, of mitochondrial NUC1, which encodes the major mitochondrial nuclease, and

POR1, the voltage-dependent anion channel of the outer mitochondrial membrane, lead to overproduction of the capsid Gag protein encoded by the L-A dsRNA helper genome of Killer virus. In this work we explore mechanisms of Nuc1 and porin regulation of the L-A dsRNA genome of yeast Killer viruses. We observed strains with NUC1 deletion have higher L-A levels than POR1 deleted strains. Furthermore, deletion of both NUC1 and POR1 have comparable L-A

iii levels to single NUC1 mutants. Plasmid expression of Nuc1 mutants that are nuclease deficient or cytosolic localized fail to rescue L-A viral loads in NUC1 deleted strains. In addition, plasmid expression of NUC1 appears to increase the L-A viral load in the NUC1 and POR1 double knockout. Nuclease protection experiments demonstrate L-A transcripts are protected in ΔNUC1 and ΔPOR1 mitochondria but not in wildtype. We propose a model where ss(+) L-A transcripts are imported into the mitochondrial IMS and are subsequently degraded by Nuc1. Meanwhile,

POR1 deletion disrupts the L-A regulating function of Nuc1. Overall these results demonstrate a role for mitochondria in regulating dsRNA viruses in yeast.

iv

The dissertation of Eric Rommel Torres is approved.

Alexander M Van der Bliek

Jorge Torres

Carla Marie Koehler, Committee Chair

University of California, Los Angeles

2019

v

DEDICATION

Friends, Family and Mentors that loved and supported me along the way to make this possible.

vi

TABLE OF CONTENTS

List of Chapters, Figures and Tables ...... vii Acknowledgements ...... x Vita ...... xiv

Chapter 1. Introduction- Mitochondrial Biogenesis and Methods for Probing Protein Import Pathways ...... 1 Chapter 2. Characterizing Utility and Mitochondrial Effects by MitoBloCK-50 and MitoBloCK-51 ...... 37 Chapter 3. Characterization of Additional Selected MPP Inhibitors ...... 94 Chapter 4. The MPP BioID Proteome, Implications in Protein Import and Respiration ...... 132 Chapter 5. Nuc1 and Porin Influence Killer Virus Loads Co-dependently in Saccharomyces Cerevisiae ...... 171

LIST OF FIGURES

Chapter 1. Figure 1-1. Outer Membrane Protein Import through the SAM Pathway ...... 19 Figure 1-2. Intermembrane Space Import through the Mia40/Erv1 Disulfide Relay ...... 20 Figure 1-3. Inner Membrane Import through the Carrier Pathway ...... 21 Figure 1-4. Presequence Pathway through the Tim23 Complex ...... 22 Figure 1-5. MPP Cleavage Mechanism ...... 24 Figure 1-6. Inhibitors of Mitochondrial Import Pathways ...... 26

Chapter 2. Figure 2-1. Model of JPT fluorogenic peptide assay and plate layout for high-throughput screen ...... 57 Figure 2-2. Structures of MitoBloCK-50, MitoBloCK-51 and associated analogs ...... 59 Figure 2-3. MB-50 and MB-51 reduce MPP cleavage activity in vitro...... 60 Figure 2-4. Import and Cleavage of [35S] Su9-DHFR with MB-50/51 SAR compounds ...... 61 Figure 2-5. MB-50 and MB-51 do not interfere with general mitochondrial function ...... 63 Figure 2-6. MB-50 and MB-51 cause pigment deficiency and other morphological defects in zebrafish ...... 65 Figure 2-7. Treatment with excess divalent cation Mn2+ does not rescue MPP activity after MB-50 or MB-51 treatment ...... 66 Figure 2-8. MB-50 and MB-51 inhibit the import of TIM23 substrates ...... 67 Figure 2-9. MB-50 and MB-51 inhibit the import of non-TIM23 substrates ...... 68 Figure 2-10. Effects of MB-50 and MB-51 on import of dual-localized Fum1 and Pink1 in yeast mitochondria ...... 69 Figure 2-11. Overexpression of MPP increases mitochondrial membrane potential ...... 70 Figure 2-12. Overexpression of MPP increases import in isolated mitochondria ...... 72 Figure 2-13. Overexpression of MPP does not rescue import defects induced by MB-50 or MB- 51...... 74

vii

Figure 2-14. Temperature sensitive MPP mutants show processing defects and no import defects ...... 76 Figure 2-15. Pre-treatment with MB-50 and MB-51 uncouples respiration in isolated mouse liver mitochondria ...... 78 Figure 2-16. Morpholino knockdown of MPP subunits in zebrafish causes pigment deficiency and other morphological defects ...... 80

Chapter 3. Figure 3-1. Structures of MPP inhibitors selected for further characterization ...... 106 Figure 3-2. Time course import [35S] Su9-DHFR with selected MPP inhibitors ...... 108 Figure 3-3. Time course import [35S] Cpn10, Cmc1 and hPink1 with selected MPP inhibitors ...... 110 Figure 3-4. Co-import of [35S] Tom40 and AAC with selected MPP inhibitors ...... 111 Figure 3-5. DisC3(5) potentiometric assay with selected MPP inhibitors ...... 112 Figure 3-6. Figure 3-6 Chelation assay of select MPP inhibitors using the YDHA7 JPT peptide and recombinant MPP ...... 113 Figure 3-7. Membrane integrity assay with selected MPP inhibitors ...... 115 Figure 3-8. IC50 analysis of recombinant MPP with selected inhibitors...... 116 Figure 3-9. Yeast cell viability assay of compounds 0136 and 875 ...... 117 Figure 3-10. Yeast cell viability assay of compounds 0022 and 0115 ...... 118 Figure 3-11. MTT toxicology assay in HeLa cells with compounds 8247, 0115, 0136 and 875 ...... 119 Figure 3-12. Agilent Seahorse respirometry assay with compounds 0115, 875 and 8247 ...... 120

Chapter 4. Figure 4-1. BioID in vivo tagging system to detect transient interactions of MPP ...... 146 Figure 4-2. MPP BioID constructs are functional and localize to the mitochondria ...... 147 Figure 4-3. Hek293 MPP BioID cell lines demonstrate transient interactions with Qcr2 of complex III and PAM motor components ...... 149 Figure 4-4. MPP is associated with higher molecular weight complexes ...... 151 Figure 4-5. MB-50 and MB-51 promote supercomplex assembly of Complexes III and IV ...... 153

Chapter 5. Figure 5-1. L-A viral levels are increased in ΔNuc1 and ΔPOR1 mutants ...... 185 Figure 5-2. Localization of pNuc1-FLAG and ΔN9pNuc1-FLAG constructs ...... 187 Figure 5-3. L-A Gag levels in Nuc1 and Por1 rescue strains ...... 189 Figure 5-4. RT-qPCR of L-A levels in Nuc1 and Por1 rescue strains ...... 191 Figure 5-5. L-A RNA is protected in mitochondrial fractions of ΔNUC1 and ΔPOR1 strains ...... 192 Sup. 5-1. Immunofluorescence of L-A infected strains with anti-dsRNA (J2) monoclonal antibody...... 194 Sup. 5-2. Nuc1 localizes to the mitochondrial IMS in wildtype and ΔPOR1 strains ...... 196 Sup. 5-3. Apoptotic stress induces Annexin V and Propidium Iodide staining in yeast .....198 Sup. 5-4. Nuc1 does not localize to the nucleus under apoptotic stress ...... 200 Sup. 5-5. Mitochondria deleted in Por1 are import deficient ...... 201 viii

LIST OF TABLES

Chapter 2. Table 2-1. Summary of [35S] Radiolabeled Import Studies with MB-50 and MB-51 ...... 81 Table 2-S1. Strains used in this study ...... 82

Chapter 3. Table 3-1. Cumulative results of import efficiencies with selected MPP inhibitors ...... 122 Table 3-2. Cumulative results of secondary assays with selected MPP inhibitors ...... 123 Table 3-S1. Strains used in this study ...... 124

Chapter 4. Table 4-1. BioID hits of PMPCA Vs. WT Control...... 155 Table 4-2. BioID hits of PMPCB Vs. WT Control ...... 157 Table 4-3. BioID hits of each MPP-BirA* subunit against the other plus the WT Control ..159 Table 4-S1. Strains used in this study ...... 161

Chapter 5. Table 5-S1 Strains used in this study...... 204

LIST OF REFERENCES

Chapter 1. References ...... 28 Chapter 2. References ...... 90 Chapter 3. References ...... 130 Chapter 4. References ...... 167 Chapter 5. References ...... 213

ix

ACKNOWLEDGEMENTS

I want to thank all my friends, family and mentors that have contributed to this challenging but incredibly rewarding journey. All my life I have respected all fields of science and how they clarify and shape the world we live in. With this dissertation I am honored for the opportunity to contribute to the scientific community. I am especially grateful for the opportunity to study at The University of California Los Angeles under the mentorship of Dr. Carla Koehler.

Over the years Carla has provided me the mentorship and support I needed to pursue the burning scientific questions that motivated me through grad school. Carla’s advice was often tough and blunt, but necessary to allow me to grow as a scientist. I would also like to acknowledge my committee members, Dr. Alex van der Bliek, Dr. Michael A Teitell, Dr. James W. Gober, and

Dr. Jorge Z. Torres for their advice and guidance over the years. I particularly want to thank Dr.

Alex van der Bliek for always being available to share his expertise on mitochondrial fusion/fission dynamics.

My graduate school experience would not have been as memorable without my fellow graduate student colleagues of the Chemistry and Biochemistry Department. Our time together, particularly within the first year, fostered incredible friendships that I will always remember. I particularly want to thank Charles Wang for being an amazing friend to hang out with, share stories, grab food, or go to the gym. Thank you for always being there for me and I hope I have done the same for you. I would also like to acknowledge my best friend in graduate school Alex

Bradley. Thank you, Alex, for deciding to rotate in the Carla Koehler lab and initiate the best friendship I have in my adult life. Graduate school has been the most challenging endeavor in my life and I would certainly not be where I am today without you. Thank you for all the engaging conversations and always making me laugh even on the worst of days. x

I would also like to extend my gratitude to past and current members of the Carla Koehler lab. Janos Steffen, Deepa Dabir, Vivian Zhang, Jesmine Cheung, Juwina Wijaya, Jisoo Han,

Eriko Shimada, Colin Douglas, Gary Schmorgan, Michael Conti, Matthew Maland, Kayla Frank,

Sean Atamdede. To Janos, thank you for all your scientific expertise and advice on my own projects and career. I also would like to thank you for our random discussions about movies and

TV shows. To Juwina and Jisoo, thank you for your friendship during my first years in the

Koehler lab. I have missed our outings to concerts and restaurants since you left. To Matthew and Michael, thank you for your friendship and the hilarious discussions in lab. To Jesmine, thank you for all the advice, conversations and being such an amazing friend. I also want to thank all the undergraduates who have worked with me over the years and for putting up with my many last-minute requests. I particularly want to thank my undergraduate Kayla Frank for being such an amazing student. I am confident Kayla will have an amazing career in science as she begins her first year at Scripps Research Institute.

I would also like to thank my family for always being supportive of my unconventional career choices. I am thankful to have all of you in my life. Lastly, I want to thank my partner

Richard Hong. Meeting you Richard has been one of the best things to come out of graduate school. I love you and I’m looking forward to many more years with you.

xi

Chapter 1 is an introduction to mitochondrial protein import pathways and processing mechanisms. Mitochondria protein import pathways as a target for high-throughput chemical biology is explored. Published small molecule probes and the associated screening methods are highlighted.

Chapter 2 summarizes past and current results of MitoBloCK-50 (MB-50) and

MitoBloCK-51 (MB-51) identified in the in vitro MPP high-throughput screen. The screen and initial characterization of MB-50 and MB-51 was done by Colin Douglas. Further characterization of MB-50 and MB-51 as well as studies involving overexpressed MPP or temperature sensitive strains was performed by me or Kayla Frank. I would like to thank Marc

Liesa for his help with respirometry experiments using the Agilent Seahorse XF Analyzer.

Chapter 3 summarizes data from 19 additional small molecules that were structurally distinct MPP inhibitors identified in the high-throughput screen. Similar methods applied to MB-

50 and MB-51 were used to characterize the small molecules for their viability as MPP inhibitors. Import experiments were done by me and Kayla Frank. Marc Liesa helped with respirometry experiments using the Agilent Seahorse XF Analyzer.

Chapter 4 highlights an in vivo tagging method known as BioID to identify transient interactions with MPP in mammalian Hek293 cells. I would like to thank the Wohlschlegel

TM TM group for the Hek293 Flp-In T-Rex cell lines and BioID construct. The Wohlchlegel group also performed the peptide digestion and mass spectrometry analysis. Melania Abrahamian performed the data analysis of the BioID hits. I would also like to thank Alejandro Martorell for his advice on the PLA experiments and Janos Steffen for his help with constructing the BioID

xii strains. The chapter also includes follow up experiments with native gel electrophoresis performed by me and Kayla Frank.

Chapter 5 is a version of manuscript in preparation to be authored by myself, Kayla

Frank, and Carla Koehler. Porin and Nuc1 mediate L-A killer virus levels in Saccharomyces

Cerevisiae. I would like to thank Kayla Frank for conducting the apoptosis experiments and

Janos Steffen for his advice and technical expertise. I would also like to thank Carol Dieckmann and Reed Wickner for their advice and knowledge of yeast double-stranded RNA viruses.

xiii

VITA

2008-2013 B.S in Biochemistry California State University San Bernardino

2013-2019 Graduate Student Researcher Department of Chemistry and Biochemistry University of California, Los Angeles

2015-2018 Teaching Assistant Department of Chemistry and Biochemistry University of California, Los Angeles

PRESENTATIONS AND PUBLICATIONS

Schall, K. A., Holoyda, K. A., Grant, C. N., Levin, D. E., Torres, E. R., Maxwell, A., Grikscheit, T. C. (2015). Adult zebrafish intestine resection: a novel model of Short Bowel Syndrome, adaptation, and intestinal stem cell regeneration. American Journal of Physiology. Gastrointestinal and Liver Physiology, 309(3), ajpgi.00311.2014.

Patananan, A. N., Budenholzer, L. M., Eskin, A., Torres, E. R., & Clarke, S. G. (2014). Ethanol-induced differential expression and acetyl-CoA in a longevity model of the nematode Caenorhabditis elegans. Experimental Gerontology, 61C, 20–30.

Patananan, A. N., Budenholzer, L. M., Pedraza, M. E., Torres, E. R., Adler, L. N., & Clarke, S. G. (2015). The invertebrate Caenorhabditis elegans biosynthesizes ascorbate. Archives of Biochemistry and Biophysics, 569, 32–44.

Torres, E.R., Douglas, C.J., Koehler, C.M. (2015). The Role of Matrix Processing Peptidase in Mitochondrial Biogenesis. Presented at Mechanisms and Regulation of Protein Translocation, EMBOL conference in Dubrovnik, Croatia.

Torres, E.R., Douglas, C.J., Koehler, C.M. (2016). The Role of Matrix Processing Peptidase in Mitochondrial Biogenesis. Presented at Protein Transport Across Cell Membranes, Gordon Research Conference in Galveston, TX.

Torres, E.R., Douglas, C.J., Koehler, C.M. (2016). Small Molecule Modulators Elucidate Role of Matrix Processing Peptidase in Mitochondrial Biogenesis. Presented at Annual MBI Retreat in Lake Arrowhead, CA.

Torres, E.R., Douglas, C.J., Ngo, J., Damoiseaux, R., Koehler, C.M. (2017). Small Molecule Modulators Elucidate Role of Matrix Processing Peptidase in Mitochondrial Biogenesis. Presented at Mitochondria Communication, Keystone Conference in Taos, NM.

xiv

Steffen, J., Torres. E.R., Filipuzzi, I., Cheung, J., Koehler, C.M. (2018) Small Molecule Modulators for Mitochondrial Protein Import. Presented at Biomedical and Life Science Innovation Day, UCLA in Los Angeles, CA.

HONORS AND AWARDS

2010 POLYED Undergraduate Award for Achievement in Organic Chemistry California State University San Bernardino San Bernardino, California

2010-2013 College of Natural Sciences MASS Scholarship California State University San Bernardino San Bernardino, California

2012 CIRM-BRIDGES Fellowship California State University San Bernardino San Bernardino, California

2013 Outstanding Undergraduate in Chemistry California State University San Bernardino San Bernardino, California

2013 Chemistry Department Honors California State University San Bernardino San Bernardino, California

2013 Competitive Edge, UCLA Fellow University of California Los Angeles Los Angeles, California

2013-2019 Cota-Robles Fellow University of California Los Angeles Los Angeles, California

2014-2017 Cellular and Molecular Biology Program Trainee University of California Los Angeles Los Angeles, California

2015-2018 PA-14-148 Ruth L. Kirschstein National Research Service Award National Institutes of Health (NIH) Bethesda, Maryland

xv

Chapter 1: Mitochondrial Biogenesis and Methods for Probing Protein Import Pathways

1.1 Introduction: Mitochondria and the Cell

The first observations of mitochondria are credited to Richard Altmann in the late 19th century. His description of granule-like filaments across all cell types initiated the study of what we know today as mitochondria. Over 100 years later, research to further characterize the filamentous symbionts revealed them as a centerpiece for essential cellular processes that include respiration, metabolism, apoptosis1–3 , metabolite transport4, maintenance of reactive oxygen species5,6, and protein turnover7,8. Mitochondria emerged from a symbiosis of prokaryote and eukaryotic lineages as a product of ecologic and atmospheric adaptation over the course of 1.45 billion years9. The result led to numerous gene transfers from symbiont to host as well as a myriad of conserved processes that maintain mitochondria integrity and function. The mitochondria possess two membranes, aptly named the outer mitochondrial membrane (OMM) and the inner mitochondrial membrane (IMM)10,11. The membranes are predominantly composed of typical membrane containing phospholipids such as phosphatidyl ethanolamine and phosphatidyl choline; additionally, mitochondria exclusively contain cardiolipin12. These lipid bilayers form two aqueous compartments; the matrix which is bordered by the IMM, and an intermembrane space (IMS) bordered by the OMM and IMM. The IMM forms numerous cristae that protrude into the matrix and contain the electron transport chain complexes I-V responsible for oxidative in the IMS. The matrix houses the , the mitochondrial genome and . Given their vital role in crucial cellular mechanisms, mitochondria are highly dynamic. They are constantly fusing, dividing and interacting with various cellular components across its interconnected network. The established ER-mitochondria encounter structure (ERMES) is an example of the interdependence of mitochondria and other organelles of the cell. The ERMES complex forms a stable interaction between the mitochondria and ER, regulating calcium signaling, mitochondria morphology, and phospholipid transport13,14.

Communication between mitochondria and the nucleus serve to induce stress pathways that maintain mitochondrial homeostasis. Specifically, the mitochondria unfolded protein response

(mitoUPR) signals transcription factors upregulating mitochondrial proteases and chaperones in response to increased ROS damage, respiratory defects or unfolded proteins7,8,15. Despite the major advances in mitochondria research, the precise mechanisms of these pathways are not well understood. However, as research continues in the diverse fields of mitochondria biology, it is overwhelming clear the role of mitochondria extends beyond the duties of respiration and metabolism. Further understanding of these mechanisms will enable broader understanding of how the mitochondria contribute to overall health and disease.

1.2 Pathways of Mitochondria Import

There are approximately 1000 (yeast) to 1500 (human) proteins within the mitochondria with only 1% of these proteins encoded by the mitochondrial genome16,17. The mitochondrial encoded genes primarily include components of the electron transport chain such as subunits I, II and II, ATPase subunits 6 and 9, as well as mitochondria rRNA 21S and

15S18. The remaining proteins are encoded within the nucleus and imported into the mitochondria, a result of endosymbiosis and lateral gene transfer16,19. Mitochondria import of all precursor proteins is initiated by a single channel known as the translocase of the outer membrane (TOM). The TOM complex is a β-barrel channel formed by the oligomerized integral membrane protein TOM40. Other TOM components include presequence recognition proteins

TOM20, TOM70 and TOM22 which collectively recognize protein presequences and guide

2 precursors to the TOM40 channel20–23. Though there are variations of signaling motifs, targeting and translocation for most soluble mitochondria proteins depend on an N-terminal targeting sequence roughly 15-40 amino acid residues in length. Though targeting sequences of over 100 amino acids have also been documented24,25. These targeting sequences are amphipathic and α- helical in nature, features that facilitate their recognition and translocation by the TOM complex26,27. After translocation through the TOM complex, the precursors are shuttled by several small chaperones in the IMS known as the Tim proteins (Tim9-10 and Tim8-13). These chaperones guide the precursor to four canonical pathways of mitochondrial protein import.

These pathways are the sorting and assembly machinery pathway of the OMM (SAM), the IMS redox-regulated mitochondrial intermembrane space assembly pathway (MIA), a presequence pathway for matrix proteins regulated by the translocon of the inner membrane (TIM23), and a pathway for carrier proteins to the IMM through the TIM22 complex28.

The path of each mitochondria precursor depends largely on their localization and function within the mitochondria. Outer membrane β-barrel proteins such as Tom40 and VDAC

(yeast Por1), as well as some α-helical anchored proteins like Tom2229, are imported through the

SAM complex. The SAM channel is composed of core protein Sam50 and signal receptor

Sam3530. Both proteins are essential for cell viability (Figure 1-1). After translocation through the TOM complex, small chaperones Tim9-10 and Tim8-13 guide the precursor to the SAM complex. For β-barrel proteins a glycine-rich internal targeting signal at the most C-terminal β strand , known as a β-signal, is recognized by Sam3531. Recognition of the β-signal by Sam35 induces a conformational change that allows lateral release of the precursor into the lipid phase of the outer membrane. Sam37 acts later than Sam35 and aides in lateral release from the SAM complex32. For α-helical proteins, the precise mechanism of membrane insertion is still elusive.

3

However, studies show the TOM, SAM and more recently TOM assembly component Mim1 have been implicated in the import of multispanning outer membrane α-helical proteins33. For instance, assembly of N-terminal α-helical anchored proteins Tom20 and Tom70 are dependent on TOM assembly component Mim134. Mutispanning outer membrane proteins such as Ugo1 are also dependent on Mim1. Proteins that are C-terminally anchored to the membrane like TOM22 requires the SAM complex rather than Mim1 for efficient insertion29.

Most soluble proteins that are targeted to the IMS are imported by the Mia40/Erv1 disulfide relay system (Figure 1-2). IMS proteins are typically smaller in size (<20kDa) and possess a helix-loop-helix motif characterized by two antiparallel α-helices with two cysteines in

35,36 each helix . These cysteines are spaced by either three (CX3C) or nine (CX9C) amino acids in the primary sequence and ultimately form a disulfide linkage in the native protein. Once the unfolded precursor enters the IMS via TOM complex, a disulfide relay between Mia40 and sulfhydryl oxidase Erv1 oxidize the internal cysteine residues that allow maturation of the precusor37–39. This process is initiated by Mia40 which binds to the cysteine residues using a

CPC motif (cysteine-proline-cysteine). Once the cysteines are oxidized, the precursor forms internal disulfide linkages that induce folding into a mature protein. The Mia40 CPC motif is regenerated by reoxidation via Erv1. Erv1 then transfers its electrons to respiratory chain component cytochrome C and ultimately O2, thus regenerating the system for precursor oxidation28,38,39. Studies using S35 radiolabeled precursor substrate shows evidence of a ternary complex containing the substrate, Mia40 and Erv1. This temporary complex appears to serve as a conduit for Erv1 to transfer disulfide linkages to the substrate through the Mia40 disulfide carrier40.

4

The Mia40/Erv1 import pathway is well conserved between yeast and with some noteworthy differences. While human Mia40 is a substrate for its own pathway to the IMS, yeast

Mia40 contains targeting sequences for both the presequence (TIM) and Mia40 pathways.

However, the N-terminal targeting sequence of yeast Mia40 appears dispensable for function and relies on the C-terminal half known as Mia40core for cell viability. Meanwhile, human Mia40 is smaller relative to the yeast homolog and does not contain the N-terminal targeting sequence.

This suggests that early import of Mia40 in lower relied on the presequence pathway before evolution drove the transition to the Mia40/Erv1 pathway41.

Mitochondria import into the mitochondria IMM has been well characterized for metabolite carrier proteins such as ATP/ADP carrier AAC. Import of carrier proteins is dependent on multiple internal targeting signals, cystosolic chaperones, and the TIM22 complex42,43 (Figure 1-

3). The entire process occurs in several stages. Initially, the precursor carrier protein is translated in the cytosol and guided to the TOM complex with the aid of chaperones Hsp70 and Hsp90. The -precursor complex can then bind to TOM complex receptor Tom70 and induce formation of Tom70 oligomers44. The precursor is then translocated through the TOM complex after release from the cytosolic chaperones. Once in the IMS, the precursor forms a complex with Tim9-Tim10 and later Tim12 at the IMM on the surface of the TIM22 complex45,46. A membrane protein of the TIM22 complex known as Tim54 allows binding of the chaperone precursor complex to the TIM22 channel47. Though the exact mechanism is not well understood, the precursor is laterally released into the inner membrane in a potential-dependent fashion by

TIM22. Other components of the TIM22 complex include Tim18 which is suggested to be essential for TIM22 complex assembly46. More recently, Tim29 has been identified as another component of the TIM22 complex in humans. Tim29 has been implicated in maintaining the

5 stability of the TIM22 complex as well as possessing contacts to the TOM complex. Interactions with the TOM complex may implicate a mechanism for shuttling hydrophobic carrier proteins across the IMS48.

Nearly all N-terminally targeted matrix proteins are imported by the presequence pathway which involves translocation through the TOM and TIM23 complexes. The TIM23 complex guides presequences into the matrix in an ATP and membrane potential dependent manner

(Figure 1-4). Given the amount of subunits involved, the TIM23 complex is the most dynamic and intricate translocon of the mitochondria import machinery49. Moreover, the TIM23 complex is also capable of lateral sorting precursors into the IMM, making it one of the exceptional translocases that can transport proteins across and into a membrane49. The complex can be distinguished by the main channel components embedded in the IMM and the subunits forming the ATP-dependent presequence associated motor (PAM). The main components embedded in the IMM include Tim23, Tim50, Tim17 and Tim21. Two homodimers of Tim23 and Tim17 form the core of the channel, also known as the 90kDa Tim17-Tim23 core complex50,51. Each

Tim23 monomer contains four transmembrane helices embedded in the IMM at the C- terminus52. Meanwhile, the N-terminus spans the IMS and contains an anchor at the OMM, tethering the two mitochondrial membranes53. Tim23 tethering of the OMM is suggested to facilitate transfer of precursor proteins coming from the TOM complex. Tim17 shows significant with Tim23 and also contains four membrane spanning domains54. Though Tim23 forms the central core of the channel, Tim17 is still required for presequence import, lateral sorting, PAM complex recruitment, formation and stability of the channel51,52,55. Tim50 associates closely with Tim23 in the IMM and is responsible for recognizing nascent precursors from the TOM channel with an IMS exposed C-terminal domain56,57. However, in vitro studies

6 suggest the Tim50 IMS domain also mediates Tim23 channel gating58. Interactions between

Tom22 with Tim23 and Tim50 have also been shown to allow for efficient coupling of precursor import56. Tim21 acts as another modulator of Tim23 participates in coupling respiratory chain

Complex III to potential dependent precursor import59. Tim21 has also been proposed to transiently interact with the Tom22 for import coupling with the TOM complex51,56,60.

The driving force behind precursor import comes from the presequence associated motor

(PAM) which associates directly with the Tim23 core complex on the matrix side. The components of the PAM complex include Tim44, mtHsp70, Pam16, Pam17 and Pam18

(Tim14)61. At the center between the Tim23 core complex and the PAM complex sits Tim44, a peripheral membrane protein which acts as a scaffold for the motor driving cochaperones. mtHsp70 is an ATPase that provides the energy required for import through nucleotide exchange by a Brownian ratchet mechanism62. Mitochondrial Grpe (Mge1) acts as a cofactor for nucleotide exchange for regenerating the ATP bound form of mtHsp7063. As precursors emerge from the

Tim23 core, outward movement is prevented by mtHsp70 molecules continually binding to the emerging polypeptide. mtHsp70 also affects polypeptide conformation through unfolding and may provide an additional pulling force for translocation64. mtHsp70 activity is stimulated by J- proteins Pam16 and Pam18 which also bind to Tim4465–70. Pam17 is anchored to the IM and binds to Tim23 while stabilizing the PAM complex65. The TIM23 complex is also capable of inner membrane insertion for proteins with hydrophobic sorting signals51,71. These sorting signals stall precursor translocation through the Tim23 channel and are then laterally inserted into the

IMM. The TIM23 complex responsible for the lateral sorting is deemed TIM23SORT and includes

Tim23, Tim17, Tim50 and Tim21. This complex can be readily separated from the 90kDa

Tim23-Tim17 Core complex on a BN-PAGE61,65. The Tim23 complex is also responsible for

7 import of certain IMS destined precursors with N-terminal targeting sequences. Specifically, cytochromes c1 and b2 are sorted into the IMS by aa stop-transfer mechanism. This mechanism involves an arrest during translocation by a stop transfer sequence and subsequent release into the IMS by proteolytic cleavage72,73.

1.3 Mitochondria Proteases and Quality Control

The mitochondria proteome is tightly regulated by various proteolytic systems regulating protein folding, turnover and oxidative damage. An estimated 70% of imported mitochondria proteins contain a presequence that is removed upon import to form the mature protein25. The proteases responsible for presequence removal are essential for maintaining mitochondria integrity and function74–76. Sequence specificity and catalytic mechanisms of these proteases have been primarily characterized in yeast but are conserved in mammalian homologs76. The most notable of these proteases is mitochondrial processing peptidase (MPP) which cleaves the majority of presequences of matrix targeted proteins77–81. Other processing proteases include inner membrane protease (IMP), intermediate cleavage peptidase (Icp55), intermediate peptidase

(Oct1), and a number of m-AAA proteases of the inner membrane25,82,83. Oct1 and Icp55 are a source of secondary processing after MPP cleavage. Oct1/MIP is a soluble matrix protein and exists as a 75 kDa monomer in rat liver mitochondria84. Oct1 is responsible for producing an octapeptide from the MPP generated N-terminus of precursor proteins. Icp55 is a peripheral membrane protein on the IMM facing the matrix and cleaves just one amino acid from a precursor after MPP processing. IMP is responsible for cleaving off hydrophobic sorting signals from proteins facing the IMS. Only a few substrates of IMP have been identified and include

Cox2, Mcr1, Gut2, Cyb2 and Cyc176. In yeast IMP consists of 3 subunits: Imp1, Imp2 and Som1.

Imp1 and Imp2 are the catalytic subunits and appear to have different substrate specificities.

8

Imp1 prefers acidic residues at the +1 position and nonaromatic hydrophobic residues at +3. In contrast, Imp2 appears to have a preference for alanine at positions +1 and +276,85.

Once presequences are released from the mature precursor, their abundance and amphiphilic nature can disrupt membrane integrity, potential and respiration86. A protease known as presequence protease (PreP/Mop112) combats the accumulation of presequence peptides by degrading them upon their removal87–89. Knockouts of Prep in yeast exhibit a strong growth phenotype in non-fermentable carbon sources90. Furthermore, the human homolog of Prep

(hPrep) has been shown to degrade amyloid-β peptides, an implication in the progression of

Alzheimer’s disease89.

The AAA ( ATPase associated with cellular activities) proteases maintain protein quality through protein processing and degradation of damaged/unfolded proteins. This group of proteases include the LON/PIM1 proteases of the MEROPS peptidase family as well as IM bound ATPases such as YME1L. Lon protease (PIM1 in yeast) is a major soluble protease in the mitochondrial matrix responsible for degradation of misfolded or damaged peptides83,91. For example, oxidatively damaged iron-sulfur protein is a substrate of Lon92. AAA proteases within the IM contain catalytic domains facing the IMS (i-

AAA) or the matrix (m-AAA), maintain protein turnover in both mitochondria compartments.

One example of AAA processing includes ribosomal protein MrpL2 which is processed by m-

AAA through proteolytic degradation from the N-terminus93. Other substrates of the AAA proteases include unassembled subunits of the respiratory chain complexes and non-native IM proteins76,83,94.

1.4 MPP Mechanisms and Insights

9

Matrix processing peptidase (MPP) is the major mitochondria presequence peptidase responsible for the initial cleavage of N-terminal targeted precursors. Recently, quantitative

ChaFRADIC (charge-based fractional diagonal chromatography) with yeast temperature sensitive mutants identified 70 substrates of MPP, though more substrates are expected95. The mas1 and mas2 genes encode both subunits of yeast MPP and were first discovered by Michael

Yaffe and Gottfried Schatz in 198481,96,97. Their findings suggest both genes were essential for survival and temperature-sensitive cause an accumulation of mitochondrial precursors.

This suggested the vital role for MPP in cellular viability and their crucial role in maintaining mitochondria homeostasis. Subsequent studies conclude that MPP is a metalloendopeptidase of the pitrilysin (M16) family assembled as a heterodimer by two subunits, βMPP (Mas1) and

αMPP (Mas2). Both subunits share 48 percent sequence identity and are approximately 50 kDa, forming a 100kDa holoenzyme98. Each subunit contains two domains of ~210 residues with near identical structure. Each domain contains three α helices positioned against a six-stranded β sheet with another five α helices positioned on one end of the sheet. The domains are connected by a linker of 16 residues in the α subunit and 22 residues in the β subunit98. Despite their similarity in structure and sequence, only βMPP possesses the inverted thermolysin HXXEH zinc binding motif necessary for cleavage while αMPP carries a glycine-rich loop formed by residues 284-301 proposed to aid in substrate binding/release98.

MPP does not cleave a specific sequence but shows a preference for common motifs.

These motifs have an arginine at positions -2, -3 or -10 or an aromatic at position +1 from the cleavage site. However, the structure of yeast MPP demonstrates that R-3 is not a cleavage site for MPP but from sequential processing by Icp55 of one amino acid, converting R-3 to an R-2 motif25,98. There are also some MPP substrates with no distinguishable motif 75,98,99. The crystal

10 structure of recombinant yeast MPP reveals a large cavity between the two subunits that binds to

MPP substrates in an extended conformation. The cavity is negatively charged due to the prevalence of aspartate and glutamate residues, an important feature for the interaction of positively charged N-terminal targeting sequences. The active site containing the inverted

HXXEHX76E motif refers to residues βHis-70, βGlu-73, βHis-74 and βGlu-150 (Figure 1-5). All residues are involved in the zinc binding and catalysis. Cleavage of peptides occurs through an analogous mechanism to thermolysin peptidase. A Zn2+ coordinates with a water molecule near the βGlu-73 from the HXXEHX76E motif. Polarization of the water molecule by βGlu-73 aligns it for nucleophilic attack on the carbonyl carbon of the peptide bond and subsequent cleavage98.

Critical residues that stabilize the substrate interaction include βGlu-160 and βAsp-164 which accommodates the -2 arginine near the zinc binding site. Both residues are required for full activity and therefore may act cooperatively. βPhe-77 appears important for interacting with the bulky +1 aromatic residue of common substrate motifs. The glycine rich loop of the α subunit is crucial for MPP activity and appears to function analogously to Tom20 with regard to substrate recognition100–102 (Figure 1-5). Though the precise mechanism is unknown, the glycine rich loop stabilizes the conserved hydrophobic residues of the presequence, orientating the more positively charged residues toward the negatively charged components of the β subunit102.

Mutations disrupting the flexible glycine rich loop abolishes enzyme activity101.

MPP has an evolutionary relationship to the respiratory chain because of its structural similarity to the Cor1 and Cor2 proteins of Complex III76,103. In other eukaryotes such as

Neurospora Crassa, βMPP is bifunctional as the Cor1 protein of the bc1 complex104. In plants

MPP is fully integrated into the membrane as the Cor1 and Cor2 proteins105. Though yeast and mammalian MPP have been well characterized as a soluble matrix protein, several questions

11 remain to how it retains its specificity without a canonical cleavage site and where it localizes in the matrix during precursor import77,106,107.

1.5 Implications of Mitochondrial Import Defects in Human Disease

Given the crucial and diverse roles of mitochondria maintaining cellular homeostasis, it is no surprise that mitochondrial defects lead to a wide array of human disease. Defects in mitochondrial homeostasis have been implicated in mitochondrial myopathies and neuropathies.

This includes Parkinson’s disease, Friedreich’s ataxia, diabetes, and cancer108. Furthermore, the prevalence of mitochondria inherited disorders is nearly 1 in 4,300, making it one of the most common types of inherited disease in adults109,110.

Specifically, mitochondria translocation defects contribute to the wide array of mitochondria inherited disease. Human deafness dystonia syndrome (Mohr-Tranenbjaerg syndrome) is a recessive, X-linked neurodegenerative disorder that causes progressive sensorineural deafness in early child hood. The disease has been shown to be caused by truncations or deletion in IMS chaperone TIMM8, leading to reduced import through the Tim23 translocon111,112. Mutations in DNAJC19 (Pam18 in yeast) lead to dilated cardiomyopathy and

Ataxia (DCMA). The disease is characterized by infant-onset-dilated cardiomyopathy and cerebellar ataxia113,114. Defects in DNAJC19 play a crucial role in Tim23 dependent import, but may also participate in other mitochondria housekeeping functions115. More recently, a novel protein of the Tim22 complex in humans known as AGK has been implicated in Sengers syndrome. Sengers syndrome is an autosomal recessive disease causes early childhood metabolic disorders and can cause death in infancy114. Tissues from patients with Sengers syndrome display a destabilized Tim22 complex and perturbed TCA cycle. This is consistent with the role of the

Tim22 complex importing proteins which contribute to metabolic flux116.

12

Disruptions in dual-localized proteins to the mitochondria have also been implicated in disease. An autosomal recessive disorder known as primary hyperoxaluria (PH1) can be caused by mutations in alanine:glyoxylate aminotransferase (AGT) that lead to altered protein targeting.

In healthy patients AGT is targeted to peroxisomes using a C-terminal targeting sequence where it can detoxify glyoxylate through conversion to alanine. PH1 patients with an AGTP11LG170R mutation lead to an N-terminal mitochondria targeting sequence that directs AGT to mitochondria117,118. Defects in glyoxylate metabolism in the liver lead to accumulation of calcium oxalate in the kidney and urinary tract. Though the severity of the disease is varied, PH1 can lead to kidney failure and death if left untreated117.

MPP has been linked to certain ataxia related diseases such as Friedreich’s ataxia and a form of non-progressive cerebellar ataxia because it is required to process the associated protein frataxin119,120. is a highly conserved mitochondrial protein with a critical role in maintaining mitochondrial iron homeostasis121,122. Defects in human frataxin leads to an autosomal recessive disease known as Friedreich’s ataxia. The disease is characterized by progressive in the spinal cord and hypertrophic cardiomyopathy123. Upon import, frataxin is cleaved twice by MPP at residues 41-42, and 55-56, yielding an intermediate and mature form121. Interestingly, a form of cerebellar ataxia have also been attributed to defects in PMPCA and PMPCB, the human genes encoding the MPP subunits120,124. It is proposed that defects in frataxin processing by MPP contributes to the MPP disease phenotype. Lastly, familial recessive Parkinson’s disease is influenced by defects in human PTEN-induced Kinase 1

(Pink1). Pink1 is recruited to the mitochondrial outer membrane upstream of Parkin induced mitophagy125,126. Pink1 is processed by MPP and attenuation of β-MPP has been shown to recruit

Pink1 to the mitochondria OM and induce mitophagy127.

13

1.6 Probing Mitochondrial Import Mechanisms with Small Molecule

Inhibitors

Studies elucidating the role of mitochondrial proteins in biogenesis have primarily relied on yeast genetic manipulations and in vitro biochemical assays. However, mutations and knockouts may have unintended side effects that are unrelated to the target. Biochemical assays with recombinant protein are done in vitro and lack cellular context. Furthermore, genetic manipulations in mammalian cells can be more challenging than yeast model systems. An alternative to probing protein function is the use of small molecules which have established themselves as powerful tools in basic research and human therapy. Probing with small molecules has the advantage of cell permeability, temporal control and selectivity for specific protein functions. Several small molecule modulators are established in crucial cellular processes that include autophagy, translation, proteolysis, and even specific complexes of the respiratory chain128–132. The Koehler lab is applying this approach by developing a “tool box” of small molecules specific for mitochondria import pathways. These small molecules coined with the name “MitoBloCK” were screened and characterized by a combination of genetic and in vitro techniques118,133,134 (Figure 1-6).

The first of these small molecules, aptly named MitoBloCK-1, is specific for carrier protein import by the Tim22 complex (Figure 1-6). Screening of carrier pathway inhibitors was done using a lethality screen against a yeast temperature sensitive mutant of Tim10, tim10-1, at the permissive temperature. Secondary screens were performed by rescuing tim10-1 mutants with an integrated Tim10 gene or a plasmid harboring wild-type TIM10. Small molecules causing lethality in only tim-10-1 were considered MitoBloCK candidates. Further testing of lead compounds insured the small molecules do not cause general mitochondrial defects in membrane

14 integrity, oxidative phosphorylation or polarization. From an initial screen of 40,000 compounds,

MitoBlock-1 was identified as a potent inhibitor of the carrier pathway. Further characterization reveal treatment of MitoBloCK-1 abolishes the Tim9-Tim10 interaction through radiolabeled import crosslinking of AAC135. An advantage of MitoBloCK-1 is that it shows strong specificity for the carrier pathway as it was screened using genetic mutants. However, MitoBloCK-1 inhibition is exclusive to the mutant strain and not wild-type mitochondria, making it difficult to study general import mechanisms.

