PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5/08 KrlPrqlAcc&ProsICIRCIDateDuo.indd COORDINATION OF DIVISION COMPLEXES ACROSS THE PLASTID ENVELOPE MEMBRANES By Jonathan Matthew Glynn A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Genetics 2009 ABSTRACT COORDINATION OF DIVISION COMPLEXES ACROSS THE PLASTID ENVELOPE MEMBRANES By Jonathan Matthew Glynn Chloroplasts are cellular organelles descended from a cyanobacterial endosymbiont and house the photochemical machinery which powers synthesis of reduced carbon compounds from carbon dioxide and water. In higher plants, the chloroplast is also the site of synthesis for a select group of lipids, amino acids, and plant hormones. This set of organelle-specific functions make the chloroplast essential to survival of land plants, the most prominent group of terrestrial primary producers. The replication and segregation of plastids within land plants occurs through binary fission and is an important part of plant cell biology, as chloroplast size and number may impact the efficiency of photosynthesis and the partitioning of chloroplasts to daughter cells during plant cell division. The apparatus that facilitates the scission of a single chloroplast into two daughter chloroplasts is a complex macromolecular machine, partly composed of a host-derived dynamin ring on the outside of the organelle and an endosymbiont-derived FtsZ ring (Z- ring) inside the organelle. The activities of these two rings must be tightly coordinated across the two envelope membranes of the chloroplast to ensure the timely progression and completion of division. Here, I show that ARC6, a known FtsZ assembly factor that promotes formation of FtsZ filaments, also specifies the mid-plastid positioning of the paralogous outer envelope proteins PDVl and PDV2, which have parallel functions in dynamin recruitment. PDV2 positioning requires a direct interaction between ARC6 and PDV2 that may be regulated by post-translational modification of ARC6 within the interrnembrane space. I also show that PARC6 (Paralog of ARC6), like ARC6, is a multifunctional inner envelope chloroplast division protein. PARC6 acts downstream of ARC6 to position PDVl at the division site, but is not required for PDV2 or ARCS localization. Arabidopsis parc6 mutants exhibit compound chloroplast division phenotypes and FtsZ filament morphology defects suggesting that PARC6 acts antagonistically to ARC6 within the chloroplast stoma as an inhibitor of FtsZ assembly. This FtsZ assembly-inhibiting activity of PARC6 may occur through interaction with ARC3, a protein with functional similarity to bacterial MinC. PARC6-OPP localization is dynamic, consistent with its complex role in division. Our findings indicate that PARC6 and ARC6 play related but distinct roles in coordinating the internal and external components of the chloroplast division complex. DEDICATION This work is dedicated to my mother and father, both of whom instilled the values and ideals that are a critical component of any scientist’s success. This work is dedicated to my grandfather, Elmer E. Lehman, whose patience, thoughtfulness, and pragmatic approach to life were an inspiration to all who knew him. This work is dedicated to Van McWilliams and Helmut Bertrand, two teachers who kept me interested in science at times when I might have followed another path. This work is dedicated to my family, whose understanding and encouragement over the years have been irreplaceable. Most importantly, this work is dedicated to Angie. Your attitude, patience, kindness, understanding, support, and smile have helped make this possible. You are the love of my life and I am thankful for every day we have together. iv ACKNOWLEDGEMENTS I wish to acknowledge my advisor, Katherine W. Osteryoung, for her support and approach to mentoring that fosters independent critical thinking. I wish to thank the members of my guidance committee for their encouragement and useful commentary: Dr. Jianping Hu Dr. Robert Last Dr. Beronda Mongomery-Kaguri Dr. Andreas Weber I wish to thank Aaron Schmitz, Deena Kadirj an—Kalbach, Shin-ya Miyagishima, Qiang Wang, and Yue Yang for helpful scientific discussions and friendship over the past several years. I wish to thank John Froehlich for insightful experimental advice, useful scientific discussions, and technical support throughout this project. TABLE OF CONTENTS LIST OF FIGURES .................................................................................. ix LIST OF ABBREVIATIONS ........................................................................ xiii CHAPTER 1 EVOLUTION AND OPERATION OF THE CHLOROPLAST DIVISION MACHINERY ...................................................................................................................... 1 Plastid Biology and Analysis of Chloroplast Division ............................................. 2 A Survey of the Molecular Biology of Plastid Division .......................................... 7 Regulating Plastid Fission: Why do Plastids Divide in Land Plants? .................... 24 Perspectives and Acknowledgements ...................................................................... 29 CHAPTER 2 IDENTIFICATION AND MODELING OF AN FTZ2-BINDIN G DOMAIN WITHIN ARC6 .................................................................................................................................. 30 Abstract ................................................................................................................. 31 Introduction ............................................................................................................ 32 Results .................................................................................................................... 35 The F tsZ2-binding domain of ARC6 occupies amino acids 351 -503 ......... 35 The C-terminus of F tsZ2 is suflicient for ARC6-FtsZ2 interaction ........... 37 ARC6ZBD can be modeled onto the structure of ZipA ZBD ........................ 39 Mutations that aflect ZipAZBD-FtsZ interaction afl'ect ARC 6 ZB D-F tsZ2 interaction .................................................................................................. 44 Analysis of the ARC6ZBD-FtsZ2 interaction by pulldown assay ............... 46 In vivo analysis of the ARC 6 F44 2 D site-directed mutation ........................ 47 In vivo analysis of chloroplast-targeted ARC6 ZBD ................................... 53 Discussion ............................................................................................................. 58 Materials and Methods .......................................................................................... 64 CHAPTER 3 ARC6 BINDS AND POSITIONS PDV2 DURING PLASTID FISSION IN ARABIDOPSIS ................................................................................................................... 70 Abstract ................................................................................................................. 71 Introduction ............................................................................................................ 72 Results .................................................................................................................... 75 PD V2 localization and topology are similar to PDVI ............................... 75 The [MS regions of ARC6 and PD V2 interact ............................................ 81 PD V2 family members possess a unique C-terminal domain .................... 83 The conserved terminal glycine of PD V2 is required for interaction with ARC6 and for PDVZfimction in vivo .............................................................. 83 vi The complete IMS region of ARC6 is not required for ARC6 localization but is required for chloroplast division activity .......................................... 90 ARC6 acts upstream of PD V2 and is required for PD V2 activity ............. 92 ARC6 is required for positioning PD V2, PD V1, and ARC5 ...................... 94 Discussion ............................................................................................................ 100 Materials and Methods ......................................................................................... 104 CHAPTER 4 CHARACTERIZATION OF THE PDV2 BINDING DOMAIN OF ARC6 ................... 110 Abstract ............................................................................................................... 111 Introduction .......................................................................................................... 1 12 Results .................................................................................................................. 114 The PDV2 binding domain of ARC6 occupies amino acids 63 6- 759 ....... 114 Structural Analysis of ARC6 p31) and PD V2IMS ...................................... 1 16 A conserved serine within ARC 61231) influences ARC 6-PD V2 interaction ........................................................................ 120 A phosphomimetic mutation in ARC6 p31) causes chloroplast division defects in vivo ........................................................................................... 124 PD V2, PD VI, ARC5, and F tsZ localize to the division site in ARC 63744 E mutants ..................................................................................................... 126 Discussion ............................................................................................................ 1 30 Materials and Methods ......................................................................................... 136 CHAPTER 5 PARC6 INFLUENCES FTSZ ASSEMBLY AND PDVl RECRUITMENT IN ARABIDOPSIS ...................................................................................................... ....... 140 Abstract ............................................................................................................... 141 Introduction ........................................................................................................... 142 Results .................................................................................................................... 145 PARC6 family members are distinct fiom ARC 6 and are unique to vascular plants ......................................................................................... 145 Mutations in PARC6 cause aberrations in chloroplast morphology and F tsZ filament morphology ........................................................................ 149 PARC6 is an inner envelope protein with localization similar to PD VI .................................................................................................... 151 PARC6 inhibits F tsZ assembly and interacts with ARC3 ........................ 155 PARC6 acts downstream of ARC 6 ........................................................... 157 PARC6 is required for PD VI localization, but not for PD V2 or ARC5 localization ............................................................................................... 161 PARC 6 binds the cytosolic domain of PD V1 in two-hybrid assays ......... 163 PD V1 and PD V2 independently localize to the division site ................... 164 Discussion ............................................................................................................ 167 Materials and Methods ......................................................................................... 173 vii CHAPTER 6 A CONSERVED ASPARTATE WITHIN ARC6 INFLUENCES PLASTID SIZE AND Z-RING POSITION .......................................................................................................... 178 Abstract ................................................................................................................ 179 Introduction .......................................................................................................... 180 Results .................................................................................................................. 183 arc6 D 20 5N is a hypomorphic allele of ARC6 that is associated with defects in chloroplast morphology and number ....................................... 183 ARC 6 D 20 5 is conserved in land plants, algae, and cyanobacteria ........ 186 arc6 D20 5N is associated with miniaturized and misplaced plastidic Z- rings, as well as a decrease in ARC6 protein levels ................................ 186 AtMinE— YF P has a unique localization pattern in Arabidopsis and ARC60205 mutants do not interact with AtMinE ....................................... 189 ARC 6 probably functions downstream of F tsZI/F tsZ2 assembly and downstream of A tMinE ............................................................................. 194 Discussion ............................................................................................................ 196 Materials and Methods ......................................................................................... 200 CHAPTER 7 CONCLUSIONS AND FUTURE DIRECTIONS ........................................................... 205 Summary of ARC6 and PARC6 Functional Analysis and Future Directions ...... 206 Unanswered Questions, Additional Observations, and New Hypotheses ............ 208 Working Model of the Coordination of the Chloroplast Divisome in Tracheophytes ....................................................................................................... 21 5 APPENDICES ................................................................................................................. 224 REFERENCES ................................................................................................................ 242 viii LIST OF FIGURES Images in this dissertation are presented in color. Figure 1.1. overview of chloroplast morphology and features of a typical chloroplast. . .3 Figure 1.2. Comparison of dividing chloroplasts and cyanobacteria. . . . . . . . . . .5 Figure 1.3. F tsZ2-1 immunolocalization in wild type Arabidopsis chloroplasts and in arc6 mutants... 15 Figure 1.4. Schematic of the Arabidopsis ARC6 protein ..................................... 16 Figure 1.5. FtsZ and ARC6 colocalize to continuous rings within the chloroplast. . .. 17 Figure 2.1. Mapping the FtsZ2-binding domain of ARC6 by yeast two-hybrid assay. . .36 Figure 2.2. Two-hybrid HIS reporter assays showing ARC6-FtsZ2 interaction ........... 38 Figure 2.3. Alignment of ARC6ZBD with ZipAZBD ........................................... 40 Figure 2.4. Comparison of the E. coli ZipAZBD crystal structure and the A. thaliana ARC6ZBD homology model ....................................................................... 42 Figure 2.5. Secondary structure predictions of ARC6ZBD .................................... 43 Figure 2.6. ARC6ZBD (1:442) site-directed mutants ............................................ 45 Figure 2.7. Pulldown of FtsZ2 by ARC6 ........................................................ 48 Figure 2.8. Demonstration of ARC6 and ARC6 F44 2D transgene expression in the arc6 (SAIL_693_G04) background .................................................................... 50 Figure 2.9. Quantitative analysis of chloroplast number in ARC6 1:44 ZD-expressing transgenics ............................................................................................ 51 Figure 2.10. FtsZ immunolocalization in ARC6F44ZD-expressing transgenics. . 5........2 Figure 2.11. Schematic of the plastid-targeting vector pCAMBIA-ZBD used for ARC6ZBD domain analysis ....................................................................... 55 Figure 2.12. Demonstration of expression and targeting of ARC6ZBD-EYFP..............56 ix Figure 2.13. Chloroplast and FtsZ morphology in ARC6ZBD-EYFP expressing lines...57 Figure 2.14. ARC6 A A1-509-GFP and ARC6AA1-331-GFP can immunoprecipitate FtsZ2 ................................................................................................. 63 Figure 3.1. Schematic comparison of PDVl and PDV2 proteins from Arabidopsis. ....76 Figure 3.2. Localization of YFP-PDV2 in Arabidopsis ......................................... 78 Figure 3.3. Colocalization of ARC6-CFP and YFP-PDV2 in Arabidopsis. . . . . . 7.9 Figure 3.4. PDV2 fractionation and topology using isolated pea chloroplasts ............. 80 Figure 3.5. The IMS-localized regions of PDV2 and ARC6 interact ...................... 82 Figure 3.6. Alignment of C-termini of PDVl and PDV2 family members ................ 84 Figure 3.7. The terminal glycine of PDV2 is important for interaction with ARC6. ....85 Figure 3.8. PD V20 307D is a loss-of—function allele ............................................. 87 Figure 3.9. PDVZG307D localization in Arabidopsis .......................................... 89 Figure 3.10. ARC6A1M3-GFP is partially functional and localizes to the division site..91 Figure 3.11. ARC6 acts upstream of PDV2 ...................................................... 93 Figure 3.12. Overexpression of PD V2 does not abrogate the arc6 phenotype ............ 95 Figure 3.13. ARC6 is required for localization of YFP-PDV2 to the division site ........ 97 Figure 3.14. PD V2 is not required for ARC6-GF P localization to equatorial rings. .....98 Figure 3.15. ARC6 is required for localization of GFP-ARCS to the division site. . . .....99 Figure 4.1. The PDV2 binding domain of ARC6 occupies amino acids 636-759 ...... 115 Figure 4.2. Pairwise alignment between the N-terrninal domain of Rst and the IMS region of PDV2 ........................................................................................ 118 Figure 4.3. Structural model of PDVZIMS ...................................................... 119 Figure 4.4. Identification of a plant-specific motif within the C-termini of ARC6/Fm2 family members ..................................................................................... 121 Figure 4.5. A phosphomimetic mutation in the IMS region of ARC6 decreases its affinity for PDV2 .............................................................................................. 123 Figure 4.6. ARC6S744E is dysfimctional ........................................................ 125 Figure 4.7. Phenotypes observed in lines expressing ARC6S744 A and ARC 657445. .....127 Figure 4.8. Localization of PDV2, ARC5, PDVl, and FtsZ in ARC6S744E~expressing transgenics ............................................................................................ 128 Figure 4.9. An unusual phenotype observed in ARC6S744E petioles phenocopies plastidic Min system defects ................................................................................ 132 Figure 5.1. Summary of PARC6 protein features and similarity to related proteins. . ...146 Figure 5.2. Phylogenetic analysis of PARC6, ARC6, and Ftn2 family members ........ 148 Figure 5.3. Diagram of the PARC6 locus and phenotypic analysis of parc6 mutants... 150 Figure 5.4. PARC6 is a protein that localizes to the division site and polar spots. . . 1 52 Figure 5.5. PARC6 is a chloroplast membrane protein ....................................... 154 Figure 5.6. Analysis of FtsZ morphology and FtsZ protein levels in parc6-1 ............ 156 Figure 5.7. PARC6 does not interact with an FtsZ protein ................................... 158 Figure 5.8. PARC6 interacts with ARC3 ........................................................ 159 Figure 5.9. PARC6 acts downstream of ARC6 ................................................ 