To develop more versatile small molecule inhibitors, an in vivo screen was developed to identify small molecules that disrupt presequence dependent import in wild-type yeast. A Su9-

Ura3 fusion protein containing the targeting sequence of subunit 9 of N. crassa F1Fo ATPase was fused to the N-terminus of genetic reporter Ura3. If mitochondrial import is attenuated in yeast growing in media lacking uracil, Su9-Ura3 will remain in the cytosol and allow growth134. As a control, a tim23-2 mutant expressing the Su9-Ura3 construct grew much faster than the isogenic wild-type strain in media lacking uracil. A screen of 30,000 compounds identified MitoBloCK compounds MB-10, MB-12 (DECA) and MB-20 (Figure 1-6)118,134,136. These small molecules were characterized as potent inhibitors of mitochondria protein import without disrupting general mitochondria function. Further characterization of these compounds shows a preference for the presequence pathway. Specifically, MB-10 and MB-20 inhibit assembly of the PAM complex during import. PAM complex inhibition likely occurs through disruption of interactions between

Tim44 and PAM complex components. The clinical applications of this screen were studied in the potential treatment of primary hyperoxaluria (PH1). As discussed previously, known causes

P11LG170R of PH1 include an AGT mutation that directs the AGT enzyme to mitochondria instead of the peroxisome. Interestingly, dequalinium chloride (DECA, aka MitoBloCK-12) is a Food

15 and Drug Administration (FDA) approved drug that proved to be a potent inhibitor of Su9-Ura3 import (Figure 1-6). DECA has been approved as an antibacterial treatment for oral and vaginal applications137. Though the mode of import inhibition is unknown, DECA treatment in isolated

CHO-K1 cells was able to redirect mutant AGTP11LG170R to the peroxisome118. This screen demonstrates a proof of concept for screening of general inhibitors of the presequence pathway.

However, unlike the genetic screen identifying MitoBloCK-1, characterizing the precise target and mechanism of inhibition remains a challenge for several Su9-Ura3 inhibitors.

To address the ambiguity of small molecule targets and mechanisms affecting import in genetic screens, in vitro screens against the recombinant target of interest were also developed in the Koehler lab. Specifically, a screen against recombinant Erv1 from the Mia40/Erv1 disulfide relay system in yeast was utilized to identify novel inhibitors133. The screen was specific for Erv1 sulfhydryl oxidase activity or recombinant Erv1 with DTT, producing hydrogen peroxide H2O2.

Production of H2O2 was then measured using a standard fluorometric assay with Amplex Red and horseradish peroxidase (HRP)38,133. A counter screen was then done to assess any inhibition of the Amplex Red-HRP reaction by the small molecule. A commercial library of 50,000 compounds was screened, yielding 29 compounds that were chosen for subsequent characterization. MitoBloCK-6 was then chosen for its potency and specificity for Erv1 without disruption of general mitochondrial function (Figure 1-6). Further studies with MitoBloCK-6 shows its disruption of the ternary complex between substrate, Mia40 and Erv1. Moreover, ALR

(human Erv1) is enriched in embryonic, neuronal and hematopoietic stem cells138. Given the high level of homology between the yeast and human mitochondria import pathways, the effects of MitoBloCK-6 on differentiated and non-differentiated human lineages was studied.

Interestingly, MitoBloCK-6 specifically induces apoptosis in hESCs and not differentiated cells.

16

This suggests a role for ALR in pluripotent stem cell maintenance133. Studies with MitoBlock-6 on zebrafish show morphologic and cardiac defects that were consistent with ALR downregulation, suggesting target specificity with in vivo models. Cumulatively, these results demonstrate the feasibility of in vitro screens for identifying novel inhibitors of mitochondria protein import.

Genome wide haploinsufficiency screens in yeast have also been used to identify novel inhibitors of mitochondria protein import. These screens work by lowering the dosage of target genes from two copies to one copy in diploid cells, yielding a heterozygote. Increased sensitivity to drugs in heterozygotes versus the wild-type can identify genetic targets139. A lipopeptide produced by Streptomyces known as stendomycin was identified in a S. cerevisiae screen with

Tim17 and Tim23 being likely targets. Studies in mammalian cells with mitochondrial targeted

GFP constructs (mito-EGFP) show import is blocked with stendomycin treatment, suggesting specificity for the Tim23 translocon in mammalian cells. Interestingly, higher concentrations of stendomycin also show perturbations in membrane potential in addition to import defects. This could be attributed to further disruptions of the Tim23 complex or to other non-specific effects on mitochondrial function140. Overall, the haploinsufficiency method of screening demonstrates both the advantages of target identification while working within a relevant cellular context.

1.7 An In Vitro Screen for MPP Inhibitors

Studies characterizing MPP primarily have relied heavily on genetic yeast manipulations and biochemical assays. However, given the intricate and dynamic processes of the mitochondrial import machinery, novel methods are needed to study the role of MPP. A small molecule screen based on an in vitro cleavage assay by recombinant MPP was reported by Colin

Douglas, a previous member of the Koehler lab141. The in vitro assay was based on signal

17 fluorescence of a substrate peptide (YDHA7) flanked by a fluorophore and quencher moiety.

Incubation of recombinant MPP protein with the synthetic peptide yields an end-point fluorescent signal relative to enzyme activity. Interestingly, MB-50 and MB- demonstrated import inhibition of several yeast precursors destined for the presequence pathway and carrier pathways141 (Figure 1-6). This contradicts with previous studies suggesting the function of MPP is solely in presequence cleavage and not import81,96,97,142. However, these studies rely on genetic depletion and temperature-sensitive mutations that may have auxiliary effects that mask potential functions of MPP. Meanwhile MB-50 and MB-51 demonstrate specific inhibition of MPP and mitochondrial import without perturbations in polarization or membrane integrity in isolated mitochondria141.

These results suggest a novel role for MPP in regulating mitochondrial protein import. In chapters 2 through 4 I will further characterize MB-50, MB-51 and other hits from the MPP screen regarding their effects on import, complex assembly and overall mitochondrial function. I will also include data independent of small molecule treatment on the potential broader roles of

MPP in mitochondrial biogenesis.

18

Figure 1-1 Outer Membrane Protein Import through the SAM Pathway. The SAM channel is composed of core protein Sam50 and signal receptor Sam35. After translocation through the

TOM complex, small chaperones in the IMS known as Tim9-10 and Tim8-13 (not shown) guide the precursors to the SAM complex. Recognition of the β-signal by Sam35 induces a conformational change that allows lateral release of the precursor into the lipid phase of the outer membrane. Sam37 assists in precursor release from the SAM complex.

19

Figure 1-2 Intermembrane Space Import through the Mia40/Erv1 Disulfide Relay.

Precursors destined for the IMS containing CX3C or CX9C motifs first enter the IMS through the

TOM complex. The disulfide relay is initiated by Mia40 which oxidizes internal cysteine residues and allows folding of the precursor into a mature protein. Mia40 is reoxidized by sulfhydryl oxidase Erv1. Erv1 then transfers its electrons to respiratory chain component cytochrome C and ultimately O2, thus regenerating the system for precursor oxidation.

20

Figure 1-3 Inner Membrane Import through the Carrier Pathway. Import of carrier proteins into the inner membrane depends on multiple internal targeting signals, cystosolic chaperones

Hsp70 and Hsp90, and the TIM22 complex. Initially, the precursor carrier protein is translated in the cytosol and guided to the TOM complex with the aid of chaperones Hsp70 and Hsp90. The chaperone-precursor complex can then bind to TOM receptor Tom70 and induce formation of

Tom70 oligomers. The precursor is then translocated through the TOM complex into the IMS.

The precursor then forms a complex with Tim9-Tim10 and later Tim12 at the IM on the surface of the Tim22 complex. Tim54 allows binding of the chaperone precursor complex to the TIM22 channel. The precursor is then laterally released into the inner membrane in a potential- dependent fashion by TIM22.

21

Figure 1-4 Presequence Pathway through the Tim23 Complex. N-terminally targeted matrix proteins take the presequence pathway through the TOM and TIM23 complexes. Translocation through the TOM complex depends on the amphipathic and α-helical nature or the N-terminal targeting sequence. After translocation through the TOM complex, the TIM23 complex can guide precursors into the matrix in a potential dependent manner. Two homodimers of Tim23 and Tim17 form the core of the Tim23 channel. The PAM complex provides the driving force behind precursor import and includes Tim44, mtHsp70, Pam16, Pam17 and Pam18 (Tim14). mtHsp70 provides the energy required for import through nucleotide exchange. ATPase activity

22 of mtHsp70 activity is stimulated by J-proteins Pam16 and Pam18 which also bind to Tim44.

Precursors are subsequently processed by specific mitochondrial proteases that facilitate maturation of the protein. Mitochondrial processing peptidase (MPP) cleaves the majority of presequences of matrix targeted proteins. Oct1/MIP is responsible for producing an octapeptide from the MPP generated N-terminus of precursor proteins. IMP is responsible for cleaving off hydrophobic sorting signals from proteins facing the IMS. Once precursors are processed

PreP/Mop112 degrades the presequence peptide.

23

Figure 1-5 MPP Cleavage Mechanism. Illustration of MPP cleavage mechanism of CoXIV (2-

25) presequence. Based on crystal structure data from Taylor et. al 200198. MPP does not cleave a specific sequence but prefers motifs with an arginine at positions -2, -3 or -10 and an aromatic or bulky hydrophobic residue at position +1 from the cleavage site. Here an arginine at position -

2 (P2) upstream of the cleavage site and a leucine is at position +1 (P1’) downstream of the cleavage site. The active site containing the inverted zinc binding motif (HXXEHX76E) is represented by residues βHis-70, βGlu-73, βHis-74 and βGlu-150. Cleavage of peptides occurs through a mechanism analogous to thermolysin peptidase. A Zn2+ coordinates with a water molecule near the βGlu-73 residue. Polarization of the water molecule by βGlu-73 aligns it for 24 nucleophilic attack on the carbonyl carbon of the peptide bond and subsequent cleavage. Critical residues that stabilize the substrate interaction include βGlu-160 and βAsp-164 which accommodates the -2 arginine near the zinc binding site. The βPhe-77 residue appears important for interacting with the bulky +1 aromatic residue of common substrate motifs.

25

Figure 1-6 Inhibitors of Mitochondrial Import Pathways. A summary of small molecules that target mitochondrial import pathways. MitoBloCK-1 is specific for import through the Tim22 complex and is effective only in a yeast temperature sensitive mutant of Tim10, tim10-1, at the permissive temperature (indicated by the asterisk)135. MitoBloCK-6 targets blocks the

Mia40/Erv1 disulfide relay by inhibiting the sulfhydryl oxidase Erv1. Stendomycin,

MitoBloCK-10 and MitoBloCK-12 (aka DECA) block Tim23 dependent import of the presequence pathway. Stendomycin was identified in a yeast haploinsufficiency screen in Tim23 and Tim17 heterozygotes140. MitoBloCK-10 and MitoBloCK-12 were identified in a gain of growth screen in yeast expressing Su9-Ura3. MitoBloCK-10 appears to inhibit import by disrupting interactions between Tim44 and components of the PAM complex134. MitoBloCK-12 26 aka DECA (dequalinium chloride) is an FDA approved antibacterial drug capable of blocking

Su9-Ura3 import118. MitoBloCK-50 and MitoBloCK-51 were potent inhibitors identified in an in vitro screen against recombinant yeast MPP141.

27

References 1. Wang, C. & Youle, R. J. The role of mitochondria in apoptosis*. Annu. Rev. Genet. 43, 95–118 (2009). 2. Kluck, R. M., Bossy-Wetzel, E., Green, D. R. & Newmeyer, D. D. The release of cytochrome c from mitochondria: a primary site for Bcl-2 regulation of apoptosis. Science 275, 1132–6 (1997). 3. Rossé, T. et al. Bcl-2 prolongs cell survival after Bax-induced release of cytochrome c. Nature 391, 496–499 (1998). 4. Duchen, M. R. Mitochondria and calcium: from cell signalling to cell death. J. Physiol. 529 Pt 1, 57–68 (2000). 5. Hekimi, S., Wang, Y. & Noë, A. Mitochondrial ROS and the Effectors of the Intrinsic Apoptotic Pathway in Aging Cells: The Discerning Killers! Front. Genet. 7, 161 (2016). 6. Lin, J. L. J. et al. Oxidative Stress Impairs Cell Death by Repressing the Nuclease Activity of Mitochondrial Endonuclease G. Cell Reports 16, (2016). 7. Shpilka, T. & Haynes, C. M. The mitochondrial UPR: mechanisms, physiological functions and implications in ageing. Nat. Rev. Mol. Cell Biol. 19, 109–120 (2017). 8. Pellegrino, M. W., Nargund, A. M. & Haynes, C. M. Signaling the mitochondrial unfolded protein response. Biochim. Biophys. Acta 1833, 410–6 (2013). 9. Martin, W. F. The Origin of Mitochondria. Nat. Educ. 3, 58 (2010). 10. Hsia, J. C., Chen, W. L., Long, R. A., Wong, L. T. & Kalow, W. Existence of phospholipid bilayer structure in the inner membrane of mitochondria. Proc. Natl. Acad. Sci. U. S. A. 69, 3412–5 (1972). 11. Zhang, Q. et al. Biosynthesis and roles of phospholipids in mitochondrial fusion, division and mitophagy. Cell. Mol. Life Sci. 71, 3767–3778 (2014). 12. Horvath, S. E. & Daum, G. Lipids of mitochondria. Prog. Lipid Res. 52, 590–614 (2013). 13. van Vliet, A. R. & Agostinis, P. Mitochondria-Associated Membranes and ER Stress. in 73–102 (Springer, Cham, 2017). doi:10.1007/82_2017_2 14. Ellenrieder, L. et al. Separating mitochondrial protein assembly and endoplasmic reticulum tethering by selective coupling of Mdm10. Nat. Commun. 7, 13021 (2016). 15. Nargund, A. M., Pellegrino, M. W., Fiorese, C. J., Baker, B. M. & Haynes, C. M. Mitochondrial import efficiency of ATFS-1 regulates mitochondrial UPR activation. Science 337, 587–90 (2012). 16. Sickmann, A. et al. The proteome of Saccharomyces cerevisiae mitochondria. Proc. Natl. Acad. Sci. U. S. A. 100, 13207–12 (2003). 17. Perocchi, F. et al. Assessing systems properties of yeast mitochondria through an interaction map of the organelle. PLoS Genet. 2, e170 (2006). 18. Borst, P. & Grivell, L. A. The Mitochondrial Genome of Yeast. Cell 15, (1978).

28

19. Palmfeldt, J. & Bross, P. Proteomics of human mitochondria. Mitochondrion 33, 2–14 (2017). 20. Perry, A. J. et al. Structure, topology and function of the translocase of the outer membrane of mitochondria. Plant Physiol. Biochem. 46, 265–274 (2008). 21. Dekker, P. J. et al. Preprotein translocase of the outer mitochondrial membrane: molecular dissection and assembly of the general import pore complex. Mol. Cell. Biol. 18, 6515–24 (1998). 22. Bausewein, T. et al. Cryo-EM Structure of the TOM Core Complex from Neurospora crassa. Cell 170, 693–700 (2017). 23. Model, K. et al. Protein translocase of the outer mitochondrial membrane: role of import receptors in the structural organization of the TOM complex. J. Mol. Biol. 316, 657–666 (2002). 24. Roise, D. & Schatz, G. Mitochondrial presequences. J. Biol. Chem. 263, 4509–11 (1988). 25. Vögtle, F.-N. et al. Global analysis of the mitochondrial N-proteome identifies a processing peptidase critical for protein stability. Cell 139, 428–39 (2009). 26. Saitoh, T. et al. Tom20 recognizes mitochondrial presequences through dynamic equilibrium among multiple bound states. EMBO J. 26, 4777–87 (2007). 27. Yamano, K. et al. Tom20 and Tom22 Share the Common Signal Recognition Pathway in Mitochondrial Protein Import. J. Biol. Chem. 283, 3799–3807 (2008). 28. Chacinska, A. et al. Importing mitochondrial proteins: machineries and mechanisms. Cell 138, 628–44 (2009). 29. Stojanovski, D., Guiard, B., Kozjak-Pavlovic, V., Pfanner, N. & Meisinger, C. Alternative function for the mitochondrial SAM complex in biogenesis of α-helical TOM proteins. J. Cell Biol. 179, 881–893 (2007). 30. Ishikawa, D., Yamamoto, H., Tamura, Y., Moritoh, K. & Endo, T. Two novel proteins in the mitochondrial outer membrane mediate β-barrel protein assembly. J. Cell Biol. 166, 621–627 (2004). 31. Kutik, S. et al. Dissecting Membrane Insertion of Mitochondrial β-Barrel Proteins. Cell 132, 1011–1024 (2008). 32. Chan, N. C. & Lithgow, T. The Peripheral Membrane Subunits of the SAM Complex Function Codependently in Mitochondrial Outer Membrane Biogenesis. Mol. Biol. Cell 19, 126–136 (2008). 33. Becker, T. et al. The mitochondrial import protein Mim1 promotes biogenesis of multispanning outer membrane proteins. J. Cell Biol. 194, 387–95 (2011). 34. Becker, T. et al. Biogenesis of the Mitochondrial TOM Complex. J. Biol. Chem. 283, 120–127 (2008). 35. Banci, L. et al. MIA40 is an oxidoreductase that catalyzes oxidative protein folding in mitochondria. Nat. Struct. Mol. Biol. 16, 198–206 (2009).

29

36. Erdogan, A. J. & Riemer, J. Mitochondrial disulfide relay and its substrates: mechanisms in health and disease. Cell Tissue Res. 367, 59–72 (2017). 37. Bourens, M. et al. Role of twin Cys-Xaa9-Cys motif cysteines in mitochondrial import of the cytochrome C oxidase biogenesis factor Cmc1. J. Biol. Chem. 287, 31258–69 (2012). 38. Dabir, D. V et al. A role for cytochrome c and cytochrome c peroxidase in electron shuttling from Erv1. EMBO J. 26, 4801–11 (2007). 39. Chacinska, A. et al. Essential role of Mia40 in import and assembly of mitochondrial intermembrane space proteins. EMBO J. 23, 3735–3746 (2004). 40. Stojanovski, D. et al. Mitochondrial protein import: precursor oxidation in a ternary complex with disulfide carrier and sulfhydryl oxidase. J. Cell Biol. 183, 195–202 (2008). 41. Chacinska, A. et al. Mitochondrial Biogenesis, Switching the Sorting Pathway of the Intermembrane Space Receptor Mia40. J. Biol. Chem. 283, 29723–29729 (2008). 42. Young, J. C., Hoogenraad, N. J. & Hartl, F. U. Molecular chaperones Hsp90 and Hsp70 deliver preproteins to the mitochondrial import receptor Tom70. Cell 112, 41–50 (2003). 43. Wiedemann, N., Pfanner, N. & Ryan, M. T. The three modules of ADP/ATP carrier cooperate in receptor recruitment and translocation into mitochondria. EMBO J. 20, 951– 60 (2001). 44. Wu, Y. & Sha, B. Crystal structure of yeast mitochondrial outer membrane translocon member Tom70p. Nat. Struct. Mol. Biol. 13, 589–593 (2006). 45. Rehling, P. et al. Protein Insertion into the Mitochondrial Inner Membrane by a Twin-Pore Translocase. Science (80-. ). 299, 1747–1751 (2003). 46. Wagner, K. et al. The Assembly Pathway of the Mitochondrial Carrier Translocase Involves Four Preprotein Translocases. Mol. Cell. Biol. 28, 4251–4260 (2008). 47. Kerscher, O., Holder, J., Srinivasan, M., Leung, R. S. & Jensen, R. E. The Tim54p– Tim22p Complex Mediates Insertion of Proteins into the Mitochondrial Inner Membrane. J. Cell Biol. 139, 1663–1675 (1997). 48. Kang, Y. et al. Tim29 is a novel subunit of the human TIM22 translocase and is involved in complex assembly and stability. Elife 5, (2016). 49. Mokranjac, D. & Neupert, W. The many faces of the mitochondrial TIM23 complex. Biochim. Biophys. Acta 1797, 1045–54 (2010). 50. Dekker, P. J. T. et al. 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 (1997). 51. Chacinska, A. et al. Mitochondrial Presequence Translocase: Switching between TOM Tethering and Motor Recruitment Involves Tim21 and Tim17. Cell 120, 817–829 (2005). 52. Milisav, I., Moro, F., Neupert, W. & Brunner, M. Modular structure of the TIM23 preprotein translocase of mitochondria. J. Biol. Chem. 276, 25856–61 (2001). 53. Donzeau, M. et al. Tim23 Links the Inner and Outer Mitochondrial Membranes. Cell 101,

30

401–412 (2000). 54. 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. 289, 28689–96 (2014). 55. Meier, S., Neupert, W. & Herrmann, J. M. Conserved N-terminal negative charges in the Tim17 subunit of the TIM23 translocase play a critical role in the import of preproteins into mitochondria. J. Biol. Chem. 280, 7777–85 (2005). 56. Tamura, Y. et al. Tim23-Tim50 pair coordinates functions of translocators and motor proteins in mitochondrial protein import. J. Cell Biol. 184, 129–41 (2009). 57. Alder, N. N., Jensen, R. E. & Johnson, A. E. Fluorescence Mapping of Mitochondrial TIM23 Complex Reveals a Water-Facing, Substrate-Interacting Helix Surface. Cell 134, 439–450 (2008). 58. Meinecke, M. et al. Tim50 Maintains the Permeability Barrier of the Mitochondrial Inner Membrane. Science (80-. ). 312, 1523–1526 (2006). 59. 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. 179, 1115–22 (2007). 60. Mokranjac, D., Popov-Celeketić, D., Hell, K. & Neupert, W. Role of Tim21 in mitochondrial translocation contact sites. J. Biol. Chem. 280, 23437–40 (2005). 61. van der Laan, M., Hutu, D. P. & Rehling, P. On the mechanism of preprotein import by the mitochondrial presequence translocase. Biochim. Biophys. Acta 1803, 732–9 (2010). 62. Yamano, K., Kuroyanagi-Hasegawa, M., Esaki, M., Yokota, M. & Endo, T. Step-size Analyses of the Mitochondrial Hsp70 Import Motor Reveal the Brownian Ratchet in Operation. J. Biol. Chem. 283, 27325–27332 (2008). 63. Liu, Q., D’Silva, P., Walter, W., Marszalek, J. & Craig, E. A. Regulated Cycling of Mitochondrial Hsp70 at the Protein Import Channel. Science (80-. ). 300, 139–141 (2003). 64. Wilcox, A. J., Choy, J., Bustamante, C. & Matouschek, A. Effect of protein structure on mitochondrial import. Proc. Natl. Acad. Sci. U. S. A. 102, 15435–40 (2005). 65. Hutu, D. P. et al. Mitochondrial protein import motor: differential role of Tim44 in the recruitment of Pam17 and J-complex to the presequence translocase. Mol. Biol. Cell 19, 2642–9 (2008). 66. Schiller, D., Cheng, Y. C., Liu, Q., Walter, W. & Craig, E. A. Residues of Tim44 involved in both association with the translocon of the inner mitochondrial membrane and regulation of mitochondrial Hsp70 tethering. Mol. Cell. Biol. 28, 4424–33 (2008). 67. Josyula, R., Jin, Z., Fu, Z. & Sha, B. Crystal Structure of Yeast Mitochondrial Peripheral Membrane Protein Tim44p C-terminal Domain. J. Mol. Biol. 359, 798–804 (2006). 68. Rassow, J. et al. Mitochondrial protein import: biochemical and genetic evidence for interaction of matrix hsp70 and the inner membrane protein MIM44. J. Cell Biol. 127, 1547–56 (1994).

31

69. Schneider, H.-C. et al. Mitochondrial Hsp70/MIM44 complex facilitates protein import. Nature 371, 768–774 (1994). 70. D’Silva, P., Liu, Q., Walter, W. & Craig, E. A. Regulated interactions of mtHsp70 with Tim44 at the translocon in the mitochondrial inner membrane. Nat. Struct. Mol. Biol. 11, 1084–1091 (2004). 71. van der Laan, M. et al. Motor-free mitochondrial presequence translocase drives membrane integration of preproteins. Nat. Cell Biol. 9, 1152–9 (2007). 72. Glick, B. S. et al. Cytochromes c1 and b2 are sorted to the intermembrane space of yeast mitochondria by a stop-transfer mechanism. Cell 69, 809–822 (1992). 73. Glick, B. S., Wachter, C., Reid, G. A. & Schatz, G. Import of cytochrome b2 to the mitochondrial intermembrane space: the tightly folded heme-binding domain makes import dependent upon matrix ATP. Protein Sci. 2, 1901–17 (1993). 74. Patron, M., Sprenger, H.-G. & Langer, T. m-AAA proteases, mitochondrial calcium homeostasis and neurodegeneration. Cell Res. 28, 296–306 (2018). 75. Gakh, O., Cavadini, P. & Isaya, G. Mitochondrial processing peptidases. Biochim. Biophys. Acta - Mol. Cell Res. 1592, 63–77 (2002). 76. Teixeira, P. F. & Glaser, E. Processing peptidases in mitochondria and chloroplasts. Biochim. Biophys. Acta 1833, 360–70 (2013). 77. Luciano, P. & Géli, V. The mitochondrial processing peptidase: function and specificity. Experientia 52, 1077–82 (1996). 78. Pollock, R. A. et al. The processing peptidase of yeast mitochondria : the two co-operating components MPP and PEP are structurally related. 7, 3493–3500 79. Yang, M. et al. The MAS-encoded Processing Protease of Yeast Mitochondria. 6416– 6423 (1991). 80. Geli, V., Yang, M. J., Suda, K., Lustig, A. & Schatz, G. The MAS-encoded processing protease of yeast mitochondria. Overproduction and characterization of its two nonidentical subunits. J. Biol. Chem. 265, 19216–22 (1990). 81. Jensen, R. E. & Yaffe, M. P. Import of proteins into yeast mitochondria: the nuclear MAS2 gene encodes a component of the processing protease that is homologous to the MAS1-encoded subunit. EMBO J. 7, 3863–71 (1988). 82. Vögtle, F.-N. et al. Mitochondrial protein turnover: role of the precursor intermediate peptidase Oct1 in protein stabilization. Mol. Biol. Cell 22, 2135–43 (2011). 83. Gerdes, F., Tatsuta, T. & Langer, T. Mitochondrial AAA proteases — Towards a molecular understanding of membrane-bound proteolytic machines. Biochim. Biophys. Acta - Mol. Cell Res. 1823, 49–55 (2012). 84. Kalousek, F., Isaya, G. & Rosenberg, L. E. Rat liver mitochondrial intermediate peptidase (MIP): purification and initial characterization. EMBO J. 11, 2803–9 (1992). 85. Luo, W., Fang, H. & Green, N. Substrate specificity of inner membrane peptidase in yeast mitochondria. Mol. Genet. Genomics 275, 431–436 (2006). 32

86. Hugosson’, M., Andreu3, D., Boman’, H. G., Glaser’, E. & Glaser, E. Antibacterial peptides and mitochondrial presequences affect mitochondrial coupling, respiration and protein import. Eur. J. Biochem 223, (1994). 87. Chow, K. M. et al. Mammalian pitrilysin: substrate specificity and mitochondrial targeting. Biochemistry 48, 2868–77 (2009). 88. Bhushan, S., Johnson, K. A., Eneqvist, T. & Glaser, E. Proteolytic mechanism of a novel mitochondrial and chloroplastic PreP peptidasome. Biol. Chem. 387, 1087–90 (2006). 89. Falkevall, A. et al. Degradation of the Amyloid beta-Protein by the Novel Mitochondrial Peptidasome, PreP. J. Biol. Chem. 281, 29096–29104 (2006). 90. Kambacheld, M., Augustin, S., Tatsuta, T., Müller, S. & Langer, T. Role of the Novel Metallopeptidase MoP112 and Saccharolysin for the Complete Degradation of Proteins Residing in Different Subcompartments of Mitochondria. J. Biol. Chem. 280, 20132– 20139 (2005). 91. Tatsuta, T. & Langer, T. Quality control of mitochondria: protection against neurodegeneration and ageing. EMBO J. 27, 306–14 (2008). 92. Bota, D. A., Van Remmen, H. & Davies, K. J. A. Modulation of Lon protease activity and aconitase turnover during aging and oxidative stress. FEBS Lett. 532, 103–6 (2002). 93. Nolden, M. et al. The m-AAA protease defective in hereditary spastic paraplegia controls assembly in mitochondria. Cell 123, 277–89 (2005). 94. Langer, T., Käser, M., Klanner, C. & Leonhard, K. AAA proteases of mitochondria: quality control of membrane proteins and regulatory functions during mitochondrial biogenesis. Biochem. Soc. Trans. 29, 431–6 (2001). 95. Burkhart, J. M., Taskin, A. A., Zahedi, R. P. & Vögtle, F.-N. Quantitative Profiling for Substrates of the Mitochondrial Presequence Processing Protease Reveals a Set of Nonsubstrate Proteins Increased upon Proteotoxic Stress. J. Proteome Res. 14, 4550–4563 (2015). 96. Yaffe, M. P. & Schatz, G. Two nuclear mutations that block mitochondrial protein import in yeast. Proc. Natl. Acad. Sci. U. S. A. 81, 4819–4823 (1984). 97. Witte, C., Jensen, R. E., Yaffe, M. P. & Schatz, G. MAS1, a gene essential for yeast mitochondrial assembly, encodes a subunit of the mitochondrial processing protease. EMBO J. 7, 1439–47 (1988). 98. Taylor, A. B. et al. Crystal Structures of Mitochondrial Processing Peptidase Reveal the Mode for Specific Cleavage of Import Signal Sequences. Structure 9, 615–625 (2001). 99. Luciano, P., Tokatlidis, K., Chambre, I., Germanique, J.-C. & Géli, V. The mitochondrial processing peptidase behaves as a zinc-metallopeptidase. J. Mol. Biol. 280, 193–199 (1998). 100. Kučera, T. et al. A Computational Study of the Glycine-Rich Loop of Mitochondrial Processing Peptidase. PLoS One 8, e74518 (2013). 101. Nagao, Y. et al. Glycine-rich region of mitochondrial processing peptidase alpha-subunit

33

is essential for binding and cleavage of the precursor proteins. J. Biol. Chem. 275, 34552– 6 (2000). 102. Dvořáková-Holá, K. et al. Glycine-Rich Loop of Mitochondrial Processing Peptidase α- Subunit Is Responsible for Substrate Recognition by a Mechanism Analogous to Mitochondrial Receptor Tom20. J. Mol. Biol. 396, 1197–1210 (2010). 103. Braun, H.-P. P. & Schmitz, U. K. Are the ‘core’ proteins of the mitochondrial bc1 complex evolutionary relics of a processing protease? Trends Biochem. Sci. 20, 171–175 (1995). 104. Schulte, U. et al. A family of mitochondrial proteins involved in bioenergetics and biogenesis. Nature 339, 147–149 (1989). 105. Emmermann, M., Braun, H.-P., Arretzs, M. & Schmitzg, U. K. THE JOURNAL OF BIOLOO~CAL CHEMISTRY Characterization of the Bifunctional Cytochrome c Reductase- processing Peptidase Complex from Potato Mitochondria*. 268, (1993). 106. Yang, M., Jensen, R. E., Yaffe, M. P., Oppliger, W. & Schatz, G. Import of proteins into yeast mitochondria: the purified matrix processing protease contains two subunits which are encoded by the nuclear MAS1 and MAS2 genes. EMBO J. 7, 3857–62 (1988). 107. Ou, W. J., Ito, A., Okazaki, H. & Omura, T. Purification and characterization of a processing protease from rat liver mitochondria. EMBO J. 8, 2605–2612 (1989). 108. DiMauro, S. & Schon, E. A. Mitochondrial disorders in the nervous system. Annu. Rev. Neurosci. 31, 91–123 (2008). 109. Ng, Y. S. & Turnbull, D. M. Mitochondrial disease: genetics and management. J. Neurol. 263, 179–91 (2016). 110. Gorman, G. S. et al. Prevalence of nuclear and mitochondrial DNA mutations related to adult mitochondrial disease. Ann. Neurol. 77, 753–9 (2015). 111. Roesch, K., Curran, S. P., Tranebjaerg, L. & Koehler, C. M. Human deafness dystonia syndrome is caused by a defect in assembly of the DDP1/TIMM8a-TIMM13 complex. Hum. Mol. Genet. 11, 477–86 (2002). 112. Jin, H. et al. A novel X–linked gene, DDP, shows mutations in families with deafness (DFN–1), dystonia, mental deficiency and blindness. Nat. Genet. 14, 177–180 (1996). 113. Davey, K. M. et al. 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. 43, 385–393 (2005). 114. Jackson, T. D., Palmer, C. S. & Stojanovski, D. Mitochondrial diseases caused by dysfunctional mitochondrial protein import. Biochem. Soc. Trans. 46, 1225–1238 (2018). 115. Hatle, K. M. et al. MCJ/DnaJC15, an endogenous mitochondrial repressor of the respiratory chain that controls metabolic alterations. Mol. Cell. Biol. 33, 2302–14 (2013). 116. Kang, Y. et al. Sengers Syndrome-Associated Mitochondrial Acylglycerol Kinase Is a Subunit of the Human TIM22 Protein Import Complex. Mol. Cell 67, 457–470.e5 (2017). 117. Danpure, C. J., Cooper, P. J., Wise, P. J. & Jennings, P. R. An enzyme trafficking defect 34

in two patients with primary hyperoxaluria type 1: peroxisomal alanine/glyoxylate aminotransferase rerouted to mitochondria. J. Cell Biol. 108, 1345–52 (1989). 118. Miyata, N. et al. Pharmacologic rescue of an enzyme-trafficking defect in primary hyperoxaluria 1. Proc. Natl. Acad. Sci. U. S. A. 111, 14406–11 (2014). 119. Koutnikova, H. et al. Studies of human, mouse and yeast homologues indicate a mitochondrial function for frataxin. Nat. Genet. 16, 345–51 (1997). 120. Jobling, R. K. et al. PMPCA mutations cause abnormal mitochondrial protein processing in patients with non-progressive cerebellar ataxia. Brain 138, awv057- (2015). 121. Cavadini, P., Adamec, J., Taroni, F., Gakh, O. & Isaya, G. Two-step processing of human frataxin by mitochondrial processing peptidase. Precursor and intermediate forms are cleaved at different rates. J. Biol. Chem. 275, 41469–75 (2000). 122. Branda, S. S. et al. Yeast and human frataxin are processed to mature form in two sequential steps by the mitochondrial processing peptidase. J. Biol. Chem. 274, 22763–9 (1999). 123. Harding, A. E. Friedreich’s ataxia: a clinical and genetic study of 90 families with an analysis of early diagnostic criteria and intrafamilial clustering of clinical features. Brain 104, 589–620 (1981). 124. Vögtle, F.-N. et al. Mutations in PMPCB Encoding the Catalytic Subunit of the Mitochondrial Presequence Protease Cause Neurodegeneration in Early Childhood. Am. J. Hum. Genet. 102, 557–573 (2018). 125. Silvestri, L. et al. Mitochondrial import and enzymatic activity of PINK1 mutants associated to recessive parkinsonism. Hum. Mol. Genet. 14, 3477–3492 (2005). 126. Mills, R. D. et al. Biochemical aspects of the neuroprotective mechanism of PTEN- induced kinase-1 (PINK1). J. Neurochem. 105, 18–33 (2008). 127. Greene, A. W. et al. Mitochondrial processing peptidase regulates PINK1 processing, import and Parkin recruitment. EMBO Rep. 13, 378–85 (2012). 128. Meric-Bernstam, F. & Gonzalez-Angulo, A. M. Targeting the mTOR signaling network for cancer therapy. J. Clin. Oncol. 27, 2278–87 (2009). 129. Mauvezin, C. & Neufeld, T. P. Bafilomycin A1 disrupts autophagic flux by inhibiting both V-ATPase-dependent acidification and Ca-P60A/SERCA-dependent autophagosome-lysosome fusion. Autophagy 11, 1437–1438 (2015). 130. Brand, M. D. & Nicholls, D. G. Assessing mitochondrial dysfunction in cells. Biochem. J. 435, 297–312 (2011). 131. Schneider-Poetsch, T. et al. Inhibition of eukaryotic translation elongation by cycloheximide and lactimidomycin. Nat. Chem. Biol. 6, 209–217 (2010). 132. Kisselev, A. F. & Goldberg, A. L. Proteasome inhibitors: from research tools to drug candidates. Chem. Biol. 8, 739–758 (2001). 133. Dabir, D. V. V et al. A small molecule inhibitor of redox-regulated protein translocation into mitochondria. Dev. Cell 25, 81–92 (2013). 35

134. Miyata, N. et al. Adaptation of a Genetic Screen Reveals an Inhibitor for Mitochondrial Protein Import Component Tim44. J. Biol. Chem. 292, 5429–5442 (2017). 135. Hasson, S. A. et al. Substrate specificity of the TIM22 mitochondrial import pathway revealed with small molecule inhibitor of protein translocation. Proc. Natl. Acad. Sci. U. S. A. 107, 9578–83 (2010). 136. Cheung, T. M. J. A small molecule dissociates the PAM motor from the TIM23 channel. (University of California Los Angeles, 2017). 137. Weissenbacher, E. R. et al. 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. 73, 8–15 (2012). 138. Phillips, R. L. et al. The Genetic Program of Hematopoietic Stem Cells. Science (80-. ). 288, 1635–1640 (2000). 139. Baetz, K. et al. Yeast genome-wide drug-induced haploinsufficiency screen to determine drug mode of action. Proc. Natl. Acad. Sci. U. S. A. 101, 4525–30 (2004). 140. Filipuzzi, I. et al. Stendomycin selectively inhibits tIM23-dependent mitochondrial protein import HHS Public Access Author manuscript. Nat Chem Biol 13, 1239–1244 (2017). 141. Douglas, Joseph, C. Investigation and Characterization of Small Molecule Modulators for Mitochondrial Processing Peptidase. (University of California, Los Angeles, 2014). 142. Yaffe, M. P., Ohta, S. & Schatz, G. A yeast mutant temperature-sensitive for mitochondrial assembly is deficient in a mitochondrial protease activity that cleaves imported precursor polypeptides. EMBO J. 4, 2069–74 (1985).

36

Chapter 2: Characterizing Utility and Mitochondrial Effects by MitoBloCK-50 and MitoBloCK-51

2.1 Introduction: Previous Reports on MitoBloCK-50 and MitoBloCK-51

An estimated 70% of imported mitochondrial proteins contain an amino terminal mitochondrial targeting sequence (MTS) and processing proteases in the matrix and intermembrane space coordinate the removal and degradation of the MTS1. The matrix processing peptidase (MPP) mediates the first cleavage of the presequence2, followed by the

Oct1 protease that removes an additional 8 amino acids. Once the cleavage occurs, the presequence peptidase/metallopeptidase 1 (PreP/MP1, Cym1 in yeast) degrades the targeting sequence3–5. If the precursor localizes to the intermembrane space, the inner membrane protease

(IMP) mediates a second cleavage of the sorting sequence. In addition, other specialized processing pathways have been developed that are substrate-specific.

MPP, assembling as a heterodimer of α-MPP (Mas2/PMPCA) and β-MPP

(Mas1/PMPCB), is a metalloendopeptidase in the pitrilysin family. Both subunits of MPP are

6–8 essential for viability in yeast . β-MPP possesses the inverted HXXEHX74-76E zinc binding motif for cleavage while α-MPP carries a glycine-rich loop proposed to aid in substrate binding/release9. Common cleavage site motifs have an arginine at position -2, -3 or -10 or an aromatic amino acid at position +1 from the cleavage site; however, other substrates have no discernible motif2,9,10.

Studies characterizing MPP rely heavily on genetic yeast manipulations and biochemical assays. However, given the intricate and dynamic processes of the mitochondrial import

37 machinery, novel methods are needed to study the role of MPP. A small molecule screen based on an in vitro cleavage assay by recombinant MPP was reported by Colin Douglas, a previous member of the Koehler lab11. The in vitro assay was based on signal fluorescence of a substrate peptide (YDHA7) flanked by fluorophore and quencher moiety (Figure 2-1A)11. Incubation of

MPP protein with the synthetic peptide yields an end-point fluorescent signal inversely proportional to the inhibition by a given compound. The screen was performed in 384-well plates with DMSO as a positive vehicle control and a combination of chelators (EDTA and o- phenanthroline) as a negative control (Figure 2-1B). Roughly 130,000 compounds were screened at the California NanoSystems Institute at UCLA, yielding 88 confirmed hits. Metrics to evaluate the validity of high-throughput screens can be determined by a low hit rate less than 0.14%, and a screening coefficient (Z’) value greater than 0.512. The screen produced a Z’ value of 0.8, with a hit rate of 0.07%. Colin “cherry picked” 28 compounds from the 88 confirmed hits for further study and grouped them based on their central scaffolding. The “cherry picked” compounds were selected based on the absence of known moieties with high reactivity or likely to disrupt membrane integrity. The “cherry picked compounds were also absent from a screen identifying inhibitors of presequence peptidase (PreP), another metalloendopeptidase from the same family as MPP. The screen identified MB-50 and MB-51 as structurally distinct and potent inhibitors of

MPP (Figure 2-2).

The IC50 for MPP inhibition in vitro for MB-50 and MB-51 were 1.32 µM and 2.99µM respectively (Figure 2-3A)11. MB-50 and MB-51 were well tolerated in MIC50 tests against

Hek293 cells. Neither MB-50 or MB-51 caused any significant decrease in cell viability in concentrations below 100 µM (Figure 2-3B)11. Mitochondrial morphology was also preserved at concentrations up to 100 µM MB-50 and MB-51 in Hela cells transfected with mitochondrial

38 localized dsRed (MLS dsRed)11. Radiolabeled import assays of several Tim23 precursors were performed with isolated yeast mitochondria treated with MB-50 and MB-51. Su9-DHFR is a canonical Tim23 destined precursor targeted to the matrix and is cleaved twice by MPP. Import of Su9-DHFR was drastically inhibited by MB-50 at 50 µM and by MB-51 at 100 µM. Other tested precursors include the CytB2(1-167)-DHFR fusion protein which is cleaved twice by MPP and is laterally sorted into the intermembrane space by the Tim23SORT complex13. Meanwhile, the CytB2(A63P)-DHFR mutant is targeted to the matrix because of a mutation in the stop

14 transfer sequence and is cleaved twice by MPP . Import of both CytB2(1-167)-DHFR and the

CytB2(A63P)-DHFR mutant were inhibited by MB-50 and MB-51 at similar concentrations to

Su9-DHFR inhibition11. Surprisingly, import of a matrix precursor Cpn10 (yeast Hsp10) was also inhibited by MB-50 and MB-51 despite not having a cleavable presequence. These results suggest a broader function for MPP in import rather than solely presequence procesing6–8,15. MB-

50 and MB-51 also displayed varying potencies toward other Tim23 destined precursors. While

MB-50 showed robust inhibition of Su9-DHFR, IMS localized CytB2(1-167)-DHFR and

CytB2(A63P)-DHFR, Colin observed a more limited inhibition for CytC1. Meanwhile MB-51

11 showed significant inhibition of CytC1 but not for CytB2(A63P)-DHFR . The mechanisms that dictate precursor specificity of MB-50 and MB-51 is still unknown.

Structural-Activity-Relationship (SAR) studies were performed on several analogs of

MB-50 and MB-51. The selected MB-50 SAR compounds were based on a para-t-Butyl phenol linked to a substituted benzene ring via a methyl imine moiety that was common among positive

MB-50 analogs (Figure 2-2A). Positive analog 50.1 substitutes the para-t-butyl group with an ethyl moiety. Negative analog 50-A1 removes the para-t-butyl group entirely while another negative analog 50-A2 removes the hydroxyl group from the phenol and replaces the para-t-

39 butyl group with a geometrically equivalent trifluoromethyl group. Notably, 50.1 is the only MB-

50 SAR compound that inhibits both MPP cleavage in vitro, and mitochondrial import (Figure 2-

4 A and C). These results suggest the ethyl moiety may be suitable to retain inhibition of MPP, but removal of the hydroxyl or the para-t-butyl group entirely, abolishes MPP inhibition. A

SAR study of MB-51 was centered on its quinone/hydroquinone moiety similar to coenzyme Q and was present in all positive MB-51 analogs (Figure 2-2B). MB-51.1 and MB-51.2 positive analogs retained the quinone moiety as well as inhibition of MPP during [35S] Su9-DHFR in vitro cleavage assays (Figure 2-4B). Negative analogs 51-A1, 51-A2 and 51-A3 were selected based on substitution of the quinone/hydroquinone moiety (Figure 2-2B). 51-A1 replaces the hydroxyl groups in the quinone for a methyl, which allows for some similarity in geometry to the original quinone but a loss of activity that the original group possessed. 51-A2 changes the quinine moiety for a phenol, removing one of the hydroxyl groups. Lastly, 51-A3 completely substitutes the quinone moiety for a benzene ring, thereby removing both hydroxyl groups. None of the negative analogs displayed inhibition of [35S] Su9-DHFR cleavage in vitro, suggesting the crucial role for the quinone moiety in MB-51 (Figure 2-4B). However, import assays of [35S]

Su9-DHFR display some import inhibition for all negative analogs. Positive analogs MB-51.1 and MB-51.2 show import inhibition comparable to MB-51 (Figure 2-4D). This suggests that

MB-51 import inhibition does not depend on inhibition of MPP cleavage activity.

Screened hits including MB-50 and MB-51 were also tested for non-specific mechanisms that perturb mitochondria function using secondary assays. The membrane integrity assay assesses the ability of small molecules to act like detergents by solubilizing membranes. The assay involves treating isolated mitochondria with small molecules before centrifugation, separating the supernatant from the membrane bound vesicles. DMSO and the ionic detergent

40

TritonX-100 were used as controls. Unlike the TritonX-100 treatment, MB-50 or MB-51 do not disrupt membrane integrity by leaking mitochondrial proteins into the supernatant (Figure 2-5A).

Non-specific effects on respiration can also disrupt mitochondria function and can be accurately measured using a Clark-type electrode16. Respiration is induced by adding an ETC substrate such as NADH to isolated mitochondria which leads to an increase in oxygen consumption. Adding an ionophore such as CCCP causes membrane depolarization and uncoupling of ATP synthesis from the proton gradient, leading to rapid oxygen consumption. Neither MB-50 or MB-51 caused the same increase in oxygen consumption relative to CCCP at concentrations where import inhibition is shown (Figure 2-5B). Lastly, small molecules that act like ionophores depolarize mitochondria which impedes potential dependent import. A membrane potential dependent fluorescent probe 3,3’-dipropylthiadicarbocyanine iodide (DiSC3(5)) can be used to assess whether small molecules decouple mitochondria17. The assay relies on the ability of mitochondria quenching the fluorescent DiSC3(5) probe in normal healthy mitochondria. Upon depolarization, DiSC3(5) is no longer quenched and fluorescence increases. MB-51 showed no significant increase fluorescence upon treatment at 100 µM relative to CCCP control (Figure 2-

5C). Interestingly, MB-50 at 100 µM showed a slight increase in fluorescence relative to DMSO control. This suggests MB-50 may partially reduce membrane potential, a mechanism that may contribute to the import inhibition. However, complete import inhibition by MB-50 is seen at concentrations at or below 50 µM. It is possible that higher concentrations of MB-50 may cause secondary deleterious effects in isolated mitochondria not related to MPP.

In recent years, zebrafish has proven to be a highly versatile organism for disease modeling and high-throughput screening. Zebrafish offers several advantages over other model organisms due to its low-cost, transparent larval stages, shorter life cycle, and ease for genetic

41 manipulation and drug treatment18–20. Zebrafish also share a high degree of homology to the human genome21. Zebrafish are particularly useful for mitochondrial studies because the organismal development and muscular systems (i.e., the cardiac system) are extremely prone to mitochondrial abnormalities due to their reliance on ATP production. Zebrafish expressing

DsRed cardiac myocyte light chain 2 (cmlc2) were treated with varying concentrations of MB-50 and MB-51 to determine how impaired MPP affects early stage body and heart development.

These treatments were applied on embryos at 3 hours post-fertilization. The zebrafish were then allowed to grow to 3 days post-fertilization before imaging. At 20 μM MB 50 the zebrafish show defects in body curvature and a slight loss of pigmentation. At 50 μM MB-50 the body curvature and pigmentation defects were more severe (Figure 2-6A)11. Negative MB-50 analogs 50-A1 and

50-A2 did not exhibit the same body curvature defects or pigmentation deficiency at similar concentrations. Unexpectedly, positive analog MB-50.1 was lethal at all concentrations which may suggest a pathway of toxicity unrelated to MPP. Pigmentation loss was also seen for MB-51 treatment at 20 µM with increased severity at higher concentrations. However, MB-51 did not cause the body curvature defects seen with MB-50 at concentrations up to 100 μM (Figure 2-

6B)11. Positive SAR analogs MB-51.1 and MB-51.2 also exhibited loss of pigmentation at concentrations of 20 µM and higher. Negative analogs 51-A1, and 51-A3 did not exhibit pigmentation defects like MB-51, though higher concentrations displayed changes in body shape and heart development. 51-A2 displayed some depigmentation at 20 µM but this phenotype was not consistent with higher concentrations. With regards to heart development, MB-50 and MB-

51 at 20 µM exhibit clear cardiac defects that include atrial distension (Figure 2-6 A and B).

The zebrafish pigmentation defects appear unique for MB-50 and MB-51 with respect to tested MitoBloCK inhibitors. However, chelation of compounds through N-phenylthiourea

42

(PTU) and bathocuproine have also been shown to cause pigmentation defects in zebrafish22–24.

To determine the lasting effects of MB-50 and MB-51, Colin performed a wash out experiment was performed with PTU and bathocuproine as a control. As expected, significant pigmentation defects were observed for MB-50, MB-51, PTU and bathocuproine treatment as well as morphological defects for MB-50 and MB-5111. However, the pigmentation appeared to return 2 days after the drugs were washed out. Given the similar pigmentation defect demonstrated by

MB-50 and MB-51 with known chelators, a mode of MPP inhibition could be through chelation of metals from the active site. However, this hypothesis is unlikely because MPP was counter- screened against PreP, another metalloprotease from the same family. To confirm that MB-50 and MB-51 do not inhibit MPP through chelation, a titration of metal ions was performed in the

YDHA7 JPT peptide cleavage assay used in the MPP screen. Recombinant MPP was pre- incubated in the presence of MB-50 and MB-51 with varying amounts of MnCl2. After incubation, the YDHA7 JPT fluorogenic peptide was added and fluorescence was monitored over time (Figure 2-7). DMSO and the chelator o-phenanthroline, were used as controls. As expected, titration of MnCl2 significantly reduced inhibition by o-phenanthroline. However, increasing amounts of MnCl2 fail to rescue MPP activity with MB-50 and MB-51 treatments.

This suggests that MB-50 and MB-51 do not inhibit MPP through chelation.

Parkinson’s associated protein PTEN induced kinase 1 (Pink1) is recruited to the mitochondrial outer membrane upstream of Parkin induced mitophagy25–29. Specifically, Pink1 is recruited to the TOM complex and processed by MPP and PARL, triggering its degradation by the cytosolic proteasome29–32. Attenuation of β-MPP has been shown to recruit Pink1 to the mitochondria OM and induce mitophagy30. Previous work by Colin Douglas hypothesized that treatment of MB-50 or MB-51 could induce mitophagy by stabilizing Pink1 recruitment to the

43 outer membrane by inhibiting MPP dependent processing. Import inhibition of radiolabeled hPink1 import in isolated yeast mitochondria was demonstrated by MB-50 and MB-51.

Additionally, MB-51 caused accumulation of a high molecular weight aggregate during hPink1 import11. This aggregate was initially thought to be a Pink1 complex at the outer membrane.

However, MB-51 failed to accumulate at the mitochondrial outer membrane in Hela cells stably expressing Parkin-EGFP. Interestingly, MB-50 was able to induce Parkin recruitment to mitochondria but only in the presence of a suboptimal concentration of CCCP11.

Cumulatively, these results suggest MB-50 and MB-51 are potent inhibitors of MPP in vitro, but their mode of mitochondrial import inhibition requires further study. In this chapter I will further characterize MB-50 and MB-51 with respect to potency and specificity for precursors of various mitochondrial import pathways (ie. Sam, Mia40, Tim22). These studies would elucidate whether the small molecules inhibit a specific pathway or mechanism of import.

Studies were also done in conjunction with genetic manipulations of MPP. This will determine whether these inhibitors behave differently under altered MPP activity. Lastly, MB-50 and MB-

51 treatment on zebrafish display a unique pigmentation defect as well as other more common mitochondrial associated defects (Figure 2-6)11. An anti-sense knockdown of MPP subunits using morpholinos will determine whether these phenotypes are specific for MPP attenuation or non-specific side effects of MB-50 and MB-51.

2.2 Import Inhibitory Potency of MB-50 and MB-51

Previous studies by Colin Douglas show import inhibition by MB-50 and MB-51 toward a wide range of Tim23 destined precursors11. A titration was performed on Su9-DHFR import and determined inhibitory concentrations of 50 µM and 100 µM for MB-50 and MB-51 respectively11. However, several other import assays were a time course with a constant

44 concentration of 100 µM for both MB-50 and MB-51. This is over ten times the reported IC50 in vitro (Figure 2-3A). Using a saturating concentration of small molecule may produce secondary effects that alter precursor import and obscure specificity for certain precursors. A range of concentrations is needed to determine the minimum inhibitory concentration of each small molecule for each precursor. Furthermore, the only non-Tim23 destined precursor reported with

MB-50 and MB-51 treatment was AAC11. AAC import inhibition by MB-50 and MB-51 was reported as not significant. These results should be repeated with a titration using other non-

Tim23 precursors to further elucidate specificity of the compounds. I expect specificity toward

Tim23 precursors because MPP resides in the matrix and cleavage likely occurs during import of

N-terminally targeted precursors. I performed titrations with a diverse set of Tim23 and non-

Tim23 destined precursors to determine pathway specificity for each molecule.

Frataxin is an N-terminally targeted MPP substrate and is a highly conserved mitochondrial protein with a critical role in maintaining mitochondrial iron homeostasis33,34.

Frataxin is imported through the Tim23 complex and is cleaved twice by MPP at residues 41-42, and 55-56, yielding an intermediate and mature form33. A titration of MB-50 and MB-51 was performed in isolated yeast mitochondria prior to import of yeast frataxin (Yfh1) (Figure 2-8A).

Consistent with the reported inhibitory concentrations of Su9-DHFR import, MB-50 treatment substantially inhibited import of the mature and intermediate peptides at 50 µM. Furthermore, a small increase in the precursor form was observed in 25 µM and 50 µM treatments of MB-50, suggesting some accumulation of the peptide before cleavage within the mitochondria. This suggests MB-50 may prevent translocation across the inner membrane specifically. Another possibility is a partial inhibition of cleavage activity in addition to import by MB-50. In addition to the small increase in precursor at lower concentrations of MB-50, there is a notable decrease

45 in the mature form while the intermediate form stays constant. This suggests the flux forming the intermediate precursor is unchanged and that both cleavage events are inhibited equally. At 100

µM the precursor form fails to accumulate and virtually no signal is seen for the processed forms, showing a more general inhibition of import. MB-51 also shows significant inhibition of import at 50 µM and 100 µM, but to a lesser extent than MB-50. No precursor accumulation is demonstrated by MB-51 and both processed forms decrease consistently with increasing concentrations. This suggests processing by MPP is unaffected relative to the import defect when mitochondria are treated with MB-51.

To determine whether this import inhibition depends on the presence of an MPP binding motif, I performed import titrations using precursors without a cleavable presequence. Cpn10 encodes the yeast Hsp10 homolog and is imported into the mitochondrial matrix without a cleavable presesquence35. A titration of MB-50 and MB-51 on Cpn10 import showed strong inhibition (>50%) at 50 µM and 100 µM respectively (Figure 2-8B). Sym1 is a yeast homolog of

MPV17, which is associated with mitochondrial DNA depletion syndrome (MDDS). Sym1 is also a Tim23 substrate without a cleavable presequence but is imported into the inner membrane by a lateral sorting mechanism36. Import of Sym1 is also inhibited by MB-50 and MB-51, but to lesser extent than Cpn10 or Yfh1 at similar concentrations (Figure 2-8C). This suggests that MB-

50 and MB-51 can inhibit import of laterally sorted precursors without cleavable presequences.

Sym1 homolog MPV17 is also responsible for the pigmentation defect in casper zebrafish37. As noted, MB-50 and MB-51 treatment also cause a pigmentation defect in zebrafish in addition to other mitochondrial associated defects (Figure 2-6). These results suggest the inhibition of

MPV17 biogenesis could be a mechanism that induces the pigmentation defect by MB-50 and

MB-51 in zebrafish.

46

To assess whether MB-50 and MB-51 import inhibition is specific for Tim23 destined precursors, I performed similar titrations with precursors of other mitochondrial import pathways. AAC encodes the ADP/ATP carrier protein and is imported into the inner membrane by the Tim22 complex38,39. Tom40 is the channel forming protein of the Tom complex and is imported into the outer membrane by the Sam complex in a potential independent manner40–42.

Lastly, Cmc1 is a copper-binding protein of the mitochondrial intermembrane space and is imported by the Mia40/Erv1 disulfide relay43,44. MB-50 and MB-51 show inhibition of Tom40,

AAC and Cmc1 import at similar concentrations where import inhibition of Tim23 destined precursors are demonstrated (Figure 2-9).

The effects of MB-50 and MB-51 on dual-localized precursors were also tested.

Fumarase, encoded by yeast Fum1, is localized to the cytosol and mitochondria. Import and localization of depends on MPP processing at residues 24 and 2545–47. Import of fumarase is also inhibited by MB-50 but not MB-51 at 100 µM (Figure 2-10A). Another protein with complex localization is Parkinson’s associated protein Pink1. Pink1 is recruited to the mitochondrial outer membrane upstream of Parkin induced mitophagy118,19. MB-50 and to a lesser extent MB-51, also appear to inhibit Pink1 import in isolated yeast mitochondria (Figure

2-10B). These results are consistent with previous studies performed by Colin Douglas.

Cumulatively, MB-50 and MB-51 reduce import of precursors from all canonical pathways of mitochondrial import. This includes the Tim23 matrix and lateral sorting pathways, the Tim22 carrier pathway, the outer membrane SAM pathway, and the Mia40/Erv1 disulfide relay pathway. MB-50 is the more potent inhibitor of all tested precursors, typically showing strong inhibition at 50 µM. MB-51showed significant inhibition as well, but typically at the higher concentration of 100 µM for all precursors. From this data, MB-50 and MB-51 can be

47 classified as general inhibitors of protein import. This is surprising given that past studies of

MPP show its role in presequence cleavage and not import. In contrast, MB-50 and MB-51 induce a substantial defect in import efficiency with little to no accumulation of precursor. Given the diversity of tested precursors across several import pathways, it is difficult to attribute the import defects solely on MPP function. Membrane uncoupling could be a potential mechanism for disrupting import of several of the tested precursors. Notably, MB-50 did have a mild effect on membrane potential from the DiSC3(5) assay at 100 µM. However much of the inhibition by

MB-50 was accomplished by 50 µM. However, MB-50 and MB-51 also display inhibition of

Tom40 and Cmc1 which are imported in a potential independent manner. This suggests that the mechanism of import inhibition may be more complicated than membrane uncoupling. Further studies will be focused on whether the import defect can be attributed to MPP function or a non- specific effect by MB-50 and MB-51.

2.3 MPP Overexpression and Import Efficiency

Due to the broad import defects caused by MB-50 and MB-51, I asked whether overexpression of MPP would offset these defects or increase general import efficiency. Over expression of MPP was accomplished by subcloning the Mas1 and Mas2 genes in 2µ plasmids pRS424 and pRS425 respectively. Plasmids were then expressed in GA74 yeast and grown in selective media. Mitochondria extracts were obtained from the wildtype GA74 strain, vector only control (pRS424/pRS425) and the strain overexpressing MPP (MPP). Western blots of isolated mitochondria show a substantial increase in antibodies detecting MPP in the MPP over WT and vector only controls. Notably, other proteins of mitochondria import components were relatively unchanged (Figure 2-11A). This includes Pam motor components Pam16, Pam18,

Tim44 and Hsp70. To determine the relative membrane potential between each strain, the

48 potentiometric DiSC3(5) probe was applied to equal amounts of mitochondria extract from each strain and their relative fluorescence was measured. Results show a slightly lower fluorescence for the overexpressed strain relative to the control strains (Figure 2-11B). Furthermore, timecourse import of Su9-DHFR and AAC show a substantial increase in import efficiency relative to wildtype in the overexpressed strain (Figure 2-12 A and B). Su9-DHFR and AAC imports with the vector control strain also showed a smaller but significant increase in import for the overexpression strain. Similar imports were performed with the IMS and matrix localized

CytB2-DHFR precursors, showing a modest 10-20% increase in processed forms (Figure 2-11 C and D). Import of the non-cleaved Tim23 precursor Cpn10 was also increased (Figure 2-12E).

To determine whether an increase in cellular MPP can lead to resistance in import defects by

MB-50 and MB-51, a titration was performed prior to precursor import in both strains. The overexpressed strain displayed a consistent increase in Su9-DHFR import in subsequent trials.

However, the overexpressed strain did not show any significant resistance to import defects in both MB-50 and MB-51 titrations (Figure 2-13A). Similar results were demonstrated with Cpn10 import titrations (Figure 2-13B). These experiments were repeated with the vector control strain yielding the same result (data not shown). In summary, these results suggest MPP overexpression leads to an increased membrane potential which may contribute to greater import efficiency. However, overexpression of MPP does not reduce sensitivity to import defects caused by MB-50 or MB-51.

2.4. Temperature Sensitive MPP Mutants with MB-50 and MB-51

Characterization of MPP function has been reported using temperature-sensitive Mas1 and Mas2 mutants. MPP temperature sensitive mutants display processing defects and precursor accumulation of mitochondrial proteins at non-permissive temperatures6–8,48. We wanted to ask if

49

MB-50 and MB-51 treatment in vivo would recapitulate precursor accumulation seen is the MPP mutants. A GA74 mutant lacking ABC transporters SNQ2 and PDR5 was used to increase intracellular concentrations of MB-50 and MB-5149–51. WT∆pdr5∆snq2 was grown in YPEG and treated with 100 µM of MB-50 and MB-51. In addition, apomorphine was identified as a potent inhibitor in the in vitro MPP screen and was included in these experiments. Apomorphine was an interesting candidate because it is an FDA approved drug used to treat Parkinson’s disease52,53.

As a control, temperature sensitive mas1 (tsMas1) and mas2 (tsMas2) strains were grown in

YPEG and switched to the non-permissive temperature of 37ºC for 16 hours prior to cell harvesting. As expected, MPP precursors Mdh1, Mge1 and ATPase subunit F1-β showed precursor accumulation in the tsMas2 mutant. In contrast, MB-50, MB-51 and apomorphine showed no precursor accumulation relative to the DMSO control for all substrates (Figure 2-

14A). Unexpectedly, the tsMas1 mutant only showed precursor accumulation for Mge1. Past studies have demonstrated the ability of the tsMas1 strain to accumulate precursors of F1-β and

MDH18,48. Adjusting the incubation time at the non-permissive temperature for the tsMas1 strain could address this issue. Overall, this data suggests MB-50, MB-51 and apomorphine lack the potency to recapitulate the precursor accumulation demonstrated by the temperature-sensitive

MPP mutants in vivo. This could be explained by overall stability of the small molecules in vivo, or an inadequate intracellular concentration.

Next, I wanted to ask if MB-50 and MB-51 would further ablate processing or impede import in the temperature-sensitive mutants. The rationale for this experiment expects an increased sensitivity to MB-50 or MB-51 in an MPP deficient background. Cytb2(1-167)-DHFR localizes to the IMS by a stop-transfer sequence after processing by MPP and IMP54. Import of

Cytb2(1-167)-DHFR in tsMas1 mitochondria exhibit precursor accumulation within the

50 mitochondria following trypsin shaving (Figure 2-14B). Treatment of MB-50 and MB-51 inhibits import of this precursor in both wildtype GA74 and tsMas1 mitochondria, suggesting

MB-50 and MB-51 affects mechanisms of import that are not linked to peptide cleavage.

ADP/ATP carrier protein AAC is imported into the inner membrane in a potential dependent manner by the Tim22 pathway and is not a substrate of MPP. The tsMas1 mutant shows no significant defect in AAC import relative to the wildtype GA74 strain. Furthermore, MB-50 and

MB-51 appear to inhibit AAC import equally in both strains (Figure 2-14C). Consistent with the

AAC titration, MB-50 is a more potent import inhibitor than MB-51. This data suggests temperature sensitive mutations inhibit MPP cleavage exclusively while MB-50 and MB-51 may inhibit broader roles of MPP function in import.

2.5 Effect on Complex I and Complex II Respiration by MB-50 and MB-51

The broad import defects demonstrated by MB-50 and MB-51 could be caused by effects on the electron transport chain which may alter mitochondria coupling and membrane potential.

Previous results using a Clark-Type oxygen electrode demonstrate little change in oxygen consumption of isolated yeast mitochondria when treated with MB-50 and MB-51 (Figure 2-5B).

Though these results suggest little to no effect on mitochondria respiration, the Clark-Type electrode can only detect severe and immediate changes in the oxygen consumption rate once the drug is added. Furthermore, the single chamber of the Clark-Type electrode allows for only sequential treatments, limiting the scope and utility of the experiments. We decided to utilize the

Agilent Seahorse XF Analyzer to determine any effects MB-50 and MB-51 may have on mitochondria respiration. Unlike the Clark-Type electrode, the Agilent Seahorse XF Analyzer can run multiple chambers in parallel, allowing a variety of respiratory conditions to be tested.

This includes the respiratory flux through Complex I and Complex II (state 2), ADP induced

51 substrate respiration (state 3ADP), oligomycin induced state 4 respiration (state 4o), and maximal respiratory capacity induced by FCCP (state 3u)55–57. Mitochondria extracts were purified from isolated mouse liver and the respiratory flux through Complex I and Complex II was measured.

Treatment with complex I substrates pyruvate and malate enabled complex I respiration.

Meanwhile, treatment with succinate and rotenone, a complex I inhibitor, enabled complex II respiration55,56. The effects of MB-50 and MB-51 on the respiratory flux through Complex I and

Complex II was measured under two conditions. The first condition deemed “inject” measures the immediate effects on respiration by injecting the small molecule after sequential addition of

ADP (state 3ADP) and oligomycin (state 4o). As a control, CCCP was injected in parallel in a separate well to measure maximum respiratory capacity (state 3u). Antimycin A was added last to inhibit Complex III dependent respiration. The second condition labeled as “Pre-treat” measures the effects on respiration if the mitochondria were pre-treated with the small molecule for an extended period. Mitochondria were pre-treated with 50 µM of MB-50 and MB-51 for 30 minutes before sequential addition of ADP (state 3ADP), oligomycin (state 4o), FCCP (state 3u) and antimycin A. Conditions measuring both complex I and complex II respiration demonstrated the expected effects of ADP, oligomycin and antimycin on oxygen consumption in the DMSO injection control. Injection of MB-50 and MB-51 showed no significant changes to respiration relative to DMSO and were also responsive to the sequential antimycin A treatment (Figure 2-

15). DMSO controls for pre-treatment display a slightly lower respiratory capacity in state 3ADP relative to injection conditions. This could be attributed to a high concentration of 1% DMSO that may alter mitochondria respiratory capacity when treated for an extended period.

Meanwhile, MB-50 pre-treatment in complex I and complex II respiration shows little response to ADP and FCCP treatment (Figure 2-15 A and C). MB-51 treatment also shows insensitivity to

52

ADP and FCCP induced respiration in both complex I and complex II (Figure 2-15 B and D). All pre-treatments with MB-50 and MB-51 show a dramatic decrease in oxygen consumption after antimycin A treatment, demonstrating the respiratory activity is still complex III dependent.

Cumulatively, these results suggest MB-50 and MB-51 show little to no immediate defects in complex I or complex II respiration when the drugs are injected. Meanwhile, pre-treatment with

MB-50 or MB-51 induces a significant decrease in respiratory capacity in both state 3ADP and state 3u respiration. This suggests the uncoupling effect demonstrated by MB-50 and MB-51 are not as immediate as an ionophore such as FCCP. Therefore, the mechanism of uncoupling may be more complex and appears to affect both complex I and complex II respiration.

2.6 MPP Morpholino Treatment of Zebrafish Phenocopy MB-50 and MB-51

Associated Defects

To determine whether the zebrafish phenotypes demonstrated by MB-50 and MB-51 treatment are specific for MPP inhibition, other methods are required to alter MPP activity in vivo. An anti-sense morpholino knockdown of MPP was used to determine whether the phenotypes are recapitulated. Zebrafish embryos were injected with morpholinos targeting αMPP and βMPP at the 1-4 cell stage. The embryos were then monitored for up to 3 months post- fertilization. At 24 hours post-fertilization, morpholino treated embryos appeared behind developmentally and exhibit a pigment deficiency (Figure 2-16). Notably, a pigment deficiency was also observed for MB-50 and MB-51 treatment (Figure 2-6). At 72 hours post-fertilization the pigmentation returned, suggesting the MPP morpholino knockdown was attenuated over time. However, severe developmental defects were observed at 72 hours and later. This includes a strong body curvature, severe craniofacial defects, an underdeveloped tail fin, and a swollen yolk sac. Fish that survived to 3 months post-fertilization typically exhibit the craniofacial

53 defects and an underdeveloped or abnormal tail fin. These results suggest early inhibition of

MPP causes permanent and severe downstream developmental defects. Furthermore, the pigmentation defect observed by MB-50 and MB-51 treatment is recapitulated in the MPP morpholino knockdown in early-stage development. This phenotype is not demonstrated with any other MitoBloCK inhibitors tested on zebrafish and appears to be specific for MPP inhibition.

2.7 Discussion

Previous reports characterizing MB-50 and MB-51 describe them as potent MPP inhibitors that alter mitochondria import in organello and produce morphological abnormalities in vivo. Key results include import inhibition toward Tim23 precursors, morphological and pigmentation defects in zebrafish, and a limited ability to recruit Parkin to mitochondrial membranes in Hela cells expressing Parkin-EGFP. However, there is little evidence to support specificity of MB-50 and MB-51 toward their intended target. Because MB-50 and MB-51 were identified using an in vitro model against recombinant yeast MPP, their utility as MPP inhibitors may be limited to isolated systems. At the minimum, MB-50 and MB-51 could help elucidate the role of MPP in processing and import in isolated yeast mitochondria. Therefore, I decided to focus on their specificity toward a diverse set of precursors in radiolabeled import studies. Tested

Tim23 precursors include Su9-DHFR, CytB2(1-167)-DHFR, CytB2(1-167)A63P-DHFR, Yfh1,

Cpn10 and Sym1. This diverse set of precursors comprise all known pathways of the Tim23 complex. Tested precursors that are not imported through the Tim23 complex include Tom40,

AAC and Cmc1. Many of these precursors were accompanied by a titration of either MB-50 or

MB-51 prior to import. Table 2.1 shows the cumulative results of all import experiments done by me and Colin Douglas. An inhibitory concentration for 50% inhibition (MIC50) is provided if a

54 titration was performed on the indicated precursor. My findings suggest MB-50 and MB-51 are general import inhibitors with no significant preference for any single import pathway.

Specifically, MB-50 appears to significantly inhibit all tested Tim23 and non-Tim23 destined precursors at 50 µM. Likewise, MB-51 inhibited all tested precursors at 100 µM with the exception of fumarase (Fum1). Inhibition of Tom40 and Cmc1 at similar concentrations is a noteworthy result because their import is not potential dependent and are localized to the outer membrane and inner membrane space respectively. Because MPP is localized exclusively to the matrix, it seems unlikely that it would have the broad role in import suggested by MB-50 and

MB-51. Another possibility is that MB-50 and MB-51 may inhibit respiration or membrane coupling to prevent import of potential-dependent precursors. Studies using the Agilent Seahorse

XF Analyzer on isolated mitochondria from mouse liver display an uncoupling effect by both

MB-50 and MB-51 (Figurer 2-15). Consistent with the import experiments, MB-50 appears to be the more potent compound with regard to mitochondria uncoupling. However, because the uncoupling is not immediate, it is unlikely that MB-50 and MB-51 function as ionophores such as FCCP. Instead, MB-50 and MB-51 may disrupt respiration through mechanisms that alter electron transport or the flow of protons across the inner membrane. Further studies are required to determine the nature of MB-50 and MB-51 induced uncoupling and whether it is linked to

MPP function or an off-target effect.

Temperature-sensitive mutants of MPP accumulate precursor in vivo as well as in radiolabeled import assays (Figure 2-14). However, whole cell extracts from yeast treated with

MB-50 and MB-51 failed to recapitulate the precursor accumulation and inhibit import rather than processing in radiolabeled import assays. Furthermore, temperature-sensitive MPP mutants do not show increased sensitivity to MB-50 or MB-51 with respect to import or processing.

55

Similarly, the overexpressed MPP strain did not display increased resistance to MB-50 or MB-51 treatment (Figure 2-13 A and B). Overall, these results fail to correlate the import inhibition displayed by MB-50 and MB-51 to MPP inhibition in isolated yeast mitochondria. However, specificity can be argued by the shared pigmentation defect demonstrated through morpholino knockdown of MPP and MB-50/51 treatment in zebrafish. One possible mechanism for this phenotype is the import inhibition of MPV17. Notably MB-50 and MB-51 appear to block import of MPV17 homolog Sym1 in yeast (Figure 2-8C). MPV17 is responsible for the pigmentation defect in casper zebrafish and the inhibition of MPV17 biogenesis could be a mechanism that causes the pigmentation defect by MB-50 and MB-5137. However, it is also possible that MB-50 and MB-51 may induce pigment deficiency through mechanisms unrelated to MPP or MPV17 deficiency.

One of the more significant findings is the increased coupling and import efficiency when

MPP is overexpressed (Figure 2-12 and Figure 2-13). Increased import was observed for Su9-

DHFR, CytB2(1-167), CytB2(1-167)A63P-DHFR, Cpn10 and AAC. This result suggests the broader role of MPP in import of potential dependent precursors. In chapter 4 I will further explore these findings and the potential links between MPP, import and mitochondria coupling.

56

Figure 2-1 Model of JPT fluorogenic peptide assay and plate layout for high-throughput screen. (A) High-throughput screen for MPP inhibitors is based on cleavage of a YDHA7 JPT

57 fluorogenic peptide (JPT Peptide Technologies Inc.) with recombinant MPP. In vitro cleavage by MPP occurs at the +2 position from the indicated arginine (red). Following cleavage, the proprietary quencher and fluorophore are separated producing a fluorescent signal with an excitation of 330 nm and an emission of 405 nm. (B) 384-well plate layout and dispensing sequence for HTS screen. The positive control (MPP + DMSO vehicle) and negative control

(MPP + EDTA & o-phenanthroline + DMSO vehicle) were placed in the indicated columns. Due to edge effects no samples were placed in the first or the last column of the plate. Each drug was pinned into the remaining wells. A solution of 128 nM MPP in screening buffer was incubated at

25°C for 1 hour with DMSO, drug, or EDTA & o-phenanthroline. A fluorescence reading was taken at time=0 after the hour incubation. 12 µM JPT synthetic peptide was then added to initiate the cleavage reaction. An end-point fluorescence reading was taken following incubation for 25 minutes at 25ºC. Figure included with permission from Colin Douglas11.

58

Figure 2-2 Structures of MitoBloCK-50, MitoBloCK-51 and associated analogs. (A)

MitoBloCK-50, positive analog 50.1 and negative analogs: 50-A1, 50-A2. (B) MitoBloCK-51, positive analogs: 51.1, 51.2 and negative analogs: 51-A1, 51-A2, 51-A3.

59

Figure 2-3 MB-50 and MB-51 reduce MPP cleavage activity in vitro. (A) IC50 analysis of recombinant MPP (128 nM) cleavage activity with the YDHA7 peptide (0.1 to 20 µM of drug) for 90 min at 25C. Average % activity ± SD of n = 3 trials. See Materials and Methods for details. (B) MIC50 analysis of MB-50 and MB-51 with Hek293 cells was determined by treating cells for 24 hours with MB-50 and MB-51. The % viability was set relative to DMSO at 100% for n = 3 trials. Figure included with permission from Colin Douglas11.

60

Figure 2-4 Import and cleavage of [35S] Su9-DHFR with MB-50 and MB-51 SAR compounds. (A-B). Time course assay with radiolabeled Su9-DHFR with recombinant MPP at

25°C. MPP were pre-incubated with 100 µM of (A) MB-50 and analogs 50.1, 50-A1 and 50-A2, or (B) MB-51 and analogs 51.1, 51.2, 51-A1, 51-A2 and 51-A3 for 15 minutes. As a control, cleavage was performed in the presence of vehicle control, 1% DMSO or 5mM EDTA. Cleavage of [35S] Su9-DHFR was performed at the indicated time points. (C-D) Time course import of

61

[35S] Su9-DHFR in GA74 mitochondria. Mitochondria were pretreated with 1% DMSO, 50 µM

CCCP or 100 µM (C) MB-50 and analogs, or (D) MB-51 and analogs for 15 minutes and the import was performed at the indicated time points. Protease treatment was used to remove non- imported precursor. As a control, import was also performed in the presence of the vehicle control, 1% DMSO. Protease treatment was used to remove non-imported precursors. The membrane potential (ΔΨ) was dissipated with 4 µg/ml valinomycin and 50 µM CCCP. p, precursor; m, mature; the % import was set at 100% for DMSO at the last time point. Figure B-C included with permission from Colin Douglas11. Figure D included with permission from Kayla

Frank.

62

Figure 2-5 MB-50 and MB-51 do not interfere with general mitochondrial function. (A)

Membrane integrity assay of isolated GA74 mitochondria. Mitochondria were incubated with 1%

DMSO, 100 µM of drug or 0.2% TritonX-100 at 25oC for 15 minutes. After centrifugation, supernatant (S) and pellet (P) were analyzed by Coomassie stain (left) or immunoblot (right).

Immunoblot was analyzed with antibodies against Cym1, Tom70, Mia40 and Cyt C. (B) Oxygen consumption assay of GA74 mitochondria. Mitochondria were treated with 1% DMSO, 50µM

MB-50, 100µM MB-51 or 20 µM CCCP at the indicated time points. (C) Membrane potential assay using potentiometric DisC3(5) probe. GA74 mitochondria were treated with 1% DMSO,

63

100 µM of drug or 40 µM CCCP for 15 minutes. Fluorescence was measured at λem=670 nm and

λex=620 nm with 3 technical replicates.

64

Figure 2-6 MB-50 and MB-51 cause pigment deficiency and other morphological defects in zebrafish. Zebrafish were grown to 3 hours post-fertilization before administering treatment, then grown for 3 days. Following treatment, embryos were imaged using a Leica S8APO microscope. Cmlc2 embryos were stained with o-dianisidine and imaged in the dark. (A) MB-50 treatment of zebrafish from 0-50 µM. Concentrations of MB-50 higher than 100µM were lethal.

(B) MB-51 treatment of zebrafish from 0-100 µM. Figure included with permission from Colin

Douglas11.

65

Figure 2-7 Treatment with excess divalent cation Mn2+ does not rescue MPP activity after

MB-50 or MB-51 treatment. Chelation assay using the YDHA7 JPT peptide and recombinant

MPP. An MPP solution with the indicated concentrations of MnCl2 was pretreated with 100 µM of o-phenanthroline, MB-50, MB-51 or an equal volume of DMSO for 1 hour. Fluorescence was measured for 90 minutes at λem=405 nm and λex=330 nm with 3 technical replicates.

66

Figure 2-8 MB-50 and MB-51 inhibit the import of TIM23 substrates. (A) Titration import assay with [35S] Yfh1 (yeast frataxin) radiolabeled precursor was performed with yeast mitochondria at 25°C. Mitochondria were pre-incubated with increasing concentrations of MB-

50 or MB-51 for 15 minutes. Radiolabeled precursor was then added and allowed to incubate for an additional 15 minutes. As a control, import was also performed in the presence of the vehicle control, 1% DMSO. Protease treatment was used to remove non-imported precursors. The membrane potential (ΔΨ) was dissipated with 4 µg/ml valinomycin and 50 µM CCCP. (B) As in

‘A’ with [35S] Cpn10. (C) As in ‘A’ with Sym1 (yeast Mpv17) and import reaction was incubated for 30 minutes. p, precursor; i, intermediate; m, mature; the % import was set at 100% for DMSO control.

67

Figure 2-9 MB-50 and MB-51 inhibit the import of non-TIM23 substrates. (A) Titration import assay with [35S] AAC radiolabeled precursor was performed with yeast mitochondria at

25°C. Mitochondria were pre-incubated with increasing concentrations of MB-50 or MB-51 for

15 minutes. Radiolabeled precursor was then added and allowed to incubate for an additional 30 minutes. As a control, import was also performed in the presence of the vehicle control, 1%

DMSO. Protease treatment was used to remove non-imported precursors. The membrane potential (ΔΨ) was dissipated with 4 µg/ml valinomycin and 50 µM CCCP. (B) As in ‘A’ with

[35S] Tom40 and import reaction was incubated for 15 minutes. As a control, mitochondria were treated with 0.2% Triton X-100 following import. (C) As in ‘A’ with Cmc1. The % import was set at 100% for DMSO control.

68

Figure 2-10 Effects of MB-50 and MB-51 on import of dual-localized Fum1 and Pink1 in yeast mitochondria. (A) Time course import assay with [35S] Fum1 (yeast fumarase) radiolabeled precursor was performed with yeast mitochondria at 25°C. Mitochondria were pre- incubated with 100 µM MB-50 or MB-51 for 15 minutes and the import was performed for the indicated time points. As a control, import was also performed in the presence of the vehicle control, 1% DMSO. Protease treatment was used to remove non-imported precursors. The membrane potential (ΔΨ) was dissipated with 4 µg/ml valinomycin and 50 µM CCCP. (B) As in

‘A’ with [35S] hPink1. The % import was set at 100% for DMSO control at the respective time point.

69

Figure 2-11 Overexpression of MPP increases mitochondrial membrane potential. (A)

Mitochondrial proteins (10, 25, 50 µg) were purified from strains WT GA74 background, expressing vector controls, and overexpressing Mas1 and Mas2 (MPP) and separated by SDS-

PAGE followed by immunoblot analysis. Antibodies for the indicated proteins were used. Note an antibody that detects Mas1 and both Mas1 and Mas2 were used. (B) Membrane potential assay using potentiometric DisC3(5) probe. Mitochondria from the indicated GA74 strains were

70 normalized by BCA assay and treated with the DisC3(5) probe. As a control, mitochondria from each strain were pre-treated with 40 µM CCCP for 15 minutes. Fluorescence was measured at

λem=670 nm and λex=620 nm with 3 technical replicates.

71

Figure 2-12 Overexpression of MPP increases import in isolated mitochondria. (A) Time course import assay with [35S] Su9-DHFR radiolabeled precursor was performed with mitochondria from GA74 background (WT) and overexpressed MPP strain (MPP) (left), or

GA74 expressing the empty vector (Vector) and the MPP strain (right). Imports were performed for the indicated time points. As a control, import was also performed in the presence of the vehicle control, 1% DMSO. Protease treatment was used to remove non-imported precursors. The membrane potential (ΔΨ) was dissipated with 4 µg/ml valinomycin and 50 µM

35 35 CCCP. (B) As in ‘A’ with [ S] AAC. (C) As in ‘A’ with [ S] CytB2 (1-167)-DHFR. (D) As in

72

35 35 ‘A’ with [ S] CytB2 (A63P)-DHFR. (E) As in ‘A’ with [ S] Cpn10. p, precursor; i, intermediate; m, mature; the % import was set at 100% for DMSO control at the last time point.

73

Figure 2-13 Overexpression of MPP does not rescue import defects induced by MB-50 or

MB-51. (A) Titration import assay with [35S] Su9-DHFR radiolabeled precursor was performed with yeast mitochondria from the WT GA74 background strain and the overexpressed MPP

(MPP) strain at 25°C. Mitochondria were pre-incubated at the indicated concentrations of

MB-50 or MB-51 for 15 minutes. Radiolabeled precursor was then added and allowed to incubate for an additional 15 minutes. As a control, import was also performed in the presence of the vehicle control, 1% DMSO. Protease treatment was used to remove non-imported

74 precursors. The membrane potential (ΔΨ) was dissipated with 4 µg/ml valinomycin and 50 µM

CCCP. Quantification of import % is shown on the right with DMSO control at 100%. p, precursor; m, mature. (B) As in ‘A’ with [35S] Cpn10.