160 Figure 5.10. Localization of PDVl, PDV2, and ARC5 in parc6 mutants .................... 162 Figure 5.11. The C-terrninus of PARC6 binds the cytosolic domain of PDVl .......... 165 Figure 5.12. PDVl and PDV2 localize independently of each other ...................... 166 Figure 6.1. arc6 D 20 5N is associated with a defect in chloroplast size and number. . l 85 Figure 6.2. Multiple sequence alignment showing conservation of ARC69205 ......... 187 Figure 6.3. F tsZ localization and mini-chloroplasts in arc6 020 5 N ......................... 188 xi Figure 6.4. Preliminary analysis of ARC6, Ftle, and FtsZZ levels in arc6Dzo5N. .....190 Figure 6.5. Localization of AtMinE-YFP in Arabidopsis ................................... 192 Figure 6.6. ARC6-GFP does not localize to rings in ftsZ2 or arc12 mutants ............. 195 Figure 7.1. Domain architecture of ARC6 and PARC6 proteins ............................ 207 Figure 7.2. Coordination of Z-ring dynamics and ARC5 recruitment through ARC6 and PARC6 ............................................................................................... 222 xii 35$ ABRC AD ARC ATP BD BiF C BLAST CD CFP Chl Col CT C-terminal DIC EMPTY EMS F ts ABBREVIATIONS Deletion Cauliflower Mosaic Virus 35S promoter Amino acid Arabidopsis Biological Resource Center Activation domain Accumulation and Replication of Chloroplasts Adenosine-5'-Triphosphate DNA binding domain Bimolecular Fluorescence Complementation Basic Local Alignment Search Tool Celcius Circular dichroism Cyan fluorescent protein Chlorophyll Columbia C-terrninus or C-terminal Carboxy terminal Differential interference contrast Empty vector Ethane methyl sulfonate Filamentation temperature sensitive xiii FZL FZO GC GDP GFP GST GTP His Hiss IEM IMS INT IP IPTG Ler MORN NADPH NMR N-terminal OEM PARC PBD PBS FZO-like Fuzzy onions Giant chloroplast Guanosine-5'-diphosphate Green fluorescent protein Glutathione-S-transferase Guanosine-5'-triphosphate Histidine 8X Histidine tag Inner envelope membrane Intermembrane space Intermediate Immunoprecipitation Isopropyl-beta-D-thiogalactopyranoside Landsberg erecta Membrane occupation and recognition nexus Nicotinamide adenine dinucleotide phosphate (reduced) Nuclear magnetic resonance spectroscopy Amino terminal Outer envelope membrane Paralog of ARC PDV2 binding domain Phosphate-buffered saline xiv PD PDB PDV pro SD SD/-ULT SDI-ULTH SOE-PCR TBS THM TM TP Ws YFP ZBD Zip Plastid dividing Protein Data Bank Plastid division Proline or promoter (subscript) Synthetic dropout Synthetic dropout media lacking uracil, leucine, and tryptophan SD media lacking uracil, leucine, tryptophan, and histidine Site overlap extension polymerase chain reaction Tris-buffered saline Thylakoid membrane Transmembrane domain Transit peptide Wassilewskija Yellow fluorescent protein FtsZ (or FtsZZ) binding domain FtsZ-interacting protein XV Chapter 1 Evolution and Operation of the Chloroplast Division Machinery Plastid Biology and Analysis of Chloroplast Division Chloroplasts are cellular organelles in land plants, responsible for photosynthesis and a number of other life-supporting functions, such as amino acid biosynthesis and lipid metabolism (Pyke 2009). In land plants like Arabidopsis thaliana, each mature leaf mesophyll cell contains 60-120 chloroplasts (Figure 1.1), depending on the accession and growth conditions (Aldridge, Maple, et al. 2005, Pyke 1999, Pyke 2009, Pyke and Leech 1994). Chloroplasts within Arabidopsis are typically ovoid-shaped structures, 5-7 microns (pm) in length, and harbor three distinct compartments: the thylakoid lumen, the stroma, and the intermembrane space (Figure 1.1). Each of these compartments is bounded by a biological membrane: the thylakoid lumen is surrounded by the thylakoid membrane, the stroma by the inner envelope membrane, and the entire organelle is bounded by the outer envelope membrane; the space between the inner and outer envelopes is called the intermembrane space (IMS). The thylakoid membrane is the site of photocollection and electron transport that drives photosynthesis, while the thylakoid lumen stores hydrogen ions generated during the photolysis of water molecules. These hydrogen ions are transferred to the stroma through an ATP synthase embedded within the thylakoid membrane to generate ATP within the stroma. The transport of electrons through the electron transport chain results in the reduction of NADPH. Within the stroma, ATP is consumed during operation of the Calvin cycle, which generates reduced carbon (triose phosphate) from carbon dioxide and NADPH. The inner and outer envelope membranes separate the chloroplast stroma from the surrounding cytosol and other organelles. These two membranes contain a variety of metabolite transporters Figure 1.1. Overview of chloroplast morphology and features of a typical chloroplast. The lefi panel shows a typical mesophyll cell from an expanded leaf of Arabidopsis thaliana accession Col-0. Using DIC microscopy, wild type chloroplasts typically appear as greenish spherical or ovoid bodies, approximately 5-7 pm in diameter. The right panel is a schematic detailing the compartments and membranes that make up a typical chloroplast. Outer Envelope Membrane (OEM); Intermembrane Space (IMS); Inner Envelope Membrane (IEM); Thylakoid Membrane (THM); and Stromal compartment (Stroma). Scale bar = 10pm. and protein translocators that allow the plastid to efficiently interface with the rest of the cell (Pyke 2009). The chloroplasts of land plants are decendants of a cyanobacterial endosymbiont that first inhabited a primitive protozoan about 1.2-1 .8 billion years ago (Dyall, Brown, et al. 2004, Yoon, Hackett, et al. 2004). Chloroplasts replicate through binary fission (Schimper 1883), dividing near their midpoint, similar to many bacteria (Figure 1.2). The process of division may appear relatively simple, but is a highly complex process that involves a number of protein components. The chloroplasts of land plants maintain a minimal genome, typically harboring 110-120 genes derived from the cyanobacterial endosymbiont (Cui, Veeraraghavan, et al. 2006); however, all known plastid division factors are encoded by nuclear genes in Arabidopsis and other land plants (Yang, Glynn, et al. 2008). Some of the genes encoding plastid division factors are inventions of the host organism and arose after a stable endosymbiotic relationship was established, while others are descendants of cyanobacterial genes that were transferred to the nucleus from the endosymbiont genome during the evolution of photosynthetic eukaryotes. Because of their common origin, several components of the chloroplast divisome share a high degree of sequence similarity with cyanobacterial division proteins from extant species (Yang, Glynn, et al. 2008). Due to the evolutionary history of the chloroplast, the study of chloroplast division has benefited greatly from the study of bacterial cell division. Surprisingly, genetic screens for cyanobacterial cell division -E Figure 1.2. Comparison of dividing chloroplasts and cyanobacteria. (A) Micrographs of chloroplasts from Arabidopsis thaliana Col-0 and (B) cells of the cyanobacterium Synechococcus elongatus PCC7942 as they progress from single individuals (upper panels) to two new entities (lowest panel). Scale bars = 5pm. 1 v components have uncovered only a few novel division genes (Koksharova and Wolk 2002, Miyagishima, Wolk, et al. 2005), but the division of cyanobacteria has been understudied in comparison to that of Escherichia coli (Lutkenhaus 2007, Margolin 2003, Margolin 2005, Rothfield, Taghbalout, et al. 2005). Consequently, models of plastid division at a molecular level are often compared to E. coli cell division models. Many of the components that make up the divisome of E. coli are also utilized by cyanobacteria (Miyagishima 2005), but cyanobacteria possess division factors that are unique to photosynthetic prokaryotes (Koksharova and Wolk 2002, Miyagishima, Wolk, et al. 2005) and are probably the most relevant prokaryotic system for comparative analysis with regard to the chloroplast. Regardless, the first known component of the plastid divisome was identified by a reverse genetic approach, parsing sequences of translated expressed sequence tags (ESTs) from Arabidopsis thaliana for similarity to the amino acid sequence of the E. coli cell division protein FtsZ (Osteryoung and Vierling 1995). Later studies identified more components of the chloroplast divisome by reverse genetics, with the initial query sequences being E. coli cell division proteins (Colletti, Tattersall, et al. 2000, Itoh, Fujiwara, et al. 2001, Maple, Fujiwara, et al. 2004). While reverse genetic screens have been useful in identifying broadly-conserved division proteins common to bacteria and plastids, they fall short in their ability to identify plastid division factors of host origin, as these factors are generally products of genes that emerged after the intial endosymbiotic event (Gao, Kadirj an-Kalbach, et al. 2003, Miyagishima, Froehlich, et al. 2006, Miyagishima, Nishida, et al. 2003a). The discovery of these host-derived components has benefited from forward genetic analysis, where the leaf cells of mutagenized lines are analyzed for aberrations in chloroplast number and/or morphology by simple light microscopy (Miyagishima, Froehlich, et al. 2006, Pyke and Leech 1992, Pyke and Leech 1994) and the causative mutation mapped by routine molecular methods (Jander, Norris, et al. 2002). Forward genetic studies have been instrumental to the discovery of plant-specific genes involved in plastid division, uncovering factors that act both inside and outside the organelle. Surprisingly, forward genetic approaches have only recently uncovered high-level regulators of plastid division, such as transcription factors and/or hormone responsive factors (Okazaki, Kabeya, et al. 2009), whose existence and involvement in plastid division and differentiation has been postulated for some time. However, the elucidation of the complete inventory of plastid division genes will not come through one mode of analysis, but through a comprehensive approach that incorporates both forward and reverse genetics, comparative genomics, transcriptome/network analysis, proteomics/interactomics, and integrative approaches that utilize all of these strategies in multiple experimental systems. A Survey of the Molecular Biology of Plastid Division F tsZ FtsZ is a conserved division protein found in most bacteria and is a structural homolog of eukaryotic tubulins (Erickson, Taylor, et al. 1996). In E. coli, FtsZ forms a ring (Z-ring) at the mid-cell which demarcates the site of cell division (Bi and Lutkenhaus 1991), probably serving as a scaffold for other division proteins and providing nominal contractile force during membrane constriction (Ghosh and Sain 2008, Lan, Daniels, et al. 2009, Lan, Wolgemuth, et al. 2007, Osawa, Anderson, et al. 2008). The biochemistry of the cyanobacterial FtsZ has yet to be rigorously examined and most of what we know of FtsZ assembly and biochemistry come from several detailed studies of F tsZ from E. coli, though the similarity between these two proteins (Stokes and Osteryoung 2003) would suggest that they behave similarly in vitro. In vitro, E. coli FtsZ assembles into a simple linear polymer (protofilament) with a diameter of about 5 nm in the presence of GTP, but can form mini-rings or other higher-order structures under certain conditions (Erickson, Taylor, et al. 1996). Hydrolysis of GTP to GDP can lead to curvature of the FtsZ polymer, and this induced curvature is thought to be one source of contractile force during cell division (Lu, Reedy, et al. 2000). The addition of calcium or low pH conditions facilitates lateral association of FtsZ protofilaments into bundles or sheets approximately 30 nm or more in diameter (Erickson, Taylor, et al. 1996). It is thought that the protofilaments serve as building blocks for the cytosolic Z-ring in vivo, undergoing further lateral association and organization in the presence of accessory factors such as FtsA (Jensen, Thompson, et al. 2005, Pichoff and Lutkenhaus 2002) and ZipA (Hale, Rhee, et al. 2000, RayChaudhuri 1999), proteins that enhance polymer bundling and anchor the developing Z-ring to the inner leaflet of the cell membrane (Osawa, Anderson, et al. 2008, Pichoff and Lutkenhaus 2005). In addition to the polymer curvature associated with GTP hydrolysis, the sliding of adjacent protofilaments has also been proposed as an alternative or additional source of contractile force during bacterial cell division (Lan, Daniels, et al. 2009, Li, Trimble, et al. 2007), but is energetically unfavorable (Erickson 2009) and has yet to be experimentally proven. In contrast to most bacteria, all plant genomes encode at least two phylogenetically-distinct families of FtsZ proteins, termed Ftle and FtsZZ (Miyagishima, Nozaki, et al. 2004, Stokes and Osteryoung 2003). Consistent with their divergent sequences, Ftle and FtsZ2 have distinct roles in plastid division (Schmitz, Glynn, et al. 2009, Stokes and Osteryoung 2003, Vitha, McAndrew, et al. 2001). Both Ftle and FtsZ2 have been shown to be part of the Z-ring, but the exact arrangement of the polymer is unclear. Ftle-FtsZ2 mixed polymers are able to form large protofilament bundles similar to the large protofilament bundles observed using bacterial FtsZ, suggesting that the plastidic Z-ring is probably a heteropolymer rather than several closely-associated Ftle and FtsZZ homopolymers (Olson 2008). While bundling is probably not critical for contractile activity (Erickson 2009), it is probably required for the formation of a functional and stable Z-ring (Hale, Rhee, et al. 2000, Low, Moncrieffe, et al. 2004, Margolin 2003). The factors that contribute to FtsZ bundling and stability in plastids are discussed later. Determining the site of F tsZ protofilament and ring assembly: The Min System Mid-cell positioning of the Z-ring in bacteria is important for symmetrical division and ensures equal partitioning of cellular contents, including chromosomes and plasmids, during fission. In E. coli, two major mechanisms drive placement of the Z-ring at the mid-cell: the Min system and the Nucleoid occlusion (Noc) system (Rothfield, Taghbalout, et al. 2005). The Min system of E. coli is composed of three proteins: MinC, MinD, and MinE (Lutkenhaus 2007). MinC binds F tsZ and inhibits Z-ring assembly by preventing lateral associations (I-Iu, Mukherjee, et al. 1999, Scheffers 2008, Shen and Lutkenhaus 2009). However, the FtsZ polymer-inhibiting activity of MinC is regulated by MinD, which is tethered to the membrane and promotes MinC activity only in the polar zones of the cell (Hu and Lutkenhaus 1999, Raskin and de Boer 1999a, Szeto, Rowland, et al. 2002). MinD is regulated by MinE, through a mechanism in which MinE binds to MinD and causes it to be released from the membrane (Hu and Lutkenhaus 2001). The maximum concentration of MinE occurs near the midcell adjacent to the membrane, thereby causing the concentration of active MinC and MinD to be highest at the poles (Hale, Meinhardt, et al. 2001). While the determinants for MinE localization are unknown, the midcell zone created by MinE activity allows for FtsZ protofilament assembly along the inner leaflet of the plasma membrane at the midcell (Rothfield, Taghbalout, et al. 2005). In some bacteria that lack MinE, DivIVA tethers MinD at the cell poles and inhibits F tsZ polymerization within the polar zone by maintaining a higher concentration of active MinCD at the poles (Marston and Errington 1999). In general, it has long been thought that the functions of the Arabidopsis orthologs of MinD (AtMinD) and MinE (AtMinE) are similar to their bacterial counterparts with respect to their functions during chloroplast division. However, recent studies have shown that AtMinD is able to rescue the cell division defects of an E. coli minD minE double mutant, indicating that AtMinD has probably acquired additional functions, relative to its bacterial counterpart (Zhang, Hu, et al. 2009a). Additionally, the localization patterns of AtMinD and AtMinE vary slightly from those of their bacterial homologs. AtMinD localizes to polar spots and equatorial structures in Arabidopsis (Nakanishi, Suzuki, et al. 2009), while AtMinE localizes to polar regions in tobacco 10 chloroplasts (Maple, Chua, et al. 2002). Intriguingly, orthologs of MinC have been lost from the genomes of vascular plants (Yang, Glynn, et al. 2008) and recent work shows that plants have acquired at least two additional factors to fine-tune the operation of the Min system: ARC3 and MCDl (Glynn, Miyagishima, et al. 2007, Maple, Vojta, et al. 2007, Nakanishi, Suzuki, et al. 2009). ARC3 is a chimeric protein, consisting of an FtsZ-like N-terminal domain, a C- terrninal PIPSK-like domain, and a conserved middle domain of unknown function (Shimada, Koizumi, et al. 2004). ARC3 is proposed as a functional replacement for MinC in plastids (Maple, Vojta, et al. 2007), as arc3 mutants possess chloroplast morphologies of varying size (Marrison, Rutherford, et al. 1999, Pyke and Leech 1992, Pyke and Leech 1994), with the enlarged plastids containing multiple FtsZ rings (Glynn, Miyagishima, et al. 2007) reminiscent of those observed within bacterial minC mutants (Levin, Shim, et al. 1998). ARC3 exhibits both polar and equatorial localization within the chloroplast (Maple, Vojta, et al. 2007, Shimada, Koizumi, et al. 2004). Furthermore, ARC3 has been shown to interact with Ftle , MinD, and MinE (Maple, Vojta, et al. 2007), suggesting that it acts as a critical interface between the Z-ring and the plastidic Min system. Based on these data, ARC3 has been hypothesized to inhibit FtsZ assembly, but it is not yet known if ARC3 is sufficient to inhibit FtsZ assembly in vitro, similar to bacterial MinC (Hu, Mukherjee, et al. 1999, Scheffers 2008). MCDl is a plant-specific inner envelope protein that regulates division site placement within the chloroplast (N akanishi, Suzuki, et al. 2009). Like ARC3 (Maple, 11 Vojta, et al. 2007, Shimada, Koizumi, et al. 2004) and AtMinD (Nakanishi, Suzuki, et al. 2009), MCDl localizes to polar zones and to equatorial structures within the chloroplast (Nakanishi, Suzuki, et al. 2009). MCDl binds AtMinD and is required for AtMinD localization in vivo, and thereby directs the position of the Z-ring and division site (N akanishi, Suzuki, et al. 2009). However, it is unclear what pathways lie upstream of MCDl to direct its localization and activity within the chloroplast. Orthologs of characterized nucleoid occlusion factors are not encoded in cyanobacterial or plant genomes, suggesting that chloroplasts do not utilize a Noe-like mechanism to prevent scission of the plastid chromosome during fission (Glynn, Miyagishima, et al. 2007). Moreover, Z-rings can form around nucleoids in cyanobacteria, indicating that a Noe-like mechanism is probably absent from this lineage (Miyagishima, Wolk, et al. 2005). Both plastids and cyanobacteria maintain multiple copies of their chromosome (F alkow, Dworkin, et al. 2006, Pyke 2009), consistent with chromosomal preservation during division being accomplished by maintaining multiple chromosomal copies (Scott and Possingham 1980). Notably, plastids and cyanobacteria possess an additional membrane system, the thylakoid membrane, that is not found in E. coli or other bacteria (Falkow, Dworkin, et al. 2006). It has been proposed that some mechanism of thylakoid partitioning might occur in plants, as the thylakoid membranes might be an impediment to the contractile apparatus and thylakoids appear to be actively redistributed during plastid fission (Boffey and Lloyd 1988, Leech, Thomson, et al. 1981, Possingham and Lawrence 1983). It is possible that a thylakoid segregation mechanism 12 might be integrated into the plastidic Min system, but work on this aspect of plastid division remains to be explored. Assembly and stabilization of the Z-ring: ARC 6 The plastidic Min system only creates a zone conducive to F tsZ polymer assembly, but mounting evidence suggests that additional factors are required to: (1) stabilize the Z-ring by promoting protofilament bundling and (2) anchor FtsZ polymers to the membrane. In E. coli, two proteins serve this purpose: ZipA and FtsA. Both ZipA and FtsA proteins elicit their effect upon the Z-ring by binding to a short conserved motif near the C-terminus of FtsZ called the core motif or core domain (Pichoff and Lutkenhaus 2002). ZipA is a bitopic membrane protein of the plasma membrane thought to enhance bundling of FtsZ filaments near the membrane (Hale, Rhee, et al. 2000, RayChaudhuri 1999). FtsA is a peripheral membrane protein that has been shown to anchor the Z-ring to membranes both in vivo and in vitro (Jensen, Thompson, et al. 2005, Osawa, Anderson, et al. 2008, Pichoff and Lutkenhaus 2005). FtsA may also be involved in protofilament bundling, as some functional redundancy between FtsA and ZipA is evident in E. coli (Geissler, Elraheb, et al. 2003). Two other division factors are known to bind the core motif in some gram positive organisms, Eer (Singh, Makde, et al. 2007) and ZapA (Low, Moncrieffe, et al. 2004); these proteins bind the C-terminus of FtsZ and somehow modulate FtsZ dynamics, but their precise effects upon FtsZ polymerization are still under investigation. Interestingly, engineering the short membrane-tethering amphipathic helix of F tsA onto the C-terrninus of FtsZ is sufficient to drive Z-ring formation along artificial membranes in vitro (Osawa, Anderson, et al. 13 2008). Cyanobacteria, green algae, and land plants do not encode orthologs of ZipA or F tsA, but do encode FtsZ molecules that bear the conserved C-terminal core motif (Miyagishima, Nishida, et al. 2003b). This suggests that other factors, possibly with an amino acid sequence dissimilar from FtsA and ZipA, might provide a similar FtsZ tethering and/or bundling function within plastids. ARC6 was discovered by a forward genetic approach and encodes a protein product that stabilizes the plastidic Z-ring, possibly by promoting FtsZ filament formation (V itha, Froehlich, et al. 2003). ARC6 may also provide both FtsZ bundling and tethering functions, despite the divergence of its sequence from ZipA or FtsA; ARC6 overexpressors have elongated FtsZ filaments and arc6 mutants have short, disorganized FtsZ filaments (Figure 1.3) (V itha, Froehlich, et al. 2003). ARC6 is a nuclear-encoded bitopic inner envelope protein of cyanobacterial origin (V itha, F roehlich, et al. 2003); the cyanobacterial ortholog of ARC6 is called Ftn2 (Koksharova and Wolk 2002). The Arabidopsis ARC6 protein is 801 amino acids in length (Figure 1.4), bears an N-terminal transit peptide that directs the protein to the chloroplast, and a putative J -domain that may facilitate interaction with Hsp70/DnaK- type chaperones. ARC6 family members also harbor a single transmembrane domain that anchors the protein in the inner envelope membrane (V itha, Froehlich, et al. 2003). Like the F tsZ proteins, ARC6 forms a continuous (non-punctate) ring at the division site (Yang, Glynn, et al. 2008); this ring co—localizes with the plastidic Z-ring (Figure 1.5). 14 Figure 1.3. F tsZ2-1 immunolocalization in wild type Arabidopsis chloroplasts and in arc6 mutants. (A) FtsZ forms equatorial rings (large arrowhead) in wild type chloroplasts and (B) forms short fragments or spots (small arrowheads) in arc6 mutants, suggesting that ARC6 stabilizes the Z-ring in vivo. Chlorophyll autofluorescence is indicated in red. Scale bar = 5pm. 15 N-terminal Conserved Region C-terminal Conserved Region / TP JD TM/v Figure 1.4. Schematic of the Arabidopsis ARC6 protein. The above panel highlights the known features of ARC6 based on previous data (Vitha, Froehlich, et al. 2003). The N-terminus of ARC6 resides in the chloroplast stroma and the C-terminus of the protein resides in the intermembrane space. The N-terminal (amino acids 86-509) and C- terminal (amino acids 683-793) conserved regions show very high similarity to other ARC6/Ftn2 family members. Transit peptide (TP, amino acids 1-67); Predicted J- domain (JD, amino acids 89-153); and Transmembrane Domain (TM, amino acids 615- 635) 16 Figure 1.5. FtsZ and ARC6 colocalize to continuous rings within the chloroplast. Immunolabeling of FtsZZ-l (A, green) and ARC6 (B, bright red color) is shown. Chlorophyll autofluorescence (B, dull red color) roughly marks the boundaries of the chloroplast. Colocalization of the two proteins (arrowhead) is indicated in panel (C) using an overlaid image from panels (A) and (B). Scale bar = 5 pm. The N-terminus of ARC6 resides within the stroma, where it interacts with FtsZ2 family members; this interaction requires the C-terminus of FtsZZ (Maple, Aldridge, et al. 2005, Schmitz, Glynn, et al. 2009). Presumably, the ARC6-FtsZ2 interaction is analogous to that observed for FtsA-FtsZ and/or ZipA-FtsZ in E. coli (Pichoff and Lutkenhaus 2002), but an FtsZ-tethering or FtsZ-bundling activity for ARC6 has yet to be shown. In chapter 2 of this work, I describe refinement of the boundaries of the stroma] domain of ARC6 responsible for binding FtsZZ, show that the C-terminus of F tsZZ is both necessary and sufficient for binding ARC6, attempt to discern the effect of ARC6- FtsZ2 interaction in vivo, and provide a homology model of ARC6 that points toward structural and functional similarities between ARC6 and ZipA, despite their divergent amino acid sequence. The function of the IMS-localized C-terminal region of ARC6 was completely unknown until a recent study that showed the C-terminus of ARC6 binds to PDV2 and is responsible for positioning both PDVl and PDV2 (described below) at the division site (Glynn, Froehlich, et al. 2008). In chapter 3, I present these results and some additional data in support of this function for the C-terminus of ARC6. Chapter 4 builds on these findings, where I present approximation of the boundaries of the PDV2-binding domain of ARC6 (ARC6pBD), characterize a predicted site of post-translational modification within this domain, and introduce a new hypothesis on the regulation of division by ARC6pBD phosphorylation and dephosphorylation. 18 ARC6/Ftn2 family members are present as single-copy genes in cyanobacteria, algae, and moss. In tracheophytes, ARC6 has undergone at least one gene duplication event, resulting in a novel plastid division gene we have termed PARC6 (Paralog of ARC6). Like ARC6, PARC6 is a multifunctional division protein that resides within the inner envelope membrane, acting downstream of ARC6 to position PDV] and possibly regulating FtsZ assembly through ARC3 — implicating PARC6 in the operation of the plastidic Min system (Glynn, Yang, et al. 2009, Zhang, Hu, et al. 2009b). In chapter 5, I summarize this work and provide further insights into the role of ARC6, PARC6, PDVI , and PDV2 during plastid division in vascular plants. While it is unknown if ARC6 and PARC6 definitively interact with each other or how they precisely influence each other’s activity in vivo, it is suspected that some functional connection between these two paralogous proteins exists due to their high degree of sequence similarity. In chapter 6, I highlight preliminary analysis of a novel ARC6 allele, arc6 D 20 5 N, that exhibits division defects reminiscent of Arabidopsis min mutants and propose a hypothesis that the arc6 D 20 5N division defect results from aberrations in the ability of ARC6 to interact with a Min-system component. Facilitating outer envelope constriction during division: ARC 5, PD VI, and PD V2 In addition to the division factors inherited from the ancestor of the chloroplast, a few factors were invented by the host following symbiosis. Notably, three of these host- derived factors operate within or upon the outer envelope membrane. ARC5 is a dynamin-like protein that localizes to both cytosolic patches and an equatorial ring on the 19 outside of the plastid. Unlike the continuous ARC6 ring (Glynn, Froehlich, et‘ al. 2008, Vitha, F roehlich, et al. 2003, Yang, Glynn, et al. 2008), the ARC5 ring has a discontinuous, or punctate, appearance in Arabidopsis and other photosynthetic eukaryotes (Gao, Kadirjan-Kalbach, et al. 2003, Miyagishima, Froehlich, et al. 2006, Miyagishima, Nishida, et al. 2003a, Yang, Glynn, et al. 2008). The significance of this punctate pattern is still unclear. It is anticipated that these foci represent discrete sites of membrane remodeling rather than a bounding contractile spiral (which would likely appear as a continuous ring in vivo), which suggests that initial constriction of the outer membrane might invove membrane removal from discrete foci rather than direct mechanical constriction of the outer envelope by ARC5. Notably, the late stages of dynamin constriction activity involve a significant change in the organization of ARC5 topology relative to the outer envelope membrane, in which ARC5 polymers come into direct contact with the cytosolic face of the outer envelope (Yoshida, Kuroiwa, et al. 2006). This end-stage change in orientation may reflect a simultaneous change in the pinchase activity of ARC5 polymers, but further investigation is required to confirm this hypothesis. ARC5 can be recruited to the division site by either PDVl or PDV2, two coiled- coil proteins that arose during land plant evolution Miyagishima, Froehlich, et al. 2006). Like ARC5, PDVl localizes to a punctate ring (Miyagishima, Froehlich, et al. 2006). PDV2 localizes to a continuous ring similar to that observed for ARC6 (see chapter 3). Interestingly, full ARC5 contractile activity requires both PDV] and PDV2, as pdvI and pdv2 mutants possess a large proportion of dumbbell-shaped chloroplasts, similar to the 20 dynamin activity defect observed in arc5 mutants (Miyagishima, Froehlich, et al. 2006). Overexpression of PDVl and/or PDV2 leads to an increase in the rate of chloroplast division in vivo (Okazaki, Kabeya, et al. 2009). Taken together, these observations indicate that the PDV proteins probably affect the rate of plastid division through modulating ARC5 contractile activity, although the PDV proteins probably influence the operation of other division components as well. It is still unclear if either of the PDV proteins interacts directly with ARC5 to mediate its recruitment and/or activity along the cytosolic surface of the chloroplast outer envelope. Other factors with notable roles in maintaining chloroplast morphology and/or division. A number of other factors have been discovered through genetic and cytological analyses. The precise role of these proteins in the process of chloroplast division remains unclear, but we will provide a brief treatment of each of these here. GCl (Giant Chloroplast 1) is distantly related to E. coli SulA and was identified by a reverse genetic approach (Maple, Fujiwara, et al. 2004). In proteobacteria, SulA expression is part of the SOS response to DNA damage (Huisman, D'Ari, et al. 1984) and delays cell division by directly binding FtsZ and inhibiting FtsZ assembly (Trusca, Scott, et al. 1998). For Arabidopsis, it is unclear if GCl binds Ftle or F tsZ2 in vivo; overexpression of GCI bypasses plastid division defects observed in AtFtsZ overexpressors (Raynaud, Cassier-Chauvat, et al. 2004), but strangely GCl does not bind F tle or FtsZZ in two-hybrid assays (Maple, Fujiwara, et al. 2004). However, GCl can self-associate (Maple, Fujiwara, et al. 2004), similar to SulA (Cordell, Robinson, et al. 21 2003). GCl is localized to the chloroplast periphery and has not been shown to interact with any other division components (Maple, Fujiwara, et al. 2004). While GCl deficiency results in enlarged chloroplasts, similar to arc6 mutants, it is unclear if GCl overexpression inhibits chloroplast division (Maple, Fujiwara, et al. 2004, Raynaud, Cassier-Chauvat, et al. 2004) in the same way that SulA overexpression leads to division arrest in cyanobacteria (Sakr, Jeanjean, et al. 2006). Further work is required to understand the inputs that lead to GC 1 activity and the precise mechanism by which GCl affects plastid morphology in vivo. MSL2 and MSL3 belong to a family of mechanosensitive ion channels that reside in the chloroplast (Haswell and Meyerowitz 2006). MSL2 and MSL3 appear to be redundant genes with respect to chloroplast morphology, as only mle msl3 double mutants exhibit plastid morphology defects (Haswell and Meyerowitz 2006). MSL3 has been shown to colocalize with AtMinE in transient expression assays conducted using tobacco leaf cells, but the significance of this finding remains unclear (Haswell and Meyerowitz 2006). It has been suggested that mechanosensitive ion channels, like MSL2 and MSL3, are a means by which chloroplasts sense volume, thereby influencing the decision to continue expanding or to initiate division (Pyke 2006). This seems reasonable, since chloroplast volume is only slightly reduced in division-impaired backgrounds like arc6 (Pyke 1998, Robertson, Pyke, et al. 1995). However, the precise mechanism by which MSL2 and MSL3 influence plastid morphology, and presumably division, remains to be elucidated. 22 F ZL (FZO-Like) was discovered based on its similarity to FZO (Gao, Sage, et al. 2006), a protein known to mediate mitochondrial membrane fusion in animals and fungi (Hales and Fuller 1997). FZL is chloroplast-targeted protein that localizes to the envelope and thylakoid membranes (Gao, Sage, et al. 2006). Unlike the other characterized members of the dynamin superfamily within plants, FZL is suspected to be a membrane remodeling GTPase that regulates organization of the thylakoid membranes, as le mutants have thylakoid morphology defects and possess aberrant vesicular inclusions within the stroma (Gao, Sage, et al. 2006). le mutants also show defects in chloroplast morphology —— possibly implicating the organization or position of the thylakoid membranes in division (Gao, Sage, et al. 2006). Curiously, FZL-GFP localizes to spots or vesicles around the periphery of the chloroplast, which seems inconsistent with its role in thylakoid biogenesis. However, it has recently been shown that tgd4 act] double mutants have enlarged plastids with thylakoid defects similar to le mutants, perhaps as a result of the loss of both sources of membrane material that partly compose the thylakoid membrane in plants (Xu, Fan, et al. 2008). Based on its mutant phenotype and localization, we suspect that FZL might be part of a vesicular transport system that aids trafficking of membrane material from outside the chloroplast (i.e. the TGD4 pathway) to the thylakoid membrane (Xu, Fan, et al. 2008), which suggests that the chloroplast morphology phenotype observed within fil mutants is a pleiotropic effect, but further work is required to rigorously test this hypothesis. Regardless, the impact of thylakoid morphology upon the process of chloroplast division is interesting and should be investigated further. 23 Some factors that make up the chloroplast divisome remain to be identified. The Plastid Dividing (PD) rings are electron-dense structures that appear after the Z-ring is established; these structures are dynamic and may be involved in mediating constriction (Miyagishima, Kuroiwa, et al. 2001). While the consituitve parts of the PD rings are unknown, the Kuroiwa laboratory is in the process of dissecting these rings and identifying their constituent parts using a synchronizable red algae, Cyanidioschyzon merolae (Yoshida, Kuroiwa, et al. 2006). It is thought that land plants will possess the same components, as the timing, size, and dynamics of the PD rings in land plants are similar to those observed in the red algae (Kuroiwa, Kuroiwa, et al. 1998). While several factors have been introduced here, much work is still required to identify the complete function of each component and the pathways that feed into their expression and activity. Furthermore, several factors are likely to be uncovered in the coming years using comparative genomics, biochemical, and genetic approaches. Throughout this text, I will attempt to overtly indicate where pieces are missing to reiterate this idea, as a final model must account for all the components. Regulating Plastid Fission: Why Do Plastids Divide in Land Plants? It is currently theorized that at least four major inputs feed into the regulation of plastid fission in higher plants: photosynthetic and photoprotective activity, metabolite transport, cellular division, and cellular differentiation. Presumably, each of these inputs 24 contributes to the fitness of a particular organism. We will briefly review each of these here. Photocollection and Photoprotection The viability of chloroplast division mutants with large chloroplasts raises the question of why plant cells evolved to have multiple chloroplasts. The answer may have to do partly with the ability of multiple small chloroplasts to redistribute more effectively than fewer large chloroplasts inside the cell. Unlike motile unicellular algae with single chloroplasts, land plants are sessile and cannot relocate in response to sudden changes in light intensity. Chloroplasts in land plants avoid potentially damaging high light by moving to the cell periphery and orienting in columns parallel to the plane of incoming light; under low-light conditions, chloroplasts gather in periclinal layers to maximize light absorption (Kasahara, Kagawa, et al. 2002). This movement is believed to be largely performed through actin reorganization and myosin motor activity (Paves and Truve 2007, Schmidt von Braun and Schleiff 2008). Consistent with the importance of plastid division for efficient plastid movement, a mutant allele of Arabidopsis Ftle shows diminished plastid movement in response to high light (Yoder, Kadirj an-Kalbach, et al. 2007). Interestingly, the thylakoid morphology of big-plastid mutants is reminiscent of that in high light-adapted plants (Austin 11 and Webber 2005), and probably arises due to the impaired movement ability of the enlarged chloroplasts within division-defective backgrounds (Jeong, Park, et al. 2002). As discussed earlier, mutations in genes influencing thylakoid organization also impair chloroplast division (Gao, Sage, et al. 2006, Xu, Fan, et al. 2008). The regulation of thylakoid morphology 25 and chloroplast movement is probably incorporated into a larger system that maximizes photosynthetic output and minimizes light-induced damage in response to changing environmental stimuli (Suetsugu and Wada 2007, Wada, Kagawa, et al. 2003). Metabolic Exchange: The Possible Effect of Organelle Area-to- Volume Ratio In addition to the potential role of photosynthesis in regulating plastid division, the shuttling of reduced carbon and metabolites between the plastid and the cytosol might also influence the division process. Presumably, an optimal ratio of organelle