75

Figure 2-14 Temperature sensitive MPP mutants show processing defects and no import defects. (A) Immunoblot of GA74 ∆pdr5∆snq2 strain grown in YPEG and treated with 100 µM

MB-50, MB-51 or apomorphine. As a control, yeast were also treated with the vehicle control,

1% DMSO. Temperature-sensitive MPP mutants (tsMas1 and tsMas2) were allowed to grow at

25ºC for 36 hours before switching to the non-permissive temperature 37ºC for 16 hours before harvesting. Protein extracts were separated by SDS-PAGE followed by immunoblot analysis.

Antibodies for the indicated proteins were used. p, precursor; m, mature. (B) Time course assay

35 with [ S] Cytb2(1-167)-DHFR was performed with mitochondria from WT GA74 and tsMas1

76 mutant15. Mitochondria were incubated at 37°C for 15 minutes prior to import. Mitochondria were then pre-incubated with 100 µM MB-50 or MB-51 for 15 min and the import was performed for the indicated time points at 25ºC. As a control, import was also performed in the presence of the vehicle control, 1% DMSO. Protease treatment was used to remove non- imported precursors. The membrane potential (ΔΨ) was dissipated with 4 µg/ml valinomycin and 50 µM CCCP. (C) As in ‘B’ with AAC. p, precursor; i, intermediate; m, mature; the % import was set at 100% for import with DMSO at the last time point.

77

Figure 2-15 Pre-treatment with MB-50 and MB-51 uncouples respiration in isolated mouse liver mitochondria. (A) An Agilent Seahorse XF96 Analyzer was used to measure oxygen consumption rate under state 2, 3 and 4o respiration in isolated mouse liver mitochondria.

Mitochondria were resuspended in buffer containing 5mM pyruvate and malate (Complex I) (A-

B) or 5mM succinate and 2µM rotenone (Complex II) (C-D). For injection conditions, ADP and oligomycin were added subsequently at the indicated time points. 50 µM MB-50 (A, C) or MB-

51 (B-D) were then added to the mitochondria at the specified time point. As a control, mitochondria were incubated in the presence of vehicle, 1% DMSO, or 4 µM FCCP. Antimycin

A was added last to cease Complex III dependent respiration. For pre-treatment, the mitochondria were treated with DMSO, MB-50 or MB-51 for 30 minutes prior to sequential

78 addition of ADP, oligomycin, FCCP and antimycin A at the indicated time points. Conditions were performed in triplicate.

79

Figure 2-16 Morpholino knockdown of MPP subunits in zebrafish causes pigment deficiency and other morphological defects. Embryos collected from zebrafish were microinjected at the 1-4 cell stage with morpholino oligos targeting to zebrafish proteins MPPα and MPPβ (Gene Tools, LLC). Zebrafish were imaged at 24 hours post-fertilization (hpf), 72 hpf and 3 months post-fertilization (mpf). Figure included with permission from Jennifer Ngo.

80

MB-50 MB-51 MPP Tested Tested MIC50 MIC50 Cleavage Conc. Conc. MB-50 MB-51 Performed Precursor Sites Pathway Localization (µM) (µM) (µM) (µM) by Eric Torres, 6.25- Colin Su9-DHFR 2 Tim23 Matrix 100 25-100 25-50 >100 Douglas Eric Torres, 6.25- Colin Cpn10 None Tim23 Matrix 50 25-100 12.5-25 25-50 Douglas Dual Yfh1 2 Tim23 Matrix/Cyt 10-100 10-100 10-25 10-25 Eric Torres Eric Torres, 1.5- 1.5- Colin Pink1 1 Tim23 OM/Matrix 100 100 50-100 50-100 Douglas Sym1 None Tim23 IM 25-100 25-100 50-75 >100 Eric Torres 12.5- 12.5- Cmc1 None Mia40 IMS 100 100 25-50 >100 Eric Torres Eric Torres, 12.5- 12.5- Colin AAC None Tim22 IM 100 100 12.5-25 25-50 Douglas Tom40 None SAM OM 25-100 25-100 25-50 75-100 Eric Torres Cytb2(1- Eric Torres, 167)- Colin DHFR 2 Tim23 IMS 100 100 <100 <100 Douglas Cytb2(1- Eric Torres, 167)A63P- Colin DHFR 2 Tim23 Matrix 100 100 <100 <100 Douglas Colin Cytc1 1 Tim23 Matrix 100 100 >100 <100 Douglas Dual No Fum1 None Tim23 Matrix/Cyt 100 100 <100 inhibition Eric Torres Table 2-1 Summary of [35S] Radiolabeled Import Studies with MB-50 and MB-51. Table shows summary of all precursors tested with MB-50 and MB-51 at the indicated concentration range. The MIC50 is an estimated range where import is reduced to at least 50% inhibition relative to DMSO control. Table is cumulative of experiments performed by me and Colin

Douglas11.

81

Materials and Methods

Yeast Strains. Standard set of genetic and molecular techniques was used to generate the strains in this study. The overexpression strain was made by first sub-cloning the Mas1 and Mas2 with their endogenous promoters into separate 2μ plasmids pRS424 and pRS425 respectively. Primers were designed with RE sites BamHI and SalI for Mas1 cDNA while primers for Mas2 were designed with NotI and SalI at 5’ and 3’ respectively. Co-transformation of plasmids was done in a GA74 background and selected under auxotrophic markers Trp1 and Leu2.

Strain Background Genotype Source GA74-1A GA74 MATa his3-11,15 leu2 (Koehler et. al 1998) ura3 trp1 ade8 rho+ mit+ WT∆pdr5∆snq2 GA74 his3 leu2 ade8 trp1 (Miyata et. al 2017) ura3 pdr5∆0::HIS3 snq2∆0::KANMX WT pRS424/pRS425 GA74 MATa his3-11,15 leu2 This study ura3 trp1 ade8 rho+ mit+:: PRS424 [Empty:Trp1 2µ] PRS425 [Empty:Leu2 2µ] MPP GA74 MATa his3-11,15 leu2 This study ura3 trp1 ade8 rho+ mit+::[Mas1:Trp1 2µ] PRS424 [Mas2:Leu2 2µ] PRS425 tsMas1 AH 216 a, leu2, his3, phoC, (Yaffe et. Al 1984) phoE Table 2-S1 Strains used in this study

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 82 appropriate amino acid dropout mixture. Antibodies against Cym1, CytC, ATPase subunit F1-β, mtHsp70, Mas1, Mas1/Mas2, Mdh1, Mge1 Mia40 Tom70, Tim50, Tim44, Tim23, Tim17,

Pam16 and Pam18 are rabbit polyclonal and originated in the Schatz/Koehler labs.

Drug IC50. The IC50 for each compound was assessed using the MPP/JPT fluorogenic peptide assay. These assays were conducted in the same manner as described for the non-automated confirmation of each compound with one alteration. For each small molecule the concentration was varied from 0.1 to 20 μM. All small molecule concentrations were done in triplicate and analyzed with GraphPad Prism 5. The fluorescence was monitored over 90 min using a

FlexStation Plate Reader (Molecular Devices) with constant shaking at an excitation of 330 nm, an emission of 405 nm, and an auto cutoff of 325 nm.

Mitochondria Purification. Crude mitochondria extracts were obtained as described in previous studies43,58. Yeast cultures were kept at 25°C with vigorous shaking during growth. Mitochondria concentration was measured by BCA assay. Aliquots of 200 µg were flash frozen in liquid nitrogen and stored at -80°C.

Membrane Integrity Assay. Yeast GA74 mitochondria were thawed on ice and resuspended in osmotic buffer BB7.4 buffer (0.6 M sorbitol, 20mM Hepes pH 7.4). Mitochondria were incubated with 1% DMSO, 100 µM of drug or 0.2% TritonX-100 at 25oC for 15 minutes.

Mitochondria were then centrifuged at 8000 x g for 6 minutes. The remaining mitochondrial pellet was resuspended in Laemmli Sample buffer and resolved on SDS-PAGE. Supernatants were TCA precipitated, resuspended in Laemmli Sample buffer and resolved on SDS-PAGE.

Gels were visualized by standard Coomassie staining and immunoblotting.

83

Yeast Crude Protein Extraction. Crude protein extracts from yeast were obtained by a NaOH and 2-mercaptoethanol lysis followed by TCA precipitation. The protocol was adapted from the

Yaffe Schatz protein extraction method8. Briefly, GA74 ∆pdr5∆snq2 cultures were grown in 4 ml YPEG in the presence of 100 µM drug until mid-log phase. Temperature-sensitive MPP mutants were first grown at 25ºC for 36 hours and then switched to non-permissive temperature

37ºC overnight for 16 hours before harvesting. Cells were resuspended in 1 ml H2O and transferred to a 1.5 ml Eppendorf tube. 150 µL of NaOH/2-ME solution (1.85M NaOH, 7.4% 2- mercaptoethanol) was added to the resuspension and incubated for 10 minutes on ice. 150 µL of

50% TCA was then added to the mixture and incubated on ice for an additional 10 minutes.

Mixture was centrifuged and washed twice with acetone. Cell pellets were allowed to dry at room temperature for 10 minutes before resuspension in Laemmli sample buffer. 2 OD equivalent was loaded on an SDS-PAGE and analyzed by immunoblot.

Oxygen electrode assay. Oxygen consumption of WT mitochondria was measured using methods previously described 59. Briefly, purified WT mitochondria (25 mg/ml) were thawed on ice and tested within 2 hours. Oxygen consumption assays were performed with a Clark-type oxygen electrode in a stirred thermostatically controlled 1.5-ml chamber at 25°C (Oxytherm,

Hansatech). State II respiration was induced in a suspension of 100 µg/ml mitochondria in 0.25

M sucrose, 20 mM KCl, 20 mM Tris-Cl, 0.25 mM EDTA, 4 mM KH2PO4, and 3 mM MgCl2, pH 7.2 after adding 2 mM NADH. Consumption rate was monitored for approximately 2 min. 50

µM of MitoBloCK-50 and 100 µM of MitoBloCK-51 were added to a final concentration of 1%

DMSO and respiration was measured for at least 1 minute. Uncoupled respiration was achieved by the addition of 20 µM carbonyl cyanide 3-chlorophenylhydrazone (CCCP) to the chamber.

84

TM DiSC3[5] Assay. Membrane potential measurement assays were conducted with a Flexstation

II fluorometer. The quenching of potentiometric probe (3,3’-dipropylthiadicarbocyanine iodide)

DisC3(5) was measured at λem=670 nm and λex=620 nm. Purified GA74 mitochondria was thawed and resuspended in respiration buffer (0.6M sorbitol, 2 mM KH2PO4, 50 mM KCl,

10mM MgCl, 2.5 mM EDTA, 1 mg/ml BSA and 50 mM Hepes pH 7.1) to a final concentration of 21 µg/ml. Either 100 µM of drug or DMSO was then added to a final vehicle concentration of

1% and allowed to sit for 15 minutes. Mitochondria were also treated with 40 µM CCCP as a control. After incubation components were added to a final concentration of 100 nM DisC3(5), 2 mM NADH, and 2 µg/ml oligomycin. Reactions were allowed to incubate for an additional 5 minutes at room temperature. Fluorescence was measured in a 384-well plate with 3 technical replicates.

Mouse liver mitochondria isolation and Seahorse Oxygen Consumption assay. A single

C57BL6 mouse was euthanized with isoflurane. The extracted liver was washed with PBS, minced with scissors and homogenized in a Potter-Elvehjem homogenizer with 10ml of ice cold

MSHE buffer (200mM mannitol, 70mM sucrose, 10mM HEPES-KOH pH7.4, 1mM EDTA.

Homogenates were centrifuged at 1000g for 10 minutes at 4°C. Supernatant was collected and centrifuged at 10,000g for 10 minutes at 4°C. The mitochondrial pellets were washed twice in

MSHE buffer supplemented with 0.5% fatty-acid free BSA. Final mitochondrial pellet was re- suspended in ice cold MSHE buffer. Mitochondrial protein content was measured with a BCA assay (Pierce). An Agilent Seahorse XF96 Analyzer was used to measure oxygen consumption rate under state 2, 3 and 4o respiration. The XF96 cartridge was loaded with a total of 5µg and

3µg of protein per well for Complex I and Complex II driven respiration respectively.

Mitochondria were resuspended in 150 µL MAS buffer (70mM sucrose, 220mM mannitol, 5mM

85

KH2PO4, 5mM MgCl2, 2mM HEPES-KOH pH7.4, 1mM EGTA and 0.1% fatty-acid free BSA) with either 5mM pyruvate and malate (Complex I) or 5mM succinate and 2µM rotenone

(Complex II). Plates were incubated at 37°C in non-CO2 incubate for 8 minutes before loading into the XF96 instrument after 30 minutes calibration. Compounds were injected at the specified time at the given concentrations: 4mM ADP, 3.5µM oligomycin, 4µM FCCP, 4µM antimycin A,

1% DMSO, 50µM MB-50 and MB-51. For pre-treatment, the mitochondria were treated with

DMSO, MB-50 or MB-51 for 30 minutes prior to sequential addition of ADP, oligomycin, FCCP and antimycin A. Conditions were performed in triplicate.

MPP Chelation assay with JPT-Peptide. The assay was performed under similar conditions to the high-throughput YDHA7 JPT-peptide screen.128 nM of recombinant MPP was incubated in screening buffer (10 mM Hepes, pH 7.4, 50mM NaCl, 0.01% BSA) with the indicated concentrations of MnCl2. MPP solutions were then treated with 100 µM of o-phenanthroline,

MB-50, MB-51 or an equal volume of DMSO for 1 hour at room temperature. After incubation,

12 µM of YDHA7 JPT peptide was added to the reaction and mixed thoroughly. Fluorescence with 330 nm excitation and 405 nm emission was monitored for 90 minutes using a FlexStation

Plate Reader. Experiment was performed with three technical replicates.

[35S]- Radiolabeled import assays. Prior to import into purified mitochondria, 35S-methionine and cysteine labeled proteins were generated with TNT Quick Coupled Transcription/Translation kits (Promega) and plasmids carrying the gene of interest. Transcription of genes was driven by either the T7 or Sp6 promoter. Import reactions were performed as previously described49.

Frozen mitochondria aliquots were thawed and added to import at a concentration of 25 µg/ml.

Drug or DMSO was added as indicated to a final concentration of 1% vehicle for all experiments. Samples with dissipated membrane potential (-Δψ) were treated with 4 µg/ml 86 valinomycin and 50 µM CCCP for at least 4 minutes at 25°C. Following a 15 minute incubation at 25°C, import reactions were initiated by the addition of 5-10µL translation mix. Aliquots were removed at intervals during the reaction time-course and stopped with cold buffer containing 50

µg/ml trypsin. Soybean trypsin inhibitor was added in excess after a 15 minute incubation on ice.

Mitochondria were then pelleted by centrifugation (8000 x g, 6min) and re-suspended in

Laemmli sample buffer. Import of membrane proteins (AAC and Tom40) were carbonate extracted with an additional step to remove proteins that have not been inserted into the membrane as described in 60. Samples were resolved on SDS-PAGE and analyzed by autoradiography.

[35S]- Radiolabeled MPP cleavage assays. [35S]-methionine and cysteine labeled Su9-DHFR was made with the TNT Quick Coupled Transcription/Translation kit (Promega) prior to MPP cleavage. Recombinant MPP was resuspended in screening buffer (10mM Hepes, pH 7.4, 50mM

NaCl, 1mM MnCl2, 0.01% BSA) in a final concentration of 128 nM. Drug or DMSO was added at 100 µM or final concentration of 1% vehicle for all experiments and incubated at room temperature for 15 minutes. MPP inhibition through chelation was accomplished by the addition of 5 mM EDTA. Reaction was initiated by the addition of 10 µL translation mix. Aliquots were removed at intervals during the reaction time-course and stopped by the addition of Laemmli sample buffer. Samples were resolved on SDS-PAGE and analyzed by autoradiography.

Zebrafish. All zebrafish (Dani rerio) were maintained in accordance with standard laboratory conditions 61. The TDL6 (Tuebingen driver line) zebrafish used in all experiments was identified in a screen for developmentally regulated enhancers that drive tissue-specific expression. Gal4- driven GFP expression marks myocardial cells (cmlc2). Mitochondria are marked by a TOL2- mediated insertion of a Gal4-UAS-MLS-dsRed pC2+ construct62. 87

Zebrafish Husbandry. Zebrafish lines were maintained in a 14-hour-light/10-hour-dark cycle and mated for 1 hour to obtain synchronized embryonic development. Embryos were grown for

4dpf in 1xE3 buffer (5 mM sodium chloride, 0.17 mM potassium chloride, 0.33 mM calcium chloride, 0.33 mM magnesium sulfate) at 28.5°C63.

Zebrafish Manipulations & Morpholino Injections. Embryos collected from zebrafish lines characterized with an AB and TDL6 background were collected and microinjected at the 1-4 cell stage with 0.02 mM and 0.04 mM solution of morpholino oligos (MO) targeting to zebrafish proteins MPPα and MPPβ (Gene Tools, LLC). 1 mM of MO solutions were diluted with 1xE3 buffer (5 mM sodium chloride, 0.17mM potassium chloride, 0.33 mM calcium chloride, 0.33 mM magnesium sulfate). The following MO solutions were used:

MPPα 5’-CATGTGAGCAGCCATGTTGGATTCT 3’

MPPβ ATG 5’-GAGGCCGCCATGCTGTTTACTGACT-3’

Zebrafish were anesthetized in 0.01% tricaine at 24 hours post-fertilization (hpf), 72 hpf and 3 months post-fertilization (mpf) and imaged using a Leica S8AP0 stereomicroscope at 1.575x magnification and a Leica MZ16F fluorescent stereoscope at 11.5x magnifications. Images were then processed with the Leica LAS EZ software then resized to 300dpi without resampling in the

Adobe Photoshop software63.

Zebrafish Drug Treatment Assay. Zebrafish were mated for 1 hour to obtain synchronized embryonic development. Embryos were grown to 3 hours post-fertilization in E3 buffer (5 mM sodium chloride, 0.17 mM potassium chloride, 0.33 mM calcium chloride, 0.33 mM magnesium sulfate) and then incubated with buffer, 1% DMSO, or drug for 3 days at 28.5°C. Following treatment, embryos were imaged using a Leica S8APO microscope at 1.575x magnification or

88

Leica MZ16F fluorescent stereoscope (TexasRed filter set) at 5x magnification. Cmlc2 embryos were stained with o-dianisidine (40% (v/v) ethanol, 0.01 M sodium acetate, 0.65% hydrogen peroxide, 0.6 mg/mL o-dianisidine) and incubated for 15 minutes in complete darkness. Embryos were then washed with E3 buffer to remove residual stain and stereoscopically imaged under white light using a Leica S8APO microscope at 1.575x magnification. Images were resized to

300 dpi without resampling using Adobe Photoshop.

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.

89

References 1. Vögtle, F.-N. et al. Global analysis of the mitochondrial N-proteome identifies a processing peptidase critical for protein stability. Cell 139, 428–39 (2009). 2. Gakh, O., Cavadini, P. & Isaya, G. Mitochondrial processing peptidases. Biochim. Biophys. Acta - Mol. Cell Res. 1592, 63–77 (2002). 3. Chow, K. M. et al. Mammalian pitrilysin: substrate specificity and mitochondrial targeting. Biochemistry 48, 2868–77 (2009). 4. Bhushan, S., Johnson, K. A., Eneqvist, T. & Glaser, E. Proteolytic mechanism of a novel mitochondrial and chloroplastic PreP peptidasome. Biol. Chem. 387, 1087–90 (2006). 5. Falkevall, A. et al. Degradation of the Amyloid beta-Protein by the Novel Mitochondrial Peptidasome, PreP. J. Biol. Chem. 281, 29096–29104 (2006). 6. Jensen, R. E. & Yaffe, M. P. Import of proteins into yeast mitochondria: the nuclear MAS2 gene encodes a component of the processing protease that is homologous to the MAS1-encoded subunit. EMBO J. 7, 3863–71 (1988). 7. Witte, C., Jensen, R. E., Yaffe, M. P. & Schatz, G. MAS1, a gene essential for yeast mitochondrial assembly, encodes a subunit of the mitochondrial processing protease. EMBO J. 7, 1439–47 (1988). 8. Yaffe, M. P. & Schatz, G. Two nuclear mutations that block mitochondrial protein import in yeast. Proc. Natl. Acad. Sci. U. S. A. 81, 4819–4823 (1984). 9. Taylor, A. B. et al. Crystal Structures of Mitochondrial Processing Peptidase Reveal the Mode for Specific Cleavage of Import Signal Sequences. Structure 9, 615–625 (2001). 10. Luciano, P., Tokatlidis, K., Chambre, I., Germanique, J.-C. & Géli, V. The mitochondrial processing peptidase behaves as a zinc-metallopeptidase. J. Mol. Biol. 280, 193–199 (1998). 11. Douglas, Joseph, C. Investigation and Characterization of Small Molecule Modulators for Mitochondrial Processing Peptidase. (University of California, Los Angeles, 2014). 12. Zhang, J.-H., Chung & Oldenburg. A Simple Statistical Parameter for Use in Evaluation and Validation of High Throughput Screening Assays. J. Biomol. Screen. 4, 67–73 (1999). 13. Glick, B. S., Wachter, C., Reid, G. A. & Schatz, G. Import of cytochrome b2 to the mitochondrial intermembrane space: the tightly folded heme-binding domain makes import dependent upon matrix ATP. Protein Sci. 2, 1901–17 (1993). 14. Beasley, E. M., Müller, S. & Schatz, G. The signal that sorts yeast cytochrome b2 to the mitochondrial intermembrane space contains three distinct functional regions. EMBO J. 12, 2303–11 (1993). 15. Yaffe, M. P., Ohta, S. & Schatz, G. A yeast mutant temperature-sensitive for mitochondrial assembly is deficient in a mitochondrial protease activity that cleaves imported precursor polypeptides. EMBO J. 4, 2069–74 (1985). 16. Silva, A. M. & Oliveira, P. J. Evaluation of Respiration with Clark-Type Electrode in Isolated Mitochondria and Permeabilized Animal Cells. in Methods in molecular biology 90

(Clifton, N.J.) 1782, 7–29 (Humana Press, New York, NY, 2018). 17. Farrelly, E., Amaral, M. C., Marshall, L. & Huang, S.-G. A High-Throughput Assay for Mitochondrial Membrane Potential in Permeabilized Yeast Cells. Anal. Biochem. 293, 269–276 (2001). 18. Lieschke, G. J. & Currie, P. D. Animal models of human disease: zebrafish swim into view. Nat. Rev. Genet. 8, 353–367 (2007). 19. Delvecchio, C., Tiefenbach, J. & Krause, H. M. The Zebrafish: A Powerful Platform for In Vivo , HTS Drug Discovery. Assay Drug Dev. Technol. 9, 354–361 (2011). 20. Zon, L. I. & Peterson, R. T. In vivo drug discovery in the zebrafish. Nat. Rev. Drug Discov. 4, 35–44 (2005). 21. Barbazuk, W. B. et al. The syntenic relationship of the zebrafish and human genomes. Genome Res. 10, 1351–8 (2000). 22. O’Reilly-Pol, T. & Johnson, S. L. Neocuproine ablates melanocytes in adult zebrafish. Zebrafish 5, 257–64 (2008). 23. Ishizaki, H. et al. Combined zebrafish-yeast chemical-genetic screens reveal gene-copper- nutrition interactions that modulate melanocyte pigmentation. Dis. Model. Mech. 3, 639– 651 (2010). 24. Karlsson, J., von Hofsten, J. & Olsson, P.-E. Generating Transparent Zebrafish: A Refined Method to Improve Detection of During Embryonic Development. Mar. Biotechnol. 3, 0522–0527 (2001). 25. Silvestri, L. et al. Mitochondrial import and enzymatic activity of PINK1 mutants associated to recessive parkinsonism. Hum. Mol. Genet. 14, 3477–3492 (2005). 26. Mills, R. D. et al. Biochemical aspects of the neuroprotective mechanism of PTEN- induced kinase-1 (PINK1). J. Neurochem. 105, 18–33 (2008). 27. Ashrafi, G. & Schwarz, T. L. The pathways of mitophagy for quality control and clearance of mitochondria. Cell Death Differ. 20, 31–42 (2013). 28. Aerts, L., Craessaerts, K., De Strooper, B. & Morais, V. A. PINK1 kinase catalytic activity is regulated by phosphorylation on serines 228 and 402. J. Biol. Chem. 290, 2798– 811 (2015). 29. Yamano, K. & Youle, R. J. PINK1 is degraded through the N-end rule pathway. Autophagy 9, 1758–69 (2013). 30. Greene, A. W. et al. Mitochondrial processing peptidase regulates PINK1 processing, import and Parkin recruitment. EMBO Rep. 13, 378–85 (2012). 31. Narendra, D. P. & Youle, R. J. Targeting mitochondrial dysfunction: role for PINK1 and Parkin in mitochondrial quality control. Antioxid. Redox Signal. 14, 1929–38 (2011). 32. 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. 117, 856–67 (2011).

91

33. Cavadini, P., Adamec, J., Taroni, F., Gakh, O. & Isaya, G. Two-step processing of human frataxin by mitochondrial processing peptidase. Precursor and intermediate forms are cleaved at different rates. J. Biol. Chem. 275, 41469–75 (2000). 34. Branda, S. S. et al. Yeast and human frataxin are processed to mature form in two sequential steps by the mitochondrial processing peptidase. J. Biol. Chem. 274, 22763–9 (1999). 35. Ryan, M. T., Hoogenraad, N. J. & Høj, P. B. Isolation of a cDNA clone specifying rat chaperonin 10, a stress-inducible mitochondrial matrix protein synthesised without a cleavable presequence. FEBS Lett. 337, 152–156 (1994). 36. Reinhold, R. et al. The Channel-Forming Sym1 Protein Is Transported by the TIM23 Complex in a Presequence-Independent Manner. (2012). doi:10.1128/MCB.00843-12 37. D’Agati, G. et al. A defect in the mitochondrial protein Mpv17 underlies the transparent casper zebrafish. Dev. Biol. 430, 11–17 (2017). 38. Wiedemann, N., Pfanner, N. & Ryan, M. T. The three modules of ADP/ATP carrier cooperate in receptor recruitment and translocation into mitochondria. EMBO J. 20, 951– 60 (2001). 39. Truscott, K. N. et al. Mitochondrial import of the ADP/ATP carrier: the essential TIM complex of the intermembrane space is required for precursor release from the TOM complex. Mol. Cell. Biol. 22, 7780–9 (2002). 40. Stojanovski, D., Guiard, B., Kozjak-Pavlovic, V., Pfanner, N. & Meisinger, C. Alternative function for the mitochondrial SAM complex in biogenesis of α-helical TOM proteins. J. Cell Biol. 179, 881–893 (2007). 41. Ishikawa, D., Yamamoto, H., Tamura, Y., Moritoh, K. & Endo, T. Two novel proteins in the mitochondrial outer membrane mediate β-barrel protein assembly. J. Cell Biol. 166, 621–627 (2004). 42. Kozjak, V. et al. An essential role of Sam50 in the protein sorting and assembly machinery of the mitochondrial outer membrane. J. Biol. Chem. 278, 48520–3 (2003). 43. Dabir, D. V. V et al. A small molecule inhibitor of redox-regulated protein translocation into mitochondria. Dev. Cell 25, 81–92 (2013). 44. Bourens, M. et al. Role of twin Cys-Xaa9-Cys motif cysteines in mitochondrial import of the cytochrome C oxidase biogenesis factor Cmc1. J. Biol. Chem. 287, 31258–69 (2012). 45. Yogev, O. et al. Fumarase: a mitochondrial metabolic enzyme and a cytosolic/nuclear component of the DNA damage response. PLoS Biol. 8, e1000328 (2010). 46. Knox, C., Sass, E., Neupert, W. & Pines, O. Import into Mitochondria, Folding and Retrograde Movement of Fumarase in Yeast. J. Biol. Chem. 273, 25587–25593 (1998). 47. Regev-Rudzki, N., Yogev, O. & Pines, O. The mitochondrial targeting sequence tilts the balance between mitochondrial and cytosolic dual localization. J. Cell Sci. 121, 2423–31 (2008). 48. Burkhart, J. M., Taskin, A. A., Zahedi, R. P. & Vögtle, F.-N. Quantitative Profiling for

92

Substrates of the Mitochondrial Presequence Processing Protease Reveals a Set of Nonsubstrate Proteins Increased upon Proteotoxic Stress. J. Proteome Res. 14, 4550–4563 (2015). 49. Hasson, S. A. et al. Substrate specificity of the TIM22 mitochondrial import pathway revealed with small molecule inhibitor of protein translocation. Proc. Natl. Acad. Sci. U. S. A. 107, 9578–83 (2010). 50. Miyata, N. et al. Adaptation of a Genetic Screen Reveals an Inhibitor for Mitochondrial Protein Import Component Tim44. J. Biol. Chem. 292, 5429–5442 (2017). 51. Duncan, M. C., Ho, D. G., Huang, J., Jung, M. E. & Payne, G. S. Composite synthetic lethal identification of membrane traffic inhibitors. Proc. Natl. Acad. Sci. U. S. A. 104, 6235–40 (2007). 52. Boyle, A. & Ondo, W. Role of apomorphine in the treatment of Parkinson’s disease. CNS Drugs 29, 83–9 (2015). 53. Deleu, D., Hanssens, Y. & Northway, M. G. Subcutaneous Apomorphine. Drugs Aging 21, 687–709 (2004). 54. Gruhler, A., Ono, H., Guiard1, B., Neupert2, W. & Stuart, R. A. A novel intermediate on the import pathway of cytochrome b2 into mitochondria: evidence for conservative sorting. EMBO J. 14, 1349–1359 (1995). 55. Divakaruni, A. S., Rogers, G. W. & Murphy, A. N. Measuring Mitochondrial Function in Permeabilized Cells Using the Seahorse XF Analyzer or a Clark-Type Oxygen Electrode. in Current Protocols in Toxicology 60, 25.2.1-25.2.16 (John Wiley & Sons, Inc., 2014). 56. Brand, M. D. & Nicholls, D. G. Assessing mitochondrial dysfunction in cells. Biochem. J. 435, 297–312 (2011). 57. Rogers, G. W. et al. High Throughput Microplate Respiratory Measurements Using Minimal Quantities Of Isolated Mitochondria. PLoS One 6, e21746 (2011). 58. Glick, B. S. & Pon, L. A. Isolation of highly purified mitochondria from Saccharomyces cerevisiae. Methods Enzymol. 260, 213–23 (1995). 59. Claypool, S. M., Oktay, Y., Boontheung, P., Loo, J. A. & Koehler, C. M. Cardiolipin defines the interactome of the major ADP/ATP carrier protein of the mitochondrial inner membrane. J. Cell Biol. 182, 937–50 (2008). 60. Koehler, C. M. et al. Tim9p, an essential partner subunit of Tim10p for the import of mitochondrial carrier proteins. EMBO J. 17, 6477–86 (1998). 61. Westerfield, M. The Zebrafish Book. A Guide for the Laboratory Use of Zebrafish (Danio rerio). (University of Oregon Press, 2000). 62. Levesque, M. P., Krauss, J., Koehler, C., Boden, C. & Harris, M. P. New Tools for the Identification of Developmentally Regulated Enhancer Regions in Embryonic and Adult Zebrafish. Zebrafish 10, 21–29 (2013). 63. Miyata, N. et al. Pharmacologic rescue of an enzyme-trafficking defect in primary hyperoxaluria 1. Proc. Natl. Acad. Sci. U. S. A. 111, 14406–11 (2014).

93

Chapter 3: Characterization of Additional Selected MPP Inhibitors

3.1 Introduction

An estimated 70% of imported mitochondrial proteins contain an amino terminal mitochondrial targeting sequence (MTS) and processing proteases in the matrix and intermembrane space coordinate the removal and degradation of the MTS1. The matrix processing peptidase (MPP) mediates the first cleavage of the presequence2, followed by the

Oct1 protease that removes an additional 8 amino acids. Once the cleavage occurs, the presequence peptidase/metallopeptidase 1 (PreP/MP1, Cym1 in yeast) degrades the targeting sequence3–5. If the precursor localizes to the intermembrane space, the inner membrane protease

(IMP) mediates a second cleavage of the sorting sequence. In addition, other specialized processing pathways have been developed that are substrate-specific.

MPP, assembling as a heterodimer of α-MPP (Mas2/PMPCA) and β-MPP

(Mas1/PMPCB), is a metalloendopeptidase in the pitrilysin family. Both subunits of MPP are

6–8 essential for viability in yeast . β-MPP possesses the inverted HXXEHX74-76E zinc binding motif for cleavage while α-MPP carries a glycine-rich loop proposed to aid in substrate binding/release9. Common cleavage site motifs have an arginine at position -2, -3 or -10 or an aromatic amino acid at position +1 from the cleavage site; however, other substrates have no discernible motif2,9,10.

Studies characterizing MPP rely heavily on genetic yeast manipulations and biochemical assays. However, given the intricate and dynamic processes of the mitochondrial import machinery, novel methods are needed to study the role of MPP. A small molecule screen based

94 on an in vitro cleavage assay by recombinant MPP was reported by Colin Douglas, a previous member of the Koehler lab11. The in vitro assay was based on signal fluorescence of a substrate peptide (YDHA7) flanked by fluorophore and quencher moiety (Figure 2-1A)11. Incubation of

MPP protein with the synthetic peptide yields an end-point fluorescent signal inversely proportional to the inhibition by a given compound. Roughly 130,000 compounds were screened at the California NanoSystems Institute at UCLA, yielding 88 confirmed hits. Metrics to evaluate the validity of high-throughput screens can be determined by a low hit rate less than 0.14%, and a screening coefficient (Z’) value greater than 0.512. The screen produced a Z’ value of 0.8, with a hit rate of 0.07%. Colin “cherry picked” 28 compounds from the 88 confirmed hits for further study and grouped them based on their central scaffolding. The “cherry picked” compounds were selected based on the absence of known moieties with high reactivity or likely to disrupt membrane integrity. The “cherry picked compounds were also absent from a screen identifying inhibitors of presequence peptidase (PreP), another metalloendopeptidase from the same family as MPP. The “cherry picked” hits included MB-50 and MB-51 as structurally distinct and potent inhibitors of MPP (Figure 2-2).

Subsequent studies demonstrate that MB-50 and MB-51 appear to cause broad import defects from a diverse array of precursors. These import defects appear to be induced in isolated mitochondria only at concentrations 5 to 10-fold higher than their reported IC50 of recombinant

MPP. This contradicts with previous studies suggesting the function of MPP is solely in presequence cleavage and not import6–8,13. In addition, MB-50 and MB-51 treatment do not recapitulate the precursor accumulation of MPP substrates demonstrated in temperature sensitive mutants in vivo and in isolated mitochondria (Figure 2-13A). Cumulatively, these results limit the application of MB-50 and MB-51 to study the role of MPP in mitochondrial biogenesis. As a

95 result, I focused on characterizing other MPP inhibitors from Colin’s screen. My goal was to identify a more potent compound that induces the precursor accumulation expected for MPP inhibition. This chapter will summarize the data characterizing the selected MPP inhibitors not included in Colin’s “cherry picked” list of compounds.

3.2 Analysis of Import and Processing Defects by MPP Inhibitors

Due to the apparent limitations of MB-50 and MB-51, I focused on screening 19 uncharacterized compounds from the list of 88 confirmed MPP inhibitors. These compounds were selected based on similar principles applied to the original 28 “cherry-picked” compounds from Colin’s screen11. The new set of compounds were grouped into 7 families based on their central scaffolding (Figure 3-1). The rationale for this was to identify an MPP inhibitor that would be more potent and recapitulate the precursor accumulation expected for MPP inhibition.

Based on this objective, I focused on the effects of Su9-DHFR import in isolated mitochondria.

At the minimum, a new compound of interest should be just as potent as MB-50 in isolated mitochondria and induce precursor accumulation. Furthermore, this assay can identify general inhibitors of mitochondria import, including molecules that solubilize membranes or dissipate membrane potential. A timecourse import of Su9-DHFR was performed on all 19 selected compounds with MB-50 and CCCP as positive controls for import inhibition (Figure 3-2). To address relative potency, I kept the concentration of small molecule at 50 µM, the lowest concentration of MB-50 where precursor import is significantly reduced. Compounds that displayed import inhibition similar to MB-50 include 0042, 0100, 0017, 0136, 1125 and 0160

(Figure 3-2 A-C). Meanwhile, compounds that caused precursor accumulation and a corresponding reduction in mature form include 0022, 0015, 8247, 603 and 875. These compounds were with a lower concentration of 10 µM (Figure 3-2D). As expected, MB-50 did

96 not significantly inhibit Su9-DHFR import at 10 µM. Meanwhile, compounds 0115, 0136 and

875 continued to display potent import inhibition at the lower concentration. Interestingly, compounds 0022, 0115, 8247, 603 and 875 failed to demonstrate the robust precursor accumulation observed at 50 µM. This suggests the compounds require a higher concentration for MPP inhibition to occur in isolated mitochondria. Potent Su9-DHFR inhibitors were also tested for their effects on Cpn10 and Cmc1 import (Figure 3-3A). Compound 0115, 0136 and

875 exhibited significant inhibition at 10 µM toward Cpn10 import while compounds 8247, 603 and 0022 displayed a milder inhibition. These results were consistent with Cmc1 import inhibition at similar concentrations except for compound 8247 which did not inhibit import of

Cmc1 at 50 µM (Figure 3-3B). Effects on the import of Parkinson’s associated Pink1 was also tested with the Su9-DHFR inhibitors (Figure 3-3C). Compounds 0022, 0115, 0136, 875, 8247,

603 along with MB-50 abolished Pink1 import at 50 µM.

Next, my goal was to determine the effects on import of hydrophobic precursors not localized to the matrix. To address this, I performed a co-import and carbonate extraction of radiolabeled Tom40 and AAC following treatment with the more potent inhibitors of Su9-DHFR

(Figure 3-4). The compounds were tested at concentrations where they exhibited potent import or processing defects toward Su9-DHFR. MB-50 was also tested at 50 µM as a control. Compounds that displayed strong inhibition of Tom40 and AAC at 10 µM include 0136, 875 and 0115.

Compound 8247 also exhibited a smaller but significant inhibition of Tom40 and AAC at 50

µM. Meanwhile, compound 0022 showed little to no inhibition of Tom40 and AAC at 50 µM.

In summary, several of the selected MPP inhibitors exhibit potent inhibition of import and processing. MPP inhibitors 0115, 875 and 8247 exhibit the most severe processing defects of

Su9-DHFR and were included in subsequent import experiments. Moreover, these compounds

97 have an equal or greater potency to MB-50 with respect to in organello radiolabeled import experiments. However, each of these compounds also exhibit some import defect in non-cleaved or non-matrix precursors that include Cpn10, Cmc1, Tom40 and AAC. This suggests a lack of specificity similar to the effects observed with MB-50. In the next section I will summarize results from secondary assays that identify potential mechanisms of import inhibition and quantify potency of the selected MPP inhibitors.

3.3 Secondary Assays of MPP inhibitors

Results from the Su9-DHFR import assays demonstrate a range of effects on processing and import from the selected MPP inhibitors. To determine if the import defects can be attributed to effects on membrane potential, mitochondria were treated with compound in the presence of a potential dependent DiSC3(5) (3,3’-dipropylthiadicarbocyanine iodide) dye. The assay relies on

14 the ability of coupled mitochondria quenching the fluorescent DiSC3(5) probe . Upon depolarization, DiSC3(5) is no longer quenched and fluorescence increases. DMSO and FCCP treatment were used as controls (Figure 3-5A). Several compounds including processing inhibitors 0115, 8247, 603 and 875 did not appear to cause a signal increase. Therefore, these compounds were not considered general mitochondria uncouplers. Meanwhile, compounds 0022,

0100, 0113, 0017, 1125, 0160 and 0146 displayed a significant fluorescence increase at 100 µM.

These compounds were retested at a lower concentration of 50 µM. Under the lower concentration, compounds 0022, 0100, 0113 and 0146 showed no significant fluorescence increase and were therefore not considered uncouplers at 50 µM. Meanwhile, compounds 0017,

1125 and 0160 displayed a significant fluorescence increase at both 100 µM and 50 µM. This result is consistent with the Su9-DHFR import inhibition without the accumulation of precursor at the same concentration. I concluded compounds 0017, 1125 and 0160 may inhibit import by

98 dissipating mitochondria membrane potential and were excluded from further studies. Based on the cumulative data from the import experiments and the DiSC3(5) probe assay, compounds

0022, 0115, 8247, 603 and 875 were selected as candidate compounds for further characterization.

MPP is a metalloprotease and therefore relies on the coordination of a metal ion at the active site for cleavage activity. Treatment with known chelators leads to sequestration of coordinating metal ions and inhibition of MPP activity15. To assess the chelation activity of the

MPP inhibitors that induce processing defects, a titration of metal ions was performed in the

YDHA7 JPT peptide cleavage assay used in the MPP screen. Recombinant MPP was pre- incubated with varying amounts of MnCl2 in the presence of DMSO, o-phenanthroline or 100

µM compound. After incubation, the YDHA7 JPT fluorogenic peptide was added and fluorescence was monitored over time (Figure 3-6). Compounds 0022, 0115, 8247, 603 and 875 were selected for this assay because they displayed precursor accumulation in the Su9-DHFR import experiments. Compound 0136 was also because it was a potent import inhibitor. The

DMSO controls showed an increase in MPP activity with increasing amounts of MnCl2.

Meanwhile, negative control o-phenanthroline showed little to no activity with no MnCl2 added but a complete rescue at 0.1 mM of MnCl2. The fluorescence was not rescued after addition of saturating amounts of MnCl2 for compounds 0022, 0115, 875 and 0136. This suggests the mechanism of import for these compounds do not depend on the presence of metal ions in solution and likely do not function as chelators. Compounds 8247 and 603 displayed some rescue of MPP activity with increasing amounts of MnCl2. However, the MPP activity relative to DMSO was consistently lower at the respective MnCl2 concentrations for 8247 and 603 treatment. Therefore, unlike the o-phenanthroline treatment, saturating amounts of MnCl2 does

99 not completely abolish 8247 or 603 inhibition. Furthermore, increasing amounts of MnCl2 increases MPP activity in the DMSO control. This result complicates the nature of MPP inhibition by 8247 and 603. One possibility could be inhibition by chelation and additional mechanisms not dependent on the presence of metal ions. Another hypothesis is the inhibition displayed by 8247 and 603 is not sufficient to offset the increased activity acquired by saturating

MnCl2. Notably, 8247 and 603 displayed weaker import inhibition relative to 0115, 0136 and

875 at 10 µM (Figure 3-2D). Furthermore, compound 8247 and 603 were not identified as general inhibitors when cross-screened against PreP, another metalloprotease. Based on this reasoning, it is unlikely that 8247 and 603 function purely as chelators to inhibit MPP activity.

However, because compound 603 showed very little potency in the Su9-DHFR import assays relative to MB-50, it was not further assessed with subsequent assays.

Next, I wanted to assess if the selected compounds behaved like detergents and inhibit import by solubilizing membranes. The mitochondrial integrity assay involves treating isolated mitochondria with small molecules before centrifugation, separating the supernatant from the membrane bound vesicles. For this assay I selected the most potent compounds that exhibit precursor accumulation in Su9-DHFR import assays; this includes compounds 0022, 0115 and

875. Though compound 0136 did not accumulate precursor, it was a potent import inhibitor and was also included in the assay. DMSO and the ionic detergent TritonX-100 were used as controls. Unlike the TritonX-100 treatment, none of the tested compounds disrupted membrane integrity by leaking mitochondrial proteins into the supernatant (Figure 3-7).

To evaluate the relative potency between processing inhibition in isolated mitochondria and

MPP inhibition in vitro, I conducted an IC50 analysis toward recombinant MPP using the

YDHA7 JPT peptide cleavage assay with the candidate compounds (Figure 3-8). The IC50 of

100 compounds 8247, 0115, 875 and 0136 roughly correlated with their relative potency in the Su9-

DHFR import assays. Compound 875 was the most potent compound with an IC50 of 7.09 µM, while 8247 was the least potent compound with an IC50 of 50.1 µM. The relative toxicity in yeast was also assessed with titrations of compounds 0022, 0115, 0136 and 875. To increase intracellular concentrations, a mutant lacking drug pumps Snq2 and Pdr5 was used for the assay.

GA74∆Snq2∆Pdr5 yeast were grown in YPEG media with compound for 24 and 48 hours before their optical density was measured (Figure 3-9 and Figure 3-10). Compound 875 was the most toxic compound with the relative growth rate dropping below 50% at 50 µM for both time points

(Figure 3-9). Meanwhile, compound 0136 displayed no significant decrease in relative growth rate at concentrations up to 100 µM. Compound 0115 was only mildly toxic at 48 hours, dropping relative growth rates to 70% at the maximum concentration (Figure 3-10). Compound

0022 also displayed some growth inhibition at 100 µM, reducing the relative growth to near 50% at both time points. Relative toxicity was also determined in Hela cells expressing EGFP-Parkin using the MTT assay. The yellow tetrazolium MTT (3-(4,5-dimethylthiazoyl-2)-2, 5- diphenyltetrazolium bromide) is reduced by metabolically active cells proportional to the rate of

NADPH production. A measure of cell viability can then be determined by quantifying the resulting intracellular purple formazan by spectrophotometric techniques. The percent viability was evaluated for compounds 8247, 0115, 0136 and 875 by normalizing to a DMSO control

(Figure 3-11). All the tested compounds exhibit strong toxicity at concentrations at or above 50

µM. Compound 875 displayed the highest toxicity, a result consistent with the yeast toxicity assay.

Mitochondrial protein import is often coupled to membrane potential and the respiratory chain.

Specifically, contacts between the Tim23 complex and complex III of the electron transport

101 chain act synergistically to promote import of certain precursors16–19. Therefore, non-specific effects on the respiratory chain could alter protein processing and import in isolated mitochondria. We decided to utilize the Agilent Seahorse XF Analyzer to determine any effects the MPP inhibitors may have on mitochondria respiration. The Agilent Seahorse XF Analyzer allows for sensitive measurements of mitochondria respiration across a wide array of conditions simultaneously. This includes the respiratory flux through Complex I and Complex II (State 2),

ADP induced substrate respiration (state 3ADP), oligomycin induced state 4 respiration (state 4o), and maximal respiratory capacity induced by FCCP (state 3u)20–22. For this assay I chose compounds 8247, 0115 and 875 because they induced the most severe processing defects in Su9-

DHFR import and did not affect general mitochondrial function in secondary assays.

Mitochondria extracts were purified from isolated mouse liver and the respiratory flux through

Complex I and Complex II was measured (Figure 3-12). Treatment with complex I substrates pyruvate and malate enabled complex I respiration. Meanwhile, treatment with succinate and rotenone, a complex I inhibitor, enabled complex II respiration20,21. Furthermore, the effects of each compound on the respiratory flux through Complex I and Complex II was measured under two conditions. The first condition deemed “inject” measures the immediate effects on respiration by injecting the small molecule after sequential addition of ADP (state 3ADP) and oligomycin (state 4o). As a control, FCCP treatment was done in parallel to measure maximum respiratory capacity (state 3u). Antimycin A was added last to inhibit all mitochondria respiration. The second condition labeled as “Pre-treat” measures the effects on respiration if the mitochondria were pre-treated with the small molecule for an extended period. Mitochondria were pre-treated with 50 µM of each compound or 1% DMSO for 30 minutes before sequential addition of ADP (state 3ADP), oligomycin (state 4o), FCCP (state 3u) and antimycin A.

102

Conditions measuring both complex I and complex II respiration demonstrated the expected effects of ADP, oligomycin and antimycin on oxygen consumption in the DMSO injection control. Notably, DMSO controls for pre-treatment displayed a slightly lower respiratory capacity in state 3ADP relative to injection conditions. This could be attributed to a high concentration of 1% DMSO that may alter mitochondria respiratory capacity when treated for an extended period. Injection of compound 0115 displayed a lower basal respiration rate for complex I respiration while complex II respiration was relatively unchanged (Figure 3-12A).

However, mitochondria that were pre-treated with compound 0115 were completely insensitive to ADP induced respiration (state 3ADP) as well as FCCP induced (state 3u) maximal respiration in both complex I and complex II conditions. Compound 875 induced maximal respiration under injection conditions similar to FCCP but retained sensitivity to antimycin A (Figure 3-12B). Pre- treatment with compound 875 resembles pre-treatment with compound 0115. The mitochondria are not responsive to ADP or FCCP induced respiration but still retain a sensitivity to antimycin

A. This result suggests compound 875 uncouples mitochondria by behaving similar to an ionophore like FCCP. Compound 8247 shows no significant change to respiration under injection conditions but shows a reduction in state 3ADP respiration under pre-treatment conditions (Figure 3.12C). Moreover, pre-treatment with 8247 affects maximal respiration induced by FCCP (state 3u) in complex I and complex II differently. State3u respiration appears to be increased in complex I but is decreased in complex II relative to DMSO. Cumulatively this data demonstrates that compounds 0115, 8247 and 875 decrease state 3 respiration with varying potencies. Compound 875 remains the most potent compound and affects respiration in both injection and pre-treatment conditions similar to FCCP. Compound 0115 does not affect mitochondria under injection conditions but demonstrates strong uncoupling under pre-treatment

103 conditions. Lastly, compound 8247 appears to be the least potent of the tested compounds. Pre- treatment with 8247 retains some sensitivity to ADP under pre-treatment conditions but appears to affect maximal respiration in complex I and complex II differently.

3.4 Discussion

MB-50 and MB-51 are limited in their utility because they lack potency in vivo and produce wide effects on import and respiration in mitochondria. To identify more ideal MPP inhibitors, I selected 19 compounds from the remaining 60 uncharacterized MPP hits for further study. The initial Su9-DHFR import study identified 5 compounds that inhibited import (0042, 0100, 0017,

0136, 1125 and 0160). Another 5 compounds that accumulated precursor and reduced formation of the mature form were also identified (875, 0015, 0022, 603 and 8247). Compounds 875 and

0115 showed the most robust precursor accumulation at 50 µM toward Su9-DHFR import.

However, these compounds also inhibited import of Cpn10, Cmc1, AAC and Tom40. None of which contain cleavable pre-sequences. Compounds 0022 and 603 accumulated Su9-DHFR precursors but lacked potency for inhibition of the mature form. 0022 and 603 also inhibited

Pink1 import but not Tom40 or AAC import. This suggests these compounds may possess some specificity for Tim23 destined precursors. However, due to their lack of potency relative to MB-

50, these compounds were not considered ideal MPP inhibitors. Compound 8247 showed significant processing defects but also partially inhibited Tom40, AAC and Pink1 import.

Compound 0136 was further characterized because of a strong potency to inhibit Su9-DHFR import relative to MB-50. However, 0136 also inhibited Pink1, Cmc1, Tom40 and AAC, suggesting import inhibition by a more general mechanism. Cumulative results for import efficiencies with the tested compounds are displayed in Table 3-1.

104

Secondary assays testing for non-specific effects on mitochondrial function were also performed on the more promising MPP inhibitors. An assay measuring membrane potential using the fluorescent DiSC3(5) probe identified compounds 0017, 1125 and 0160 as membrane uncouplers and were eliminated from further characterization. Meanwhile, compounds 875, 0115 and 8247 showed little effect from the DiSC3(5) assay and did not permeabilize membranes of isolated mitochondria. An IC50 analysis against recombinant MPP was consistent with the potency demonstrated in the Su9-DHFR import assays. However, respirometry studies with the

Agilent Seahorse XF Analyzer identified an uncoupling effect when mouse liver mitochondria were pre-treated with 875, 0115 or 8247. This result seems to contradict the DiSC3(5) assay which shows little effect on membrane potential by these compounds. One hypothesis is the compounds uncouple ATP synthesis from O2 production while retaining membrane potential.

Another possibility is the subtle changes in oxygen consumption detected by the Agilent

Seahorse XF Analyzer may be more sensitive to changes in membrane potential than the

DiSC3(5) assay. Cumulative results for secondary assays with the tested compounds are displayed in Table 3-2.

Cumulatively, these studies failed to identify an ideal candidate for MPP inhibition.

Compounds 875, 0115 and 8247 were identified as the most robust inhibitors of Su9-DHFR processing while displaying equal or greater potency than MB-50. However, these compounds also displayed general import inhibition of non-cleaved and non-matrix precursors and uncouple

ATP synthesis from O2 production.

105

Figure 3-1 Structures of MPP inhibitors selected for further characterization. Compounds were selected from the list of 88 confirmed MPP inhibitors based on similar principles applied to

106 the original 28 “cherry-picked” compounds from Colin’s screen11. The new set of compounds were grouped into 7 families based on their central scaffolding.

107

Figure 3-2 Time course import [35S] Su9-DHFR with selected MPP inhibitors. (A-C) Time course import assay with Su9-DHFR radiolabeled precursor was performed in yeast

108 mitochondria at 25ºC. Mitochondria were pre-incubated with 50 µM of small molecule for 15 min and the import was performed for the indicated time points. As a control, import was also performed in the presence of the vehicle control, 1% DMSO. Protease treatment was used to remove non-imported precursors. The membrane potential (ΔΨ) was dissipated with 4 µg/ml valinomycin and 50 µM CCCP. (D) As in ‘A-C’ with 10 µM of the indicated small molecule. p, precursor; m, mature; the % import was set at 100% for DMSO control at the last time point.

109

Figure 3-3 Time course import [35S] Cpn10, Cmc1 and hPink1 with selected MPP inhibitors. (A) Time course import assay with Cpn10 radiolabeled precursor was performed in yeast mitochondria at 25ºC. Mitochondria were pre-incubated with 10 µM of small molecule for

15 min and the import was performed for the indicated time points. As a control, import was also performed in the presence of the vehicle control, 1% DMSO. Protease treatment was used to remove non-imported precursors. The membrane potential (ΔΨ) was dissipated with 4 µg/ml valinomycin and 50 µM CCCP. (B) As in ‘A’ with [35S] Cmc1 and the indicated concentration of small molecule. (C) As in ‘A’ with [35S] hPink1 and 50 µM of the indicated small molecule.

The % import was set at 100% for DMSO control at the last time point. 110

Figure 3-4 Co-import of [35S] Tom40 and AAC with selected MPP inhibitors. (A) Time course import assay of both Tom40 and AAC radiolabeled precursor was performed in yeast mitochondria at 25ºC. Mitochondria were pre-incubated with either 50 µM or 10 µM of small molecule for 15 min and the import was performed for the indicated time points. As a control, import was also performed in the presence of the vehicle control, 1% DMSO. Protease treatment was used to remove non-imported precursors. The membrane potential (ΔΨ) was dissipated with

4 µg/ml valinomycin and 50 µM CCCP. The Triton-X control was permeabilized with 0.2%

Triton X-100 following import. (B) Lower exposure of ‘A’ that shows AAC import without saturation. The % import was set at 100% for DMSO control at the last time point.

111

Figure 3-5 DisC3(5) potentiometric assay with selected MPP inhibitors. (A) GA74 mitochondria were treated with 1% DMSO, 100 µM of drug or 40 µM CCCP for 15 minutes. (B)

Same as “A” at 50 µM of drugs that showed membrane uncoupling in A. Fluorescence was measured at λem=670 nm and λex=620 nm with 3 technical replicates.

112

113

Figure 3-6 Chelation assay of select MPP inhibitors using the YDHA7 JPT peptide and recombinant MPP. An MPP solution with the indicated concentrations of MnCl2 was pretreated with 100 µM of o-phenanthroline, drug or an equal volume of DMSO for 1 hour. Fluorescence was measured for 90 minutes at λem=405 nm and λex=330 nm with 3 technical replicates.

114

Figure 3-7 Membrane integrity assay with selected MPP inhibitors. (A) Mitochondria were incubated with 1% DMSO, 100 µM of drug or 0.2% TritonX-100 at 25oC for 15 minutes. After centrifugation, supernatant (S) and pellet (P) were analyzed by Coomassie stain (A) or immunoblot (B). Immunoblot was analyzed with antibodies against Cym1, Mas1, AAC and Cyt

C.

115

Figure 3-8 IC50 analysis of recombinant MPP with selected inhibitors. (A) Recombinant

MPP (128 nM) cleavage activity with the YDHA7 peptide (0.006 to 100 µM of drug) for 90 min at 25C. Average % activity ± SD of n = 3 trials. See Materials and Methods for details.

116

Figure 3-9 Yeast cell viability assay of compounds 0136 and 875.

Yeast GA74 ∆pdr5∆snq2 strain was treated with varying concentrations of small molecule from

0.01 µM to 100 µM. OD600 was measured at 24 and 48 hours. % survival was normalized to 1%

DMSO treatment. Average percentage of survival ± S.D. of n = 3 trials.

117

Figure 3-10 Yeast cell viability assay of compounds 0022 and 0115.

Yeast GA74 ∆pdr5∆snq2 strain was treated with varying concentrations of small molecule from

0.01 µM to 100 µM. OD600 was measured at 24 and 48 hours. % survival was normalized to 1%

DMSO treatment. Average percentage of survival ± S.D. of n = 3 trials.

118

Figure 3-11 MTT toxicology assay in HeLa cells with compounds 8247, 0115, 0136 and 875.

HeLa cells were treated for 24 hours with small molecule at the indicated concentrations, and then viability was measured. 100% was defined as the signal from cells treated with 1% DMSO.

119

Figure 3-12 Agilent Seahorse respirometry assay with compounds 0115, 875 and 8247. (A)

An Agilent Seahorse XF96 Analyzer was used to measure oxygen consumption rate under state

2, 3 and 4o respiration in isolated mouse liver mitochondria. Mitochondria were resuspended in buffer containing 5mM pyruvate and malate (Complex I) (left) or 5mM succinate and 2µM rotenone (Complex II) (right). For injection conditions, ADP and oligomycin were added

120 subsequently at the indicated time points. 50 µM of 0115 was then added to the mitochondria at the specified time point. As a control, mitochondria were incubated in the presence of vehicle,

1% DMSO, or 4 µM FCCP. Antimycin A was added last to cease Complex III dependent respiration. For pre-treatment, the mitochondria were treated with DMSO or 0115 for 30 minutes prior to sequential addition of ADP, oligomycin, FCCP and antimycin A at the indicated time points. (B) as in ‘A’ with compound 875. (C) as in ‘A” with compound 8247. Conditions were performed in triplicate.

121

9%

13%

29%

(10µM)

Efficiency

AAC Import Import AAC

0%

1%

34%

(10µM)

Efficiency

Tom40 Import Import Tom40

7%

27%

81%

93%

33%

87%

(10µM)

Efficiency

Cpn10 Import Import Cpn10

25%

75%

28%

72%

11%

100%

Import Import

(10µM)

Efficiency

Su9-DHFR Su9-DHFR

efficiency was normalized to 1% DMSO. 1% to was normalized efficiency

Table shows import efficiencies of various import shows of Table efficiencies

92%

66%

(50µM)

Efficiency

AAC Import Import AAC

64%

111%

(50µM)

Efficiency

Tom40 Import Import Tom40

0%

6%

1%

0%

4%

2%

(50µM)

Efficiency

Pink1 Import

0%

3%

13%

102%

(50µM)

Efficiency

Cmc1 Import Cmc1Import

89%

17%

15%

88%

28%

33%

70%

81%

86%

83%

92%

12%

93%

70%

75%

45%

32%

103%

126%

Import Import

(50µM)

Efficiency

Su9-DHFR Su9-DHFR

No

No

No

No

No

No

No

No

No

No

No

No

No

Yes

Yes

Yes

Yes

Yes

Yes

(50µM)

Precursor

Su8-DHFR Su8-DHFR

Accumulation

Vendor

Molport

Molport

Molport

Molport

Hit2lead

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

298.3

282.34

411.51

349.44

312.33

459.55

451.78

341.31

305.38

321.38

372.47

348.38

430.72

434.54

375.37

465.34

(g/mol)

Weight

357.409

385.463

398.418

Molecular

1 Cumulative results of import efficiencies with selected MPP inhibitors. MPP efficiencies selected with import of results Cumulative 1

-

5925539

Catalog#

F3328-0146

F2359-0160

F1967-1125

F1798-0479

F1094-0017

F1023-0113

F1023-0100

F0815-0042

F0499-0032

F1430-0136

F3282-0003

F1023-0022

F1430-0115

F3225-8247

002-558-125

001-019-706

001-485-603

002-161-875

tested precursors against selected MPP against import concentration. given the inhibitors % selected precursors at tested out. blacked are tested not Conditionswere that Table3 122

(µM)

10-50

10-50

10-50

10-50

Hela MTT HelaMTT

Assay MIC50 MIC50 Assay

>100

>100

>100

12.5-25

MIC50 (µM) MIC50

Yeast Toxicity Yeast

IC50 IC50

analysis

7.09±1.1

3.27±0.31

6.25±0.33

50.1±0.19

in vitro in

Results that display a range for for range and IC50 a Resultsdisplay that

No

No

No

No

Assay

Integrity

Membrane Membrane

No

No

No

No

assay

Partial

Partial

Conditions that were not tested are blacked out. blacked tested not Conditionsare were that

peptide

YDHA7 YDHA7

Chelation

(50µM)

Relative

0.84±0.04

5.76±0.06

2.95±0.06

4.19±0.10

1.61±0.04

0.86±0.04

0.96±0.06

to DMSO toDMSO

DiSC3 Assay DiSC3Assay

with selected MPP inhibitors. selected with

Fluorescence

0.03

0.05

±

±

±0.06

±0.02

0.9

1.23

Relative

(100µM)

tested tested range. concentration

2.98±0.02

4.76±0.08

4.87±0.16

1.28±0.06

2.53±0.13

2.43±0.09

3.18±0.05

1.06±0.02

0.89±0.02

1.00±0.05

1.02±0.07

0.66±0.03

1.00±0.03

0.99±0.07

0.85±0.04

2.44

0.53

to DMSO toDMSO

DiSC3 Assay DiSC3Assay

Fluorescence

of secondary assays secondary of

Vendor

Molport

Molport

Molport

Molport

Hit2lead

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

LifeChemicals

298.3

282.34

411.51

349.44

312.33

459.55

451.78

341.31

305.38

321.38

372.47

348.38

430.72

434.54

375.37

465.34

(g/mol)

Weight

357.409

385.463

398.418

Molecular

Cumulative results results Cumulative

2

-

5925539

Catalog#

F3328-0146

F2359-0160

F1967-1125

F1798-0479

F1094-0017

F1023-0113

F1023-0100

F0815-0042

F0499-0032

F1430-0136

F3282-0003

F1023-0022

F1430-0115

F3225-8247

002-558-125

001-019-706

001-485-603

002-161-875

MIC50 values are estimates based on the the on estimates are values MIC50 based Table3 123

Materials and Methods

Yeast Strains. Standard set of genetic and molecular techniques was used to generate the strains in this study. The overexpression strain was made by first sub-cloning the Mas1 and Mas2 with their endogenous promoters into separate 2μ plasmids pRS424 and pRS425 respectively. Primers were designed with RE sites BamHI and SalI for Mas1 cDNA while primers for Mas2 were designed with NotI and SalI at 5’ and 3’ respectively. Co-transformation of plasmids was done in a GA74 background and selected under auxotrophic markers Trp1 and Leu2.

Strain Background Genotype Source GA74-1A GA74 MATa his3-11,15 leu2 (Koehler et. al 1998) ura3 trp1 ade8 rho+ mit+ WT∆pdr5∆snq2 GA74 his3 leu2 ade8 trp1 (Miyata et. al 2017) ura3 pdr5∆0::HIS3 snq2∆0::KANMX Table 3-S1 Strains used in this study

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. Antibodies against Cym1, CytC, ATPase subunit F1-β, mtHsp70, Mas1, Mas1/Mas2, Mdh1, Mge1 Mia40 Tom70, Tim50, Tim44, Tim23, Tim17,

Pam16 and Pam18 are rabbit polyclonal and originated in the Schatz/Koehler labs.

Drug IC50. The IC50 for each compound was assessed using the MPP/JPT fluorogenic peptide assay. These assays were conducted in the same manner as described for the non-automated

124 confirmation of each compound with one alteration. For each small molecule the concentration was varied from 0.006 to 100 μM. All small molecule concentrations were done in triplicate and analyzed with GraphPad Prism 5. The fluorescence was monitored over 90 min using a

FlexStation Plate Reader (Molecular Devices) with constant shaking at an excitation of 330 nm, an emission of 405 nm, and an auto cutoff of 325 nm.

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 small molecule for 24 hours. After treatment, cells were incubated with MTT solution for additional 4 hour as described in the manufacturer protocols.

Yeast MIC50 was determined in the GA74 ∆pdr5∆snq2 strain. A dilution series was performed in clear 96-well plates with 100 µL YPEG yeast cultures. The concentration of each small molecule varied from 0.01 µM to 100 µM with a final 1% DMSO concentration. The OD600 was measured after vigorous mixing at 24 and 48 hours. % survival was normalized to 1% DMSO treatment. Each condition was performed in triplicate.

Mitochondria Purification. Crude mitochondria extracts were obtained as described in previous studies23,24. Yeast cultures were kept at 25°C with vigorous shaking during growth. Mitochondria concentration was measured by BCA assay. Aliquots of 200 µg were flash frozen in liquid nitrogen and stored at -80°C.

Membrane Integrity Assay. Yeast GA74 mitochondria were thawed on ice and resuspended in osmotic buffer BB7.4 buffer (0.6 M sorbitol, 20mM Hepes pH 7.4). Mitochondria were incubated with 1% DMSO, 100 µM of drug or 0.2% TritonX-100 at 25oC for 15 minutes.

Mitochondria were then centrifuged at 8000 x g for 6 minutes. The remaining mitochondrial pellet was resuspended in Laemmli Sample buffer and resolved on SDS-PAGE. Supernatants 125 were TCA precipitated, resuspended in Laemmli Sample buffer and also resolved on SDS-

PAGE. Gels were visualized by standard Coomassie staining and immunoblotting.

Oxygen electrode assay. Oxygen consumption of WT mitochondria was measured using methods previously described 25. Briefly, purified WT mitochondria (25 mg/ml) were thawed on ice and tested within 2 hours. Oxygen consumption assays were performed with a Clark-type oxygen electrode in a stirred thermostatically controlled 1.5-ml chamber at 25°C (Oxytherm,

Hansatech). State II respiration was induced in a suspension of 100 µg/ml mitochondria in 0.25

M sucrose, 20 mM KCl, 20 mM Tris-Cl, 0.25 mM EDTA, 4 mM KH2PO4, and 3 mM MgCl2, pH 7.2 after adding 2 mM NADH. Consumption rate was monitored for approximately 2 min. 50

µM of MitoBloCK-50 and 100 µM of MitoBloCK-51 were added to a final concentration of 1%

DMSO and respiration was measured for at least 1 minute. Uncoupled respiration was achieved by the addition of 20 µM carbonyl cyanide 3-chlorophenylhydrazone (CCCP) to the chamber.

TM DiSC3[5] Assay. Membrane potential measurement assays were conducted with a Flexstation

II fluorometer. The quenching of potentiometric probe (3,3’-dipropylthiadicarbocyanine iodide)

DisC3(5) was measured at λem=670 nm and λex=620 nm. Purified GA74 mitochondria was thawed and resuspended in respiration buffer (0.6M sorbitol, 2 mM KH2PO4, 50 mM KCl,

10mM MgCl, 2.5 mM EDTA, 1 mg/ml BSA and 50 mM Hepes pH 7.1) to a final concentration of 21 µg/ml. Either 100 µM of drug or DMSO was then added to a final vehicle concentration of

1% and allowed to sit for 15 minutes. Mitochondria were also treated with 40 µM CCCP as a control. After incubation components were added to a final concentration of 100 nM DisC3(5), 2 mM NADH, and 2 µg/ml oligomycin. Reactions were allowed to incubate for an additional 5 minutes at room temperature. Fluorescence was measured in a 384-well plate with 3 technical replicates. 126

Mouse liver mitochondria isolation and Seahorse Oxygen Consumption assay. A single

C57BL6 mouse was euthanized with isoflurane. The extracted liver was washed with PBS, minced with scissors and homogenized in a Potter-Elvehjem homogenizer with 10ml of ice cold

MSHE buffer (200mM mannitol, 70mM sucrose, 10mM HEPES-KOH pH7.4, 1mM EDTA.

Homogenates were centrifuged at 1000g for 10 minutes at 4°C. Supernatant was collected and centrifuged at 10,000g for 10 minutes at 4°C. The mitochondrial pellets were washed twice in

MSHE buffer supplemented with 0.5% fatty-acid free BSA. Final mitochondrial pellet was re- suspended in ice cold MSHE buffer. Mitochondrial protein content was measured with a BCA assay (Pierce). An Agilent Seahorse XF96 Analyzer was used to measure oxygen consumption rate under state 2, 3 and 4o respiration. The XF96 cartridge was loaded with a total of 5µg and

3µg of protein per well for Complex I and Complex II driven respiration respectively.

Mitochondria were resuspended in 150 µL MAS buffer (70mM sucrose, 220mM mannitol, 5mM

KH2PO4, 5mM MgCl2, 2mM HEPES-KOH pH7.4, 1mM EGTA and 0.1% fatty-acid free BSA) with either 5mM pyruvate and malate (Complex I) or 5mM succinate and 2µM rotenone

(Complex II). Plates were incubated at 37°C in non-CO2 incubate for 8 minutes before loading into the XF96 instrument after 30 minutes calibration. Compounds were injected at the specified time at the given concentrations: 4mM ADP, 3.5µM oligomycin, 4µM FCCP, 4µM antimycin A,

1% DMSO, 50µM MB-50 and MB-51. For pre-treatment, the mitochondria were treated with

DMSO, MB-50 or MB-51 for 30 minutes prior to sequential addition of ADP, oligomycin, FCCP and antimycin A. Conditions were performed in triplicate.

MPP Chelation assay with JPT-Peptide. The assay was performed under similar conditions to the high-throughput YDHA7 JPT-peptide screen.128 nM of recombinant MPP was incubated in screening buffer (10 mM Hepes, pH 7.4, 50mM NaCl, 0.01% BSA) with the indicated

127 concentrations of MnCl2. MPP solutions were then treated with 100 µM of o-phenanthroline, drug or an equal volume of DMSO for 1 hour at room temperature. After incubation, 12 µM of

YDHA7 JPT peptide was added to the reaction and mixed thoroughly. Fluorescence with 330 nm excitation and 405 nm emission was monitored for 90 minutes using a FlexStation Plate Reader.

Experiment was performed with three technical replicates.

[35S]- Radiolabeled import assays. Prior to import into purified mitochondria, 35S-methionine and cysteine labeled proteins were generated with TNT Quick Coupled Transcription/Translation kits (Promega) and plasmids carrying the gene of interest. Transcription of genes was driven by either the T7 or Sp6 promoter. Import reactions were performed as previously described26.

Frozen mitochondria aliquots were thawed and added to import at a concentration of 25 µg/ml.

Drug or DMSO was added as indicated to a final concentration of 1% vehicle for all experiments. Samples with dissipated membrane potential (-Δψ) were treated with 4 µg/ml valinomycin and 50 µM CCCP for at least 4 minutes at 25°C. Following a 15 minute incubation at 25°C, import reactions were initiated by the addition of 5-10µL translation mix. Aliquots were removed at intervals during the reaction time-course and stopped with cold buffer containing 50

µg/ml trypsin. Soybean trypsin inhibitor was added in excess after a 15 minute incubation on ice.

Mitochondria were then pelleted by centrifugation (8000 x g, 6min) and re-suspended in

Laemmli sample buffer. Import of membrane proteins (AAC and Tom40) were carbonate extracted with an additional step to remove proteins that have not been inserted into the membrane as described in Koehler et. al 199827. Samples were resolved on SDS-PAGE and analyzed by autoradiography.

Miscellaneous. Western blotting was performed using standard protocols. Proteins were transferred to nitrocellulose membranes and immune complexes were visualized with HRP 128 labeled goat secondary antibody against rabbit or mouse IgG. Chemiluminescent and autoradiographic imaging was performed on film unless otherwise noted.

129

References 1. Vögtle, F.-N. et al. Global analysis of the mitochondrial N-proteome identifies a processing peptidase critical for protein stability. Cell 139, 428–39 (2009). 2. Gakh, O., Cavadini, P. & Isaya, G. Mitochondrial processing peptidases. Biochim. Biophys. Acta - Mol. Cell Res. 1592, 63–77 (2002). 3. Chow, K. M. et al. Mammalian pitrilysin: substrate specificity and mitochondrial targeting. Biochemistry 48, 2868–77 (2009). 4. Bhushan, S., Johnson, K. A., Eneqvist, T. & Glaser, E. Proteolytic mechanism of a novel mitochondrial and chloroplastic PreP peptidasome. Biol. Chem. 387, 1087–90 (2006). 5. Falkevall, A. et al. Degradation of the Amyloid beta-Protein by the Novel Mitochondrial Peptidasome, PreP. J. Biol. Chem. 281, 29096–29104 (2006). 6. Jensen, R. E. & Yaffe, M. P. Import of proteins into yeast mitochondria: the nuclear MAS2 gene encodes a component of the processing protease that is homologous to the MAS1-encoded subunit. EMBO J. 7, 3863–71 (1988). 7. Witte, C., Jensen, R. E., Yaffe, M. P. & Schatz, G. MAS1, a gene essential for yeast mitochondrial assembly, encodes a subunit of the mitochondrial processing protease. EMBO J. 7, 1439–47 (1988). 8. Yaffe, M. P. & Schatz, G. Two nuclear mutations that block mitochondrial protein import in yeast. Proc. Natl. Acad. Sci. U. S. A. 81, 4819–4823 (1984). 9. Taylor, A. B. et al. Crystal Structures of Mitochondrial Processing Peptidase Reveal the Mode for Specific Cleavage of Import Signal Sequences. Structure 9, 615–625 (2001). 10. Luciano, P., Tokatlidis, K., Chambre, I., Germanique, J.-C. & Géli, V. The mitochondrial processing peptidase behaves as a zinc-metallopeptidase. J. Mol. Biol. 280, 193–199 (1998). 11. Douglas, Joseph, C. Investigation and Characterization of Small Molecule Modulators for Mitochondrial Processing Peptidase. (University of California, Los Angeles, 2014). 12. Zhang, J.-H., Chung & Oldenburg. A Simple Statistical Parameter for Use in Evaluation and Validation of High Throughput Screening Assays. J. Biomol. Screen. 4, 67–73 (1999). 13. Yaffe, M. P., Ohta, S. & Schatz, G. A yeast mutant temperature-sensitive for mitochondrial assembly is deficient in a mitochondrial protease activity that cleaves imported precursor polypeptides. EMBO J. 4, 2069–74 (1985). 14. Farrelly, E., Amaral, M. C., Marshall, L. & Huang, S.-G. A High-Throughput Assay for Mitochondrial Membrane Potential in Permeabilized Yeast Cells. Anal. Biochem. 293, 269–276 (2001). 15. Becker, D., Richter, J., Tocilescu, M. A., Przedborski, S. & Voos, W. Pink1 kinase and its membrane potential (Deltaψ)-dependent cleavage product both localize to outer mitochondrial membrane by unique targeting mode. J. Biol. Chem. 287, 22969–87 (2012). 16. Mokranjac, D. & Neupert, W. The many faces of the mitochondrial TIM23 complex. Biochim. Biophys. Acta 1797, 1045–54 (2010). 130

17. Wenz, L.-S., Opaliński, Ł., Wiedemann, N. & Becker, T. Cooperation of protein machineries in mitochondrial protein sorting. Biochim. Biophys. Acta - Mol. Cell Res. 1853, 1119–1129 (2015). 18. Kulawiak, B. et al. The mitochondrial protein import machinery has multiple connections to the respiratory chain. Biochim. Biophys. Acta 1827, 612–26 (2013). 19. 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. 179, 1115–22 (2007). 20. Divakaruni, A. S., Rogers, G. W. & Murphy, A. N. Measuring Mitochondrial Function in Permeabilized Cells Using the Seahorse XF Analyzer or a Clark-Type Oxygen Electrode. in Current Protocols in Toxicology 60, 25.2.1-25.2.16 (John Wiley & Sons, Inc., 2014). 21. Brand, M. D. & Nicholls, D. G. Assessing mitochondrial dysfunction in cells. Biochem. J. 435, 297–312 (2011). 22. Rogers, G. W. et al. High Throughput Microplate Respiratory Measurements Using Minimal Quantities Of Isolated Mitochondria. PLoS One 6, e21746 (2011). 23. Glick, B. S. & Pon, L. A. Isolation of highly purified mitochondria from Saccharomyces cerevisiae. Methods Enzymol. 260, 213–23 (1995). 24. Dabir, D. V. V et al. A small molecule inhibitor of redox-regulated protein translocation into mitochondria. Dev. Cell 25, 81–92 (2013). 25. Claypool, S. M., Oktay, Y., Boontheung, P., Loo, J. A. & Koehler, C. M. Cardiolipin defines the interactome of the major ADP/ATP carrier protein of the mitochondrial inner membrane. J. Cell Biol. 182, 937–50 (2008). 26. Hasson, S. A. et al. Substrate specificity of the TIM22 mitochondrial import pathway revealed with small molecule inhibitor of protein translocation. Proc. Natl. Acad. Sci. U. S. A. 107, 9578–83 (2010). 27. Koehler, C. M. et al. Tim9p, an essential partner subunit of Tim10p for the import of mitochondrial carrier proteins. EMBO J. 17, 6477–86 (1998).

131

Chapter 4: The MPP BioID Proteome, Implications in Protein Import and Respiration

4.1 Introduction

An estimated 70% of imported mitochondrial proteins contain an amino terminal mitochondrial targeting sequence (MTS) and processing proteases in the matrix and intermembrane space coordinate the removal and degradation of the MTS1. The matrix processing peptidase (MPP) mediates the first cleavage of the presequence2, followed by the

Oct1 protease that removes an additional 8 amino acids. Once the cleavage occurs, the presequence peptidase/metallopeptidase 1 (PreP/MP1, Cym1 in yeast) degrades the targeting sequence3–5. If the precursor localizes to the intermembrane space, the inner membrane protease

(IMP) mediates a second cleavage of the sorting sequence. In addition, other specialized processing pathways have been developed that are substrate-specific.

MPP, assembling as a heterodimer of α-MPP (Mas2/PMPCA) and β-MPP

(Mas1/PMPCB), is a metalloendopeptidase in the pitrilysin family. Both subunits of MPP are

6–8 essential for viability in yeast . β-MPP possesses the inverted HXXEHX74-76E zinc binding motif for cleavage while α-MPP carries a glycine-rich loop proposed to aid in substrate binding/release9. Common cleavage site motifs have an arginine at position -2, -3 or -10 or an aromatic amino acid at position +1 from the cleavage site; however, other substrates have no discernible motif2,9,10. MPP has an evolutionary relationship with the respiratory chain because of its structural similarity to Cor1 and Cor2 proteins of Complex III11,12. N. crassa β-MPP is

13 bifunctional: Cor1 of the bc1 complex and cleavage of the MTS . In plants MPP is fully integrated as Cor1 and Cor2 proteins14.

132

Studies characterizing MPP rely heavily on genetic yeast manipulations and biochemical assays. However, given the intricate and dynamic processes of the mitochondrial import machinery, novel methods are needed to study the role of MPP. A small molecule screen based on an in vitro cleavage assay by recombinant MPP was reported by Colin Douglas, a previous member of the Koehler lab. The in vitro assay was based on signal fluorescence of a substrate peptide (YDHA7) flanked by fluorophore and quencher moiety (Figure 2-1A)15. This screen yielded two potent inhibitors of MPP (MB-50 and MB-51) which attenuated import of several precursors15. Notably, MB-50 and MB-51 also appeared to affect respiratory capacity of isolated liver mitochondria. These results suggest possible roles of MPP in respiration and protein import.

To elucidate the complex role of MPP in mitochondrial biogenesis, I employed an in vivo tagging system known as BioID to identify potential interacting partners. In this chapter I will present the results from the MPP BioID proteomic analysis and complementary experiments with

MB-50 and MB-51. Results show an interaction between MPP and core subunit Qcr2 of

Complex III, suggesting an evolutionary conserved interaction with the respiratory chain. Follow up experiments using native gel electrophoresis demonstrate MPP assembling in higher molecular weight complexes, broadening the role of MPP in mitochondria biogenesis.

4.2 Identifying MPP Interacting Proteins using BioID

BioID is a method used to screen for protein interactions in living cells. The technique involves expression of bait protein fused to a 35-kDa promiscuous biotin ligase BirA* which permits labeling of adjacent proteins. The promiscuous BirA* carries an R118G mutation that reduces affinity for the labile intermediate biotinoyl-5’-AMP (bioAMP) formed from biotin and

ATP. The bioAMP intermediate is normally kept within the active site until it reacts with the specific biotin acceptor tag (BAT)16,17. However, the R118G mutation forces early release of

133 bioAMP and biotinylation of primary amines within 10nm18. Interacting proteins can be identified by streptavidin affinity capture followed by mass spectrometry (Figure 4-1A)16,19–22.

BioID and other in vivo tagging systems have been widely implemented in mammalian and other eukaryotic species to develop interactomes of several compartmentalized proteins19,22,23. Given the strong covalent modification of the tag, these methods are particularly useful for detecting weak transient interactions that may have solubility issues in standard immunoprecipitation methods21.

The HEK Flp-InTM T-RexTM 293 system was used to generate stable cell lines expressing the

MPP-BirA* fusion proteins24. This host cell line contains a single pFRT/lacZeo target site under an SV40 early promoter for integration. Human αMPP or βMPP (PMPCA and PMPCB) C- terminally tagged with a GGSG-FLAG-BirA* was sub-cloned into a pcDNA5/FRT/TO expression vector (Figure 4-1B). As a control, a matrix-localized EGFP-BirA* fusion protein with a human COX8 leader sequence was sub-cloned into the same expression vector. In this system the constructs are expressed under a tetracycline-inducible human CMV promoter25. The vector also contains a hygromycin resistance gene embedded downstream of an FRT recombination site25. The vector was then co-transfected with the pOG44 plasmid that constitutively expresses the Flp recombinase into the host cell line. Homologous recombination between FRT sites of the expression vector and the host cell line integrates the hygromycin resistance in frame with the SV40 promotor while inactivating a lacZ-ZeocinTM fusion gene.

Selection for hygromycin resistance and ZeocinTM sensitivity allows for stable expression of the

BirA* fusion proteins.

4.3 The MPP-BirA* Fusion Construct is a Functional Biotin Ligase

134

Due to the large size of the BirA* tag (35 kDa), care must be taken to ensure proper function and localization of the bait proteins. Levels of biotinylated proteins were analyzed in whole cell extracts with SDS-PAGE followed by streptavidin-HRP. The presence of specific biotinylated proteins was dependent on doxycycline and supplemental biotin treatment (Figure 4-

2A). Notably, the MTS-EGFP-BirA* was highly active relative to MPP BioID strains even in the absence of doxycycline. This may have been due to leaky expression and a more uniform distribution across the mitochondrial matrix. Subcellular fractionation and immunoblot analysis show expression of the BirA* constructs and biotinylated proteins enriched in the mitochondrial pellet (Figure 4-2 B and C). Furthermore, immunofluorescence of the MPP fusion constructs demonstrate colocalization of Flag and mitochondrial outer membrane protein Tomm20 (Figure

4-2D). Colocalization between GFP fluorescence and mitotracker red was demonstrated in the

MTS-EGFP strain (data not shown). Cumulatively these results verify the BioID fusion constructs properly localize in the mitochondria and are biologically active.

4.4 MPP BioID Mass Spectrometry Analysis

BioID experiments were performed using three biological replicates of Hek293 Flp-InTM T-

RexTM background (WT), PMPCA-GGSG-FLAG-BirA* (αMPP), PMPCB-GGSG-FLAG-BirA*

(βMPP) and MTS-EGFP-GGSG-FLAG-BirA*. Cells were treated with doxycycline and D-

Biotin 24 hours prior to harvest. A mitochondrial extract was isolated for each strain and normalized for affinity capture on GE Streptavidin SepharoseTM High Performance agarose beads followed by mass spectrometry analysis. An interactome of each MPP BioID strain was then generated against the WT and MTS-EGFP controls. Spectral analysis of the cumulative data sets was performed using Saint Express algorithms. We accepted the minimum AvgP score of

0.65 with false discovery rate of <0.05 minimum. Each sample had three biological replicates.

135

To reduce noise, contaminating proteins such as keratin isoforms, biotin containing carboxylases and self-biotinylation of MPP subunits, were removed before identifying significant hits. The analysis yielded 385 hits for PMPCA and 328 hits for PMPCB relative to the WT control alone.