surface area to internal volume is maintained in chloroplasts. A larger plastid, such as those in the arc6 background, has a disproportionate surface to volume ratio relative to smaller plastids (Pyke and Leech 1992, Pyke and Leech 1994, Pyke, Rutherford, et al. 1994), and as a result may have insufficient transporter capabilities. Maintaining large numbers of small plastids may allow for greater membrane surface area and prevent bottlenecks of metabolic flux between the plastid and the cytosol or between the plastid and other organelles. However, investigation of this potential regulatory input into the division process has not been explored. Cell Division and Cell Expansion Chloroplast division is intimately coupled to cell expansion and cellular division in algae, and to some degree in plants. In dividing meristematic cells and juvenile leaf cells, the pace of plastid division must keep pace with cellular division to ensure that each resulting cell has one of these essential organelles (Boffey and Lloyd 1988, Possingham 26 and Lawrence 1983), though aplastidic cells can reside within the leaves of plants (Chen, Asano, et al. 2009, Forth and Pyke 2006, Pyke, Rutherford, et al. 1994). At least one factor, CDTl , has been implicated in coordinating cell division with plastid division in plant cells (Raynaud, Perennes, et al. 2005). CDTl proteins are known to be part of the pre-replication complex, required for origin licensing during nuclear DNA replication in eukaryotes (DePamphilis 2003, Thomer, May, et al. 2004), including Arabidopsis (Castellano, Boniotti, et al. 2004, Masuda, Ramos, et al. 2004). However, in Arabidopsis, RN Ai knockdown lines that diminish expression of CDTl paralogs have profound defects in chloroplast morphology and CDTl has been shown to interact with ARC6 (Raynaud, Perennes, et al. 2005). The precise details of how CDTl impacts chloroplast division are unknown, but the dual function of CDTl proteins in nuclear DNA replication and plastid division should be examined further. As cells expand, the number and/or volume of chloroplasts increases proportionately in mesophyll cells, so that ~70% of the cell volume is consistently occupied by chloroplasts, even in division-impaired mutants (Pyke 1998, Robertson, Pyke, et al. 1995). Cell expansion is part of leaf cell development in plants and is influenced by brassinosteroids (Hu, Poh, et al. 2006, Nakaya, Tsukaya, et al. 2002), auxin (Chen, Shimomura, et al. 2001, Jones, Irn, et al. 1998), and cytokinin (Baskin, Cork, et al. 1995, Beemster and Baskin 2000). Of these, only cytokinins have been shown to influence plastid differentiation (Reski, Wehe, et al. 1991) and plastid division (Okazaki, Kabeya, et al. 2009). In Arabidopsis, application of exogenous cytokinin or overexpression of CRF 2 (Cytokinin-Responsive Factor 2) leads to an increased rate of 27 plastid division as a result of increased expression of PDV] and PDV2 (Okazaki, Kabeya, et al. 2009), two factors involved in modulating dynamin (ARC5) recruitment and activity (Miyagishima, Froehlich, et al. 2006). Strangely, while cytokinin does upregulate the rate of chloroplast division through the PDV proteins, expression levels of other division factors such as FtsZ or ARC6 are unchanged in response to cytokinin (Okazaki, Kabeya, et al. 2009) — suggesting that the protein level of PDVl and PDV2 can stimulate the activity of both cytosolic and stromal division factors from their position within the outer envelope membrane (Okazaki, Kabeya, et al. 2009). While it is clear that cell expansion and cytokinin certainly influence plastid differentiation and division, the full scope of the gene regulatory network that controls plastid fission remains to be determined. Controls on Division During Cellular Development and Difikrentiation In addition to binary fission of chloroplasts within leaf cells, plastids can undergo developmentally-regulated division and differentiation during plant development. For example, plastid differentiation during fruit development in tomato occurs through a unique budding/fragmentation mechanism where even enlarged chloroplasts are capable of differentiating into normal chromoplasts during fruit ripening (Forth and Pyke 2006). The bypass of the block in plastid division within sufl'ulta mutants that occurs during tomato fruit development is not observed with the enlarged chloroplasts of ftsZ, arc6, or arc12 mutants (although Arabidopsis itself does not possess chromoplasts); once an oversized chloroplast is generated, it is presumed to have lost its polarity and can no longer divide using an FtsZ-based mechanism (Schmitz, Glynn, et al. 2009). It is 28 possible that the pathway that produces stromules, stroma-containing extrusions of the plastid, might be involved in these fission events, as stromules are generated even in the absence of a ftmctional Z-ring (Holzinger, Kwok, et al. 2008). The molecular identification of suflulta and the study of plastid fission in the big-plastid backgrounds (i.e. arc6, arc12, and others) will likely provide critical insights into FtsZ-independent pathways of plastid partitioning. Perspectives and Acknowledgements Previous to this body of work, several issues in the field of plastid division were unclear. Within the following chapters, I attempt to address at least three major questions: (1) what discrete region of ARC6 binds FtsZ2 family members and what is the purpose of this interaction in vivo? (2) How does ARC6 participate in the coordination of the stromal and cytosolic plastid division machineries? (3) What is the function of Arabidopsis ARC6 paralog (PARC6) and how might PARC6 and ARC6 link into the plastidic Min system? I follow these chapters with some general conclusions and suggested future directions for research. Additionally, this volume contains two appendices that highlight peripheral projects undertaken during the last few years. I hope it is a clear and fascinating read. Please note that portions of this chapter are reproduced from (Glynn, Miyagishima, et al. 2007) (http://wwwiafficdkn and is Copyright Blackwell Publishing. Blackwell grants authors the freedom to reuse their own articles in new publications, provided that they are the editor of the new publication. 29 Chapter 2 Identification and Modeling of an FtsZZ—Binding Domain within ARC6 30 Abstract The plastidic Z-ring is not a static structure, but probably undergoes remodeling in response to various input signals. The regulation of FtsZ filament dynamics is an important part of remodeling the Z-ring during its assembly and constriction. ARC6 has been shown to be required for assembly and/or stabilization of the Z-ring in vivo. This stabilizing activity has been proposed to occur through F tsZZ family members, as ARC6 has been shown to bind FtsZZ, but not Ftle. However, the precise effect of ARC6 upon Z-ring assembly through its interaction with FtsZ2 has not been demonstrated experimentally, nor have the exact boundaries of the FtsZ2-binding domain (ZBD) of ARC6 been identified. To provide insight into the function of ARC6ZBD, we fine- mapped the boundaries of this domain using two hybrid assays and used this information to guide analysis of the in vivo function of ARC6ZBD. Our results indicate that ARC6ZBD occupies amino acids 351-503. This domain is conserved amongst the Chlorophytes and cyanobacteria, but bears no sequence similarity to other characterized protein sequences. We generated a homology model of ARC6ZBD based on the crystal structure of the FtsZ-binding domain of E. coli ZipA, an FtsZ-bundling protein. ARC6ZBD was fit onto the known structure of ZipAZBD, despite the large degree of sequence divergence between ARC6ZBD and ZipAZBD, suggesting that ARC6ZBD might aid bundling of F tsZ filaments through a ZipA-like mechanism, but rigorous in vivo analysis and structural data is still needed to validate this model. 31 Introduction Plastids are essential organelles. In plant cells, binary fission is a mechanism by which each cell receives its required complement of plastids and likely is influenced by development, environmental cues, and cellular metabolism (Boffey and Lloyd 1988, Leech, Thomson, et al. 1981, Possingham and Lawrence 1983). Plastids evolved from a cyanobacterium that took up residence within a primitive protozoan (Cavalier-Smith 2000). Like many free-living cyanobacteria and proteobacteria, chloroplasts divide by binary fission in a process that requires FtsZ (Osteryoung, Stokes, et al. 1998, Strepp, Scholz, et al. 1998, Vitha, McAndrew, et al. 2001). FtsZ is a filament-forming GTPase with structural similarity to tubulin (Erickson, Taylor, et al. 1996, Mukherjee, Dai, et al. 1993). E. coli FtsZ forms protofilaments that undergo organized assembly into a structure at the mid-cell called the Z-ring, which acts as a dynamic scaffold for other division components (Margolin 2005) and probably generates the force that contributes to cellular constriction (Erickson 2009, Li, Trimble, et al. 2007, Osawa, Anderson, et al. 2008). Several systems contribute to controlling both FtsZ protofilament assembly and Z-ring assembly. The Min system, composed of MinCDE, prevents protofilament assembly at the poles, but allows bulk protofilament assembly to occur at the mid-cell, ensuring that the cell divides near its midpoint (Lutkenhaus 2007). ZipA is a bitopic membrane protein that binds a conserved C-terminal segment of FtsZ in E. coli and is 32 thought to be important for protofilament bundling and organization adjacent to the inner leaflet of the cell membrane at the midcell (Hale, Rhee, et al. 2000, Mosyak, Zhang et al. 2000, RayChaudhuri 1999). The conserved C-terminal motif of bacterial FtsZ consists of ~12 amino acids and is required for FtsZ functidn (Ma and Margolin 1999). Besides ZipA, the conserved C-terminus of FtsZ is a target for FtsA, another protein involved in cell division (Pichoff and Lutkenhaus 2002, Pichoff and Lutkenhaus 2005, Pichoff and Lutkenhaus 2007, Ricard and Hirota 1973). FtsA is a membrane-associated protein that acts as a membrane anchor for the Z-ring (Osawa, Anderson, et al. 2008, Pichoff and Lutkenhaus 2005) and binds the C-terminal tail of FtsZ through a domain that is structurally distinct from the F tsZ-binding domain of ZipA (Mosyak, Zhang, et al. 2000, van den Ent and Lowe 2000). While the coordinated action of MinCDE, ZipA, and F tsA contribute to FtsZ ring formation in proteobacteria, the precise roles of proteins influencing assembly and dynamics of the Z-ring in plastids is still somewhat unclear. In chloroplasts, FtsZ assembles into ring-shaped structures near the middle of the organelle, analogous to the bacterial protein (V itha, McAndrew, et al. 2001). In the case of plants, however, FtsZ gene duplication and divergence have generated two distinct families of FtsZ protein (termed F tle and FtsZ2), both of which are required for chloroplast division (Osteryoung, Stokes, et al. 1998, Schmitz, Glynn, et al. 2009). The most notable difference between these two families occurs at their C-terrnini, where FtsZZ family members possess a C-terminal motif similar to the ZipA- and FtsA- binding motif of bacterial FtsZ proteins; this motif is absent from plant F tle family members (Miyagishima, Nishida, et al. 2003b, Stokes and Osteryoung 2003). 33 While ZipA is an essential gene in some bacteria, orthologs of ZipA are absent from cyanobacteria and plants (Miyagishima, Nishida, et al. 2003b). However, we have hypothesized the existence of an FtsA-like or ZipA-like protein in these lineages because FtsZZ family members have retained the conserved C-terminal extension present in most bacterial F tsZ proteins. Despite the inability to identify a ZipA-like protein in plants by BLAST (Altschul, Madden, et al. 1997), analysis of ARC6 suggests that it may fulfill a similar functional role within the plastid. ARC6 promotes Z-ring assembly, possibly by bundling FtsZ filaments within the plastid since loss-of-function arc6 mutants have short F tsZ filaments distributed throughout the stroma and transgenic lines overexpressing ARC6 have elongated F tsZ filaments, some of which are spiraled or branched (V itha, Froehlich, et al. 2003). A previous report revealed an interaction between ARC6 and FtsZZ-l that requires the conserved C-terrninal extension of F tsZ2-l and proposed a functional role analogous to that of ZipA or FtsA in E. coli; although the precise boundaries for the FtsZ2 binding domain were not identified in this study (Maple, Aldridge, et al. 2005). The results we show here define the FtsZ2 binding domain (ZBD) of Arabidopsis ARC6 and show that the ARC6ZBD shares features with the FtsZ-binding domain of E. coli ZipA (Mosyak, Zhang, et al. 2000). We then used site-directed mutagenesis to identify critical residues within the ARC6ZBD that are probably sites of intermolecular contact between ARC6 and FtsZZ based on our structural model. Further, we introduced transgenes into Arabidopsis that are predicted ARC6ZBD loss-of-function or gain-of- 34 function alleles based on our protein interaction analyses in yeast. Unfortunately, these transgenes were uninformative with regard to ARC6ZBD function in vivo and further analysis will be required to validate the boundaries of ARC6ZBD and provide insight into its structure and function in vivo. Results The F tsZZ—binding domain of ARC 6 (ARC 6231)) occupies amino acids 351-503. Previously, our colleagues discovered that a portion of the conserved stromal region of ARC6 (AA 154-509) specifically bound FtsZZ-l in two-hybrid experiments (Maple, Aldridge, et al. 2005, Maple and Moller 2006). Recently, our laboratory showed that both Arabidopsis FtsZ2 isoforrns (FtsZ2—1 and FtsZ2-2) bind the same region of ARC6 with similar affinity and are functionally redundant (Schmitz, Glynn, et al. 2009); therefore, we used only FtsZZ-l (hereforth FtsZZ) in our analyses for simplicity. To delineate the precise boundaries of the ARC6ZBD, we used a two-hybrid- based domain-mapping approach, using exon junctions as the boundaries for our initial constructs (Figure 2.1) since critical domains might be less likely to be interrupted by introns during genome evolution (de Souza, Long, et al. 1996, Fodor and Aldrich 2009). The conserved stromal region of ARC6 is encoded within four exons. We generated 9 constructs representing different portions of ARC6 and assayed these for interaction with FtsZZ. The results of these assays indicate that ARC6ZBD is contained within AA 351- 35 TP JD TM FtsZZ Interaction AA 154-509 _— n. AA 170-509 i AA 245-509 E AA 332-509 AA 154-331 1‘ - — AA 344-509 — +++ AA 351-503 i AA 358-509 AA 344-490 Figure 2.1. Mapping the FtsZ2-binding domain of ARC6 by yeast two-hybrid assay. Constructs shown are GAL4-AD fusions of ARC6, with FtsZZ-l present as a GAL4-BD fiision. Interaction strength is reported as the ratio of growth (-HIS : +HIS) of a transformant on supplemented synthetic dropout media (+—++ growth ratio is greater than or equal to 0.67; - growth ratio is less than or equal to negative control). The minimal portion of ARC6 required for interaction with FtsZZ is boxed. Transit peptide (TP); J - domain (JD); transmembrane domain (TM); N-terminus (N); and C-terminus (C). 36 503 (Figure 2.1); the degree of reporter activation from AA 351-503 was similar to that of the original boundaries described previously (Maple, Aldridge, et al. 2005). The C-terminus of F tsZZ is sufi’icient for ARC 6-F tsZZ interaction. Previously it was shown that the C-terminus of FtsZ2 is necessary for ARC6- FtsZZ interaction (Maple, Aldridge, et al. 2005), but it was unclear if this short C- terrninal motif was sufficient for ARC6-FtsZ2 interaction, or if other features within FtsZZ might be required for ARC6-FtsZZ interaction. To address this question, we generated two-hybrid constructs encoding the C-terminal 19 amino acids of Arabid0psis FtsZZ as a GAL4 fusion and transformed them into yeast carrying plasmids encoding either ARC6154-509 01' ARC6ZBD. Our results indicate that the C-terminus of F tSZZ is sufficient for ARC6-FtsZ2 interaction, as GAL4 fusions to both the mature FtsZZ protein and 19 amino acid motif at the C-terminus of FtsZZ (F tsZ2459_473) generated comparably high levels of reporter activation (Figure 2.2), though the presence of the full-length FtsZZ protein was associated with slightly higher levels of HIS3 reporter activation. To gain insight into the structure and function of the ARC6ZBD, we wanted to generate a testable structural model of this protein domain. There is currently no experimentally-tested structural data available for ARC6, so our initial approaches involved the use of BLAST-based sequence queries (Altschul, Madden, et al. 1997) to identify similar domains with solved crystal or NMR structures. Using this approach, we were unable to identify similar sequences with solved three-dimensional structures. 37 INTERACTION PAIRING SD/-ULT SD/-ULTH ARC6154-509 VS. FI82249-478 ARC6351-503 VS~ Ftszz49—478 ARC6154-509 VS- FtSZ7-459-478 ARC6351-503 VS- FtSZ2459-478 ARC6154-509 VS- FtSZ249-462 ARC6351_503 vs. FtsZZ49_462 Empty vector vs. FtsZ249-473 Empty vector vs. FtsZZ459_478 Empty vector vs. FtsZ249_462 ARC6154_ 509 vs. Empty vector ARC6351_503 vs. Empty vector - Empty vector vs. Empty vector 1/10 1/100 1/10 1/100 Figure 2.2. Two-hybrid HIS reporter assays showing ARC6-FtsZ2 interaction. Yeast harboring pairings of AD-ARC6 and BD-FtsZZ are shown following 48 hours of growth on the synthetic dropout (SD) media indicated. ARC6-FtsZ2 interaction requires only ARC6351_503 (ARC6ZBD) and F tsZ2459_47g. Removal of the C-terminus (AA463- 478) of FtsZ2 abolishes ARC6-FtsZ2 interaction. Dilutions from a stationary phase culture at OD600 = 1.0 are indicated 1, 1/10, and 1/ 100. ARC 6231) can be modeled onto the crystal structure of ZipA ZBD Because ARC6ZBD binds the short conserved C-terminal motif of F tsZ2 family members and this motif shares conservation with an analogous motif found at the C- terrninus of bacterial FtsZ sequences (hereforth referred to FtsZCT), we compared the ARC6ZBD amino acid sequence to sequences of bacterial proteins known to bind F tsZCT, as some of these have solved crystal structures. At least three proteins bind bacterial FtsZCT: F tsA (Pichoff and Lutkenhaus 2007), ZipA (Hale, Rhee, et al. 2000), and Eer (Singh, Makde, et al. 2007). The F tsZ-binding domain of FtsA is at least 288 amino acids in length (Pichoff and Lutkenhaus 2007) and there is no crystal structure available for Eer. The F tsZ-binding domain of ZipA (ZipAZBD) has a solved structure (Mosyak, Zhang, et al. 2000) and is similar in length (144 amino acids) to the ARC6ZBD (153 amino acids) and was chosen to generate a structural model of ARC6ZBD. We generated a sequence alignment between ARC6ZBD and ZipAZBD to identify regions of similarity between these two domains (Figure 2.3). The homology-modeling server at http://proteins.msu.edu/ was then used to superimpose the amino acid sequence of ARC6ZBD onto the main chain structure of ZipAZBD (Protein Data Bank Identification Ntunber 1F47) (Mosyak, Zhang, et al. 2000), using the identical residues (Figure 2.3) between the two sequences as anchor points. Following this, we modeled the sidechains of ARC6ZBD using SCWRL (Wang, Canutescu, et al. 2008) to get the lowest energy conformation of ARC6ZBD with minimal conflict. 39 ARC6ZBD EEALALVAQSEGKKEHLLOEE QFQQLQ ZipAZBD . DKPKRK IMN GSELNG ARcézBD M EIPA L EIDFG RGL L ZipAZBD NS QQAGFE HRHLSP GSGP F AncszBD IG DECRMWLG ELEN ZipAZBD s. . ANMVKPGT MIS vector-only negative control (+); and growth ratio < negative control (-). 45 (Moreira, Femandes, et al. 2006) and suggest that F tsZ2 might bind ARC6ZBD within a hydrophobic pocket, similar to ZipA-FtsZ interaction (Moreira, F emandes, et al. 2006, Mosyak, Zhang, et al. 2000). From this analysis, I conclude that the structures of the ARC6ZBD and ZipAZBD might be similar, with the C-terminus of FtsZ2 binding to ARC6ZBD within a hydrophobic pocket, but further analysis of ARC6ZBD site-directed mutants and real structural analyses (i.e. X-ray crystallography or NMR) are required to fully test this structural model. Analysis of the ARC 6 ZBD‘F tsZ2 Interaction by Pulldown Assay. To confirm our two-hybrid results, I attempted to use a pulldown approach to compare the Ftszz-binding abilities of ARC6ZBD and ARC6ZBD (F442D)- I expressed the C-terminus of F tsZZ (AA 459-478) as a C-terminal fusion to GST (GST-FtsZ2CT). Both ARC6ZBD and ARC6ZBD (F442D) were expressed with a polyhistidine tag at their N-termini (Hi83-ARC6ZBD and Hi83-ARC6ZBD (F442D) respectively). All fusions were expressed under control of an IPTG-inducible promoter in E. coli. After immobilizing and purifying the polyhistidine-ARC6 fusion proteins using Ni-sepharose beads, I treated Hiss-ARC6ZBD or Hiss-ARC6ZBD (124429) with clarified cell extracts fi'om lines expressing GST-FtsZZCT. After a series of washes, bound protein was eluted from the beads using concentrated imidazole the eluates were probed for GST-FtsZZC—r. In preliminary results, I observed similar GST-FtsZZC-r binding characteristics for both 46 Hiss-ARC6ZBD and Hiss-ARC6ZBD(F4421)) (Figure 2.7A). I repeated this protocol with more stringent wash conditions and observed that neither ARC6 fusion was able to retain GST-ZZCT (not shown). To gain insight into the root cause of this result, I performed a pulldown between Hiss-ARC6ZBD and either GST-FtsZZCT or GST-FtsZ2CT (F466 A); the FtsZ2F466 A mutation was previously shown to abolish interaction between ARC6 and F tsZZ using a two-hybrid based approach (Maple, Aldridge, et al. 2005). Surprisingly, both FtsZZCT and F tsZZCT(p466 A) interact with ARC6ZBD using pulldowns (Figure 2.7B). Curiously, the results of the in vitro pulldowns between site-directed mutants of ARC6 and FtsZ2 (Figure 2.7) are inconsistent with two-hybrid assays (Maple, Aldridge, et al. 2005) (Figure 2.6), leading us to question the overall importance of ARC6F442 and FtSZZF466 in vivo. In vivo analysis of the ARC6 F44 2 D site-directed mutation. Because ARC6F442 was shown to be critical for ARC6-FtsZZ interaction in yeast, but not for in vitro interaction, I aimed to test one of these results in vivo and clarify the role of ARC6ZBD by analyzing a prospective ZBD loss-of-ftmction mutant. To do this, I generated transgenes encoding the native ARC6 protein and an ARC6F44213 mutant protein. Both of these transgenes were placed under control of the native ARC6 promoter (V itha, Froehlich, et al. 2003). Constructs were introduced into Col-0 and arc6 (SAIL_693_G04) backgrounds by Agrobacterium-mediated 47 A 1 2 3 B 1 2 3 a-HIS a-GST mam-m air-a, - - a-GST a—GST Figure 2.7. Pulldown of FtsZZ by ARC6. (A) Anti-HIS inununoblot (upper panel) showing eluted bait protein and anti-GST (lower panel) showing eluted GST-FtsZZC-r fusions. Pulldown eluates from immobilized Hiss-ARC6ZBD treated with GST-FtsZ2C-r (lane 1); pulldown eluates from immobilized Hi83-ARC6ZBD(F44ZD) treated With GST- FtsZZCT (lane 2); and pulldown eluates from beads treated with GST-FtsZZCT extract. (B) Anti-GST immunoblot showing eluted GST—FtsZ2CT fusions (upper panel) and input proteins (lower panel). Pulldown eluates from immobilized Hiss-ARC6ZBD treated with GST-FtsZ2CT (upper panel, lane 1); pulldown eluates from Hiss-ARC6ZBD treated with GST-FtsZ2C1-(p466 A) (upper panel, lane 2); and pulldown eluate fi'om beads treated with only Hiss-ARC6ZBD (upper panel, lane 3). Pulldown inputs (0.5 pg total protein/lane) for GST-FtsZ2C-r (lower panel, lane 1); GST-FtsZ2C-r (p466 A) (lane 2); and Hiss- ARC6ZBD (lower panel, lane 3). The molecular weight of the Hiss-ARC6ZBD and Hiss- ARC6ZBD (F442D) fusions are ~19 kD. The molecular weight of GST-FtsZZCT and GST-FtsZZCT(F466 A) are ~28 kD. 48 transformation and transgenic individuals were selected using hygromycin. Once plants were approximately 5 weeks old, T1 individuals and controls were examined for ARC6 or ARC6 F44 2 D transgene expression by immunoblotting (Figure 2.8); only transformants from the SAIL_693_G04 background (arc6) were analyzed for ARC6 protein levels, as it would be impossible to differentiate the native and transgenic ARC6 protein products using our polyclonal ARC6 antibody. Transformed (T1) arc6 individuals with ARC6 F44 2 D transgene expression near the range of expression yielding wild type chloroplast phenotypes were quantitatively assayed for chloroplast number and morphology (Figure 2.9), though all T] lines were subjected to microsc0pic examination. Consistent with pulldown results that hinted that ARC6ZBD (F442D) mutant proteins might be fully functional (Figure 2.7A), the ARC6 F44 ZD transgene was able to complement the arc6 mutant (Figure 2.9D-E). Further, I observed very few defects in Col—0 plants transformed with ARC6 F44 2 D, (not shown) indicating that this transgene probably does not impart any dominant negative effect on chloroplast morphology, consistent with its ability to fully rescue an arc6 mutant (Figure 2.9). To determine if the ARC6 F44 2 D-expressing lines have subtle defects in FtsZ assembly or disassembly, we examined Z-ring morphology within chloroplasts of juvenile leaves (Figure 2.10). Consistent with the ability of the ARC 6 F44 2D transgene to complement the chloroplast morphology defect of arc6 mutants (Figure 2.9D-E), FtsZ 49 a-ARC6 195kD> 117kD> 97kD> 50 kD > Ponceau Figure 2.8. Demonstration of ARC6 and ARC6F442D transgene expression in the arc6 (SAIL_693_G04) background. Exposure from an anti-ARC6 immunoblot (top panel) and Ponceau stain of the corresponding membrane (lower panel). Whole cell extracts from controls and transformed arc6 mutants: molecular weight marker (lane 1); untransformed SAIL_693_G04 (lane 2); untransformed Col-0 (lane 3); ARC6 F44 2 D' expressing transgenic line #1 (lane 4); ARC 6 F44 2 D-expressing transgenic line #2 (lane 5); ARC 6-expressing transgenic line #1 (lane 6); ARC 6-expressing transgenic line #2 (lane 7); ARC6-expressing transgenic line #3 (lane 8) ; and ARC6-expressing transgenic line #4 (lane 9). Two milligrams of whole-cell extract from flower buds are loaded in each lane. 50 I... E Chloroplasts per cell Col-0 arc6 arc6 arc6 +ARC6 (T1) + F442D (T1) Figure 2.9. Quantitative analysis of chloroplast number in ARC6F442D-expressing transgenics. Micrographs of mesophyll cells from: Col-0 (A); arc6 (B); ARC6- expressing transgenic line #3 (C); and ARC6 F44 2D-expressing line #2 (D). Scale bars = 10 um. Quantitative analysis of chloroplast number within lines shown in A-D and Figure 2.8. Chloroplasts were counted using single-plane images from mesophyll cells with average areas of 3838 i 713 umz. Ten cells were analyzed for each line using leaf tip samples from plants ~5 weeks post-germination. Error bars = standard deviation from the mean. 51 A _ A Figure 2.10. F tsZ immunolocalization in ARC6F4420-expressing transgenics. FtsZ2- 1 localization (green) in wild type Col-0 (A); arc6 (B); ARC6-expressing transgenic line #3 (C); ARC 61:44 ZD-expressing line #2 (D). FtsZ localizes to equatorial rings in wild type chloroplasts (large arrowheads, panels A, C, and D) and disorganized fragments (small arrowheads, panel B) in arc6 loss-of—function mutants. Chlorophyll autofluorescence (red) marks the approximate boundaries of the chloroplast. Scale bars = 5 pm. 52 morphology within the chloroplasts of ARC6F44ZD-expressing lines (Figure 2.10D) was clearly different from that in untransformed arc6 mutants (Figure 2.10B) and was indistinguishable from that in wild type Col-0 (Figure 2.10A) and transgene- complemented arc6 mutants (Figure 2.10C). From these results, I conclude that ARC6F442 is either not involved in mediating ARC6-FtsZ2 interaction in vivo, or is only one of several residues that stabilize ARC6- FtsZ2 interaction in vivo. These results are congruent with the comparable FtsZ2-binding abilities of both ARC6ZBD and ARC6ZBD (F442D) in pulldown assays (Figure 2.7). However, these results are inconclusive with regard to the function of ARC6ZBD, as the prospective ARC6 F44 2 D transgene was fully functional and therefore did not provide further insight into the role of this proposed domain. In vivo analysis of chloroplast targeted ARC6 ZBD As a secondary approach to elucidating the function of the FtsZZ-binding domain of ARC6, I generated a novel expression vector that targets the desired protein to the chloroplast stroma using the transit peptide of RecA (Kohler, Cao, et al. 1997). In addition to the N-terminal chloroplast transit peptide of RecA, the vector also encodes a C-terminal EYFP to verify protein expression and localization to the chloroplast. We placed the coding sequence of ARC6ZBD between that of the RecA transit peptide and EYFP; this transgene is expressed under control of the CaMV 358 promoter (Figure 53 2.11). The transgene was introduced into Ws-2 and arc6-I backgrounds by Agrobacterium-mediated transformation. To assess localization of the ARC6ZBD-EYFP fusion, I examined mesophyll cells from young leaves of several lines by fluorescence microscopy (Figure 2.12A-D). I observed EYFP signals that overlay with chlorophyll autofluorescence in transformed Ws-2 and arc6-I backgrounds (Figures 2.12B and 2.12D), indicating that the ARC6ZBD-EYFP was successfully imported into the chloroplast. Unlike ARC6, which forms a mid-plastid ring, I only observed plastid- targeted ARC6ZBD-EYFP as a diffuse signal that was distributed throughout the chloroplast stroma (Figures 2.123 and 2.12D). Following examination by fluorescence microscopy, I verified transgene expression using an anti-YFP antibody. Protein extracts from some individuals within this T1 population included a cross-reacting protein of ~47 kD (Figure 2.12B), which corresponds to the molecular weight of imported (processed) ARC6ZBD-EYFP. To determine if the plastid-targeted ARC6ZBD-EYFP expressing lines have defects in FtsZ assembly, I examined Z-ring morphology within dividing chloroplasts of juvenile leaves (Figure 2.13, insets). In each case, FtsZ morphology was similar to untransformed plants. To determine if the plastid-targeted ARC6ZBD-EYFP had any gross effect upon chloroplast division, we examined fully expanded mesophyll cells from the same lines. Consistent with our observations of FtsZ morphology (Figure 2.13, insets), expression of ARC6ZBD-EYFP had no effect upon chloroplast morphology in Ws-2 nor arc6-I backgrounds (Figure 2.13A-D, large panels). 54 NcoI ApaI BgIII Xmal BstEII pCAMBIA-ZBD Figure 2.11. Schematic of plastid—targeting vector pCAMBIA-ZBD used for ARC6ZBD domain analysis. To gain insight into the role of the FtsZZ binding domain of ARC6, the coding sequence for ARC6ZBD was cloned into a modified plant transformation vector derived from pCAMBIA-l302 (Hajdukiewicz, Svab, et al. 1994). Expression of the integrated transgene is driven by the CaMV 35$ promoter (grey box, left side). The inserted coding sequence for the protein or domain of interest (light blue, ARC6ZBD) is directed to the chloroplast by the transit peptide of Arabidopsis RecA (coding sequence shown in light green) (Kohler, Cao, et al. 1997). Protein expression and localization is confirmed by fusion to EYFP (coding sequence is shown in yellow). The nopaline synthase terminator sequence is used for transcriptional termination (grey box, right side). The left and right border sequences of the T-DNA are indicated by the large black arrowheads. Restriction sites are indicated above the schematic. Markers for selection of transformants are not shown. Drawing is not to scale. 55 50 kD> 38 kD> 117kD>—- . . 97kD>- '5“ ‘1’.“ Figure 2.12. Demonstration of expression and targeting of ARC6ZBD-EYFP. Merged fluorescence micrographs from young leaves of Ws-2 (A); Ws-2 expressing ARC6ZBD-EYFP (B); arc6-I (C); and arc6-1 expressing ARC6ZBD-EYFP (D). Exposure times in all four panels are identical. (E) a-YF P immunoblot (upper panel) and Ponceau stain (lower panel) from arc6 lines expressing pCAMBIA-ZBD-derived ARC6ZBD-EYFP. Molecular weight marker (lane 1); protein extracts from three different transformant lines (lanes 2-4); Ws-2 (lane 5); and arc6-1 (lane 6). Extracts in lane 3 were taken from the transgenic individual shown in panel D. Scale bar = 5 pm. 56 Figure 2.13. Chloroplast and FtsZ morphology in ARC6ZBD~EYFP expressing lines. Terminal chloroplast phenotypes in expanded leaf cells from Ws-2 (A); arc6-1 (B); Ws-2 expressing plastid-targeted ARC6ZBD-EYFP (C); arc6-I expressing plastid-targeted ARC6ZBD-EYFP (D). Insets in each panel show FtsZ2-1 localization (green) in chloroplasts (red) within young leaves of the same lines shown in the large panels of A- D. The inset micrographs are the same scale as the large panels. Scale bars = 10 um. 57 Taken together, from these results I conclude that either ARC6ZBD does not function in plastid division or that ARC6ZBD cannot function outside of the context of the full-length ARC6 protein. Discussion Here I have identified a putative FtsZ2-binding domain of ARC6 (ARC6ZBD) and attempted to characterize its function in vivo. From our results, I conclude that the conserved ARC6ZBD resides within the context of a larger stromally-localized conserved region, occupying amino acids 351-503 (Figure 2.1). Multiple sequence alignment between ARC6 orthologs of higher plants reveals that ARC6ZBD is conserved amongst this group of organisms (Glynn, Yang, et al. 2009) and our pairwise alignment (Figure 2.3) suggests that ARC6ZBD is distantly related to the FtsZ binding domain of ZipA (ZipAZBD) (Figure 2.3). Despite these differences in amino acid sequence, these divergent domains are similar in length and all bind the core motif of FtsZ proteins. Using sequence alignment sofiware, a homology-modeling platform, and side—chain positioning algorithms, we generated a model of ARC6ZBD (Figure 2.4) based on the known crystal structure of ZipAZBD (Mosyak, Zhang, et al. 2000). This model highlights the potential importance of at least 3 key residues within ARC6ZBD and provides a testable foundation for structural analysis of ARC6ZBD by crystallographic or 58 NMR-based methods. Similar to the ZipA-FtsZ interaction (Moreira, Femandes, et al. 2006), site-directed mutagenesis of a conserved phenylalanine residue abolished ARC6ZBD interaction with FtsZZ in yeast two-hybrid assays (Figure 2.6), suggesting that this residue of ARC6ZBD might secure the core motif of FtsZ2 within a hydrophobic pocket. To test this, I generated a site-directed mutation in recombinant ARC6ZBD protein, ARC6ZBD(F44215), that I predicted to be unable to bind FtsZ2 in pulldowns, but I could not validate two-hybrid results using this assay; both ARC6ZBD and ARC6ZBD (F442D) had similar affinity for FtsZZ in vitro (Figure 2.7). This result conflicts with the structural model and hypothesis that ARC6F442 is critical for ARC6 function in vivo. To determine the functional relevance of ARC6F442 and to assess which of the interaction assays was yielding valid information, I made a construct that represents an ARC6ZBD loss-of-function mutant (ARC6pr0-ARC6 F44 2 D), based on the two-hybrid results (Figure 2.6). ARC6pro-ARC 6 F44 2 D was able to fully complement an arc6 T-DNA insertion mutation (Figure 2.9) and F tsZ ring morphology in complemented individuals was indistinguishable from that in wild type chloroplasts (Figure 2.10), suggesting that the two-hybrid tests between ARC6ZBD (F442D) and FtsZ2 may have yielded false negative results. However, the in vivo functionality of the ARC6F442D protein is completely consistent with our pulldown assays that show that ARC6ZBD (F442D) is able to coprecipitate the C-terminus of FtsZZ (Figure 2.7). It is possible that site-directed 59 substitutions of ARC6 residues 1362 and R422 (both individually and in combination with ARC6F442 substitutions) might provide sufficient evidence to validate our structural model (Figure 2.4) if those substitutions significantly disrupt ARC6-FtsZ2 interaction both in vitro and in vivo. However, in the absence of these additional site-directed mutants and the lack of an experimentally-detennined structure for ARC6ZBD, our ZipA- based structural model should be considered preliminary. While our work highlights a prospective functional domain within ARC6, the conflicting conclusions drawn from the two-hybrid and in vivo results does lead to some uncertainty regarding the structure and function of ARC6ZBD. Our results here reiterate the previous findings of our colleagues (Maple, Aldridge, et al. 2005), who used yeast two-hybrid assays to identify the ARC6-FtsZ2 interaction. They claimed to confirm this two-hybrid result using BiFC between ARC6 and FtsZZ in tobacco (Maple, Aldridge, et al. 2005, Maple and Moller 2006), using C-terminal BiFC tags for ARC6 and F tsZ2 to reconstitute YFP in vivo. The bitopic topology of ARC6 (V itha, Froehlich, et al. 2003) and the stromal localization of FtsZ2 (McAndrew, Froehlich, et al. 2001) conflict with this result, because the C-terminus of ARC6 has been shown to reside within the intermembrane space; BiFC-based assays will not work across membrane boundaries, as the two halves of the fluorescent moiety cannot physically interact when separated by a membrane barrier (Bracha-Drori, Keren, et al. 2004). In this case, the topology of ARC6 in these assays was probably not the native topology (perhaps due to overexpression or the presence of the C-terminal tag) or some FtsZZ was present within the intermembrane 60 space. We also analyzed a mutation in FtsZZ (FtsZ2F466 A) that was shown to impede ARC6-FtsZ2 interaction by yeast two-hybrid (Maple, Aldridge, et al. 2005), but in our hands FtsZZF466 A was coprecipitated by ARC6ZBD (Figure 2.7), suggesting that previous results (Maple, Aldridge, et al. 2005) might be two-hybrid false negatives. Strangely, expression of plastid-targeted ARC6ZBD did not modify the chloroplast morphology of F tsZ immunolocalization patterns of wild type or arc6 mutants (Figure 2.13), leading us to question the overall relevance of ARC6ZBD in the process of chloroplast division; perhaps the FtsZ2-binding domain of ARC6 has some other function in vivo. Further confounding the issue, another group has demonstrated that interaction between the cyanobacterial ortholog of ARC6 (Ftn2) and FtsZ requires the J -domain of Ftn2, suggesting that an interaction between ARC6 and FtsZZ might occur directly through the putative J-domain of ARC6 — or may be indirect, perhaps occuring through a DnaK-like chaperone (Mazouni, Domain, et al. 2004). When combined with our negative results obtained using a prospective ARC6ZBD gain-of-fimction mutant (Figure 2.13), we believe more work is needed to show that ARC6ZBD directly binds FtsZ2 within the stroma and to identify the functional implications of that interaction, if it does, in fact, occur. Regardless, the overall effect of ARC6 upon Z—ring formation is profound, as arc6 mutants possess serious defects in