Meanwhile, normalization against the matrix localized MTS-EGFP control yielded no significant mitochondrial hits in either MPP subunit. Greater expression and BirA* activity of the MTS-

EGFP construct relative to αMPP and βMPP may have contributed to this result, masking potential interactions (Figure 4-2 A-C). Top hits for both subunits against the WT control include: a leucine-rich PPR motif-containing protein (LRPPRC), Acyl-coenzyme A thioesterase

1 and 2 (ACOT1, ACOT2), isoleucyl-tRNA synthetase (IARS2) and ATP synthase subunits alpha and beta (ATP5A1 and ATP5B). Utilizing a computational method known as MitoFates, significant hits were analyzed to determine the presence of a predicted N-terminal mitochondrial presequence26. The rate of significant hits with predicted mitochondrial presequences were 68% and 72% for αMPP and βMPP respectively. The top 50 significant hits by fold change with the

MitoFates analysis are listed in Table 4-1 (αMPP) and Table 4-2 (βMPP). This analysis lacks a proper matrix-compartmentalized BirA* and therefore does not account for non-specific interactions with the ligase. To address this issue the cumulative data from each MPP subunit was normalized against the other subunit and the WT background control. The rationale for this was to identify potential interactions exclusive to each subunit while normalizing against another matrix-localized BirA*. Results identify cytosolic LDHB (L-lactate dehydrogenase B chain) as the only single significant hit for αMPP. However, LDHB is strictly a cytosolic enzyme and yielded the minimum AvgP value for significance (0.65). It is therefore unlikely that LDHB is a genuine interactor with αMPP and was disregarded. Meanwhile, βMPP yielded three mitochondrial hits: Uqcrc2 (cytochrome b-c1 complex subunit 2), Timm13 (mitochondrial

136 import inner membrane translocase subunit Timm13), and Trnt1 (mitochondrial tRNA- nucleotidyltransferase 1) (Table 4-3).

Uqcrc2 was an interesting find because MPP shares an evolutionary relationship to the core proteins of Complex III11,12. Specifically, the Cor1 and Cor2 subunits of Complex III share structural similarities with αMPP and βMPP respectively and are argued to be remnants of the

MPP subunits2,11,12,27. Interaction of Uqcrc2 with βMPP was confirmed by a streptavidin-HRP immunoblot from the BioID pull-down (Figure 4-3A). Notably, Tim44 and Timm14 were import components that were significant in the data analysis normalized to the WT control but not the

MTS-EGFP control. The western blot confirmed those findings, displaying more enrichment in the MTS-EGFP control than either MPP subunit. To further validate these results, an in situ detection method for protein interactions known as the proximity ligation assay (PLA) was applied to the BioID strains. This technique relies on close-proximity of hybridized oligo probes that bind to the host epitope of primary antibodies. The signal is then amplified by DNA polymerase, producing a rolling circle amplicon, followed by hybridization of a fluorophore probe. The tagged BirA* strains were also transfected with a plasmid expressing mito-GFP prior to fixation. PLA signals were detected with pair antibodies of mouse and rabbit origin. The MPP

BirA* constructs were probed with an anti-Flag mouse antibody while antibodies against βMPP,

Tim44, and Uqcrc2 were from rabbit origin (Figure 4-3B). As a negative control the anti-Flag mouse antibody without a corresponding rabbit antibody was applied to the αMPP-BirA* strain.

As expected αMPP-BirA* strain displayed numerous interactions between the anti-Flag and anti-

βMPP antibodies. Conversely, αMPP-BirA* with anti-Flag alone produced no PLA signal.

Antibody pairs between anti-Flag, anti-Tim44 or anti-Uqcrc2 produced no PLA signal in the

αMPP-BirA* strain. Meanwhile, βMPP-BirA* displayed interactions with Uqcrc2 and to a lesser

137 extent Tim44. Notably, interactions between βMPP-BirA* and Timm44 were not significant from the BioID mass spectrometry analysis when normalized to the αMPP-BirA* strain. Based on this data, only Uqcrc2 was identified as a genuine interacting partner with βMPP.

To determine if the Uqcrc2 interaction with MPP is conserved in yeast, Ni-NTA affinity capture coupled with DSP (dithiobis-(succinimidyl propionate)) cross-linking was performed in strains endogenously expressing Mas1 (βMPP) and Mas2 (αMPP) C-terminally tagged with polyhistidine (Figure 4-3C). Crosslinking with DSP permits detection of weak protein-protein interactions through an amine-reactive N-hydroxysuccinimide (NHS) ester that reacts with primary amines. Amine-containing molecules are stably bound by a 12.0Ǻ spacer and can be readily released with reductant28. Tim44-6XHis was included as a control for stable interactions with the Tim23 translocon and PAM complex components. As expected, Tim44-6XHis displayed interactions with Pam18 (hTimm14), Tim17, mtHsp70 and Tim23. Interestingly, Mas1 showed a weak interaction with mtHsp70, the motor subunit of the PAM complex29–35. Mas1 or

Mas2 did not show any interaction with Cor2 or the other Tim23 subunits. These results suggest yeast MPP does not interact with Cor2 or components of the Tim23 translocon but may show weak interactions with PAM complex component mtHsp70.

4.5 MPP Assembles in High Molecular Weight Protein Complexes

Results from the MPP BioID pulldowns suggest a possible interaction between the βMPP subunit and the Cor2 protein of Complex III. Moreover, import inhibition of Tim23 destined precursors by MB-50 and MB-51 suggest a role for MPP in mediating protein import. To elucidate broader roles of MPP in mitochondrial function, I asked if MPP assembles in higher molecular weight complexes. I hypothesized that MPP may mediate protein import by associating with the Tim23 translocon. Two homodimers of Tim23 and Tim17 form the core of

138 the Tim23 channel, also known as the 90kDa TIM23CORE36. The TIM23SORT complex is responsible for lateral sorting into the inner membrane and includes Tim23, Tim17, Tim50 and

Tim2129,37. This complex can be readily separated from the 90kDa TIM23CORE with a mild detergent lysis followed by blue native gel electrophoresis (BN-PAGE)29,37. To determine if

MPP co-migrates with the Tim23 complexes GA74 mitochondrial extracts were analyzed by two-dimensional native electrophoresis (2D Native PAGE). Protein complexes were first resolved by BN-PAGE in the first dimension followed by an SDS-PAGE to resolve the individual protein subunits in the second dimension. Antibodies against Tim23 and Tim17 display protein complexes between 90-232 kDa, representative of the TIM23CORE and

TIM23SORT complexes (Figure 4-4A). Antibodies against both subunits of MPP and βMPP alone migrated predominately as a heterodimer at 60-100 kDa. However, larger complexes up to 232 kDa were also observed for both antibodies. These complexes are uniformly increased in the

MPP overexpression strain relative to WT (Figure 4-4B). Meanwhile, distribution of the

Tim23SORT and Tim23CORE complexes appear unaffected with MPP overexpression. High molecular weight complexes of MPP are uncharacterized and their migration patterns differ from the Tim23SORT and Tim23CORE complexes, making it unclear if MPP associates with the translocon. To determine whether MB-50 or MB-51 inhibits import by altering assembly of

Tim23 complexes, two-dimensional SDS-PAGE was applied to GA74 mitochondrial extract pre- treated with DMSO, MB-50 or MB-51 (Figure 4-4C). Migration patterns for Tim17 and Mas1 were not altered by drug treatment, suggesting the import inhibition is not caused by disruption of Tim23 or MPP associated complexes.

4.6 MB-50 and MB-51 Promote Assembly of Respiratory Chain

Supercomplexes

139

The mitochondria respiratory chain assembles into distinct supramolecular structures composed of isolated and combined complexes. These supramolecular structures can be preserved with a mild detergent lysis and resolved by BN-PAGE38–41. In Saccharomyces cerevisiae, Complex III assembles as a dimer (500 kDa) and forms supercomplexes with one or two complex IV monomers (III2IV1 or III2IV2). Coupling of these complexes is proposed to allow for efficient substrate channeling of quinol and cytochrome C during respiration40. In addition, protein import through the Tim23 complex is mediated by multiple contacts with the core subunits of Complex III suggesting translocation and respiration are interconnected42,43. I hypothesized MB-50 and MB-51 could inhibit protein import by altering assembly of the respiratory complexes. Yeast mitochondria pretreated with MB-50 and MB-51 were lysed by digitonin and analyzed by BN-PAGE. Respiratory chain complexes were visualized with antibodies against Cor2 (Uqcrc2) and CoxIV. Coomassie staining and immunoblot analysis of

WT GA74 mitochondria treated with 50µM MB-50 shows an accumulation of III2IV1 and III2IV2 supercomplexes (Figure 4-5 A and B). Notably, MB-50.1 treatment also accumulates Complex

III,IV supercomplexes and is the only MB-50 SAR compound that inhibits both MPP cleavage in vitro and mitochondrial import. MB-51 treatment leads to a similar accumulation at higher concentrations of 100 µM.

Cumulatively, these results suggest MB-50 and MB-51 may disrupt oxidative phosphorylation by altering assembly of respiratory chain complexes. I then asked if MPP overexpression alters Complex III,IV supercomplex assembly and whether this effect would be attenuated by MB-50 or MB-51 treatment. BN-PAGE analysis of mitochondrial extract from vector and overexpressed MPP strains exhibit similar amounts of III2IV1 and III2IV2 complexes

(Figure 4-5C). Moreover, pretreatment with MB-50 and MB-50.1 induces a comparable

140 accumulation of the supercomplexes relative to DMSO or no treatment (NT) controls. Based on these results I concluded MB-50, MB-50.1 and MB-51 may alter mitochondrial protein import by altering ratiometric amounts of Complex III,IV supercomplexes. Meanwhile, overexpression of MPP does not appear to alter respiratory chain assembly or diminish MB-50 induced accumulation of the supercomplexes.

4.7 Discussion

The high-throughput screen for MPP inhibitors identified several candidate molecules that were further characterized by me and Colin Douglas15. MB-50 and MB-51 were the most promising molecules due to their potent inhibition in vitro and minimal effects on general mitochondrial function. The most notable attribute from these molecules is their ability to inhibit protein import of several mitochondrial precursors, suggesting broader roles for MPP in mitochondrial biogenesis. However, respirometry studies using the Agilent Seahorse XF

Analyzer suggest the general import inhibition induced by MB-50 and MB-51 pretreatment can be attributed to a weakened respiratory capacity. Other potent MPP inhibitors that cause general import inhibition also induced significant respiratory defects (Figure 3-12). Furthermore, overexpression of MPP causes an increase in mitochondrial coupling and import efficiency.

These results provided the impetus for analyzing the MPP interactome using the in vivo tagging system known as BioID. Identification of MPP interactors may help elucidate whether

MPP mediates import directly by associating with components of the Tim23 translocon, or indirectly by regulating respiration and membrane potential through association with the electron transport chain. Results demonstrated a putative interaction between βMPP and the Cor2

(Uqcrc2) protein of electron transport Complex III. MPP has an evolutionary relationship to the respiratory chain because of its structural similarity to the Cor1 and Cor2 proteins of Complex

141

11,12 III . It is proposed that the core proteins of the bc1 complex once displayed MPP-like processing activity, coupling the respiratory chain directly to precursor processing12,44. Previous studies have also shown coupling of complex III to the Tim23 channel supports potential dependent import27,42. I speculated that MPP may support this coupling through transient interactions with Complex III and regulate the import of potential dependent precursors.

Native gel electrophoresis of yeast mitochondrial extract identified MPP in protein complexes at 232 kDa. This was unexpected because MPP is described as a soluble matrix protein in Saccharomyces Cerevisiae and higher level eukaryotes44–46. MPP did not co-migrate with the Tim23SORT or Tim23CORE complexes in yeast and Ni-NTA affinity capture of strains expressing polyhistidine tagged MPP failed to pulldown any component of the Tim23 translocon. However, pulldowns followed by DSP crosslinking identified weak interactions between Mas1 (βMPP) and PAM complex component mtHsp70. ATPase mtHsp70 is the motor component of the PAM complex which drives import through nucleotide exchange by a

Brownian ratchet mechanism37,47. mtHsp70 activity is stimulated by J-proteins Pam16 and

Pam18 which also bind to Tim4429–35. Consequently, transient interactions with mtHsp70 may promote recruitment of MPP to emerging nascent peptides and subsequent processing. However,

DSP crosslinking did not identify interactions between MPP subunits and other PAM complex components, leaving MPP as a genuine component of the PAM complex uncertain. Furthermore, association with the PAM complex does not account for inhibition by MB-50 toward non-Tim23 precursors such as AAC and Cmc1. Therefore, MPP may regulate import indirectly by playing a role in respiration and membrane potential.

PAM complex associated mtHsp70 is also identified in a complex with CoxIV of electron transport Complex IV and nucleotide exchange factor Mge148. A mtHsp70 mutation that disrupts

142 the mtHsp70-Mge1-Cox4 ternary complex decreases assembly of Complex IV and electron transport chain supercomplexes III2IV1 and III2IV2. Interestingly, BN-PAGE analysis shows an accumulation of these supercomplexes after pretreatment with MB-50 and MB-51. I speculate

MPP mediates assembly of the electron transport chain supercomplexes through a mechanism dependent on interactions between MPP and mtHsp70. Consequently, MB-50 and MB-51 may prohibit broader functions of MPP that mediate supercomplex assembly, thereby disrupting membrane coupling and potential dependent import.

These conclusions are based on cumulative data from the MPP BioID experiments, native gel analysis, yeast genetic manipulations, and the effects of MB-50 and MB-51. Collectively these studies provide reasonable support for broader roles of MPP in mitochondrial biogenesis.

However, there is enough experimental evidence that supports careful skepticism. For instance, the transient interaction between βMPP and Cor2 (Uqcrc2) was not detected by Ni-NTA affinity capture coupled with DSP crosslinking in yeast. It is possible the MPP-Uqcrc2 interaction is unique to mammalian lineages or the BioID and PLA techniques provides a more proper cellular context than pulldowns with mitochondrial extract. There is also no direct evidence that links the

MPP-mtHsp70 interaction with import inhibition or supercomplex accumulation from MB-50 or

MB-51 treatment. Furthermore, MPP overexpression alone does not alter assembly of the electron transport supercomplexes or offset any effects by MB-50 or MB-51. These results obscure an association of MPP with the respiratory chain in yeast and question whether the effects of MB-50 and MB-51 can be linked to MPP inhibition. Additional studies are required to elucidate the biological significance of MPP interactions with the respiratory chain and whether

MB-50 or MB-51 reduces import efficiency by disrupting these interactions. The answer to these

143 enquiries would help clarify MPP functions that extend beyond precursor processing and determine the utility of MB-50 and MB-51 to modulate MPP activity.

4.8 Future Perspectives

The work presented in Chapters 2-4 characterize the effects of small molecule inhibitors, identified from an in vitro screen against MPP, on cell viability and mitochondrial function15.

The most pertinent conclusion from the small molecules suggest a role for MPP in mitochondrial protein import by supporting mitochondrial respiration through interactions with Complex III.

Past studies that modulate MPP activity typically rely on chelation, depletion, or siRNA knockdown49–51. However, these methods can be prone to off-target effects that may obscure the role for MPP in mitochondrial biogenesis. For instance, treatment with chelators would theoretically inhibit several mitochondrial metallopeptidases that possess distinct functions crucial for maintaining mitochondrial homeostasis. siRNA knockdown can also be prone to off- target effects caused by the transfection or partial complementarity to unintended targets.

Furthermore, MPP activity is crucial for proper processing and maturation of several mitochondrial proteins. Therefore, prolonged attenuation would obscure direct functions of MPP by disrupting all mitochondrial functions that rely on MPP substrates. A small molecule that is specific for MPP avoids these drawbacks because the treatment is cell permeable, potent and immediate. MPP inhibitors MitoBloCK-50 and MitoBloCK-51 join a growing list of small molecules that target various mechanisms of mitochondrial biogenesis52–55. However, the structural and mechanistic basis for MB-50 and MB-51 inhibition of MPP remains unclear.

Crystallography or in silico modeling that characterize interactions between MPP and the small molecule would shed light on the mechanism of MPP inhibition. Moreover, small molecule

144 binding pockets or structural changes induced by MB-50 or MB-51 may serve as a starting point for studying MPP functions in protein import or respiration.

Additionally, the discovery of potent and specific MPP inhibitors has implications in human disease research. MPP has been linked to certain neurodegenerative diseases such as

Friedreich’s ataxia and Parkinson’s disease49,56–59. Friedreich’s ataxia is linked to defects in mitochondrial frataxin, a mitochondrial protein with a critical role in maintaining mitochondrial iron homeostasis60,61. Frataxin is imported and cleaved twice by MPP at residues 41-42, and 55-

56, yielding an intermediate and mature form60. It is proposed that defects in frataxin processing by MPP contributes to the MPP disease phenotype. Interestingly, MB-50 and MB-51 inhibits import of yeast frataxin homolog YFH1 (Figure 2-8). Consequently, application of MB-50 or

MB-51 in relevant cellular models may help elucidate the role of MPP in cerebellar ataxias.

Parkinson’s associated protein Pink1 is recruited to the mitochondrial outer membrane upstream of Parkin induced mitophagy. Pink1 is processed by MPP and attenuation of β-MPP has been shown to recruit Pink1 to the mitochondria OM and induce mitophagy49. However, these experiments were performed using siRNA knockdown against βMPP, limiting studies to in vitro cultures and potential off-target effects. Meanwhile, small molecules such as MB-50 and MB-51 could be applied to both in vitro and in vivo Parkinson’s disease models as well as allow temporal control for treatment. Such studies would help clarify the role of MPP in Pink1 processing and determine if MPP inhibition would be a suitable mode for mitochondrial clearance in Parkinson’s disease models.

145

Figure 4-1 BioID in vivo tagging system to detect transient interactions of MPP. (A) Model of BioID tagging system with MPP-BirA* fusion construct. Hek293 cells expressing the construct are supplemented with biotin for 24 hours. MPP interactors are biotinylated in a proximity dependent fashion. Biotinylated proteins are then identified by subsequent streptavidin pulldown and mass spectrometry. (B) MPP BioID constructs. Genes encoding MPP subunits

PMPCA (αMPP) and PMPCB (βMPP) were C-terminally tagged to the R118G BirA* biotin ligase with a GGSG-FLAG leader sequence. As a control, the leader sequence of human Cox8 fused to EGFP was also tagged with BirA*. Model is drawn to scale. 146

Figure 4-2 MPP BioID constructs are functional and localize to the mitochondria. (A)

Immunoblot analysis using streptavidin-HRP of whole cell extracts from Hek293 BioID cell lines. Cells were treated for 24 hours with 50 µM biotin and 0.5 µg/ml doxycycline. (B)

Subcellular fractionation of cell lines Hek293 Flp-InTM T-RexTM (WT), expressing EGFP-BirA*

(MTS-EGFP), αMPP-BirA* (αMPP) or βMPP-BirA* (βMPP). Fractions containing heavy membrane supernatant (S) and heavy membrane pellet (P) were analyzed by immunoblot with antibodies against Flag, tubulin and Tomm40. (C) As in ‘B’, fractions were analyzed by immunoblot using a streptavidin-HRP antibody. (D) TOMM20 and BirA*-tagged MPP co- localize. Hek293 cells stably expressing PMPCA-FLAG-BirA* or PMPCB-FLAG-BirA* were

147 grown on sterile coverslips. After 24 hrs, cells were fixed and incubated with mouse anti-FLAG and rabbit anti-TOMM20 antibodies. Scale bar is 10 µm.

148

149

Figure 4-3 Hek293 MPP BioID cell lines demonstrate transient interactions with Qcr2 of complex III and PAM motor components. (A) Streptavidin immunoprecipitation of Hek293

BioID cell lines. Heavy membrane fractions of cells treated with biotin and doxycycline for 24 hours were lysed and immunoprecipitated using streptavidin agarose beads. Total, bound and flow-through (FT) were resolved by SDS-PAGE and blotted using antibodies against Flag,

Uqcrc2, Timm44, and Timm14. (B) Proximity ligation analysis (PLA) shows interactions among

αMPP (left) and βMPP (right) BioID strains. Hek293 cells were transiently transfected with

Mito-GFP and PMPCA-FLAG-BirA*. After 24 hrs, cells were fixed and incubated with the indicated antibodies: mouse anti-Flag, rabbit anti-βMPP, rabbit anti-TIMM44 and rabbit anti-

UQCRC2 (Complex III). (C) DSP Crosslinking and pulldown of mitochondria from GA74 strains with 6XHis-tagged Mas1, Mas2 and Tim44. 50 µg of lysate was loaded for input and flow-through (Flow). Immunoblot analysis was performed with antibodies against 6XHis,

Tim44/His, Pam18, Tim17, Hsp70, Tim23 and Cor2 (Uqcrc2).

150

Figure 4-4 MPP is associated with higher molecular weight complexes. (A) 2-D BN gel analysis was performed with wildtype GA74 mitochondria, followed by immunoblot analysis with antibodies that recognize βMPP, α/βMPP, Hsp70, Tim23 and Tim17. The Tim23CORE and

Tim23SORT complexes are noted. (B) as in ‘A’ with WT and MPP mitochondria followed by

151 blotting with antibodies against Hsp70, Mas1, and Tim17. Note that low (LE) and high (HE) exposure panels are presented for the Tim17 analysis. (C) as in ‘B’ with 100 µM treatment of

MB-50 or MB-51 of WT GA74 mitochondria. As a control, mitochondria were treated with vehicle, 1% DMSO.

152

Figure 4-5 MB-50 and MB-51 promote supercomplex assembly of Complexes III and IV.

(A) GA74 mitochondria were pretreated with 1%DMSO, MB-50 or MB-51 for 30 minutes at the indicated concentrations. Extracts were ran on a BN-PAGE and stained with Coomassie Brilliant

Blue. Complexes III2IV2, III2IV1, III2, and III are noted. (B) Isolated GA74 mitochondria were not treated (WT) or pretreated with DMSO, or 50 µM of MB-50 and analogs 50.1, 50-A1 and

50-A2 for 15 minutes. Extracts were also treated with MB-51 at 50 µM and 100 µM for 15 minutes. Samples were ran on a BN-PAGE and analyzed by immunoblot with antibodies against

Cor2 and CoxIV. (C) GA74 mitochondria containing vector control (Vector) and mitochondria overexpressing MPP (MPP) were not treated (WT) or pretreated with DMSO, or 50 µM of 153

MB-50 and positive analog 50.1 for 15 minutes. Samples were ran on a BN-PAGE and analyzed by immunoblot with antibodies against Cor2 and CoxIV. As a control, a separate SDS-PAGE gel containing an equivalent amount of extract was analyzed by immunoblot with anti-Porin antibody.

154

BioID PMPCA Vs. WT control Cleavage site Fold Probability of Mitochondrial (processing Net Hit # Gene Name AvgP Change presequence Presequence enzyme) charge 58(MPP), 1 LRPPRC 1 1227 0.999 Yes 59(Icp55*) 0.138 25(MPP), 2 ACOT1 1 1183 0.01 No 26(Icp55*) 0 3 IARS2 1 1167 0.922 Yes 44(MPP) 0.182 4 SHMT2 1 747 0.996 Yes 26(MPP) 0.154 5 AFG3L2 1 720 0.988 Yes 42(MPP) 0.119 19(MPP), 6 ATP5B 1 713 0.996 Yes 20(Icp55*) 0.158 7 KIAA0564 1 697 0.967 Yes 28(MPP) 0.179 8 ACOT2 0.67 690 0.099 No 63(MPP) 0.127 9 LETM1 1 687 0.974 Yes 67(MPP) 0.104 10 C17orf80 1 513 0.045 No 21(MPP) 0.143 35(MPP), 11 ATP5A1 1 500 1 Yes 43(Oct1) 0.171 12 RG9MTD1 1 463 0.577 Yes 39(MPP) 0.179 13 POLRMT 1 457 0.996 Yes 41(MPP) 0.171 14 MRPL37 1 443 0.5 Yes 29(MPP) 0.138 15 DHX30 1 433 0.012 No 64(MPP) 0.125 16 CCT8 1 433 0.013 No 45(MPP) 0.044 17 DBT 1 390 0.989 Yes 22(MPP) 0.227 18 TACO1 1 390 0.954 Yes 49(MPP) 0.122 58(MPP), 19 TRAP1 1 377 0.936 Yes 59(Icp55*) 0.19 20 TIMM44 1 373 0.993 Yes 17(MPP) 0.235 21 VARS2 1 363 0.387 Yes 25(MPP) 0.16 22 MTHFD1L 1 353 0.996 Yes 31(MPP) 0.258 23 ALDH18A1 1 333 0.854 Yes 39(MPP) 0.103 24 FLNA 1 327 0 No 64(MPP) 0.016 71(MPP), 25 GLS 1 327 0.998 Yes 79(Oct1) 0.141 18(MPP), 26 PYCR2 1 327 0.019 No 19(Icp55*) 0.056 27 NFS1 1 320 0.939 Yes 28(MPP) 0.179 29(MPP), 28 ACAD9 1 317 0.991 Yes 37(Oct1) 0.172 29 C1QBP 1 317 0.999 Yes 30(MPP) 0.133 32(MPP), 30 SDHA 1 317 0.995 Yes 40(Oct1) 0.188 31 CLPX 1 313 0.41 Yes 27(MPP) 0.074 32 SUPV3L1 1 303 0.506 Yes 17(MPP) 0.176 33 ME2 1 303 0.996 Yes 18(MPP) 0.167 34 MRPS31 1 300 0.224 No 59(MPP) 0.136 35 SLC25A5 1 290 0.007 No 81(MPP) 0.062 14(MPP), 36 NDUFAF2 1 290 0.252 No 15(Icp55*) 0.071

155

16(MPP), 37 MDH2 1 283 0.999 Yes 24(Oct1) 0.188 38 MRPS9 1 280 0.137 No 49(MPP) 0.122 52(MPP), 39 GLUD1 1 277 0.965 Yes 53(Icp55*) 0.058 40 GRSF1 1 277 0.012 No 65(MPP) 0.123 41 GFM1 1 273 0.838 Yes 35(MPP) 0.2 42 PTCD1 1 273 0.291 No 31(MPP) 0.129 43 FASTKD5 1 267 0.438 Yes 27(MPP) 0.185 44 NNT 1 267 0.975 Yes 35(MPP) 0.143 42(MPP), 45 TUFM 1 263 0.996 Yes 43(Icp55*) 0.119 46 CARS2 1 257 0.662 Yes 33(MPP) 0.121 27(MPP), 47 MRPS24 1 253 0.922 Yes 35(Oct1) 0.074 48 OAT 1 247 0.996 Yes 17(MPP) 0.176 49 ATAD3A 1 243 0 No 36(MPP) 0 50 PHB2 1 243 0.806 Yes 89(MPP) 0.067 77 PMPCB 1 183 0.996 Yes 29(MPP) 0.138 Table 4-1 BioID hits of PMPCA Vs. WT Control. Table shows top 50 hits by fold change of

Hek293 PMPCA-BirA* versus WT parental strain. The βMPP subunit (PMPCB) is included as hit #77 and is highlighted in yellow. FDR (false discovery rate) for all proteins listed is 0. Hits were also analyzed using the MitoFates algorithm to determine the probability of a mitochondrial presequence and putative processing sites26. Hits were determined using the SAINTexpress analysis with a minimum AvgP score of 0.65 and false discovery rate of <0.05 minimum. n = 3 trials.

156

BioID PMPCB Vs. WT control Cleavage site Fold Probability of Mitochondrial (processing Net Hit # Gene Name AvgP Change presequence Presequence enzyme) charge 58(MPP), 1 LRPPRC 1 980 0.999 Yes 59(Icp55*) 0.138 25(MPP), 2 ACOT1 1 966.67 0.01 No 26(Icp55*) 0 3 IARS2 1 846.67 0.922 Yes 44(MPP) 0.182 19(MPP), 4 ATP5B 1 683.33 0.996 Yes 20(Icp55*) 0.158 5 CCT8 1 623.33 0.013 No 45(MPP) 0.044 6 LETM1 1 616.67 0.974 Yes 67(MPP) 0.104 7 AFG3L2 1 603.33 0.988 Yes 42(MPP) 0.119 8 KIAA0564 1 583.33 0.967 Yes 28(MPP) 0.179 9 C17orf80 1 563.33 0.045 No 21(MPP) 0.143 10 ACOT2 0.67 540 0.099 No 63(MPP) 0.127 58(MPP), 11 TRAP1 1 530 0.936 Yes 59(Icp55*) 0.19 12 SHMT2 1 503.33 0.996 Yes 26(MPP) 0.154 13(MPP), 13 UQCRC2 1 480 0.996 Yes 14(Icp55*) 0.231 14 CLPX 1 446.67 0.41 Yes 27(MPP) 0.074 15 RG9MTD1 1 440 0.577 Yes 39(MPP) 0.179 35(MPP), 16 ATP5A1 1 423.33 1 Yes 43(Oct1) 0.171 17 DBT 1 360 0.989 Yes 22(MPP) 0.227 71(MPP), 18 GLS 1 340 0.998 Yes 79(Oct1) 0.141 14(MPP), 19 NDUFAF2 1 333.33 0.252 No 15(Icp55*) 0.071 20 TIMM44 1 313.33 0.993 Yes 17(MPP) 0.235 21 DHX30 1 303.33 0.012 No 64(MPP) 0.125 22 TACO1 1 296.67 0.954 Yes 49(MPP) 0.122 23 VARS2 1 280 0.387 Yes 25(MPP) 0.16 24 PHB2 1 270 0.806 Yes 89(MPP) 0.067 42(MPP), 25 TUFM 1 260 0.996 Yes 43(Icp55*) 0.119 26 MTHFD1L 1 253.33 0.996 Yes 31(MPP) 0.258 27 ME2 1 253.33 0.996 Yes 18(MPP) 0.167 28 POLRMT 1 246.67 0.996 Yes 41(MPP) 0.171 29 SLC25A5 1 236.67 0.007 No 81(MPP) 0.062 30 PMPCA 1 236.67 0.999 Yes 33(MPP) 0.182 31 ALDH18A1 1 233.33 0.854 Yes 39(MPP) 0.103 32 NFS1 1 230 0.939 Yes 28(MPP) 0.179 52(MPP), 33 GLUD1 1 230 0.965 Yes 53(Icp55*) 0.058 34 NNT 1 230 0.975 Yes 35(MPP) 0.143 32(MPP), 35 SDHA 1 226.67 0.995 Yes 40(Oct1) 0.188 36 GRSF1 1 223.33 0.012 No 65(MPP) 0.123 16(MPP), 37 MDH2 1 223.33 0.999 Yes 24(Oct1) 0.188

157

38 SUPV3L1 1 220 0.506 Yes 17(MPP) 0.176 39 MRPS31 1 216.67 0.224 No 59(MPP) 0.136 40 GFM1 1 213.33 0.838 Yes 35(MPP) 0.2 41 FLNA 1 213.33 0 No 64(MPP) 0.016 42 C1QBP 1 210 0.999 Yes 30(MPP) 0.133 43 PHB 1 210 0.01 No 44(MPP) 0.045 44 TRMT5 1 206.67 0.065 No 16(MPP) 0.125 27(MPP), 45 MRPS24 1 203.33 0.922 Yes 35(Oct1) 0.074 46 MRPS9 1 203.33 0.137 No 49(MPP) 0.122 47 PTCD1 1 203.33 0.291 No 31(MPP) 0.129 18(MPP), 48 PYCR2 1 203.33 0.019 No 19(Icp55*) 0.056 49 ERAL1 1 196.67 0.854 Yes 43(MPP) 0.14 50 ETFA 1 193.33 0.997 Yes 19(MPP) 0.211 Table 4-2 BioID hits of PMPCB Vs. WT Control. Table shows top 50 hits by fold change of

Hek293 PMPCB-BirA* versus WT parental strain. The αMPP subunit (PMPCA) and Uqcrc2 hits are highlighted in yellow. FDR (false discovery rate) for all proteins listed is 0. Hits were also analyzed using the MitoFates algorithm to determine the probability of a mitochondrial presequence and putative processing sites26. Hits were determined using the SAINTexpress analysis with a minimum AvgP score of 0.65 and false discovery rate of <0.05 minimum. n = 3 trials.

158

PMPCA Versus PMPCB and WT control Cleavage site Gene Fold Probability of Mitochondrial (processing Net Hit # Name AvgP Change FDR presequence Presequence enzyme) charge

1 LDHB 0.65 26.67 0 0 No 16(MPP) -0.125 PMPCB Versus PMPCA and WT control Cleavage site Gene Fold Probability of Mitochondrial (processing Net Hit # Name AvgP Change FDR presequence Presequence enzyme) charge 13(MPP), 1 UQCRC2 0.94 3.51 0.01 0.996 Yes 14(Icp55*) 0.231

2 TIMM13 0.92 26.67 0.03 0 No 83(MPP) 0

3 TRNT1 0.87 23.33 0.04 0.523 Yes 68(MPP) 0.074 Table 4-3 BioID hits of each MPP-BirA* subunit against the other plus the WT Control.

Table shows top significant hits by fold change for each analysis. Hits were also analyzed using the MitoFates algorithm to determine the probability of a mitochondrial presequence and putative processing sites26. Hits were determined using the SAINTexpress analysis with a minimum AvgP score of 0.65 and false discovery rate of <0.05 minimum. n = 3 trials.

159

Materials and Methods

Plasmids and Strains. Standard set of genetic and molecular techniques was used to generate the strains in this study. The overexpression strain was made by first sub-cloning the Mas1 and

Mas2 with their endogenous promoters into separate 2μ plasmids pRS424 and pRS425 respectively. Primers were designed with RE sites BamHI and SalI for Mas1 cDNA while primers for Mas2 were designed with NotI and SalI at 5’ and 3’ respectively. Co-transformation of plasmids was done in a GA74 background and selected under auxotrophic markers Trp1 and

Leu2. The Mas1-His10 and Mas2-His10 strain was generated using PCR-based integration at the

3’ end of open reading frame and integration was selected using the HISMX6 selection cassette49. For the BioID plasmids, the alpha and beta MPP subunits were C-terminally tagged

TM with a linker GGSG, FLAG and BirA* sequence using a pcDNA5 Frt/To expression vector. A

TM pcDNA5 Frt/To plasmid containing a GGSG-FLAG-BirA* between the XhoI and ApaI restriction sites was given to us as a gift from the Wohlschlegel lab at UCLA. The MPPα and

MPPβ subunits were subcloned into the BamHI-XhoI and BamHI-NotI restriction sites respectively. A mitochondrial matrix targeted EGFP was made by fusing the leader sequence of hCox8 to the N-terminus of EGFP. The construct was subcloned into the pcDNA5TM Frt/To

BirA* plasmid at the KpnI-XhoI restriction sites. Stable expression of the BioID constructs was achieved the in Flp-InTM T-RexTM 293 Cell line. Cells were co-transfected with pOG44 and pcDNA5TM Frt/To BioID expression vectors using Bioland Scientific BioT transfection reagent following standard user guidelines and selected on 10µg/ml hygromycin and blasticidin.

160

GA74-1A GA74 MATa his3-11,15 leu2 (Koehler et. al 1998) ura3 trp1 ade8 rho+ mit+ WT pRS424/pRS425 GA74 MATa his3-11,15 leu2 This study ura3 trp1 ade8 rho+ mit+:: PRS424 [Empty:Trp1 2µ] PRS425 [Empty:Leu2 2µ] MPP GA74 MATa his3-11,15 leu2 This study ura3 trp1 ade8 rho+ mit+::[Mas1:Trp1 2µ] PRS424 [Mas2:Leu2 2µ] PRS425 Mas1HIS10 GA74 his3 leu2 ade8 trp1 This study ura3 Mas1- HIS10:HISMX6 Mas2HIS10 GA74 his3 leu2 ade8 trp1 This study ura3 Mas2- HIS10:HISMX6 TIM44HIS10 GA74 his3 leu2 ade8 trp1 Miyata e. Al 2017 ura3 TIM44- HIS10:HISMX6 Table 4-S1 Strains used in this study

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. Hek293 cells were cultured in complete DMEM with

1% penicillin/streptomycin solution (Invitrogen) and 10% FBS. Antibodies against Mas1,

Mas1/Mas2, Tom70, Tim50, Tim44, Tim23, Tim17, Pam16, Pam18, Mia40, mtHsp70, Cor2,

CoxIV, Cym1, Cyt C and human Tomm40 are rabbit polyclonal and originated in the

Schatz/Koehler labs. Rabbit polyclonal anti-Uqcrc2 (14742-1-AP), anti-Timm44 (13859-1-AP) and anti-Tim14/DNAJC19 (12096-1-AP) were acquired from Proteintech. The anti-Flag M2

161 mouse monoclonal antibody was obtained from Sigma Aldrich. The anti-tubulin and anti-6XHis antibodies are from Covance. Streptavidin conjugated to HRP was obtained from Roche.

Affinity capture of biotinylated proteins. Protocol was adapted from 16. Briefly, Hek293

Frt/To BioID cell lines were grown in DMEM complete and treated with 0.5 µg/ml Doxycycline and 50 µM D-Biotin 24 hours prior to harvest. Cells were washed with PBS and resuspended in homogenization buffer (20 mM HEPES-KOH pH7.5, 220 mM mannitol, 70 mM sucrose, 0.5 mM PMSF) with 0.2% BSA (HB+). Cells were lysed with a Teflon downse and centrifuged at

800 x g for 5 minutes at 4°C. Supernatant was collected and centrifuged for 10,000 x g for 10 minutes at 4°C. Pellet was resuspended in homogenization buffer without BSA (HB) and centrifuged again to obtain the heavy membrane fraction. Pellets were then lysed with 8 M urea,

2% SDS and 0.1 M Tris, pH 8. Samples were briefly sonicated to disrupt mtDNA in the sample before centrifugation at 20,000 x g for 15 minutes. Supernatant was collected and normalized using the Pierce BCA kit. 5 µg of lysate was used as a reference for a Total (T). Before the immunoprecipitation, GE Streptavidin SepharoseTM High Performance agarose beads were subjected to reductive alkylation by Hampton Research Protocol HR2-434 to reduce signal derived from streptavidin peptides during the mass spectrometry analysis. 1.5 mg of lysate were loaded on a 60 µL bed volume of reductive alkylated Streptavidin SepharoseTM and incubated for

2 hours at room temperature with gentle agitation. After incubation, 5 µg of lysate was used as a reference for a flow-through (FT). Lysates were washed three times each with lysis buffer, wash buffer I (lysis buffer with 0.5M NaCl), wash buffer II (lysis buffer with 0.2% SDS, 0.5M NaCl,

5% ethanol and 5% isopropanol) and digestion buffer (8 M Urea, 0.1M Tris, pH 8.5). Beads were resuspended in 50 µl digestion buffer before on-bead digestion prior to mass spectrometry analysis. For immunoblots, beads were treated with elution buffer (2% SDS, 3 mM D-Biotin and

162

8 M Urea in PBS) and incubated for 15 minutes at room temperature and 15 minutes at 95°C.

Samples were loaded on 10% SDS-PAGE using Laemmli sample buffer.

Mass spectrometry of biotinylated proteins. The protein mixtures were reduced, alkylated, and digested by the sequential addition of trypsin and Lys-C proteases50. Samples were then desalted using Pierce C18 tips, eluted in 40% acetonitrile, and dried and resuspended in 5% formic acid.

Desalted samples were fractionated using a C18 reversed phase (1.9 μM, 100A pores, Dr.

Maisch GmbH) column, packed with 25 cm of resin in a 75μM inner diameter fused silica capillary. This separation was performed online using a 140-minute water-acetonitrile gradient with 3% DMSO and ionized using electrospray ionization by application of a distal 2.2kV.

Ionized peptides were analyzed by tandem mass spectrometry (MS/MS) in a Thermo Orbitrap

Fusion Lumos mass spectrometer51. Data was acquired using a Data-Dependent Acquisition

(DDA) method that utilized a MS1 scan resolution of 120,000, MS2 resolution of 15,000, and a cycle time of 3 seconds. Data analysis was carried out using the ProLuCID and DTASelect2 implemented in the Integrated Proteomics Pipeline - IP2 (Integrated Proteomics Applications)52–

54. Protein and peptide identifications were filtered using DTASelect and required at least two unique peptides per protein with a peptide-level false positive rate of 5% as estimated by a decoy database strategy55. We used SAINTexpress version 3.6 (default parameters) as statistical tool and calculated the probability value of each putative protein–protein interaction56. We compared counts of each target (PMPCA/PMPCB) to average counts of control and the nontarget subunit.

We accepted the minimum AvgP score of 0.65 with false discovery rate of <0.05 minimum.

Samples had three biological replicates.

Cell Lysis and Subcellular fractionation. Hek293 BioID cell lines were grown to 70% confluency and treated with doxycycline and D-Biotin as in the BioID IP. For total cell lysate, 163 cells were lysed with 1X RIPA buffer (50mM Tris pH7.5, 150mM NaCl, 1% NP-40, 0.5%

Sodium deoxycholate, 0.1% SDS, 1mM EDTA, 1mM PMSF) and incubated on ice for 30 minutes with periodic vortexing. Lysate was centrifuged at 21,000 x g for 10 minutes and supernatants were normalized with BCA. Lysates were loaded on an SDS-PAGE with 25

µg/lane. For subcellular fractionation of mitochondria and cytosolic fractions, cells were harvested as in the BioID immunoprecipitation for the heavy membrane fraction. Heavy membrane supernatant was saved after 10,000 x g spin, precipitated by TCA and normalized with the heavy membrane pellets with a BCA assay. Samples were loaded on SDS-PAGE with

30 µg/lane. Immunoblot was visualized using standard techniques.

Mitochondria Purification. Crude mitochondria extracts were obtained as described in previous studies57,58. Yeast cultures were kept at 25°C with vigorous shaking during growth. Mitochondria concentration was measured by BCA assay. Aliquots of 200 µg were flash frozen in liquid nitrogen and stored at -80°C.

Crosslinking. 400 µg of isolated yeast mitochondria treated with 2mM DSP

(dithiobis(succinimidyl propionate)) in 1ml of BB7.4 buffer (0.6 M sorbitol, 20mM Hepes pH

7.4). Reactions were incubated for 1 hour at room temperature and quenched with 50 mM Tris pH 7.4. Crosslinked mitochondria were then lysed with buffer containing 0.2% TritonX-100,

20mM Hepes pH 7.4, 80 mM KCl, 10% glycerol and 1mM PMSF) followed by centrifugation at

10,000 x g for 10 minutes.