Z-ring assembly (V itha, Froehlich, et al. 2003). While recent data clearly provide evidence of a complex containing ARC6 and FtsZZ, it 61 does not resolve the issue of direct ARC6-FtsZZ interaction (McAndrew, Olson, et al. 2008). Moreover, preliminary data suggest that ARC6 can immunoprecipitate F tsZZ-l , even in the absence of the FtsZZ-binding domain of ARC6, indicating that there may an indirect linkage between these two proteins in vivo (Figure 2.14). Further work will be required to sort out the details of the molecular arrangement within the ARC6-FtsZ2 complex and determine how ARC6 might modulate Z—ring formation in vivo. 62 50 kD> 3 7 kD> a-FtsZZ Figure 2.14. ARC6 A A 1-509-GFP and ARC6“ 1-331-GFP can immunoprecipitate FtsZ2. Immunoblot showing eluates from immunoprecipitation reactions using an immobilized anti-GFP antibody incubated with whole-cell extracts from arc6 lines expressing a 35S—ARC6AA 1_509-GFP transgene (lane 1); 35S—ARC6 AA 1-331-GFP (lane 2); or 35S-GFP (lane 3). Blot was probed with an FtsZ2-1-specific antibody. The molecular weight of processed F tsZZ-l is predicted by ChloroP (Emanuelsson, Nielsen, et al. 1999) to be ~45 .3 kD (upper band). The lower molecular weight band in lane 1 (*) is probably derived from FtsZ2-l, due to the high specificity of the FtsZ2-1 antibody. It is not clear if this lower molecular weight band represents an FtsZ2-l breakdown product or if ARC6 AA 1-509 might somehow affect FtsZZ-l translation or processing in vivo. 63 Materials and Methods Yeast two-hybrid vector construction and reporter assays. All GAL4-ARC6 fusion-encoding plasmids were generated by PCR using primers that correspond to the desired coding sequences, with Ndel and Xmal adapters to facilitate directional ligation. PCR products were generated using the ARC6 cDNA clone available from ABRC, cut, and ligated into Ndel-Xmal digested pGADT7 (Clontech). All GAL4-ARC6 site-directed mutants (F442A, F442D, F 442K, and F 442Y) were generated by SOE-PCR (Heckman and Pease 2007) and cloned into pGADT7 as described above. All GAL4-FtsZ2 fusions were generated using primers corresponding to the desired coding sequence, with Ndel and BamHI adapters to facilitate the desired directional ligation of the insert. PCR products were generated from an FtsZZ-l clone, cut, and ligated into Ndel-BamHI digested pGBKT7 (Clontech). All clones were sequenced prior to transforming yeast to ensure that the inserts were free of coding errors. HIS3 reporter assays were performed in line with the manufacturer’s (Clontech) recommendations using an established protocol (Maple, Aldridge, et al. 2005). Relative interaction strengths were computed by comparison of the ratio of yeast growth on SD/- ULTH media to yeast grth on SD/-ULT media after ~48 hours of growth at 28-30 0C. Multiple sequence alignments and homology model. Multiple sequence alignments were performed using ClustalW (EBI) using the BLOSUM62 substitution matrix with all other parameters set to their default values. The alignment was formatted using ESPript (http://esnrint.ibcp.fr/ESPript/cgi- 64 bin/ESPript.cgi/). For homology modeling, a pairwise alignment between ZipA A A 135- 323 and ARC6A A 3 5 1-503 was performed and used to make a primary homology model from PDB file 1F47 (Mosyak, Zhang et al. 2000) (http://wwwpdbcgh with all identical residues from the ZipA-ARC6 pairwise alignment as anchor points using a homology modeling program (http://proteins.msu.edu/). We performed refinement of side chain positions using SCWRL3 (httn://dunbrack.fccc.edu/SCWRL3.php). Cartoon representations of the homology model were generated using Pymol (http://pymolsourcefogenefl). In vitro pulldowns. All fusion proteins were expressed in E. coli BL21 (DE3) Codon Plus cells (Stratagene) induced at OD600 = 0.8 with 2 mM IPTG for 2 hours at 37°C. 750 ug of protein from induced cell extract was used for each bait/prey combination. Pulldowns between Hi83-ARC613D and GST-FtsZZ-ICT or Hi83-ARC6ZBD (F442D) and GST- F tsZZ-ICT were performed as in previous experiments (Glynn, Froehlich, et al. 2008), except Triton X-100 was present at 0.1% in all wash buffers to prevent clumping of the sepharose beads. Analysis of chloroplast morphology and number. Light micrographs depicting chloroplast morphology in expanded leaf cells were taken using DIC Optics on a Leica DMIBOOOB Inverted Microscope outfitted with a Leica DFC3 20 Camera. Samples for chloroplast morphology and quantitation were 65 prepared and analyzed using established protocols (Pyke and Leech 1991). Fluorescence micrographs were taken using a Leica DMRA2 using Q-Capture Camera Control Software (Q-Imaging) and the filter sets indicated (Leica) as previously described (V itha, Froehlich, et al. 2003). Image analysis and RGB composites were made using Image] v1.37 (NIH) (Bearer 2003). F tsZ immunofluorescence Tissue preparation, fixation, and immunofluorescence analysis were carried out as described previously (Miyagishima, Froehlich, et al. 2006, Vitha, Froehlich, et al. 2003, Vitha, McAndrew, et al. 2001). Construction of pCAMBIA-ZBD. pCAMBIA-ZBD was constructed from pCAMBIA-1302 (Hajdukiewicz, Svab, et al. 1994) by removing an NcoI-BstEII fiagment from pCAMBIA-1302 and performing a four-point ligation (see Figure 2.11) with: (1) NcoI/ApaI digested PCR products amplified from a RecA cDNA using primers TTTT'ITCCATGGATTCACAGCTAGTCT TGTC and TTT'I‘TTGGGCCCTCTGTCATCGAATTCAGAACTGAT'T; (2) ApaI/BgllI digested PCR products amplified from an ARC6 cDNA using primers TTTTTTGGG CCCGAAGTTGCACTTGCTCTTGTGGCT and TTTI‘TTAGATCTAACTACCTCCA CTCTTTCCAAGT; and (3) BglII/BstEII digested PCR products amplified from an EYFP-coding plasmid (F. Brandizzi Laboratory, MSU DOE-PRL) using primers TTTTTTAGATCTCCCGGGATGGTGAGCAAGGGCGAGGAGCT and I I I I I IGGTTACCTTACTTGTACAGCTCGTCCATGCC. Following transformation, 66 clones were selected on LB plates containing kanamycin (50 ug/mL) and screened by PCR. Clones containing an insert of the expected size were miniprepped (Promega) and sequence-verified at the MSU-RTSF facility before transforming Agrobacterium and Arabidopsis. Fluorescence microscopic analysis of ARC 6 ZBD'E YF P expressing plants. Fluorescence micrographs were taken using a Leica DMRA2 using Q-Capture Camera Control Software (Q-lmaging) and the filter sets indicated (Leica) as previously described (V itha, Froehlich, et al. 2003). Image analysis and RGB composites were made using Image] v1.37 (NIH) (Bearer 2003). Protein extraction and immunoblotting. Proteins from whole-cell extracts (equivalent to ~2-3 mg liquid-nitrogen ground tissue) were prepared as previously described (Wiegel and Glazebrook 2002), separated by SDS-PAGE, and transferred to nitrocellulose. Anti-GF P Irnmunoblots were performed using Clontech JL-8 anti-GFP monoclonal antibody at 1:1000 in 5% nonfat dry milk in TBS-T, pH 7.4 (Miyagishima, Froehlich, et al. 2006). Transgenic plant material and microscopic analyses. Untagged ARC6 and ARC6 F44 2 D transgenes were generated by SOB-PCR and ligated into a modified pCAMBIA-l 302 (Hajdukiewicz, Svab, et al. 1994) using NcoI and BstEII; BstEII removes the mGFP tag from pCAMBIA-1302 and retains the NOS terminator sequence. All clones were sequence-verified and transformed into 67 Agrobacterium tumefaciens GV3101 by electroporation. Overnight cultures of A grobacterium carrying the T—DNA construct were used to transform Ws-2 and arc6-1 plants (Clough and Bent 1998). Following hygromycin selection, putative transgenic T1 plants were screened for transgene expression by Western blotting of whole-cell extracts from various individuals and probing with an anti-GFP antibody (BD-Biosciences J L-8 monoclonal antibody) and by fluorescence microscopy. Immunofluorescence analysis was carried out as previously described (V itha, McAndrew, et al. 2001) using expanded leaf tissue. Immunoprecipitation from Homogenized Arabidopsis Leaf Tissue. Transgenic arc6 plants expressing GFP-fusion proteins were grown in soil for 21 days. Approximately 150-200 mg of tissue was collected from whole leaves and ground to a powder in liquid nitrogen. Four (4) volumes of IP buffer (25 mM Tris, pH 7.50; 150 mM NaCl; 0.1% Triton X-100; and 1X Roche Protease Inhibitor Cocktail) were added to the frozen powder and gently mixed by pipeting on ice over several minutes. The homogenized slurry was spun through a Miracloth filter to remove large debris. 2.5 pg of BD Living Colors A.v. Peptide Antibody (Cat. # 632377) was gently mixed into to each filtered extract and incubated overnight. 50 uL of pre-conditioned protein A beads were added to each sample and incubated for 2 hours at room temperature with gentle mixing by inversion every 10 minutes. The beads were washed 4 times with 500 uL of IP buffer (each wash was 500 uL). Immune complexes were eluted from the beads using 100 uL of elution buffer (0.2M Glycine, pH 1.85) and then buffered with 20 uL of l M Tris, pH 8.5. 29 uL of 6X SDS-PAGE buffer were added to each buffered eluate, samples were 68 vortexed briefly, boiled, collected by centrifugation, and loaded for SDS-PAGE analysis. Transfer to nitrocellulose was performed using a GENIE Blotting Apparatus (Idea Scientific) according to the manufacturer’s recommendations. Rabbit-derived anti- FtsZZ-l antibodies were used as described previously (Vitha, McAndrew, et al. 2001 ). Acknowledgements I thank Bill Wedemeyer for instruction and assistance in generating the structural model of ARC6ZBD and for all the excellent structural analysis tools he has freely provided at his website. 69 Chapter 3 ARC6 Binds and Positions PDV2 During Plastid Fission in Arabidopsis. 70 Abstract Chloroplasts arose from a free-living cyanobacterial endosymbiont and divide by binary fission. Division involves the assembly and constriction of the endosymbiont- derived, tubulin-like F tsZ ring on the stromal surface of the inner envelope membrane and the host-derived, dynamin-like ARC5 ring on the cytosolic surface of the outer envelope membrane. Despite identification of many proteins required for plastid division, the factors coordinating the internal and external division machineries are unknown. Here, we provide evidence that this coordination is mediated in Arabidopsis by an interaction between ARC6, an F tsZ-assembly factor spanning the inner envelope membrane, and PDV2, an ARC 5 recruitment factor spanning the outer envelope membrane. ARC6 and PDV2 interact via their C-terminal domains within the intermembrane space of the chloroplast, consistent with their in vivo topologies. ARC6 acts upstream of PDV2 to localize PDV2 (and hence ARC5) to the division site. We present a model whereby ARC6 relays information on stromal FtsZ ring positioning through PDV2 to the chloroplast surface to specify the site of ARC5 recruitment. Because orthologs of ARC6 occur in land plants, green algae, and cyanobacteria whereas PDV2 occurs only in land plants, the connection between ARC6 and PDV2 represents the evolution of a plant-specific adaptation to coordinate the assembly and activity of the endosymbiont- and host-derived plastid division components. 71 Introduction The plastids of plant cells arose from cyanobacteria by endosymbiosis and, like cyanobacteria, replicate by binary fission. This process requires the coordinated action of at least two macromolecular complexes, one composed of the tubulin-like cytoskeletal protein FtsZ and the other of the dynamin-related protein ARC5. These proteins assemble into mid-plastid ring-shaped structures on opposite sides of the two envelope membranes to mediate constriction of the organelle (Osteryoung and Nunnari 2003). Two nuclear-encoded plant F tsZ paralogs, Ftle and FtsZ2, function within the chloroplast stroma. Both evolved from cyanobacterial FtsZ and have unique non- overlapping functions in plastid division (Schmitz, Glynn, et al. 2009, Stokes and Osteryoung 2003, Vitha, McAndrew, et al. 2001). Analogous to-bacterial FtsZ (Bi and Lutkenhaus 1991, Goehring and Beckvvith 2005), the plastidic FtsZ proteins assemble at an early step in division to form an equatorial ring, the Z-ring, on the stromal face of the inner envelope membrane (V itha, McAndrew, et al. 2001). In bacteria and presrunably in plastids, the Z-ring probably functions both as a scaffold for the recruitment of other division proteins and to provide the contractile force needed to pull the membrane inward during constriction (Ghosh and Sain 2008, Lan, Daniels, et al. 2009, Lan, Wolgemuth, et al. 2007, Osawa, Anderson, et al. 2008). In contrast, ARC5 was a post-symbiotic adaptation of the eukaryotic host and functions outside the chloroplast (Gao, Kadirjan-Kalbach, et al. 2003, Miyagishima, Nishida, et al. 2003a). ARC5 is a member of the dynamin family of membrane 72 “pinchases,” best characterized for their roles in endocytic vesicle-budding and mitochondrial fission in eukaryotes (Cerveny, Tamura, et al. 2007, Hoppins, Lackner, et al. 2007, McNiven 1998, Shaw and Nunnari 2002, Ungewickell and Hinrichsen 2007), though dynamins probably evolved from a prokaryotic ancestor (Low and Lowe 2006). ARC5 in Arabidopsis and its orthologue in the red alga Cyaniodioschyzon merolae function late in chloroplast division by assembling on the cytosolic surface of the outer envelope membrane, where they are thought to perform the final squeeze that aids partitioning of the two daughter organelles (Yoshida, Kuroiwa, et al. 2006). Both plastidic FtsZ and ARC5 arose early in the evolution of the chloroplast division machinery, as indicated by their occurrence in both the red and green lineages (Miyagishima 2005). The assembly of the division machinery appears to occur in a linear order, with F tsZ assembly initiating the process and ARC5/dynamin mediating late-stage organelle constriction (Miyagishima, Nishida, et al. 2003a). However, the mechanisms coordinating these evolutionarily and compartrnentally disparate events are unknown. Previous studies in Arabidopsis have implicated three plastid division proteins as candidate mediators of this coordination: ARC6, PDVl and PDV2 (Miyagishima, Froehlich, et al. 2006, Vitha, Froehlich, et al. 2003). ARC6 is a bitopic transmembrane protein of the inner envelope membrane with its larger N-terminus protruding into the stroma and smaller C-terminus residing within the intermembrane space (IMS). ARC6 was inherited from the cyanobacterial endosymbiont and is localized to the mid-plastid . division site. The chloroplasts of Arabidopsis arc6 mutants, which have one or two 73 oversized chloroplasts per mesophyll cell (Pyke, Rutherford, et al. 1994), possess many short disorganized filaments, while ARC6 overexpressors have excessively long F tsZ filaments (V itha, F roehlich, et al. 2003). These findings, along with the fact that the N- terminus of ARC6 interacts specifically with FtsZ2 (Maple, Aldridge, et al. 2005), suggest that ARC6 facilitates F tsZ polymer assembly and regulates FtsZ ring dynamics through FtsZZ (Maple, Aldridge, et al. 2005, McAndrew, Olson, et al. 2008, Vitha, Froehlich, et al. 2003). The function of the C-terminal IMS region of ARC6 is unknown, but it probably has a significant role in plastid replication based on the breadth of its conservation amongst plants, algae, and cyanobacteria (Glynn, Yang, et al. 2009, Vitha, Froehlich, et al. 2003). ARC6 is one of the few division proteins known to connect the stromal Z-ring to the IMS, suggesting it could play a critical role in coordinating the inner and outer subassemblies of the division machinery. PDVl and PDV2 are paralogous plastid division proteins identified based on the similarity of the pdvl mutant phenotype to that of arc5; in both mutants, chloroplasts are enlarged and dumbbell-shaped (Miyagishima, Froehlich, et al. 2006). PDVl is a bitopic outer envelope membrane protein that localizes to the division site, with its N-terminus residing in the cytosol and its C-terrninus extending into the IMS. PD VI and PD V2 have partially redundant functions in recruiting ARC5 to the chloroplast. In pdvl and pdv2 mutants, ARC5 localizes to the central constriction in the enlarged chloroplasts, but pdvI pdv2 double mutants fail to recruit ARC5 to the chloroplast (Miyagishima, Froehlich, et al. 2006). PDVl and PDV2 proteins share some degree of sequence similarity and domain arrangement, though the localization and detailed function of PDV2 was not 74 determined previously (Miyagishima, Froehlich, et al. 2006). PDV] and PDV2 have no significant sequence similarity to known proteins and are only evident in land plants, suggesting they represent an evolutionary development in the transition of plants to terrestrial habitats. Here, we show that PDV2 has a localization and topology similar to that of PDVl in the outer envelope membrane. However, PDV2 family members have a unique C- terminal extension in their IMS regions that is lacking in PDVl proteins. We further show that the C-terminal IMS regions of ARC6 and PDV2 interact and that this interaction is required for full chloroplast division activity in Arabidopsis. Using genetic analysis, we demonstrate that ARC6 is required for positioning of PDV2 and ARC5, but PDV2 is not required for mid-plastid localization of ARC6. Our results establish a physical link across the envelope membranes at the division site and suggest that ARC6, through interaction with PDV2 within the IMS, coordinates Z-ring and ARC5 activity to synchronize scission of the envelope membranes. We present a model for the physical arrangement and interaction of these proteins in the chloroplast membranes. Results PD V2 localization and topology are similar to PD VI . The similarity of PDV2 to PDVl in sequence and domain arrangement (Figure 3.1) suggested its localization and topology might be similar to those of PDVl 75 PDV] N l D c — — TM i 272 AA PDV2 N l D c — _ TM I 7 307 AA Figure 3.1. Schematic comparison of PDVl and PDV2 proteins from Arabidopsis. Shaded lines below each protein indicate regions of high similarity between the two PDV paralogs. A comprehensive alignment and phylogeny of these two proteins that includes sequences from several plant species is published elsewhere (Glynn, Froehlich, et al. 2008). Amino terminus (N); carboxy terminus (C); transmembrane domain (TM); and amino acids (AA). 