Pulldown assays. 50 μg lysate was removed from the supernatant to use as a reference for the input. The remaining 350 μg was diluted to 0.35 μg/μl of lysis buffer. Pulldowns were performed for 2 hr at 4˚C with rotation using 30 ul of HisPur Ni2+ beads (Thermo Scientific).

After binding, 50 μg of lysate was removed as a reference for the material that did not bind 164

(denoted Flow). The beads were washed 4 times with buffer. The elution and crosslinking reversal were done by the addition of Laemmli sample buffer supplemented with 15% β- mercaptoethanol and 300 mM imidazole. The bound, input, and flow-through (Flow) were separated by SDS-PAGE and analyzed by immunoblotting.

1D and 2D BN-Page. BN-PAGE experiments were adapted from39. Briefly 400µg of isolated yeast mitochondria were lysed with 1% digitonin lysis buffer (20mM HEPES pH 7.4, 10%

o glycerol, 50mM NaCl, 0.1mM EDTA, 2.5mM MgCl2 and 1mM PMSF) for 1 hour at 4 C. The lysate was then centrifuged for 30 minutes at 20,000 x g at 4°C. The supernatant was loaded on a

BN-PAGE Tris/Tricine 6-16% gradient gel for native complex resolution. For the second dimension, lanes were excised and reduced in a 1%SDS, 1%β-ME solution for 40 minutes at

60°C. Excised lanes were embedded in a 12.5% acrylamide gel and resolved by electrophoresis.

50 µg of WT GA74 mitochondria was loaded as a standard.

Cell manipulations. For microscopy experiments, HEK293 cells were fixed with 4% formaldehyde and permeabilized with ice-cold methanol. Immunostaining was performed with a rabbit anti-Tomm20 antibody (Santa Cruz) and the mouse monoclonal M2 anti-FLAG antibody

(Sigma). Primary antibodies were visualized with Alexa Fluor 594 goat anti-rabbit IgG

(Invitrogen) and Alexa Fluor 488 donkey anti-mouse 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).

165

Proximity Ligation Assay. PLA studies were performed using the Duolink® In Situ Red

Mouse/Rabbit kit from Sigma-Aldrich following standard user guidelines. Hek293 Flp-InTM T-RexTM

293 BioID cell lines were transfected with a mitochondrial localized Mito-GFP plasmid (BioT,

Bioland) and induced with 0.5µg/ml doxycycline 24 hours prior to fixation and probe treatment.

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.

166

References 1. Vögtle, F.-N. et al. Global analysis of the mitochondrial N-proteome identifies a processing peptidase critical for protein stability. Cell 139, 428–39 (2009). 2. Gakh, O., Cavadini, P. & Isaya, G. Mitochondrial processing peptidases. Biochim. Biophys. Acta - Mol. Cell Res. 1592, 63–77 (2002). 3. Chow, K. M. et al. Mammalian pitrilysin: substrate specificity and mitochondrial targeting. Biochemistry 48, 2868–77 (2009). 4. Bhushan, S., Johnson, K. A., Eneqvist, T. & Glaser, E. Proteolytic mechanism of a novel mitochondrial and chloroplastic PreP peptidasome. Biol. Chem. 387, 1087–90 (2006). 5. Falkevall, A. et al. Degradation of the Amyloid beta-Protein by the Novel Mitochondrial Peptidasome, PreP. J. Biol. Chem. 281, 29096–29104 (2006). 6. Jensen, R. E. & Yaffe, M. P. Import of proteins into yeast mitochondria: the nuclear MAS2 gene encodes a component of the processing protease that is homologous to the MAS1-encoded subunit. EMBO J. 7, 3863–71 (1988). 7. Witte, C., Jensen, R. E., Yaffe, M. P. & Schatz, G. MAS1, a gene essential for yeast mitochondrial assembly, encodes a subunit of the mitochondrial processing protease. EMBO J. 7, 1439–47 (1988). 8. Yaffe, M. P. & Schatz, G. Two nuclear mutations that block mitochondrial protein import in yeast. Proc. Natl. Acad. Sci. U. S. A. 81, 4819–4823 (1984). 9. Taylor, A. B. et al. Crystal Structures of Mitochondrial Processing Peptidase Reveal the Mode for Specific Cleavage of Import Signal Sequences. Structure 9, 615–625 (2001). 10. Luciano, P., Tokatlidis, K., Chambre, I., Germanique, J.-C. & Géli, V. The mitochondrial processing peptidase behaves as a zinc-metallopeptidase. J. Mol. Biol. 280, 193–199 (1998). 11. Teixeira, P. F. & Glaser, E. Processing peptidases in mitochondria and chloroplasts. Biochim. Biophys. Acta 1833, 360–70 (2013). 12. Braun, H.-P. P. & Schmitz, U. K. Are the ‘core’ proteins of the mitochondrial bc1 complex evolutionary relics of a processing protease? Trends Biochem. Sci. 20, 171–175 (1995). 13. Schulte, U. et al. A family of mitochondrial proteins involved in bioenergetics and biogenesis. Nature 339, 147–149 (1989). 14. Emmermann, M., Braun, H.-P., Arretzs, M. & Schmitzg, U. K. THE JOURNAL OF BIOLOO~CAL CHEMISTRY Characterization of the Bifunctional Cytochrome c Reductase- processing Peptidase Complex from Potato Mitochondria*. 268, (1993). 15. Douglas, Joseph, C. Investigation and Characterization of Small Molecule Modulators for Mitochondrial Processing Peptidase. (University of California, Los Angeles, 2014). 16. Roux, K. J., Kim, D. I., Raida, M. & Burke, B. A promiscuous biotin ligase fusion protein identifies proximal and interacting proteins in mammalian cells. J. Cell Biol. 196, 801–10 (2012). 167

17. Lane, M. D., Rominger, K. L., David, L. & Lynen, F. The Enzymatic Synthesis of Holotranscarboxylase from Apotranscarboxylase and (+)-Biotin: II. Investigation of the Reaction Mechanism. The Journal of Biological Chemistry 239, (1964). 18. Choi-Rhee, E., Schulman, H. & Cronan, J. E. Promiscuous protein biotinylation by Escherichia coli biotin protein ligase. Protein Sci. 13, 3043–3050 (2008). 19. Trinkle-Mulcahy, L. Recent advances in proximity-based labeling methods for interactome mapping. F1000Research 8, 135 (2019). 20. Roux, K. J., Kim, D. I., Burke, B. & May, D. G. BioID: A Screen for Protein-Protein Interactions. Curr. Protoc. protein Sci. 91, 19.23.1-19.23.15 (2018). 21. Varnaitė, R. & MacNeill, S. A. Meet the neighbors: Mapping local protein interactomes by proximity-dependent labeling with BioID. Proteomics 16, 2503–2518 (2016). 22. Kim, D. I. et al. Probing nuclear pore complex architecture with proximity-dependent biotinylation. Proc. Natl. Acad. Sci. 111, E2453–E2461 (2014). 23. Liyanage, S. U. et al. Characterizing the mitochondrial DNA polymerase gamma interactome by BioID identifies Ruvbl2 localizes to the mitochondria. Mitochondrion 32, 31–35 (2017). 24. Ward, R. J., Alvarez-Curto, E. & Milligan, G. Using the Flp-InTM T-RexTM System to Regulate GPCR Expression. in 21–37 (Humana Press, Totowa, NJ, 2011). doi:10.1007/978-1-61779-126-0_2 25. Yao, F. et al. Tetracycline Repressor, tetR, rather than the tetR–Mammalian Cell Transcription Factor Fusion Derivatives, Regulates Inducible Gene Expression in Mammalian Cells. Hum. Gene Ther. 9, 1939–1950 (1998). 26. Fukasawa, Y. et al. MitoFates: Improved Prediction of Mitochondrial Targeting Sequences and Their Cleavage Sites. Mol. Cell. Proteomics 14, 1113–1126 (2015). 27. Kulawiak, B. et al. The mitochondrial protein import machinery has multiple connections to the respiratory chain. Biochim. Biophys. Acta 1827, 612–26 (2013). 28. Zlatic, S. A., Ryder, P. V, Salazar, G. & Faundez, V. Isolation of labile multi-protein complexes by in vivo controlled cellular cross-linking and immuno-magnetic affinity chromatography. J. Vis. Exp. (2010). doi:10.3791/1855 29. Hutu, D. P. et al. Mitochondrial protein import motor: differential role of Tim44 in the recruitment of Pam17 and J-complex to the presequence translocase. Mol. Biol. Cell 19, 2642–9 (2008). 30. Schiller, D., Cheng, Y. C., Liu, Q., Walter, W. & Craig, E. A. Residues of Tim44 involved in both association with the translocon of the inner mitochondrial membrane and regulation of mitochondrial Hsp70 tethering. Mol. Cell. Biol. 28, 4424–33 (2008). 31. Josyula, R., Jin, Z., Fu, Z. & Sha, B. Crystal Structure of Yeast Mitochondrial Peripheral Membrane Protein Tim44p C-terminal Domain. J. Mol. Biol. 359, 798–804 (2006). 32. Liu, Q., D’Silva, P., Walter, W., Marszalek, J. & Craig, E. A. Regulated Cycling of Mitochondrial Hsp70 at the Protein Import Channel. Science (80-. ). 300, 139–141 (2003).

168

33. Rassow, J. et al. Mitochondrial protein import: biochemical and genetic evidence for interaction of matrix hsp70 and the inner membrane protein MIM44. J. Cell Biol. 127, 1547–56 (1994). 34. Schneider, H.-C. et al. Mitochondrial Hsp70/MIM44 complex facilitates protein import. Nature 371, 768–774 (1994). 35. D’Silva, P., Liu, Q., Walter, W. & Craig, E. A. Regulated interactions of mtHsp70 with Tim44 at the translocon in the mitochondrial inner membrane. Nat. Struct. Mol. Biol. 11, 1084–1091 (2004). 36. Dekker, P. J. T. et al. 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 (1997). 37. van der Laan, M., Hutu, D. P. & Rehling, P. On the mechanism of preprotein import by the mitochondrial presequence translocase. Biochim. Biophys. Acta 1803, 732–9 (2010). 38. Schägger, H. & Pfeiffer, K. Supercomplexes in the respiratory chains of yeast and mammalian mitochondria. EMBO J. 19, 1777–83 (2000). 39. Wittig, I., Braun, H.-P. & Schägger, H. Blue native PAGE. Nat. Protoc. 1, 418–428 (2006). 40. Díaz, F., Barrientos, A. & Fontanesi, F. Evaluation of the mitochondrial respiratory chain and oxidative phosphorylation system using blue native gel electrophoresis. Curr. Protoc. Hum. Genet. Chapter 19, Unit19.4 (2009). 41. Schägger, H., Cramer, W. A. & von Jagow, G. Analysis of molecular masses and oligomeric states of protein complexes by blue native electrophoresis and isolation of membrane protein complexes by two-dimensional native electrophoresis. Anal. Biochem. 217, 220–30 (1994). 42. 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. 179, 1115–22 (2007). 43. Saddar, S., Dienhart, M. K. & Stuart, R. A. The F1F0-ATP synthase complex influences the assembly state of the cytochrome bc1-cytochrome oxidase supercomplex and its association with the TIM23 machinery. J. Biol. Chem. 283, 6677–86 (2008). 44. Luciano, P. & Géli, V. The mitochondrial processing peptidase: function and specificity. Experientia 52, 1077–82 (1996). 45. Yang, M., Jensen, R. E., Yaffe, M. P., Oppliger, W. & Schatz, G. Import of proteins into yeast mitochondria: the purified matrix processing protease contains two subunits which are encoded by the nuclear MAS1 and MAS2 genes. EMBO J. 7, 3857–62 (1988). 46. Ou, W. J., Ito, A., Okazaki, H. & Omura, T. Purification and characterization of a processing protease from rat liver mitochondria. EMBO J. 8, 2605–2612 (1989). 47. Yamano, K., Kuroyanagi-Hasegawa, M., Esaki, M., Yokota, M. & Endo, T. Step-size Analyses of the Mitochondrial Hsp70 Import Motor Reveal the Brownian Ratchet in Operation. J. Biol. Chem. 283, 27325–27332 (2008). 169

48. Böttinger, L. et al. A complex of Cox4 and mitochondrial Hsp70 plays an important role in the assembly of the cytochrome c oxidase. 24, (2013). 49. Longtine, M. S. et al. Additional modules for versatile and economical PCR-based gene deletion and modification in Saccharomyces cerevisiae. Yeast 14, 953–961 (1998). 50. Heinzelman, P., Powers, D. N., Wohlschlegel, J. A. & John, V. Shotgun Proteomic Profiling of Bloodborne Nanoscale Extracellular Vesicles. in 403–416 (2019). doi:10.1007/978-1-4939-8935-5_32 51. Kelstrup, C. D., Young, C., Lavallee, R., Nielsen, M. L. & Olsen, J. V. Optimized Fast and Sensitive Acquisition Methods for Shotgun Proteomics on a Quadrupole Orbitrap Mass Spectrometer. J. Proteome Res. 11, 3487–3497 (2012). 52. Tabb, D. L., McDonald, W. H., Yates, J. R. & III. DTASelect and Contrast: tools for assembling and comparing protein identifications from shotgun proteomics. J. Proteome Res. 1, 21–6 (2002). 53. Xu, T. et al. ProLuCID, a fast and sensitive tandem mass spectra-based protein identification program. Mol. Cell. Proteomics 5, (2006). 54. Cociorva, D., Tabb, D. & Yates, J. Validation of Tandem Mass Spectrometry Database Search Results Using DTASelect. Current protocols in bioinformatics / editoral board, Andreas D. Baxevanis ... [et al.] Chapter 13, (2007). 55. Elias, J. E. & Gygi, S. P. Target-decoy search strategy for increased confidence in large- scale protein identifications by mass spectrometry. Nat. Methods 4, 207–214 (2007). 56. Teo, G. et al. SAINTexpress: improvements and additional features in Significance Analysis of INTeractome software. J. Proteomics 100, 37–43 (2014). 57. Glick, B. S. & Pon, L. A. Isolation of highly purified mitochondria from Saccharomyces cerevisiae. Methods Enzymol. 260, 213–23 (1995). 58. Dabir, D. V. V et al. A small molecule inhibitor of redox-regulated protein translocation into mitochondria. Dev. Cell 25, 81–92 (2013).

170

Chapter 5: Nuc1 and Porin Influence Killer Virus Loads Co-dependently in Saccharomyces Cerevisiae

5.1 Introduction

Yeast carry several indigenous viruses, many of which serve as models for eukaryotic host-viral interactions. The yeast double-stranded RNA (dsRNA) killer viruses of the Totiviridae family are well-characterized and one of the most prolific in laboratory yeast strains. Yeast expressing the killer phenotype secrete protein toxins into the extracellular space that kill toxin- sensitive strains of similar genera1–8. The viruses are highly abundant in the cell but are surprisingly well tolerated suggesting a mutual symbiotic relationship. Indeed, killer yeast display a competitive advantage over invading toxin-sensitive strains in natural environments where resources are limited9. Furthermore, killer viruses are not considered infectious as they are not transmitted through any extracellular route. Instead the killer genome is passed down vertically during mating, producing “virus-like” particles (VLPs)10,11. The toxic protein responsible for the killer phenotype is encoded by one of several distinct dsRNA M genomes

(M1, M2, M28 and Mlus) which also confer immunity by the host12. The M genomes are maintained and replicated with the help of the larger 4.6 kb dsRNA L-A helper genome which encodes the major capsid protein, Gag, and Gag RNA polymerase fusion protein, Gag-Pol, through a -1 ribosomal frameshift11,13–15. Both the L-A and satellite M genomes are encapsulated in separate VLPs containing 60 asymmetric dimers of the Gag protein15. Replication of the dsRNA L-A genome occurs by extrusion of a newly transcribed L-A (+)ssRNA strand from theVLP14. The released (+)ssRNA strand is either translated yielding the Gag and Gag-Pol viral

171 proteins, or packaged into another VLP12. Replication of the satellite M genomes is identical to

L-A, except plus strands are only extruded when two dsRNA M genomes per viral particle are present10. Several genes that influence killer virus expression were identified through random mutagenesis and complementation experiments. These genes can be classified into two groups: maintenance of killer plasmid (mak) genes3,16–19, and superkiller (ski) genes which repress killer virus expression2,20,21. Notable mak genes required for L-A and M genome stability include:

MAK3, MAK10 and MAK31. Together these genes form an N-acetyltransferase complex that acetylates the initial methionine of L-A Gag which is required for particle assembly4,22–24. In contrast, ski genes enable host defense against the killer virus through repression of viral RNAs which lack a 5’ cap structure and 3’ poly(A) tail2,20. Superkiller genes SKI2, SKI3, SKI4, SKI6,

SKI7 and SKI8 are involved in the exosome and ski complexes which mediate 3’-5’ degradation of viral RNAs21,25–27. Mutations in any of the superkiller genes lead to overproduction of the L-A and satellite M genomes. Work by Carol Dieckmann and Melitta Dihanich identified two other genes, NUC1 and POR1, as repressors of Killer virus production28–30. NUC1 encodes the major mitochondrial nuclease that digests both DNA and RNA31,32, and POR1 encodes the voltage- dependent anion channel porin (human VDAC) of the outer mitochondrial membrane33,34. Both genes are nonessential for mitochondrial respiration and cell viability. Currently, Nuc1 and porin are the only mitochondrial localized proteins known to regulate killer virus expression, implicating unique mitochondrial mechanisms in viral-host symbiosis. The mechanisms behind

L-A overproduction in NUC1 or POR1 mutants are not clear, though there is some evidence that link mitochondrial function and viral particle production. For instance, VLP levels are increased when strains are grown on non-fermentable media, implicating a role for respiration.

Interestingly, POR1 mutants display growth defects in glycerol but are able to adapt after several

172 days35. NUC1 mutants have no growth defects and the increase in the L-A dsRNA and Gag protein appears to be dependent on the absence of an M genome30. In this study we explore the roles of Nuc1 and Por1 in regulating VLP production. We propose a model where porin mediates import of L-A RNA transcripts into the mitochondrial IMS where they are degraded by Nuc1.

5.2 L-A Viral Production is Influenced by NUC1 and POR1 Independently of Respiratory Capacity

Work by Carol Dieckmann (A12 strain) and Melitta Dihanich (HR125 strain) identified an 86-kDA protein that was overexpressed in NUC1 and POR1 mutant strains respectively29,30,35.

Upon further study, both strains were found to contain VLPs and the 86-kDA protein was identified as the major capsid Gag protein encoded by the L-A helper genome of killer virus. We obtained Carol Dieckmann’s and Melitta Dihanich’s wildtype and mutant strains to further study the relationship between NUC1, POR1 and the L-A viral genome. DsRNA extraction followed by agarose gel electrophoresis confirmed the increase of the 4.6 Kb L-A genome in NUC1 and

POR1 mutants relative to wildtype controls (Figure 5-1A). To observe the L-A genome levels within cells we employed an antibody specific for dsRNA. The J2 monoclonal antibody recognizes dsRNA over 40bp in length and has been widely used in detecting dsRNA viruses36.

Immunofluorescence with an L-A infected strain (WT) shows strong cytosolic puncta indicative of the virus-like particles (Sup. 5-1A). This signal was absent in an L-A null strain or when the antibody was pre-incubated with a synthetic dsRNA analog (poly(I:C)). We applied the J2 antibody to ΔNUC1 and ΔPOR1 strains which showed a uniform increase of dsRNA in both strains (Sup. 5-1B). Work by Dieckmann and Dihanich discovered L-A Gag expression can be further stimulated when yeast are grown on non-fermentable media, suggesting a link between mitochondrial respiration and killer virus production29,30. Notably, POR1 mutants show an initial

173 growth defect in glycerol, implicating a weakened ability to respire, but are able to adapt after several days29. We repeated the experiments in dextrose and glycerol containing medium in each set of strains. Indeed, the L-A Gag protein was increased in both wildtype and mutant strains when grown in glycerol (Figure 5-1B). This suggests NUC1 and POR1 mutations stimulate L-A production by mechanisms independent of mitochondrial respiration. To determine if the mitochondrial genome or encoded proteins influences L-A production, yeast wildtype strains from both backgrounds were made rho null by ethidium bromide (EtBr) treatment. Clearance of the mitochondrial genome was confirmed by colony PCR of mitochondrial gene Cox2 (Figure 5-

1C) and impaired growth on YPEG (data not shown). Immunoblot of wildtype and rho null strains displayed no difference in L-A Gag expression (Figure 5-1C). This suggests L-A production is not influenced by respiratory capacity but may be stimulated by other metabolic factors from non-fermentable carbon sources. We then asked if NUC1 and POR1 act synergistically, thereby further increasing L-A production when both genes are deleted. A

ΔNUC1 and ΔNUC1ΔPOR1 double knockout was generated from HR125 wildtype and ΔPOR1 strains respectively through stable integration of a KanMX6 cassette at the Nuc1 . We then performed immunoblot analysis to determine relative L-A Gag levels of HR125 single and double knockout mutants (Figure 5-1D). L-A Gag levels in the ΔNUC1 and ΔNUC1ΔPOR1 strains were comparable and significantly increased over both wildtype and ΔPOR1 strains. This suggests deletion of both genes are not synergistic and that NUC1 deletion contributes significantly more to L-A production than POR1 deletion.

5.3 NUC1 Localizes to the Mitochondrial IMS in Wildtype and POR1 Deleted Strains

174

The NUC1 mammalian ortholog EndoG is known for its role in pathways of mitochondrial associated apoptosis. Similar to cell-death regulators cytochrome c and apoptosis inducing factor (AIF), apoptotic stress induces EndoG translocation from the mitochondria inner membrane space (IMS) to the nucleus where it contributes to chromosomal DNA degradation37–

40. Similar apoptosis pathways have been proposed in yeast including a mechanism for NUC1 translocation into the nucleus41–44. Human VDAC also participates in mechanisms of mammalian apoptosis including Bax/Bak dependent release of cytochrome c and, though controversial, as a member of the mitochondrial permeability transition pore (MPTP)45–47. We hypothesized that

Nuc1 may regulate L-A expression by degrading nascent single-stranded transcripts extruded from the viral particles. We deduced that Nuc1 most likely regulates the viral load by translocation into the cytosol where viral particle replication occurs. There is some evidence that porin immunoprecipitates with Nuc1 and may facilitate Nuc1 translocation to the nucleus upon apoptotic stress44. We asked if there are steady state levels of cytosolic Nuc1 in L-A infected yeast and if Nuc1 localization is restricted to the mitochondria in porin deleted strains. We C- terminally tagged endogenous Nuc1 with a 3HA (hemagglutinin) epitope and performed subcellular fraction to obtain cytosolic (S) and mitochondrial enriched pellets (P). Immunoblot analysis detected HA tagged Nuc1 solely in the mitochondrial enriched pellet in both wildtype and the ΔPOR1 strain (Sup. 5-2A). Next, we conducted an in vivo approach to determine Nuc1 localization using a TEV protease cleavage assay. In this system we C-terminally tagged Nuc1 with a 3XFLAG and 3HA epitope linked by a TEV cleavage sequence. Yeast were then transformed with a plasmid expressing an IMS localized CytB2 [1-220]-TEV protease in wildtype and ΔPOR1 strains. As a control, strains were also transformed with a plasmid expressing a matrix localized Su9-TEV protease (Sup. 5-2B). Immunoblot analysis with FLAG

175 antibodies reveal the full length Nuc1-3XFlag-TEV-3HA and cleaved Nuc1-3XFLAG in both

WT and ΔPOR1 strains. Moreover, steady state levels of the full length and cleaved bands were similar between both strains. This data suggests Nuc1 resides in the mitochondrial IMS and that porin deletion does not alter Nuc1 localization in vivo.

One point of criticism is the release and detection of Nuc1 and other apoptotic factors relies on an external apoptotic stress. We measured Nuc1 release in isolated mitochondria to determine if Nuc1 retro-translocation is feasible outside of an apoptotic cell. An in organello protein release assay that tracks retro-translocation of apoptotic factor Cyc1 and other IMS proteins has been demonstrated previously48. We performed the in organello protein release assay on isolated mitochondria from GA74 wildtype and ΔPOR1 strains with endogenously tagged Nuc1-3HA. Mitochondria were incubated in protein release buffer for 30 minutes prior to centrifugation. As a control, 50 mM DTT was supplemented to reduce disulfide bonds and induce retro-translocation of Ccp1. Immunoblot analysis of the mitochondrial pellet and supernatant reveals detectable levels of released Ccp1 (Sup. 5-2C). Moreover, Ccp1 release was increased with DTT treatment. Conversely, HA tagged Nuc1 was not detected in the supernatant regardless of DTT treatment.

These results challenge a previous hypothesis proposed by the Madeo group that Nuc1 could is released from the mitochondria and is translocated to the nucleus upon apoptotic stress.

To test this result, we repeated a similar experiment reported by the Madeo group showing Nuc1 colocalization with the nucleus by immunofluorescence upon apoptotic stress. In our experiment,

ΔNUC1 and ΔNUC1ΔPOR1 strains transformed with a plasmid expressing wildtype pNuc1-

FLAG were treated with H2O2 or acetic acid for 48 hours. To confirm apoptotic stress, live cells were stained with apoptosis markers Annexin V and propidium iodide. Results show a

176 significant increase in Annexin V and Annexing V plus PI positive cells under acetic acid treatment for both strains (Sup 5-3). However, immunofluorescence using FLAG antibodies and

DAPI staining displays no colocalization of Nuc1 with the nucleus in either strain upon apoptotic stress (Sup. 5-4). Cumulatively, this data shows no evidence for extra-mitochondrial localization of Nuc1, suggesting Nuc1 may regulate the killer viral levels from within the mitochondria.

4.5 POR1 Mutants Have Reduced Import Efficiency

We asked if porin deletion may affect Nuc1 biogenesis by altering mitochondrial import and processing. We performed import assays using [35S] radiolabeled Nuc1 precursor into isolated mitochondria from wildtype and ΔPOR1 strains. Results show Nuc1 precursor is imported without further processing and is potential dependent (Sup. 5-5A). Furthermore, Nuc1 import efficiency is drastically reduced in ΔPOR1 mitochondria. Import assays of the matrix localized Su9-DHFR (presequence of Fo-ATPase subunit 9 fused to DHFR) and IMS localized

AAC (ADP/ATP Carrier) precursors also show a decrease in import efficiency in the ΔPOR1 strain (Sup. 5-5B-C). To determine whether the decrease in import efficiency may be linked to a deficiency in import complexes, an immunoblot analysis was performed to analyze steady state levels of components from the TOM and Tim23 complexes (Sup. 5-5D). Antibodies against

Tom70, Tim23, Tim44 and Pam16 showed no change between wildtype and ΔPOR1 mitochondria. These results suggest porin deficiency lead to a general decrease in import of potential dependent precursors including Nuc1. However, experiments examining steady state levels of endogenously tagged Nuc1 in ΔPOR1 mitochondria show no significant decrease in detectable Nuc1 relative to wildtype (Sup. 5-2). Therefore, import deficiency in a ΔPOR1 background does not significantly alter the biogenesis of Nuc1 in vivo. This suggests porin does not regulate L-A expression by altering Nuc1 function.

177

4.6 Nuc1 and Por1 Regulate L-A expression Co-dependently

Our localization experiments determined Nuc1 was strictly confined into the mitochondrial IMS in wildtype and ΔPOR1 strains. To assess how mitochondrial targeting of

Nuc1 affects L-A viral amounts, we expressed a C-terminal FLAG tagged mutant ΔN9pNuc1 which contains a partially deleted mitochondrial localization sequence (MLS)44. We then performed subcellular fractionation of the ΔNUC1 and double knockout ΔNUC1ΔPOR1 strains containing the wildtype pNuc1 and ΔN9pNuc1 plasmids. Immunoblot analysis confirms enrichment of the ΔN9pNuc1 construct in the cytosolic HMS (heavy membrane supernatant) fraction (Figure 5-2A). These results were confirmed by immunofluorescence with FLAG antibodies. Strains expressing the ΔN9pNuc1 construct displayed cytosolic staining while the pNuc1 construct colocalizes with mitochondrial Mia40 (Figure 5-2B). We then performed immunoblot analysis to determine relative L-A Gag levels in each strain. Rescue by the wildtype pNUC1 in a ΔNUC1 background significantly reduces L-A Gag levels (Figure 5-3A). In contrast, expression of the ΔN9pNuc1 showed no significant change in L-A Gag levels relative to the vector control. In addition, we expressed a H138ApNuc1 construct containing a mutation in the putative nuclease domain44. Similar to the ΔN9pNuc1 construct, expression of the mutant

H138ApNuc1 showed no significant reduction in L-A Gag levels. This suggests Nuc1 nuclease activity and mitochondrial localization is required for L-A attenuation. To assess if the changes in L-A Gag levels correspond to the L-A genome levels, quantitative real-time PCR (qRT-PCR) with L-A specific primers was performed on whole cell RNA extracts. Results show rescue of L-

A genome levels with wildtype pNuc1 in a ΔNuc1 background (Figure 5-4A). Meanwhile, rescue of L-A genome expression was significantly reduced by the mutant H138ApNuc1 construct while the cytosolic ΔN9pNuc1 showed an increase in L-A genome levels. The Nuc1

178 constructs were also expressed in the ΔNUC1ΔPOR1 strain to determine if porin is required for

Nuc1 rescue of L-A viral loads. L-A Gag expression was significantly increased in the double knockout strain relative to the single ΔPOR1 knockout (Figure 5-3B). This result was also confirmed by RT-qPCR analysis (Figure 5-4B). Interestingly, immunoblot and RT-qPCR analysis shows expression of wildtype pNuc1 in the ΔPOR1ΔNUC1 strain increased the viral load relative to the vector control (Figure 5-3B and Figure 5-4B). Meanwhile, expression of the cytosolic localized ΔN9pNuc1 showed no significant difference in L-A Gag levels while the mutant H138ApNuc1 demonstrated only a slight increase (Figure 5-3B). These results demonstrate the requirement of porin for Nuc1 rescue of L-A viral loads. Furthermore, the increase of L-A Gag protein induced by wildtype Nuc1 in the double knockout strain suggests the absence of porin may alter the L-A regulating function of Nuc1. We also performed the reciprocal experiment where a wildtype porin construct (pPor1) was expressed in the ΔPOR1 and

ΔPOR1ΔNUC1 double knockout. Expression of the pPor1 construct reduced L-A Gag and L-A genome expression in the ΔPOR1 strain (Figure 5-3C and Figure 5-4B). However, expression of pPor1 did not rescue L-A Gag levels when expressed in the double knockout (Figure 5-3C).

Cumulatively this data demonstrates phenotypic rescue of L-A viral loads in the double knockout strain requires the presence of both Nuc1 and Porin. This suggests Nuc1 and Porin are co- dependent in regulating L-A expression.

4.7 The L-A Genome is Protected in Mitochondrial Fractions

Increased L-A viral expression from extra-mitochondrial localization of Nuc1 allowed us to postulate L-A genomes may be imported into the mitochondrial IMS. We proposed a subpopulation of (+)ssRNA L-A genomes that are extruded from the viral particle are imported into the mitochondrial IMS and degraded by Nuc1. To assess the presence of L-A genomes

179 within the mitochondria, we performed a micrococcal nuclease protection assay of mitochondrial enriched fractions from wildtype, ΔNUC1 and ΔPOR1 strains. To remove contaminating viral particles protecting L-A genomes, we isolated mitochondria at low centrifugation (8000 x g) and destabilized viral particles by resuspension with an osmotic buffer containing low salt (5mM

10,15 CaCl2) . Immunoblot analysis of the subcellular fractionation showed the L-A Gag protein is absent in the mitochondrial enriched heavy membrane pellet (HMP) (Figure 5-5A). We then incubated the mitochondria with a micrococcal nuclease (MNase) that possesses calcium dependent endo-exonuclease activity of double stranded and single stranded nucleic acids. After incubation, RNA was isolated by phenol-chloroform extraction and L-A expression was analyzed by RT-qPCR. As a control, we also measured expression of mitochondrial targeted mRNAs Atp2 and MDM10. These mRNAs are associated with the mitochondrial outer membrane for co-translation translocation and are vulnerable to MNase degradation49. A 2^(-

ΔΔCt) analysis of L-A, Atp2 and MDM10 was normalized to the mitochondrial gene Cox1 and the untreated control sample (Figure 5-5B). Expression of ATP2 and MDM10 was significantly decreased in MNase treated samples relative to untreated controls for all strains. Meanwhile, L-A viral amounts show protection in the ΔNUC1 and ΔPOR1 strains but not in the wildtype. We hypothesize the absence of L-A degradation in ΔNUC1 strains increases sequestration of ss(+) L-

A genome levels in the mitochondria. Notably, L-A levels in the ΔPOR1 strain are also protected, suggesting porin deletion stabilizes or increases import of the L-A genome. In summary, Nuc1 and porin appear to regulate L-A genomes present within the mitochondria, implicating a role of yeast mitochondria in host-defense against L-A killer viruses.

4.8 Discussion

180

Early work by Carol Dieckmann and Melita Dihanich on yeast dsRNA genomes discovered deletion of either NUC1 or POR1 genes lead to overproduction of the dsRNA Killer virus28–30. Further studies reveal both genes encode mitochondrial proteins. NUC1 encodes the major mitochondrial nuclease Nuc1 and is an ortholog of the apoptosis associated mammalian

EndoG31,32, while POR1 encodes the voltage-dependent anion channel porin (human VDAC) of the outer mitochondrial membrane33,34. The yeast killer viruses are well-characterized and one of the most prolific in laboratory strains. Two double-stranded genomes are associated with these viruses. The 4.6kb L-A helper genome encodes the major capsid Gag and Gag-Pol fusion protein, and is required for viral maintenance11,13–15. Meanwhile the smaller 1.5-2kb M genome encodes the toxic protein responsible for the killer phenotype12. Several aspects of the killer viruses have been thoroughly investigated including the viral particle structure, replication cycle and mode of toxicity2,11,15,50,51. Moreover, several genes involved in viral repression (ski) or maintenance (mak) have been characterized16,21,22,25,27,52–54. However, mechanisms that govern killer virus overproduction from NUC1 or POR1 deletion are still not understood. We first hypothesized porin may mediate release of Nuc1 into the cytosol where it contributes to the degradation of ss(+) L-A strands extruded from viral particles. This theory is based on an apoptotic pathway of the Nuc1 mammalian ortholog EndoG which translocates to the nucleus upon apoptotic stress37–40. Notably, human porin (VDAC) is also involved in apoptosis by participating in the release of cytochrome c45–47. Furthermore, studies in yeast shows Nuc1 participates in an analogous apoptosis pathway to EndoG and that porin immunoprecipitates with

Nuc144. Based on these findings, and the cytosolic replication cycle of L-A killer viruses, we speculated Nuc1 release from the mitochondria would be the most likely scenario. However, our localization experiments were unable detect any extra-mitochondrial localization of Nuc1 in L-A

181 infected strains (Sup. 5-2 and Sup. 5-4). Furthermore, a cytosolic localized ΔN9pNuc1 failed to rescue L-A viral loads in the ΔNuc1 strain. Together, these results indicate Nuc1 mediates L-A viral levels within the mitochondria. We then speculated a mechanism for L-A RNA import into the mitochondria. This theory is not implausible as mitochondria have been known to import several noncoding RNA species55–57. Examples of imported RNA molecules include tRNAs,

RNA components of RNase P, and some species of 5s rRNA58–60. Yeast in particular are known to import tRNA(Lys)(CUU), tRK1, as an adaptive stress mechanism where the mitochondrial encoded tRNA(Lys)(UUU) fails to decode the lysine AAG codon61. Notably, the VDAC (porin) in plants is identified as the main channel for tRNA translocation across the mitochondrial outer membrane62,63. We hypothesized that certain ss(+) L-A strands extruded from the viral particle are imported into the mitochondrial IMS through porin and are subsequently degraded by Nuc1.

Interactions between RNA viruses and mitochondria have been documented in other species. For instance, the Flock House virus (FHV) of the animal nodaviruses replicates its genome within vesicular invaginations between the inner and outer mitochondrial membranes64,65. Furthermore, mammalian mitochondria possess mechanisms that restrict levels of harmful nucleotide species.

Specifically, the mitochondrial RNA degradosome, composed of RNA helicase SUV3 and polynucleotide phosphorylase PNPase, prevents accumulation and release of mitochondrial self- encoded dsRNA66. However, there is evidence here that contradict the model of porin mediated

RNA translocation. For instance, MNase protection experiments show an increase in L-A protection in the ΔPOR1 strain. This result challenges the model of RNA import through porin as we would predict the L-A genome would be occluded from mitochondria and therefore more vulnerable to MNase degradation (Figure 5-5B). Arguably, cytosolic contamination combined with an increased viral load in ΔNUC1 and ΔPOR1 strains could produce the observed results.

182

To prevent this we thoroughly washed the mitochondrial fractions with a low salt buffer known to destabilize viral particles15. RNA import studies with in vitro transcribed ss(+) L-A strands would help confirm this result. The mechanism of L-A regulation by Nuc1 and porin becomes less clear when analyzing the ΔNUC1ΔPOR1 double knockout. For instance, plasmid expression of pNUC1 rescues L-A expression in the Nuc1 knockout but increases L-A levels in the

ΔNUC1ΔPOR1 strain (Figure 5-3B and Figure 5-4B). Meanwhile, expression of the

H138ApNuc1 mutant only slightly increases L-A amounts in the ΔNUC1ΔPOR1 strain, while cytosolic ΔN9pNuc1 shows no change in L-A levels (Figure 5-3B). Expression of a wildtype pPor1 also failed to rescue L-A levels in the ΔNUC1ΔPOR1 background (Figure 5-3C). One possibility is the high L-A abundance and replication rate in the ΔNUC1ΔPOR1 background nullifies rescue by Nuc1 or porin alone. However, this theory does not address the increase in L-

A viral loads from wildtype pNuc1 expression in the ΔNUC1ΔPOR1 strain. We speculate deletions in both NUC1 and POR1 genes somehow alters the symbiotic relationship between host and virus; such that Nuc1 serves to promote L-A production rather than inhibit it.

Unfortunately, more precise models are impossible without understanding the full scope of the host-viral relationship. Due to the overwhelming abundance of VLPs in the cells, accounting for

10-20% of total protein in porin knockouts, we speculate host-virus interactions depend on tightly regulated metabolic mechanisms that allow for unhindered growth of L-A infected yeast35. Consequently, metabolic changes stemming from mitochondrial defects may alter viral expression. A study analyzing yeast transcriptional response to loss of dsRNA L-A and M genomes identified 715 genes with altered expression67. These genes are related to several cellular functions that include: mitochondrial function, stress response and metabolism.

Currently, there are no links between Nuc1 function and metabolism. Porin however serves as

183 the primary channel of metabolites, nucleotides and ions across the outer mitochondrial membrane. Furthermore, the ΔPOR1 yeast are defective in mitochondrial protein import and have an initial growth defect in nonfermentable media (Sup. 5-5)68. We speculate porin deletion may disrupt several mitochondrial metabolic functions that alter rates of VLP production.

Meanwhile, Nuc1 appears to regulate L-A viral particles more directly by contributing to the degradation of L-A RNA transcripts.

184

Figure 5-1 L-A viral levels are increased in ΔNUC1 and ΔPOR1 mutants. (A) Agarose gel electrophoresis of dsRNA extracts. From left to right: HR125 (ΔPOR1), wildtype HR125-2B

(Por+), A12 ΔNUC1 and A12 wildtype. The L-A dsRNA genome and rRNAs are denoted. (B)

Wildtype and mutant strains of HR125 and A12 backgrounds were grown in YPD (dextrose) or

YPEG (glycerol) to mid-log phase. Whole cell protein extracts were then separated by SDS-

PAGE followed by immunoblot analysis. Arrows indicate bands representing L-A Gag and porin proteins. (C) (Upper) Immunoblot analysis with L-A Gag and Tim23 antibodies of whole cell extracts from (left to right) HR125 ΔPOR1, HR125-5D wildtype, HR125-5D rho null, wildtype

HR125-2B (Por+), HR125-2B (Por+) rho null and HR125-5D ΔNUC1 strains grown in YPD.

(Lower) PCR confirmation of the mitochondrial encoded Cox2 gene in ethidium bromide treated 185

HR125-5D (WT) and wildtype HR125-2B (Por+) strains. (D) HR125-5D wildtype, ΔNUC1,

ΔPOR1 and ΔNUC1ΔPOR1 strains were grown in YPGal (galactose) to mid-log phase. Whole cell extracts were analyzed by immunoblot with L-A Gag and Tim23 antibodies.

186

187

Figure 5-2 Localization of pNuc1-FLAG and ΔN9pNuc1-FLAG constructs. (A) Subcellular fractionation of HR125 strains ΔNUC1 (ΔN) and ΔNUC1ΔPOR1 (ΔNΔP) strains expressing pNuc1-FLAG (wt) or ΔN9pNuc1-FLAG (Δ9). Total, cytosolic enriched heavy membrane supernatant (HMS) and mitochondrial enriched heavy membrane pellet (HMP) were analyzed by immunoblot with the indicated antibodies. Arrows in FLAG panel indicate wildtype Nuc1-FLAG and ΔN9Nuc1-FLAG bands. (B) Immunofluorescence of vector control and strains indicated in

‘A’ with FLAG and Mia40 antibodies. Scale bar is 5 µm.

188

Figure 5-3 L-A Gag levels in Nuc1 and Por1 rescue strains. (A) HR125-5D wildtype containing vector and HR125-5D ΔNUC1 containing vector, pNuc1, ΔN9pNuc1 or

H138ApNuc1 plasmids were grown in SC-Ura galactose to mid-log phase. Immunoblot analysis

(left) was performed on whole cell extracts with L-A Gag, FLAG and porin antibodies.