76 (Miyagishima, Froehlich, et al. 2006). To verify localization of PDV2 in vivo, we expressed a YFP-PDV2 fusion protein from the PDV2 promoter (PD V2pm- YFP-PD V2) in wild-type Arabidopsis plants. Similar to PDVl , the YFP signal localized to the mid- plastid in young emerging leaves. The YFP-PDV2 signal appeared as a continuous ring (Figure 3.2) rather than as a series of equatorial spots as was previously noted for PDVl (Miyagishima, Froehlich, et al. 2006). When YFP-PDV2 was co-expressed with an ARC6-CFP fusion protein, their fluorescence signals colocalized within mesophyll chloroplasts (Figure 3.3). To confirm the predicted topology of PDV2, we carried out in vitro chloroplast import and protease-protection assays (Figure 3.4, left panel). Following incubation with isolated pea chloroplasts, radiolabled PDV2 produced by in vitro translation (Figure 3.4, lane 1) was retained in the membrane fraction (Figure 3.4, lane 2). The protein was susceptible to degradation by therrnolysin, consistent with outer envelope localization. However, a portion of PDV2, roughly corresponding to the size of the predicted C-terminal IMS domain (predicted size ~10 kD), was protected from therrnolysin degradation (Figure 3.4, lane 4, arrowhead), but was sensitive to trypsin, which can penetrate the outer envelope membrane and enter the IMS (Jackson, Froehlich, et al. 1998, Tranel, Froehlich, et al. 1995). ARC6, used as an inner envelope control in these experiments (Figure 3.4, right panel), behaved as shown previously (V itha, Froehlich, et al. 2003). Protease protection assays on wild type Arabidopsis chloroplasts using antibody-based detection of PDV2 showed the N-terminus of PDV2 to be susceptible to thermolysin and trypsin treatment (Ronit Knopf and Zach Adam, personal communication), consistent with in vitro assays. These results confirm that PDV2, like PDV] , is a mid-plastid-localized, bitopic protein of the chloroplast outer envelope 77 Figure 3.2. Localization of YFP-PDV2 in Arabidopsis. A PD VZPrO transgene was expressed in Arabidopsis Col-0 plants and tissue samples from emerging leaves were examined for YFP-PDV2 localization. Signals observed from: Chlorophyll autofluorescence (Chl, panel A); (B) Yellow Fluorescent Protein (YF P, panel B); and merged image (panel C). YFP-PDV2 was typically observed as a continuous ring at the mid-chloroplast (arrowhead, panel C). Scale bars = 10 um. - YFP-PD V2 78 . f; .- B . l) a .3 .31} a , Figure 3.3. Colocalization of ARC6-CFP and YFP-PDV2 in Arabidopsis. ARC6- CF P and YFP-PDV2 were expressed under control of their native promoters and examined in emerging leaves. Micrographs from brightfield (A); CFP channel (B); YFP channel (C); chlorophyll autofluorescence (D); and a merged image from CFP, YF P and chlorophyll channels (E). Scale bar = 10 um. 79 --++-- Thermolysin - - + + ----++ Trypsin ----++ PS PSPS P S P S P S 50kD> 100m> 37“» 75kD> 25kD> 20kD> kD> ‘ 50 1234567 891011121314 PDV2 ARC6 Figure 3.4. PDV2 fractionation and topology using isolated pea chloroplasts. In vitro chloroplast import of [3H-leucine]-labeled PDV2 or [ 5S-methionine]-labeled ARC6 and fractionation of chloroplasts following protease treatment and hypotonic lysis. In vitro- transcribed translation products are shown for PDV2 (lane 1) and ARC6 (lane 8). Pellet (P) fractions are shown in lanes 2, 4, 6, 9, 11, and 13. Supernatant (S) fractions are shown in lanes 3, 5, 7, 10, 12, and 14. The arrowhead points to the IMS-localized C- terminal fragment of PDV2 that remains following thermolysin treatment. 80 membrane, with its N-terminus exposed to the cytosol and its smaller C-terminus residing within the intermembrane space. The IMS Regions of ARC6 and PD V2 Interact. The localization of PDVl (Miyagishima, Froehlich, et al. 2006), PDV2 (Figure 3.2) and ARC6 (V itha, Froehlich, et al. 2003) at the division site, along with their demonstrated topological orientations in the envelope membranes, suggested the possibility that ARC6 might interact with PDVl and/or PDV2 in the IMS. To test this, we carried out yeast two-hybrid assays with constructs encoding the C-terminal IMS domains of ARC6 and PDVI or PDV2. PDV2 strongly and specifically activated the HIS3 reporter in the presence of ARC61MS (Figure 3.5A, third row). No interaction was observed between PDVIIMS and ARC61MS using two hybrid assays (Figure 3.5A, first row). To confirm the ARC6IMs-PDV21MS interaction we carried out a pulldown assay (Figure 3.5B). A GST-PDVZIMS fusion protein was precipitated from crude E. coli extracts with Ni-Sepharose beads coated with His-ARC61MS (Figure 3.5B, lane 1), but no pulldown of GST-PDV21MS was observed using uncoated Ni-Sepharose beads (Figure 3.5B, lane 2). These results support an ARC6-PDV2 protein-protein interaction within the intermembrane space of dividing chloroplasts. 81 A SD/-ULT SD/—ULTH PDVIIMS vs. EMPTY e PDV2,M S.vs EMPTY B 1/10 1/100 1/10 1/100 or-HIS or-GST Figure 3.5. The IMS-localized regions of PDV2 and ARC6 interact. (A) Two-hybrid assays between PDV proteins and ARC6. The IMS regions of the proteins indicated were fused to either GAL4-BD (PDVl and PDV2) or GAL4-AD (ARC6). Dilutions (indicated at bottom of panel) from a starting culture of OD600 = 1.0 are spotted onto synthetic dropout (SD) media containing (left) or lacking (right) histidine. (B) Immunoblots of eluates from pulldown assays using either immobilized Hiss-ARC6IMS treated with GST-PDVZIMS (lane 1) or naked beads treated with GST-PDVZIMS (lane 2). 82 PD V2 Family Members Possess a Unique C-Terminal Domain. To gain insight into features that distinguish the C-terminal IMS domains of PDVl and PDV2 from one another and therefore might be important for the PDV2- specific interaction with ARC6, we generated sequence alignments between PDVl and PDV2 proteins from several land plants using a CLUSTALW identity scoring matrix (Larkin, Blackshields, et al. 2007) to clearly define PDVl-specific and PDV2-specific features (Glynn, Froehlich, et al. 2008); the C-terminal segment of this alignment is shown in Figure 3.6. As noted previously, all PDV] proteins have a conserved C- terrninal glycine residue, and mutation of this residue impairs PDVl function in Arabidopsis (Miyagishima, Froehlich, et al. 2006). A conserved terminal glycine is also found in all known PDV2 family members (Figure 3.6, red asterisk). However, PDV2 (294 at 15 amino acids, n = 7 sequences) proteins are generally longer than PDVl (261 :t 14 amino acids, n = 7 sequences); PDV2 family members harbor a conserved extension at their C-terminus. This C-terminal extension, which includes the terminal glycine residue, might be involved in mediating PDV2 interaction with ARC6. The Conserved Terminal Glycine of PD V2 is Required for Interaction with ARC6 and PD V2 Function in vivo. To ask whether the C-terminal glycine of PDV2 is important for the ARC6-PDV2 interaction, we engineered a missense mutation in the PDVZIMS two-hybrid plasmid that changes this glycine to aspartate. Following co-transformation of the resulting construct (PDV21MS(G307D)) and ARC6IMS into yeast and selection for both plasmids, we observed no grth on medium lacking histidine (Figure 3.7, row 3), indicating that 83 AtPDVl MA ............................ GhPDVl LA ............................ HtPDVl SA ............................ OsPDVl LA ............................ PtPDVl LA ............................ VvPDVl YA ............................ ZmPDVl LA ............................ AOPDVZ E . E QPRCLVKERVEI PFDLDVSAPKINYGFG ABPDVZ E . EARCLVKERLEI PFDPVVRTPNVNYGCG AtPDVZ E EARCLVKERVEI PFDSVVAKRDVTYGYG LBPDVZ E . ESRCLVKERVKIPFKSVVTLPDVNYGCG OsPDVZ E . RAHCVVKERVEIPFDTNLASPNASYGLG PtPDV2 E. EVRCVVKERVAVPFNSVAGKPDVNYGSG TaPDVZ GGD RAHCVVKERVEIPFGSSLDAPNASYGLG — * **** ** *4: * Figure 3.6. Alignment of C-termini of PDVl and PDV2 family members. All PDV2 proteins possess a C-terminal extension that is not found in PDVl family members. The dark line (bottom) corresponds to to the C-terminal conserved segment shown in Figure 3.1. Boxed regions indicate similarity between PDVl and PDV2; boxed shaded regions represent identity between PDVl and PDV2. Asterisks (*) represent identical residues within the C-terminal extension of all PDV2 family members. The red asterisk highlights the conserved glycine that ends all PDV2 proteins. Arabidopsis thaliana PDVl (AtPDVl), Gossypium hirsutum PDVl (GhPDVl), Helianthus tuberosus PDVl (HtPDVl), Oryza sativa PDVl (OsPDVl), Populus trichocarpa PDVl (PtPDVl), Vitis vinifera PDVl (VvPDVl), Zea mays PD V1 (ZmPDVl), Arabidopsis thaliana PDV2 (AtPDV2), Asparagus ojj‘icinalis PDV2 (AoPDV2), Aquilegia species PDV2 (AsPDV2), Lactuca sativa PDV2 (LsPDV2), Oryza sativa PDV2 (OsPDV2), Populus trichocarpa PDV2 (PtPDV2), and Triticum aestivum PDV2 (TaPDV2). The full-length alignment is published elsewhere (Glynn, Froehlich, et al. 2008). 84 A SD/-ULT SD/—ULTH PDV2 1M, vs. ARC6 ,M, m _ 1 PDV2 IMSvs. EMPTY _ _ 2 PDV2IMS (mmvs. ARC6,MS n - 3 PDV21M3(G3O7D,VS. EMPTY _ -. EMPTY vs. ARC6,MS m -5 1/10 1/100 1/10 moo B 1 2 3 4 ..-ns r: a—GST Figure 3.7. The terminal glycine of PDV2 is important for interaction with ARC6. (A) Yeast two-hybrid assays between PDV2 proteins and ARC6. The IMS regions of the proteins indicated were fused to either GAL4-BD (PDV2 and PDV20307D) or GAL4-AD (ARC6). Dilutions (indicated at bottom of panel) from a starting culture of OD600 = 1.0 are spotted onto synthetic dropout (SD) media containing (left) or lacking (right) histidine. (B) Immunoblots of eluates from pulldown assays using either immobilized Hiss-ARC6IMS treated with GST-PDVZIMS (lane 1); naked beads treated with GST- PDVZIMS (lane 2); Hi83-ARC6IMS treated with GST-PDVZIMS (6307p) (lane 3); or naked beads treated with GST-PDV2IMS ((33071)) (lane 4). 85 mutation of the terminal glycine in PDV2 diminishes the ARC6IMs-PDV21MS interaction in yeast. We incorporated the same mutation into PDV2IMS to create GST-PDVZIMS (G 307D) for use in pulldown assays and observed that the terminal G307D mutation reduced the affinity of PDVZIMS for ARC6IMS (Figure 3.7B, lane 3). The higher affinity of GST-PDV21MS for His-ARC61MS (Figure 3.7B, lane 1) was not a result of biased prey input into the reaction, as equal amounts of total protein were applied to each pulldown reaction and we verified equal input levels of GST-PDVZIMS and GST-PDV2IMS (03071)) in the reactions by anti-GST immunoblotting. We conclude from these results that the IMS-localized domains of ARC6 and PDV2 interact with each other and that the conserved terminal glycine of PDV2 is an important mediator of this interaction. To assess the potential importance of the terminal missense mutation in vivo, we compared the relative abilities of PD V2pm—PD V2 and PD VZPm-PD V20 307D transgenes encoding full-length proteins to complement pdv2-l , a line carrying a T-DNA insertion allele of PD V2 in which chloroplasts are frequently constricted and larger than in wild type (Miyagishima, Froehlich, et al. 2006) (Figure 3.8B). We reasoned that a PDV2 protein incapable of interaction with ARC6 would not be able to rescue the pdv2 phenotype. The wild type transgene fully complemented pdv2 in the majority of selected lines (Figure 3.8C). In contrast, no modification of the pdv2 phenotype occurred in lines transformed with PD V2pro-PD V20 30 7D (Figure 3.8D) despite evidence based on 86 . £3 "a Ponceau Figure 3.8. PD V2030”) is a loss-of function allele. Chloroplast phenotypes from leaf cells of Col-0 (A); pdv2-I (B); pdv2-I transformed with a PD V2pro-PDV2 transgene (C); and pdv2-I transformed with a PD V2pm-PD VZG 307D transgene (D). An anti-PDV2 immunoblot is shown in (E) carrying recombinant PDV2 (lane 1); extract from plant expressing a PD VZPm-PDVZG 307D transgene (lane 2); extract from Col-0 (lane 3); and extract from pdv2-1 (lane 4). Chloroplast phenotype from a plant expressing 3 5S-PD V2 (F). An immunoblot showing PDV2 expression levels is shown in (G) carrying extract from plant expressing 35S-PDV2 (lane 1); extract from Col-0 (lane 2); and extract from pdv2-I (lane 3). Scale bars = 10 um. 87 immunoblotting with a PDV2-specific antibody that PDV20307D protein levels were equivalent to or greater than PDV2 levels in wild type (Figure 3.8E). Lack of complementation by PD V2pm-PD V20 307D was not a consequence of overexpression, as overexpression of PD V2 does not impede chloroplast division, but rather accelerates this process (Figure 3.8F-G) (Okazaki, Kabeya, et al. 2009). To determine if the defect in PDV2g307D might be due to mislocalization or perhaps a loss in activity, we examined YFP-PDVZG307D localization in Arabidopsis. The YFP-PDV20307D signal was observed as scattered fragments around the periphery of the chloroplast in the wild type background (Figure 3.9B), in contrast to the rings observed in YFP-PDV2 expressing lines (Figure 3.9A), likely due to the inability of PDVZG307D and ARC6 to interact with each other. While YFP-PDV2 was observed forming rings in the wild type plastids, we did not observe YFP-PDV23307D associated with chloroplasts in the pdv2-I background, but it did localize to small epidermal plastids (compare Figures 3.9C-D); the significance of this observation is not yet clear. We conclude that PDV2 interaction with ARC6 occurs through the conserved C- terminal glycine of PDV2 and that the major role of this protein-protein interaction is to localize PDV2 to the division site in vivo; loss of this interaction leads to a defect in plastid division that is reminiscent of pdv2 and arc5 loss-of-function mutants. 88 Col-ll ("HI-ll ptli'Z-I b Figure 3.9. PDV2G307D localization in Arabidopsis. YFP-PDV2 and YFP- PDVZG307D were expressed under control of the native promoter in both Col-0 (A, B) and pdv2-I (C, D) backgrounds. YFP-PDV2 localizes to continuous rings around plastids (arrowheads) in Col—0 (A) and pdv2-1 mutants (C). In contrast, YFP- PDV20307D fails to localize to organized structures in either background (B, D). Scale in all images is identical. Scale bar = 2 pm. 89 The Complete IMS Region of ARC6 is not required for ARC6 Localization, but is required for Chloroplast Division Activity. To determine if the C-terminal IMS-localized region of ARC6 is required for plastid division in vivo, as suggested by its interaction with PDV2, we expressed a truncated ARC6-GFP fusion protein lacking much of its C-terminal IMS region (ARC6A1MS-GFP) in the arc6 background under control of the ARC6 promoter (Figure 3.10A). A full-length ARC6-GFP protein, shown previously to complement the severe plastid division defect in arc6 mutants (V itha, Froehlich, et al. 2003, Vitha, Holzenburg, et al. 2005) was expressed as a control. Accumulation of the fusion proteins in transgenic individuals was confirmed by immunoblotting (Figure 3.10B), and chloroplast morphology was examined in fixed leaf cells. Similar to previous results (V itha, Froehlich, et al. 2003), we observed rescue of the arc6 phenotype (Figure 3.10D) in lines expressing full-length ARC6-GFP (Figure 3.10B), as indicated by the increase in ntunber and decrease in size of the chloroplasts (25-60 cps/cell; Figure 3.10D) relative to those in the parent arc6 plants (1-2 cps/cell; Figure 3.10C). A change in chloroplast number and size was also observed in lines expressing ARC6A1Ms-GFP, but the extent to which the arc6 mutation was complemented (4-10 cps/cell; Figure 3.10E) was less than observed in arc6 lines expressing the full-length ARC6-GFP control. These results show that the IMS region of ARC6 is required for full plastid division activity in vivo. Moreover, arc6 plants expressing either ARC6pm-ARC6AIMS-GFP or 35Spm-ARC6 A IMS-GFP exhibit GFP rings localized to sites of constriction in chloroplasts of young leaves (Figure 3.10F), indicating that the C-terminal region of ARC6 is not required for mid-plastid 90 ARC6-GFP -; TP ARC6A1Ms-GFP -i Figure 3.10. ARC6AIMS-GFP is partially functional and localizes to the division site. (A) Schematic of GFP fusions used for analysis. (B) Anti-YFP immunoblot showing expression of ARC6-GFP (lane 1) and ARC6AIMS-GFP (lane 2) in Arabidopsis arc6 mutants. Chloroplast phenotypes of arc6 (C); arc6 complemented with ARC6!”- ARC 6-GFP (D); and arc6 expressing ARC6pr0—ARC6AIWGFP (E) in leaf mesophyll cells. Localization of ARC6-GFP (arrowheads) in young leaf cells of an arc6 mutant expressing 35S-ARC 6 A [MS-GFP (F) suggests that only stromal components are required to localize ARC6 to sites of constriction. ARC6AIMS-GFP encodes amino acids 1-682 of ARC6 fused to GFP. Transit peptide (TP); transmembrane domain (TM); green fluorescent protein (GFP); kilodaltons (kD). Scale bars = 10 pm. 91 localization of ARC6 or for constriction of the plastid during division. ARC6 acts upstream of PD V2 and is required for PD V2 activity. Because ARC6A1Ms-GFP localizes to constrictions at the plastid division site and arc6 plants expressing ARC6A1MS-GF P phenocopy the terminal chloroplast number within mesophyll cells of pdv2 mutants (Miyagishima, Froehlich, et al. 2006), we hypothesized that ARC6 acts upstream of PDV2. Because the phenotypes of arc6 (1-2 chloroplasts/cell) and pdv2 (4-8 chloroplasts/cell) mutants are easily distinguished by microscopic observation (compare Figures 3.11C-D), we made reciprocal crosses to test for epistasis between the arc6-I and pdv2-1 alleles. F1 individuals were allowed to self- fertilize, and the chloroplast phenotypes in leaves of 105 F2 individuals were examined (Figure 3.11). We observed all the expected phenotypes in the F2 population, including individuals with wild-type (WT) phenotypes, intermediate (INT) phenotypes consistent with PD V2/pdv2 heterozygotes (Miyagishima, Froehlich, et al. 2006), pdv2-like phenotypes, and arc6-like phenotypes in a ~9:3:4 ratio. Genotype analysis of individuals with arc6-like phenotypes confirmed the presence of arc6 pdv2 double mutants within the F2 population (Figure 3.11F). The overrepresentation of arc6 mutant phenotypes in the F 2 population and the arc6-like phenotype of the arc6 pdv2 double mutant are consistent with ARC6 acting upstream of PD V2. 