189

Densitometry quantification is displayed on the right. Arrows in FLAG panel indicate wildtype

Nuc1-FLAG and ΔN9Nuc1-FLAG bands. (B) As in ‘A’ with HR125 ΔPOR1 and

ΔNUC1ΔPOR1 strains. (C) As in “B” with ΔPOR1 and ΔNUC1ΔPOR1 expressing vector and pPor1 plasmids. Quantification was normalized to Tim44 or Porin. Normalized L-A Gag ratio ±

SD of n replicates in gel. Not significant; NS, P > 0.05; *, P ≤ 0.05; **, P ≤ 0.005, P ≤ 0.0005.

190

Figure 5-4. RT-qPCR of L-A levels in Nuc1 and Por1 rescue strains. (A) qRT-PCR analysis of RNA extractions obtained from strains in Fig. 3A. A 2^-(ΔΔCt) relative expression analysis of

L-A was normalized to RDN18 internal control and the wildtype vector control strain. (B) As in

‘A’ with ΔPOR1 containing vector and pPor1 plasmids and ΔNUC1ΔPOR1 containing vector and pNuc1 plasmids. Relative L-A expression ± SD of n = 9 (3 biological and technical replicates). Not significant; NS, P > 0.05; *, P ≤ 0.05; **, P ≤ 0.005, P ≤ 0.0005.

191

Figure 5-5 L-A RNA is protected in mitochondrial fractions of ΔNUC1 and ΔPOR1 strains.

(A) Subcellular fractionation of wildtype HR125-5D, ΔNUC1 and ΔPOR1 strains grown in

YPGal. Total, cytosolic enriched heavy membrane supernatant (HMS) and mitochondrial enriched heavy membrane pellet (HMP) were analyzed by immunoblot with the indicated antibodies. (B) Mitochondrial extracts obtained from strains in “A” were treated with 150 U

MNase for 20 minutes. Reactions were stopped by the addition of Trizol reagent followed by phenol chloroform RNA extraction. RT-qPCR analysis of L-A, ATP2 and MDM10 was

192 normalized to internal control Cox1 and untreated samples. Relative expression ± SD of n = 2 experiments (3 technical replicates each). Significance was determined from a one sample t test with a hypothetical value of 1. Not significant; NS, P > 0.05; *, P ≤ 0.05; **, P ≤ 0.005, P ≤

0.0005.

193

Sup. 5-1 Immunofluorescence of L-A infected strains with anti-dsRNA (J2) monoclonal antibody. (A) HR125-5D wildtype and the A12 L-A Null strain were grown in YPGal to mid- log phase before fixation and incubation with mouse anti-dsRNA (J2) and rabbit anti-Mia40

194 antibodies. The J2 antibody was preincubated overnight in 0.2 mg/ml Poly (I:C) prior to immunostaining. (B) As in ‘A’ with HR125-5D wildtype, ΔNUC1 and ΔPOR1 strains.

Immunofluorescent signal was obtained with mouse anti-dsRNA (J2) and rabbit anti-Tim44 antibodies.

195

Sup. 5-2 Nuc1 localizes to the mitochondrial IMS in wildtype and ΔPOR1 strains. (A)

Subcellular fractionation of wildtype HR125-2B (Por+) and HR125 ΔPOR1 strains. Cytosolic

(S) supernatant and mitochondrial enriched (P) pellet were analyzed by immunoblot with

196 hemagglutinin (HA), Nac1 (Cytosol), outer mitochondrial membrane (OMM) porin and mitochondrial matrix (Matrix) Cym1 antibodies. Asterisk denotes non-specific band in HA analysis. (B) HR1 wildtype and ΔPOR1 strains endogenously tagged with Nuc1-3XFLAG-TEV-

3HA were transformed with plasmids expressing: vector control, matrix localized Su9-TEV, or

IMS localized CYB2[1-220]-TEV. Strains were grown in SC-Ura galactose and harvested in mid-log phase. Whole cell extracts were analyzed by immunoblot with FLAG, HA, Tim44 and porin antibodies. Note that high (HE) and low (LE) exposure panels are shown for FLAG analysis. Arrows indicate uncut and TEV protease cut Nuc1-3XFLAG-TEV-3HA proteins. (C)

Protein release assay of isolated mitochondria. Mitochondrial extracts were collected from GA74 wildtype and ΔPOR1 strains with endogenously tagged Nuc1-3HA. Isolated mitochondria were resuspended in protein release buffer with and without 50mM DTT. Aliquots of mitochondria at time 0 minutes (P0) and supernatants at time 30 minutes (S30min) were analyzed by immunoblot with the indicated antibodies.

197

Sup. 5-3 Apoptotic stress induces Annexin V and Propidium Iodide staining in yeast. (A)

HR125 strains ΔNUC1 pNuc1-FLAG and ΔNUC1ΔPOR1 pNuc1-FLAG were grown in SC-Ura

Gal for 48 hours in H2O2 or Acetic acid. Cells were then stained with Annexin V and Propidium iodide before visualizing under a fluorescence microscope. Scale bar is 5 µm. (B) Quantification of positive stained cells after apoptotic induction. Percent positive cells from total number of

198 cells counted ± SD of duplicate experiments with n ≥ 300 cells each. Not significant; NS, P >

0.05; *, P ≤ 0.05; **, P ≤ 0.005, P ≤ 0.0005. Figure included with permission from Kayla Frank.

199

Sup. 5-4 Nuc1 does not localize to the nucleus under apoptotic stress. HR125 strains ΔNUC1 pNuc1-FLAG and ΔNUC1ΔPOR1 pNuc1-FLAG were grown in SC-Ura Gal for 48 hours in

H2O2 or Acetic acid. Cells were then fixed and stained with FLAG antibodies and DAPI. Scale bar is 5 µm. Figure included with permission from Kayla Frank.

200

Sup. 5-5 Mitochondria deleted in Por1 are import deficient. (A) Time course assay with radiolabeled Nuc1 was performed with wildtype HR125-2B (Por+) and HR125 ΔPOR1 mitochondria at 25°C. Mitochondria were incubated with the radiolabeled precursor for the indicated time points and protease treatment was used to remove non-imported precursors. The membrane potential (ΔΨ) was dissipated with 50 µM CCCP and 4 µg/ml valinomycin. (B) As in

‘A’, Su9-DHFR was imported. (C) As in ‘B’ AAC was imported. (D) Mitochondrial proteins

(10, 25, 50 µg) were purified from WT and ΔPOR1 mitochondria and separated by SDS-PAGE followed by immunoblot analysis. Antibodies for the indicated proteins were used. p, precursor; m, mature; the % import was set at 100% for import with wildtype at the last time point.

201

Materials and Methods

Plasmids and Strains. Standard set of genetic and molecular techniques was used to generate the strains in this study. Porin and Nuc1 deletion strains were generated through integration of the KanMX unit from the pUG6 gene disruption cassette69. The KanMX gene was amplified using primers complementary to 45 nt upstream of the ATG and 45 nt downstream of the stop codon of the target gene.

Por1 pUG6 deletion primers:

Fwd 5’-AACAGCCAAGCGTACCCAAAGCAAAAATCAAACCAACC

TCTCAACACAGCTGAAGCTTCGTACGC-3’

Rev 5’-CACATATATGGTATATAGTGAACATATATATATTAGATAT

ATACGT GCATAGGCCACTAGTGGATCTG-3’

Nuc1 pUG6 deletion primers:

Fwd 5’- ATTTAGTTAAGTTCCCCACTTATTAGATCCCAAAGTTAAAGTTATC

CAGCTGAAGCTTCGTACGC-3’

Rev 5’-ATCGTTGTGTTATGAATTAATTTATATTTACAGTTTTTCAGTACAT

GCATAGGCCACTAGTGGATCTG-3’

The PCR product was purified and integrated into the parental strains (GA74, HR125-5D and

HR125-2A Por-). Mutant colonies were selected on G418 YPD plates. Por1 deletion was confirmed by colony PCR and immunoblot with Por1 antibodies. Nuc1 deletion was confirmed by colony PCR. Rho null strains were generated by growing cells in YPD medium with 10 µg/ml ethidium bromide and plating on YPD for individual colonies70. Loss of mitochondrial DNA was confirmed by observing growth defects on YPEG plates and by colony PCR with COX2 specific primers. Nuc1 was C-terminally tagged with 3HA using the pFA6a-3HA-KanMX module for

202

PCR-based gene targeting71. A PCR insert of the 3HA-KanMX module was generated with ends complementary to 40 nt of the 3’ end of the Nuc1 gene before and after the stop codon.

Primers for PCR amplification for the 3HA-KanMX tag:

Fwd 5’- GAAAGATGTGAAATTGTTACCTCCTCCAAAAAAAAGGAATCGGATCC

CCGGGTTAATTAA-3’

Rev 5’- GTGTTATGAATTAATTTATATTTACAGTTTTTCAGTACAT

GAATTCGAGCTCGTTTAAAC-3

Colonies were selected by growth on G418 plates. Integration of the tag was confirmed by PCR and immunoblot with anti-HA antibody. Strains with C-terminally tagged 3XFLAG-TEV-3HA were generated from the same pFA6a-3HA-KanMX backbone following insertion of the

3XFLAG-TEV sequence at the SalI and SmaI restriction sites. A Nuc1 targeted insert of the

3XFLAG-TEV-3HA-KanMX module was generated with the following forward primer and the reverse primer used to generate the 3HA tag.

5’-GAAAGATGTGAAATTGTTACCTCCTCCAAAAAAAAGGAAT

GAAGCTTCGTACGCTGCAGG-3’

Plasmids of the p416 backbone containing a matrix localized Su9-TEV Protease and IMS localized CYB2[1-220]-TEV Protease were provided by John Thatcher and Noriko Kondo respectively72. The centromeric pRS316 vector was used for plasmid expression of Por1 and

Nuc1 proteins. The wildtype Por1 sequence was amplified from wildtype yeast genomic DNA extract and subcloned into the BamHI and SalI sites of the pRS316 GPD PGK vector containing the constitutive GPD PGK (promoter and terminator) sequences flanking the multiple cloning site. For C-terminally tagged Nuc1 plasmids, a 3XFLAG-Stop sequence was first cloned at the

HindII and SalI restriction sites in the empty pRS316 vector or pRS316 GPD PGK. The wildtype

203

Nuc1 sequence was amplified from wildtype yeast genomic DNA extract and subcloned at the

BamHI and HindIII sites in frame with the 3XFLAG-Stop sequence. Nuc1 expression under the endogenous promoter was accomplished by subcloning Nuc1 with 600bp upstream of the start codon at the SacII and SalI restriction sites in frame with the 3XFLAG-Stop sequence. The

H138A and ΔN9 Nuc1 mutants were generated by site-directed mutagenesis of the pRS316 vector expressing wildtype Nuc1-3XFLAG. Primer pairs expressing the H138A or ΔN9 deletion were used to PCR amplify the vector expressing the mutant. Subsequent digestion by DpnI removed the original vector backbone.

Primer pairs used to generate the H138A Nuc1 mutation:

Fwd 5’-GTAAACTAAGAGACTATTTTAGGTCGGGCTATGATCGAGGCGCTC-3’

Rev 5’- GAGAAAATTTTGCGTCTGCAGCTGGGGCTTGAGCGCCTCGATCAT-3’

Primer pairs used to generate the ΔN9 Nuc1 deletion:

Fwd 5’-CCAAAGTTAAAGTTATCATGTTAGTCGGACTGGGTGCTGG-3’

Rev 5’-CCAGCACCCAGTCCGACTAACATGATAACTTTAACTTTGG-3’

Strain Background Genotype Source GA74-1A GA74 MATa his3-11,15 leu2 (Koehler et. al 1998) ura3 trp1 ade8 rho+ mit+ GA74-1A ΔPor1 GA74 MATa his3-11,15 leu2 (Koehler et. al 1998) ura3 trp1 ade8 rho+ mit+, por::URA3 GA74-1A GA74 MATa his3-11,15 leu2 This study Nuc1-3XFLAG-TEV- ura3 trp1 ade8 rho+ 3HA mit+, Nuc1-3XFLAG- TEV-3HA::KanMX6 GA74-1A ΔPor1 GA74 MATa his3-11,15 leu2 This study Nuc1-3XFLAG-TEV- ura3 trp1 ade8 rho+ 3HA mit+, por::URA3, Nuc1-3XFLAG-TEV- 3HA::KanMX6

204

HR125-5D (WT) HR125 MATα gal2, his3, his4, (Dihanich et. al leu2, trp1, ura3 1987) HR125-2B Por+ (WT) HR125 MATa gal2, his3, his4, (Dihanich et. al leu2, trp1, ura3 1988) HR125-5D (ΔNUC1) HR125 MATα gal2, his3, his4, (Dihanich et. al leu2, trp1, ura3, 1987) Nuc1::KanMX6 HR125-5D (ΔNUC1), HR125 MATα gal2, his3, his4, This study pRS316 (Vector) leu2, trp1, ura3, Nuc1::KanMX6, pRS316: Ura3 HR125-5D (ΔNUC1), HR125 MATα gal2, his3, his4, This study pNuc1-3XFLAG leu2, trp1, ura3, Nuc1::KanMX6, pRS316 Nuc1- 3XFLAG: Ura3 HR125-5D (ΔNUC1), HR125 MATα gal2, his3, his4, This study ΔN9pNuc1-3XFLAG leu2, trp1, ura3, Nuc1::KanMX6, pRS316 GPDPGK ΔN9pNuc1-3XFLAG: Ura3 HR125-5D (ΔNUC1), HR125 MATα gal2, his3, his4, This study H138ApNuc1- leu2, trp1, ura3, 3XFLAG Nuc1::KanMX6, pRS316 H138ANuc1- 3XFLAG: Ura3 HR125-2A Por- HR125 MATα gal2, his3, his4, (Dihanich et. al (ΔPOR1) leu2, trp1, ura3 1988) por::LEU2 HR125-2A Por- HR125 MATα gal2, his3, his4, This study (ΔPOR1), leu2, trp1, ura3 pRS316 (Vector) por::LEU2, pRS316: Ura3 HR125-2A Por- HR125 MATα gal2, his3, his4, This study (ΔPOR1) leu2, trp1, ura3 ΔN9pNuc1-3XFLAG por::LEU2, pRS316 GPDPGK ΔN9pNuc1-3XFLAG: Ura3 HR125-2A Por- HR125 MATα gal2, his3, his4, This study (ΔPOR1) leu2, trp1, ura3 H138ApNuc1- por::LEU2, 3XFLAG pRS316 H138ANuc1- 3XFLAG: Ura3

205

HR125-2A Por- HR125 MATα gal2, his3, his4, This study Nuc1::KanMX6 leu2, trp1, ura3 (ΔNUC1ΔPOR1) por::LEU2, Nuc1::KanMX6 HR125-2A Por- HR125 MATα gal2, his3, his4, This study Nuc1::KanMX6 leu2, trp1, ura3 (ΔNUC1ΔPOR1), por::LEU2, pRS316 (Vector) Nuc1::KanMX6, pRS316: Ura3 HR125-2A Por- HR125 MATα gal2, his3, his4, This study Nuc1::KanMX6 leu2, trp1, ura3 (ΔNUC1ΔPOR1), por::LEU2, ΔN9pNuc1-3XFLAG Nuc1::KanMX6, pRS316 GPDPGK ΔN9pNuc1-3XFLAG: Ura3 HR125-2A Por- HR125 MATα gal2, his3, his4, This study Nuc1::KanMX6 leu2, trp1, ura3 (ΔNUC1ΔPOR1), por::LEU2, H138ApNuc1- Nuc1::KanMX6, 3XFLAG pRS316 H138ANuc1- 3XFLAG: Ura3 A12 L-A (ΔNUC1) A12 [rho+] trp1, ura3-52, (Dieckmann et. al leu2-3, nuc1-1::LEU2, 1989) L-A only A12 WT L-A (WT) A12 [rho+] trp1, ura3-52, (Dieckmann et. al leu2-3, WT Nuc1, L-A 1989) Only A12 θ (L-A null) A12 [rho+] trp1, ura3-52, (Dieckmann et. al leu2-3, no L-A or M 1989) genome HR125-2B Por+ Rhoo HR125 MATa gal2, his3, his4, This study leu2, trp1, ura3 [ρ0] HR125-5D Rhoo HR125 MATα gal2, his3, his4, This study leu2, trp1, ura3 [ρ0] HR125-2B Por+ HR125 MATa gal2, his3, his4, This study Nuc1-3HA leu2, trp1, ura3, Nuc1-3HA::KanMX6 HR125-2A Por- HR125 MATα gal2, his3, his4, This study (ΔPOR1) leu2, trp1, ura3 Nuc1-3HA por::LEU2, Nuc1- 3HA::KanMX6 HR125-2B Por+ HR125 MATa gal2, his3, his4, This study Nuc1-3XFLAG-TEV- leu2, trp1, ura3, 3HA, pRS316 (Vector) Nuc1-3XFLAG-TEV-

206

3HA::KanMX6, pRS316: Ura3 HR125-2B Por+ HR125 MATa gal2, his3, his4, This study Nuc1-3XFLAG-TEV- leu2, trp1, ura3, 3HA, pSu9-TEV Nuc1-3XFLAG-TEV- Protease 3HA::KanMX6, p416 Su9-TEV Protease: Ura3 HR125-2B Por+ HR125 MATa gal2, his3, his4, This study Nuc1-3XFLAG-TEV- leu2, trp1, ura3, 3HA, pCYB2[1-220]- Nuc1-3XFLAG-TEV- TEV Protease 3HA::KanMX6, p416 CYB2[1-220]-TEV Protease: Ura3 HR125-2A Por- HR125 MATα gal2, his3, his4, This study (ΔPOR1) leu2, trp1, ura3 Nuc1-3XFLAG-TEV- por::LEU2, Nuc1- 3HA, pRS316 (Vector) 3XFLAG-TEV- 3HA::KanMX6, pRS316:Ura3 HR125-2A Por- HR125 MATα gal2, his3, his4, This study (ΔPOR1) leu2, trp1, ura3 Nuc1-3XFLAG-TEV- por::LEU2, Nuc1- 3HA, pSu9-TEV 3XFLAG-TEV- Protease 3HA::KanMX6, p416 Su9-TEV Protease: Ura3 HR125-2A Por- HR125 MATα gal2, his3, his4, This study (ΔPOR1) leu2, trp1, ura3 Nuc1-3XFLAG-TEV- por::LEU2, Nuc1- 3HA, pCYB2[1-220]- 3XFLAG-TEV- TEV Protease 3HA::KanMX6, CYB2[1-220]-TEV Protease: Ura3 Table 5-S1 Strains used in this study

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. Antibodies against Tom70, Tim50, Tim44, Tim23,

207

Tim17, Pam16, Pam18, Mia40, Por1 are rabbit polyclonal and originated in the Schatz/Koehler labs. The anti-Flag M2 mouse monoclonal antibody was obtained from Sigma Aldrich. The J2 anti-dsRNA mouse monoclonal antibody was obtained from SCICONS. Antibodies against the

L-A Gag protein were obtained from dsRNA extraction. Crude dsRNA extracts were obtained as described in previous studies73

Protein extraction. Crude protein extracts were obtained using a mechanical glass bead lysis in

SUTE buffer (1% SDS, 8 M Urea, 10 mM Tris pH 7.5, 10 mM EDTA, 0.5 mM PMSF). Cultures were grown in 4 ml medium to mid log phase. 5 OD600 worth of cells were then collected through centrifugation and resuspended in 100 µl of SUTE buffer and 100 µl of glass beads.

Cells were vortexed at full speed for 3 minutes. After vortexing, 100 µl of Laemmli Sample buffer with 15% β-mercaptoethanol was added to the lysate and allowed to incubate at 55ºC for

10 minutes. Lysate was transferred to a new 1.5 ml Eppendorf tube and centrifuged again at

10,000 x g for 5 minutes. Lysates were then resolved by SDS-PAGE and visualized by immunoblot using standard techniques.

Mitochondria Purification and Subcellular fractionation. Crude mitochondria extracts were obtained as described in previous studies74,75. Yeast were grown to mid-log phase in YPGal at

30ºC with vigorous shaking. Cells were pelleted and washed in sterile water and resuspended in

Tris-DTT Buffer (10 mM dithiothreitol (DTT), 0.1 M Tris-SO4, pH 9.4). Cells were then incubated for 15 minutes at 30ºC with shaking. Cells were then washed with Buffer A (1.2 M sorbitol, 20 mM potassium phosphate, pH 7.4) before resuspension in Buffer A with Zymolyase

20T corresponding to 2.5 mg per gram of yeast. Cells were incubated for 1 hour at 30ºC with shaking. After incubation, cells were washed twice with Buffer A and resuspended in Buffer B

(0.6 M sorbitol, 20 mM K+ MES, pH 6.0, 0.5 mM PMSF). Cells were then resuspended by a 208 glass Dounce homogenizer and an aliquot of the resuspension was taken as total lysate (Total).

The remaining lysate was centrifuged at 1700 x g for 5 minutes and the supernatant was collected in a separate tube. The remaining pellet was resuspended in Buffer B. The homogenization and centrifugation were repeated, and the supernatants were combined and centrifuged again at 8000 x g for 8 minutes. The supernatant was taken as heavy membrane supernatant (HMS) and the pellet was washed twice with Buffer C (0.6 M sorbitol, 20 mM K+

HEPES, pH 7.4). The crude mitochondrial extract or heavy membrane pellet (HMP) was washed twice in Buffer C before a final resuspension in Buffer C. The Total, HMS and HMP fractions were normalized by BCA. Normalized amounts of total and HMS fractions were TCA precipitated before loading on an SDS-PAGE. Mitochondrial extracts (HMP) were either flash frozen in liquid nitrogen in 200 µg aliquots and stored at -80°C or proceeded directly to the

MNase protection assay without freezing.

[35S]- Radiolabeled import assays. Prior to import into purified mitochondria, 35S-methionine and cysteine labeled proteins were generated with TNT Quick Coupled Transcription/Translation kits (Promega) and plasmids carrying the gene of interest. Transcription of genes was driven by either the T7 or Sp6 promoter. Import reactions were performed as previously described76.

Frozen mitochondria aliquots were thawed and added to import at a concentration of 25 µg/ml.

Samples with dissipated membrane potential (-Δψ) were treated with 4µg/ml valinomycin and

50µM CCCP for at least 4 minutes at 25°C. Following a 15-minute incubation at 25°C, import reactions were initiated by the addition of 5-10µL translation mix. Aliquots were removed at intervals during the reaction time-course and stopped with cold buffer containing 50 µg/ml trypsin. Soybean trypsin inhibitor was added in excess after a 15-minute incubation on ice.

Mitochondria were then pelleted by centrifugation (8000 x g, 6min) and re-suspended in

209

Laemmli Sample buffer. Import of membrane proteins (AAC) were carbonate extracted with an additional step to remove proteins that have not been inserted into the membrane as described in77. Samples were resolved on SDS-PAGE and analyzed by autoradiography.

In Organello Protein Release Assay. The in organello protein release assay was performed as described previously48. 1.1 mg of isolated mitochondria were resuspended in release buffer (250 mM sucrose, 5 mM MgCl2, 80 mM KCl, 10 mM MOPS/KOH, 5 mM methionine, and 10 mM

KH2PO4/K2HPO4 pH 7.2) with 25 µg/mL proteinase K and incubated at 30ºC for 15 minutes.

Mitochondria were then centrifuged for 10,000 x g for 8 minutes and resuspended in release buffer containing 2 mM PMSF and with or without 50 mM DTT. At 0 and 30-minute timepoints an aliquot was transferred to a separate tube and centrifuged immediately at 10,000 x g for 8 minutes. Pellets was resuspended in Laemmli Sample buffer. Supernatants were TCA precipitated before resuspension in Laemmli Sample buffer. 12.5 µg equivalent of each sample was resolved by SDS-PAGE.

MNase Protection assay. Mitochondrial extracts were normalized by BCA and 100 µg aliquots were resuspended in Buffer C with 5 mM CaCl2. Solubilized samples were treated with 0.2%

TritonX-100 and vortexed thoroughly. Extracts were treated with 150 units of Micrococcal

Nuclease (Thermofisher) and incubated at 37ºC for 20 minutes. Reactions were stopped by the addition Trizol reagent and the RNA was purified by phenol-chloroform extraction.

RT-qPCR. RNA extracts were obtained as described in73. Residual genomic DNA was removed with the TURBO DNA-free Kit (Thermofisher). RNA was reverse transcribed into cDNA using the SuperScript III First-Strand Synthesis System (Thermofisher) according to the manufacturer’s protocols. The PCR was performed with the 2x SYBR Green qPCR Master Mix

210

(Biomake) on the CFX Connect Real-Time PCR detection system (Biorad). Primer sequences were obtained from previous studies and ordered from Integrated DNA technologies78–80.

Annexin V and Propidium Iodide Staining. Apoptosis induction and staining was performed by using methods previously described81. Yeast cells were grown in a SC-Ura dextrose starter culture overnight. The next day, cells were transferred to a 40 ml culture of SC-Ura galactose in either 0.6 mM H2O2 or 40 mM acetic acid to 0.2 OD. Cells were harvested 48 hours later and stained with Annexin V and propidium iodide using an ApoAlert™ Annexin V-FITC Apoptosis kit from Takara Bio.

Microscopy. For microscopy experiments, yeast were fixed with 4% formaldehyde and immunostained as described in82. For Poly (I:C) treatment, a 1:1000 dilution of the dsRNA (J2) monoclonal antibody was incubated overnight in 0.2 mg/ml Poly (I:C) in PBST (Dulbecco’s phosphate-buffered supplemented with 0.1% Tween-20 and 3% BSA). Primary (1:500) Sigma

M2 mouse anti-FLAG antibodies and (1:1000) dsRNA (J2) antibodies were probed using secondary (1:1000) donkey anti-mouse IgG Alexa Fluor 488 antibody (Invitrogen). Primary

(1:100) rabbit anti-Tim44 and (1:400) rabbit anti-Mia40 antibodies were probed with secondary

(1:1000) goat anti-rabbit IgG Alexa Fluor 568 antibody (Invitrogen). Nuclear staining was accomplished using Vectashield® Antifade Mounting Medium with DAPI. Cells were visualized with a microscope 9Axiovert-200M (Carl Zeiss) using a Plan-Fluor 100x 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).

Densitometry analysis. Densitometry analysis was performed using Quantity 1 software (Bio-

Rad Laboratories) or image J. 211

Miscellaneous. Agarose gel electrophoresis was performed using standard protocols.

Nucleotides were visualized by ethidium bromide staining. 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.

212

References 1. Bostian, K. A., Hopper, J. E., Rogers, D. T. & Tipper, D. J. Translational analysis of the killer-associated virus-like particle dsRNA genome of S. cerevisiae: M dsRNA encodes toxin. Cell 19, 403–14 (1980). 2. Wickner, R. B., Fujimura, T. & Esteban, R. Viruses and prions of Saccharomyces cerevisiae. Adv. Virus Res. 86, 1–36 (2013). 3. Wickner, R. B. "Killer Character" of Saccharomyces cerevisiae: Curing by Growth at Elevated Temperature. JOURNAL OF BACTERIOLOGY (1974). 4. Sommer, S. S. & Wickner, R. B. Double-stranded RNAs that encode killer toxins in Saccharomyces cerevisiae: unstable size of M double-stranded RNA and inhibition of M2 replication by M1. Mol. Cell. Biol. 4, 1747–53 (1984). 5. Wickner, R. B. Killer of Saccharomyces cerevisiae: a double-stranded ribonucleic acid plasmid. Bacteriol. Rev. 40, 757–73 (1976). 6. Wickner, R. B. The killer double-stranded RNA plasmids of yeast. Plasmid 2, 303–22 (1979). 7. Vodkin, M. H. & Fink, G. R. A nucleic acid associated with a killer strain of yeast. Proc. Natl. Acad. Sci. U. S. A. 70, 1069–72 (1973). 8. Bevan, E. A., Herring, A. J. & Mitchell, D. J. Preliminary Characterization of Two Species of dsRNA in Yeast and their Relationship to the “Killer” Character. Nature 245, 81–86 (1973). 9. Starmer, W. T., Ganter, P. F., Aberdeen, V., Lachance, M. A. & Phaff, H. J. The ecological role of killer in natural communities of yeasts. Can. J. Microbiol. 33, 783–96 (1987). 10. Esteban, R. & Wickner, R. B. Three different M1 RNA-containing viruslike particle types in Saccharomyces cerevisiae: in vitro M1 double-stranded RNA synthesis. 6, (1986). 11. Schmitt, M. J. & Breinig, F. The viral killer system in yeast: from molecular biology to application. FEMS Microbiol. Rev. 26, 257–276 (2002). 12. Schmitt, M. J. & Breinig, F. Yeast viral killer toxins: lethality and self-protection. Nat. Rev. Microbiol. 4, 212–221 (2006). 13. Sommer, S. S. & Wickner, R. B. Yeast L dsRNA consists of at least three distinct RNAs; evidence that the non-mendelian genes [HOK], [NEX] and [EXL] are on one of these dsRNAs. Cell 31, 429–441 (1982). 14. Dinman, J. D., Icho, T. & Wickner, R. B. A -1 ribosomal frameshift in a double-stranded RNA virus of yeast forms a gag-pol fusion protein. Proc. Natl. Acad. Sci. U. S. A. 88, 174–8 (1991). 15. Castón, J. R. et al. Structure of L-A virus: a specialized compartment for the transcription and replication of double-stranded RNA. J. Cell Biol. 138, 975–85 (1997). 16. Wickner, R. B. & Leibowitz, M. J. Mak mutants of yeast: mapping and characterization. J. Bacteriol. 140, 154–60 (1979). 213

17. Wickner, R. B. Twenty-six chromosomal genes needed to maintain the killer double- stranded RNA plasmid of Saccharomyces cerevisiae. Genetics 88, 419–425 (1978). 18. Fink, G. R. & Styles, C. A. Curing of a killer factor in Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. U. S. A. 69, 2846–9 (1972). 19. Wickner, R. B. & Leibowitz, M. J. Chromosomal genes essential for replication of a double-stranded RNA plasmid of Saccharomyces cerevisiae: the killer character of yeast. J. Mol. Biol. 105, 427–43 (1976). 20. Toh-E, A., Guerry, P. & Wickner, R. B. Chromosomal superkiller mutants of Saccharomyces cerevisiae. J. Bacteriol. 136, 1002–7 (1978). 21. Ridley, S. P., Sommer, S. S. & Wickner, R. B. Superkiller mutations in Saccharomyces cerevisiae suppress exclusion of M2 double-stranded RNA by L-A-HN and confer cold sensitivity in the presence of M and L-A-HN. Mol. Cell. Biol. 4, 761–70 (1984). 22. Tercero, J. C., Riles, L. E. & Wickner, R. B. Localized mutagenesis and evidence for post- transcriptional regulation of MAK3. A putative N-acetyltransferase required for double- stranded RNA virus propagation in Saccharomyces cerevisiae. J. Biol. Chem. 267, 20270– 6 (1992). 23. Tercero, J. C., Dinman, J. D. & Wickner, R. B. Yeast MAK3 N-acetyltransferase recognizes the N-terminal four amino acids of the major coat protein (gag) of the L-A double-stranded RNA virus. J. Bacteriol. 175, 3192–4 (1993). 24. Toh-e, A. & Sahashi, Y. ThePET18 locus ofSaccharomyces cerevisiae: A complex locus containing multiple genes. Yeast 1, 159–171 (1985). 25. Brown, J. T., Bai, X. & Johnson, A. W. The yeast antiviral proteins Ski2p, Ski3p, and Ski8p exist as a complex in vivo. RNA 6, 449–57 (2000). 26. Mitchell, P., Petfalski, E., Shevchenko, A., Mann, M. & Tollervey, D. The exosome: a conserved eukaryotic RNA processing complex containing multiple 3’-->5’ exoribonucleases. Cell 91, 457–66 (1997). 27. Araki, Y. et al. Ski7p G protein interacts with the exosome and the Ski complex for 3’-to- 5’ mRNA decay in yeast. EMBO J. 20, 4684–93 (2001). 28. Dihanich, M., van Tuinen, E., Lambris, J. D. & Marshallsay, B. Accumulation of viruslike particles in a yeast mutant lacking a mitochondrial pore protein. Mol. Cell. Biol. 9, 1100–8 (1989). 29. Dihanich, M., Suda, K., Schatz Biocenter, G. & Schatz, G. A yeast mutant lacking mitochondrial porin is respiratory-deficient, but can recover respiration with simultaneous accumulation of an 86-kd extramitochondrial protein. EMBO J. 6, 723–728 (1987). 30. Liu, Y. & Dieckmann, C. L. Overproduction of Yeast Viruslike Particles by Strains Deficient in a Mitochondrial Nuclease. Mol. Cell. Biol. 9, 3323–3331 (1989). 31. Ruiz-Carrillo, A. & Renaud, J. Endonuclease G: a (dG)n X (dC)n-specific DNase from higher eukaryotes. EMBO J. 6, 401–7 (1987). 32. Schäfer, P. et al. Structural and Functional Characterization of Mitochondrial EndoG, a

214

Sugar Non-specific Nuclease which Plays an Important Role During Apoptosis. J. Mol. Biol. 338, 217–228 (2004). 33. Palmieri, F. & De Pinto, V. Purification and properties of the voltage-dependent anion channel of the outer mitochondrial membrane. J. Bioenerg. Biomembr. 21, 417–425 (1989). 34. Benz, R. Porin from bacterial and mitochondrial outer membranes. CRC Crit. Rev. Biochem. 19, 145–90 (1985). 35. Dihanich, M., Van Tuinen, E., Lambris, J. D. & Biocenter, B. M. Accumulation of Viruslike Particles in a Yeast Mutant Lacking a Mitochondrial Pore Protein. Mol. Cell. Biol. 9, 1100–1108 (1989). 36. Weber, F., Wagner, V., Rasmussen, S. B., Hartmann, R. & Paludan, S. R. Double- stranded RNA is produced by positive-strand RNA viruses and DNA viruses but not in detectable amounts by negative-strand RNA viruses. J. Virol. 80, 5059–64 (2006). 37. Parrish, J. et al. Mitochondrial endonuclease G is important for apoptosis in C. elegans. Nature 412, 90–94 (2001). 38. Li, L. Y., Luo, X. & Wang, X. Endonuclease G is an apoptotic DNase when released from mitochondria. Nature 412, 95–99 (2001). 39. Arnoult, D. et al. Mitochondrial release of AIF and EndoG requires caspase activation downstream of Bax/Bak-mediated permeabilization. EMBO J. 22, 4385–99 (2003). 40. van Loo, G. et al. Endonuclease G: a mitochondrial protein released in apoptosis and involved in caspase-independent DNA degradation. Cell Death Differ. 8, 1136–1142 (2001). 41. Madeo, F. et al. Oxygen stress: a regulator of apoptosis in yeast. J. Cell Biol. 145, 757–67 (1999). 42. Carmona-Gutierrez, D. et al. Apoptosis in yeast: triggers, pathways, subroutines. Cell Death Differ. 17, 763–773 (2010). 43. Eisenberg, T. et al. The mitochondrial pathway in yeast apoptosis. Apoptosis 12, 1011– 1023 (2007). 44. Büttner, S. et al. Endonuclease G Regulates Budding Yeast Life and Death. Mol. Cell 25, 233–246 (2007). 45. Shoshan-Barmatz, V. & Mizrachi, D. VDAC1: from structure to cancer therapy. Front. Oncol. 2, 164 (2012). 46. Adachi, M. et al. Bax interacts with the voltage-dependent anion channel and mediates ethanol-induced apoptosis in rat hepatocytes. Am. J. Physiol. Liver Physiol. 287, G695– G705 (2004). 47. Zamzami, N., Larochette, N. & Kroemer, G. Mitochondrial permeability transition in apoptosis and necrosis. 12, 1478–1480 (2005). 48. Bragoszewski, P. et al. Retro-translocation of mitochondrial intermembrane space proteins. Proc. Natl. Acad. Sci. U. S. A. 112, 7713–8 (2015). 215

49. Gadir, N., Haim-Vilmovsky, L., Kraut-Cohen, J. & Gerst, J. E. Localization of mRNAs coding for mitochondrial proteins in the yeast Saccharomyces cerevisiae. RNA 17, 1551– 65 (2011). 50. Wickner, R. B. Double-Stranded and Single-Stranded RNA Viruses of Saccharomyces Cerevisiae. Annu. Rev. Microbiol. 46, 347–375 (1992). 51. Becker, B. & Schmitt, M. J. Yeast Killer Toxin K28: Biology and Unique Strategy of Host Cell Intoxication and Killing. Toxins (Basel). 9, (2017). 52. Benard, L., Carroll, K., Valle, R. C. & Wickner, R. B. Ski6p is a homolog of RNA- processing enzymes that affects translation of non-poly(A) mRNAs and 60S ribosomal subunit biogenesis. Mol. Cell. Biol. 18, 2688–96 (1998). 53. Wickner, R. B. Plasmids controlled exclusion of the K2 killer double-stranded RNA plasmid of yeast. Cell 21, 217–26 (1980). 54. Terceros, J. C. & Wickner, R. B. MAK3 Encodes an N-Acetyltransferase Whose Modification of the L-A gag NH2 Terminus Is Necessary for Virus Particle Assembly*. J. Biol. Chem. 267, (2027). 55. Sieber, F., Duchêne, A.-M. & Maréchal-Drouard, L. Mitochondrial RNA Import. in International review of cell and molecular biology 287, 145–190 (2011). 56. Wang, G., Shimada, E., Koehler, C. M. & Teitell, M. A. PNPASE and RNA trafficking into mitochondria. Biochim. Biophys. Acta - Gene Regul. Mech. 1819, 998–1007 (2012). 57. Kim, K. M., Noh, J. H., Abdelmohsen, K. & Gorospe, M. Mitochondrial noncoding RNA transport. BMB Rep. 50, 164–174 (2017). 58. Magalhães, P. J., Andreu, A. L. & Schon, E. A. Evidence for the Presence of 5S rRNA in Mammalian Mitochondria. Mol. Biol. Cell 9, 2375 (1998). 59. Wang, G. et al. PNPASE Regulates RNA Import into Mitochondria. Cell 142, 456–467 (2010). 60. Schneider, A. & Maréchal-Drouard, L. Mitochondrial tRNA import: are there distinct mechanisms? Trends Cell Biol. 10, 509–13 (2000). 61. Kamenski, P. et al. Evidence for an Adaptation Mechanism of Mitochondrial Translation via tRNA Import from the Cytosol. Mol. Cell 26, 625–637 (2007). 62. Salinas, T. et al. The voltage-dependent anion channel, a major component of the tRNA import machinery in plant mitochondria. Proc. Natl. Acad. Sci. 103, 18362–18367 (2006). 63. Salinas, T. et al. Molecular basis for the differential interaction of plant mitochondrial VDAC proteins with tRNAs. Nucleic Acids Res. 42, 9937–9948 (2014). 64. Short, J. R. et al. Role of Mitochondrial Membrane Spherules in Flock House Virus Replication. J. Virol. 90, 3676–83 (2016). 65. den Boon, J. A., Diaz, A. & Ahlquist, P. Cytoplasmic Viral Replication Complexes. Cell Host Microbe 8, 77–85 (2010). 66. Dhir, A. et al. Mitochondrial double-stranded RNA triggers antiviral signalling in humans.

216

Nature 560, 238–242 (2018). 67. Lukša, J. et al. Different Metabolic Pathways Are Involved in Response of Saccharomyces cerevisiae to L-A and M Viruses. Toxins (Basel). 9, (2017). 68. Dihanich, M., Suda, K. & Schatz, G. A yeast mutant lacking mitochondrial porin is respiratory-deficient, but can recover respiration with simultaneous accumulation of an 86-kd extramitochondrial protein. EMBO J. 6, 723–8 (1987). 69. Gueldener, U., Heinisch, J., Koehler, G. J., Voss, D. & Hegemann, J. H. A second set of loxP marker cassettes for Cre-mediated multiple gene knockouts in budding yeast. Nucleic Acids Res. 30, e23 (2002). 70. Dirick, L. et al. Metabolic and Environmental Conditions Determine Nuclear Genomic Instability in Budding Yeast Lacking Mitochondrial DNA. G3 Genes, Genomes, Genet. 4, (2014). 71. Bähler, J. et al. Heterologous modules for efficient and versatile PCR-based gene targeting inSchizosaccharomyces pombe. Yeast 14, 943–951 (1998). 72. Kondo-Okamoto, N., Shaw, J. M. & Okamoto, K. Mmm1p spans both the outer and inner mitochondrial membranes and contains distinct domains for targeting and foci formation. J. Biol. Chem. 278, 48997–9005 (2003). 73. Fried, H. M. & Fink, G. R. Electron microscopic heteroduplex analysis of "killer" double-stranded RNA species from yeast. Proc. Natl. Acad. Sci. U. S. A. 75, 4224–8 (1978). 74. Glick, B. S. & Pon, L. A. Isolation of highly purified mitochondria from Saccharomyces cerevisiae. Methods Enzymol. 260, 213–23 (1995). 75. Dabir, D. V. V et al. A small molecule inhibitor of redox-regulated protein translocation into mitochondria. Dev. Cell 25, 81–92 (2013). 76. Hasson, S. A. et al. Substrate specificity of the TIM22 mitochondrial import pathway revealed with small molecule inhibitor of protein translocation. Proc. Natl. Acad. Sci. U. S. A. 107, 9578–83 (2010). 77. Koehler, C. M. et al. Tim9p, an essential partner subunit of Tim10p for the import of mitochondrial carrier proteins. EMBO J. 17, 6477–86 (1998). 78. Teste, M.-A., Duquenne, M., François, J. M. & Parrou, J.-L. Validation of reference genes for quantitative expression analysis by real-time RT-PCR in Saccharomyces cerevisiae. BMC Mol. Biol. 10, 99 (2009). 79. Nakayashiki, T., Kurtzman, C. P., Edskes, H. K. & Wickner, R. B. Yeast prions [URE3] and [PSI+] are diseases. Proc. Natl. Acad. Sci. U. S. A. 102, 10575–80 (2005). 80. Garcia, M. et al. Mitochondria-associated yeast mRNAs and the biogenesis of molecular complexes. Mol. Biol. Cell 18, 362–8 (2007). 81. Madeo, F., Fröhlich, E. & Fröhlich, K. U. A yeast mutant showing diagnostic markers of early and late apoptosis. J. Cell Biol. 139, 729–34 (1997). 82. Amberg, D. C., Burke, D. J. & Strathern, J. N. Yeast immunofluorescence. CSH Protoc. 217

2006, pdb.prot4167 (2006).

218