92 pdv2-l F2 Phenotype Obs. WT/IN T: 63 arc6 25 pdv2 17 TOTAL 105 PD was pdv2 arc6 arc6-I Exp. 59 26 20 105 Figure 3.11. ARC6 acts upstream of PDV2. Chloroplast phenotypes of Col-0 (A); Ws- 2 (B); pdv2-I (C); arc6-I (D); PDV2/pdv2 heterozygotes (E); and pdv2 arc6 double mutants. Phenotypic distribution amongst an F2 population originating from arc6-I x pdv2-I crosses (G) with observed (Obs.) phenotypes and expected (Exp.) values for ARC6 acting upstream of PDV2 (x2 E 0.7; d.f. = 2; P 2 0.9). Scale bars = 10 um. 93 Interestingly, overexpression of PD V2 alone leads to an increase in the rate of chloroplast division (Figure 3.8) (Okazaki, Kabeya, et al. 2009). Because ARC6 acts upstream of PD V2, we suspected that PDV2 function might be fully dependent upon the presence of ARC6. To test this, we introduced a T-DNA carrying 35Spm-PD V2 into the arc6 mutant to ask whether overexpression of PD V2 would abrogate the arc6 phenotype. Following selection, we assayed T1 individuals for PDV2 protein levels by immunoblotting and identified multiple lines with elevated PDV2 levels (Figure 3.12, lower right). In none of these PD V2-overexpressing lines did we observe significant modification of the arc6 phenotype (Figure 3.12, upper right); mesophyll cells from all PD V2-overexpressors still contained only 1-2 oversized chloroplasts (Figure 3.12, lower left). This result indicates that ARC6 is required for PDV2 function in vivo. Moreover, overexpression of PD V2 in pdvl mutants (J .M. Glynn and K.W. Osteryoung, unpublished) does not bypass the division defect in the pdvl background, consistent with non-overlapping fimctions for PDV2 and PDVl. ARC6 is Required for Equatorial Positioning of PD V2, PD V1, and ARC5. We showed that ARC6 acts upstream of PDV2 (Figure 3.11 and Figure 3.12) and that ARC6 is probably responsible for positioning PDV2 at the division site by binding the C- terminus of PDV2 (Figure 3.7, Figure 3.8, and Figure 3.9). To confirm this prediction, we introduced a transgene carrying PDVZPm- YFP-PD V2 into Col-0 and into an arc6 T- DNA insertion mutant (SAIL_693_G04) and examined YFP localization in young leaves 94 (it-PDV2 Figure 3.12. Overexpression of PDV2 does not abrogate the arc6 phenotype. Chloroplast phenotypes within leaf mesophyll cells of Ws-2 (A); arc6-l (B); and arc6-I expressing a 35Spm-PD V2 transgene (C). An anti-PDV2 immunoblot (D) demonstrating relative PDV2 in the PD V2 overexpressor shown in panel C and the arc6 mutant shown in panel B. Scale bar = 10 um. 95 of T1 individuals. In the wild type background, we observed equatorial rings of YFP- PDV2 in 34% of chloroplasts (Figure 3.13A; n = 200 chloroplasts from 20 independent T1 individuals). In contrast, arc6 lines expressing YFP-PDV2 exhibited diffuse YFP localization around the periphery of the chloroplasts, but no YFP-PDV2 rings were observed (Figure 3.138; n = 200 chloroplasts from 16 T1 individuals); this result is consistent with the diffuse localization of the dysftmctional YFP-PDV20307D fusion around plastids (Figure 3.9). In the converse experiment, we examined ARC6- GFP localization in wild-type and pdv2 backgrounds (Figure 3.14). In both cases, ARC6- GFP was observed in equatorial rings, indicating that PDV2 does not influence ARC6 localization, even in the enlarged chloroplasts of a pdv2 mutant. We conclude that PDV2 is targeted to the outer envelope in an ARC6-independent manner, but absolutely requires ARC6 for its localization to the division site. To extend this analysis and further elucidate the role of ARC6 in mediating dynamin recruitment through PDVl-PDV2, we examined GFP-ARC5 localization in wild-type cells and arc6 mutants (Figure 3.15). In young wild-type leaf cells, GFP- ARCS was frequently observed in a punctate pattern around the division site (Figure 3.15A). In contrast, we observed GFP-ARC5 in patches within the arc6 mutant that do not appear to be associated with chloroplasts (Figure 3.15B), similar to reported GFP- ARCS localization in the pdvl pdv2 double mutant; these patches are probably cytosolic (Miyagishima, F roehlich, et al. 2006). Because GFP-ARC5 still localizes to the division site in a pdv2 mutant (Miyagishima, Froehlich, et al. 2006), we conclude that ARC6 96 Figure 3.13. ARC6 is required for localization of YFP-PDV2 to the division site. YFP-PD V2 was expressed from its native promoter in Col-0 (A) and arc6 mutant (B) backgrounds. YFP-PDV2 localizes to central rings within chloroplasts of wild type cells (arrowheads), but exhibits diffuse localization around the periphery of the plastid in arc6 mutants. A merged image showing signals fiom both chlorophyll autfluorescence (red) and YFP (green) is shown in each panel. Scale bars = 10 um. 97 (0L0 Figure 3.14. PDV2 is not required for ARC6-GFP localization to equatorial rings. ARC 6-GF P was expressed from its native promoter in Col—0 (A) and pdv2-I (B) backgrounds. ARC6-GFP localizes to central rings within chloroplasts of wild type cells and in pdv2 mutants (arrowheads), suggesting that equatorial localization of ARC6 is independent of PDV2. A merged image showing signals from both chlorophyll autfluorescence (red) and GFP (green) is shown in each panel. Scale bars = 10 um. 98 ( 'uI-II Figure 3.15. ARC6 is required for localization of GFP-ARC5 to the division site. GFP-ARC5 was expressed from its native promoter in Col-0 (A) and arc6 (B) backgrounds. GFP-ARC5 localizes to punctate equatorial rings within chloroplasts of wild type cells (arrowheads), but is only observed in cytosolic patches in arc6 mutants. A merged image showing signals from both chlorophyll autofluorescence (red) and GF P (green) is shown in each panel. Scale bars = 10 um. 99 mediates dynamin recruitment and patterning through both PDVl and PDV2, but ARC6 might mediate PDVl function through an indirect mechanism. Discussion Chloroplast division involves the consecutive formation and simultaneous constriction of the mid-plastid F tsZ and ARC5/dynamin rings on the stromal and cytosolic surfaces of the envelope membranes, respectively. ARC6 in the inner envelope interacts directly with FtsZ via its stroma-exposed N-terminus and is required for Z-ring assembly, while PDV2, shown here to be a transmembrane protein of the outer envelope, mediates recruitment of ARC5 to the chloroplast surface (Miyagishima, Froehlich, et al. 2006). Our results place ARC6 upstream of PDV2 in the chloroplast division process, identify a physical linkage between the C-terminal IMS domains of ARC6 and PDV2, and reveal that an important function of the ARC6IMS domain is to direct localization of PDV2 to the division site. These findings establish a role for a key inner envelope plastid division protein in organizing components of the outer envelope division machinery. Because the N-terminus of ARC6 plays a role in organizing FtsZ in the stroma (V itha, Froehlich, et al. 2003), these findings suggest a model wherein interaction between ARC6 and PDV2 within the IMS links FtsZ assembly with ARC5 recruitment, thereby promoting coordinated fission of the two envelope membranes. Our observation that ARC6 is able to localize to the division site without most of its IMS region (Figure 3.10) suggests that stromal factors are (sufficient to organize ARC6 100 at the division site within chloroplasts. The FtsZ ring is a strong candidate as an ARC6- positioning factor. The recently demonstrated ability of recombinant E. coli FtsZ to assemble into rings inside liposomes suggests that FtsZ proteins, which are highly conserved across kingdoms, require only a membrane tether and GTP for ring assembly in vivo (Osawa, Anderson, et al. 2008). We suspect that self-assembly of the Z-ring at the mid-plastid division site in plants, controlled by MinD, MinE, and ARC3 (Colletti, Tattersall, et al. 2000, Fujiwara, Nakamura, et al. 2004, Glynn, Miyagishima, et al. 2007, Maple, Chua, et al. 2002, Maple, Vojta, et al. 2007, Reddy, Dinkins, et al. 2002), establishes the site for ARC6 localization, presumably via direct interaction with FtsZ2 (Maple, Aldridge, et al. 2005). From our results, we propose that ARC6 transduces positional information from the Z-ring to the intermembrane space and serves as a landmark for PDV2 and hence ARC5 recruitment (Figure 3.13 and Figure 3.15). Consistent with this idea, arc6 plants expressing ARC6A1MS-GFP (Figure 3.10) phenocopy pdv2 and arc5 mutants with respect to their terminal chloroplast number (Gao, Kadirjan-Kalbach, et al. 2003, Miyagishima, Froehlich, et al. 2006), suggesting an equivalent block in the division process in these three genetically-distinct backgrounds. Additionally, the loss of PDV2 or ARC5 results in the formation of multiple adjacent Z- rings on the stromal side of the inner envelope membrane (Miyagishima, Froehlich, et al. 2006), suggesting that information may also travel inward from the outer envelope to the Z—ring through ARC6. Based on the collective data, we hypothesize that ARC6 has evolved to organize and coordinate the stromal and cytosolic components of the division complex and relay information about the status of each across compartment boundaries. 101 Though PDVl and PDV2 have partially redundant functions in ARC5 recruitment (Miyagishima, Froehlich, et al. 2006), their exact functional relationship in the plastid division complex is not yet clear. In contrast to ARC6 and PDV2, we did not detect an interaction between ARC6 and PDVl in yeast two-hybrid assays (Figure 3.5, row 1). However, preliminary experiments suggest that ARC6 is required for localization of PDVl (Glynn, Froehlich, et al. 2008) and PDV2 (Figure 3.13), consistent with the lack of ARC5 recruitment in an arc6 mutant (Figure 3.15). We do not believe that ARC6- dependent localization of PDVl is mediated directly by PDV2 because GFP-ARC5 localizes properly in pdv2-I (Miyagishima, Froehlich, et al. 2006). Rather, the data suggest that another factor acts downstream of ARC6 to position PDV] , and hence ARC5, independently of PDV2 (Glynn, Yang, et al. 2009). In this context, it is interesting to note that PDVl and PDV2 appear to have distinct localization patterns at the division site: GFP-PDVl is observed in discrete foci (Miyagishima, Froehlich, et al. 2006, Yang, Glynn, et al. 2008), whereas YFP-PDV2 is observed as a continuous ring (Figure 3.2). The significance of this difference is unknown, but it could reflect their interactions with different positioning factors — ARC6 in the case of PDV2 and another factor in the case of PDV]. The PDV2-interacting IMS region of ARC6 is mostly conserved with the corresponding region of the ARC6 cyanobacterial ortholog Ftn2, yet PDV2 is not found in cyanobacteria or green algae. Sequence alignments suggest the presence of land-plant- specific motifs within the IMS region of ARC6 that may have evolved in parallel with the emergence of PDVl and PDV2 in land plants (Glynn, Yang, et al. 2009). Ftn2 is 102 localized to the division site in cyanobacteria (Mazouni, Domain, et al. 2004) and presmnably has a topology similar to that of ARC6 (i.e. its C—terminus protrudes into the cyanobacterial periplasm). This leads us to ask: did PDV2 replace a periplasmic component of the cyanobacterial divisome (for example, a component of the peptidoglycan synthesis machinery or the murein layer itself) or does the ARC6-PDV2 interaction represent a completely new plant-specific function for ARC6? Identification of the precise boundaries of the PDV2-binding domain of ARC6 and rigorous sequence comparisons of ARC6 and Ftn2 family members will likely provide further insight into the evolution and operation of the plastid division machinery in land plants. 103 Materials and Methods Plant transformation vectors and analysis ARC6-GFP and ARC6AIMS-GFP were cloned into pCAMBIA-l302 (Hajdukiewicz, Svab, et al. 1994) or a derivative of pCAMBIA-1302 containing the ARC6 promoter region (Vitha, Froehlich, et al. 2003) using NcoI-BglII sites for insertion. PD V2pr0- YFP-PD V2 was generated by removing the 35S promoter and mGFP-His DNA sequences from pCAMBIA-1302 and replacing them with a fragment carrying the PD V2 promoter region and E YFP (Clontech) fused to the PD V2 coding sequence. For ARC6-CFP, ARC6 was cloned into a derivative of pCAMBIA-l 302 containing the ARC6 promoter region (V itha, F roehlich, et al. 2003) and ECFP coding sequence (Clontech). Lines expressing proteins from both PD V2pro- YFP-PD V2 and ARCopm-ARC6-CF P were generated by crossing independently transformed T1 lines (with demonstrated transgene expression) and selecting progeny with both YFP and CFP expression by epifluorescence microscopy. PD V2pro-PD V2 and PD VZPm-PD V20307D were generated by PCR and cloned into a derivative of pCAMBIA-l 302 that lacks the 35S promoter and mGF P coding sequence. All plant transformations were performed as previously described (Clough and Bent 1998) using A grobacterium tumefaciens GV3101. Selection of T1 individuals was performed on MS containing hygromycin (20-25ug/mL). 104 Microscopy and Image Analysis Light micrographs depicting chloroplast morphology in expanded leaf cells were taken using DIC Optics on a Leica DMI3000B Inverted Microscope outfitted with a Leica DFC320 Camera. Samples for chloroplast morphology and quantitation were prepared and analyzed using established protocols (Pyke and Leech 1991). Fluorescence micrographs were taken using a Leica DMRA2 using Q-Capture Camera Control Software (Q-Imaging) and the filter sets indicated (Leica) as previously described (V itha, Froehlich, et al. 2003). Image analysis and RGB composites were made using ImageJ v1.37 (NIH) (Bearer 2003). T wo-Hybrid Analysis Yeast strain AH109 (Clontech) was cultured and transformed as recommended by the manufacturer using standard Synthetic Dropout (SD) Media (Clontech) as indicated. ARC6 A A 637-801 was cloned into pGADT7 using Ndel-Xmal and PDV2 A A233-307 was cloned into pGBKT7 using Ndel-Xmal sites. PDV2 A A 233-307 ((33079) was generated by PCR-based mutagenesis and cloned into pGBKT7 using Ndel-Xmal sites. Pulldown Assays Recombinant His-ARC6 (encoding ARC6 A A 637-801 with an N-terminal 8X His- tag) was generated in pHIS8-3 (Salk Institute, La Jolla, CA) and expressed in BL21 (DE3) Codon Plus Cells (Stratagene). 2mM IPTG was applied to the cells at OD600 ~ 0.8 and incubated for 4h to generate the His-ARC6IMS fusion protein. Total protein was 105 extracted from the induced cells by sonication in Buffer A (1X TBS, 40mM imidazole, and 1 Roche EDTA-free protease inhibitor cocktail tablet #11836170001 per 100 mL Buffer A) and treatment of the sonicated material with 0.5% Triton X-100 for 30 minutes at room temperature. Following centrifugation at 18000 x g, the supernatant was collected and analyzed by Bradford assay. 750 ug of total protein was applied to 50 uL of Ni-Sepharose beads equilibrated in Buffer A (N i-Sepharose 6 Fast Flow, GE Healthcare, Uppsala, Sweden). Beads were washed 4 times in Buffer A to enrich for His- ARC61MS. Negative controls (beads only) were simply washed in Buffer A following equilibration. Production of crude GST-PDV2 (encoding PDV2 A A 233-307 with an N- terrninal GST tag) and GST-G307D (encoding GST-PDV2 A A 233-307(G307D)) was performed using the methods described above, except pGEX-4T-2 (GE Healthcare) was used as the parent vector for cloning and expression of recombinant PDV2 proteins. The cell extracts containing the GST fusions were prepared by sonication and treatment with Triton X-100 and then centrifuged to yield a crudely purified sample. Total protein was measured by Bradford assay and 750 ug of total protein was applied to the naked Ni- Sepharose beads or Ni-Sepharose beads coated with His-ARC61MS. Following a 2.5 h incubation on a rocking platform, the samples were washed several times with Buffer A and then eluted into 250 pL Buffer B (Buffer A containing 1M imidazole). Approximately 240 uL of eluate was recovered following a brief spin and total eluted material was precipitated in acetone and reconstituted in 6X sample buffer prior to SDS- PAGE. Approximately 12.5% of the eluted material was applied in each lane for separation on 15% SDS—Polyacrylarnide gels. Separated proteins were transferred to 106 nitrocellulose (GE Water and Process Technologies) and blotted with anti-His (Pierce, #15165) or anti-GST (Sigma, #G-l417) antibodies as recommended by the manufacturer and detected by standard chemilluminescence methods using the manufacturer’s HRP- based detection protocol (Pierce). Immunoblotting of Plant Material Proteins from whole-cell extracts (equivalent to ~2-3 mg liquid-nitrogen ground tissue) were prepared as previously described (Wiegel and Glazebrook 2002), separated by SDS-PAGE, and transferred to nitrocellulose. Anti-GFP Immunoblots were performed using Clontech JL-8 anti-GFP monoclonal antibody at 1:1000 in 5% nonfat dry milk in TBS-T, pH 7.4 (Miyagishima, Froehlich, et al. 2006). Anti-PDV2 Immunoblots were performed using 1 mg/mL Protein-A purified IgG’s from New Zealand White rabbit sera (CRP, Inc.) directed against PDV2AA1_212 at 1:15000 in 5% nonfat dry milk in TBS-T, pH7.4. Recombinant PDV2 (rPDV2) used to generate the immunogen used for anti-PDV2 antibody production was generated by cloning the coding sequence for PDV2 A A 1-212 into pHIS8-3 (Salk Institute, La Jolla, CA), expressing in BL21 (DE3) Codon Plus Cells (Stratagene), and purifying the resulting fusion protein from cell extracts using Ni Sepharose 6 Fast Flow (GE Healthcare, Uppsala, Sweden) according to the manufacturer’s recommendations. Phylogenetic Analysis ClustalW2 was used to generate a multiple sequence alignment using PDVl and PDV2 sequences from the organisms shown. Scoring was based on an identity (ID) 107 matrix with default settings: Protein Gap Open Penalty = 10.0; Protein Gap Extension Penalty = 0.2; Protein ENDGAP = -1; Protein GAPDIST = 4. Phylogeny of the PDV proteins (Glynn, Froehlich, et al. 2008) was calculated using MEGA 4 (Tamura, Dudley, et al. 2007). Chloroplast Import, Fractionation, and Protease Protection Assays The cDNA for PDV2 was subcloned into pDEST14 (Invitrogen) according to the manufacturer’s protocol prior to in vitro transcription translation. In vitro-produced protein was generated using a Promega TnT® Coupled Reticulocyte Lysate System (Promega, Madison, WI) according to the manufacturer’s protocol using 3H—labeled leucine (Perkin-Elmer) as a marker for PDV2 and 35S-labeled methionine (Perkin-Elmer) to track ARC6 protein. Pea chloroplasts were isolated, and import assays and fractionation were performed according to previously established protocols (Jackson, Froehlich, et al. 1998, Tranel, Froehlich, et al. 1995, Vitha, Froehlich, et al. 2003). Accession Number Genbank accession numbers used for multiple sequence alignments and phylogenetic analysis are as follows: Arabidopsis thaliana PDVl (AAM64850), Gossypium hirsutum PDVl (DW511385 and DW512027), Helianthus tuberosus PDVl (EL464676), Oryza sativa PDVl (NP_001042451), Populus trichocarpa PDVl (ABK94742), Vitis vinifera PDVl (CAO69353), Zea mays PDVl (EE161990 and DR959634), Arabidopsis thaliana PDV2 (NP_028242), Asparagus oflicinalis PDV2 (CV287540), Aquilegia species PDV2 (DT728284 and DT749563), Lactuca sativa PDV2 108 (DY983 634), Oryza sativa PDV2 (EA202618), Populus trichocarpa PDV2 (a GENSCAN based prediction (http://genes.mit.edu/GENSCAN.html) of Populus trichocarpa Scaffold LG__IX 3439000-3442500), and Triticum aestivum PDV2 (CK209373 and BJ311269). Acknowledgements I thank John Froehlich and Shin-ya Miyagishima for their advice dining the early phases of this project. I thank John Froehlich for persevering through the difficulties we experienced in translating PDV2 in vitro and for his outstanding work that helped us confirm PDV2 topology in isolated chloroplasts. I also thank Ronit Knopf and Zach Adam of The Hebrew University in Israel for sharing data with us on native PDV2 topology in Arabidopsis prior to publication. Most of this chapter and the figures therein are reproduced from (Glynn, Froehlich, et al. 2008) (http://www.plantcell.orgD and is Copyright American Society of Plant Biologists. ASPB grants to authors whose work has been published in The Plant Cell the royalty-free right to reuse images, portions of an article, or full articles in any book, book chapter, or journal article of which the author is the author or editor. 109 Chapter 4 Characterization of the PDV2 Binding Domain of ARC6. 110 Abstract During plastid division, interaction between ARC6 and PDV2 is required to position PDV2 within the outer envelope membrane. PDV2 helps recruit the dynamin- like protein, ARC5, to the division site. Previously, we showed that a carboxy-terminal glycine of PDV2 resides within a conserved motif and is required for ARC6-PDV2 interaction and PDV2 function in vivo. However, the specific domains or motifs within ARC6 that contribute to PDV2 binding were not determined. Here, we isolate the PDV2- binding domain of ARC6 (ARC6pBD) and perform comparative sequence and structural analysis of both PDV2IMS and ARC6pBD. Multiple sequence alignments and structural models highlight a plant-specific serine residue within ARC6pBD whose modification might be important for controlling operation of the chloroplast divisome in vivo. Further, we show that single amino acid substitution at this conserved serine within ARC6 affects its ability to bind PDVZIMS, leads to arc5-like chloroplast morphology, and upregulates Z-ring assembly within the stroma. Our results hint at a post-translational modification of ARC6 that fine-tunes coordination of the stroma] and cytosolic plastid division machineries in land plants. 111 Introduction The posttranslational modification of proteins can have profound consequences on their function (Huber and Hardin 2004). Several types of modification can occur in plants, including ubiquitylation (Dreher and Callis 2007), acetylation (He and Amasino 2005), methylation (Fischer, Hofrnann, et al. 2006), lipoylation(Ewa1d, Kolukisaoglu, et al. 2007), and phosphorylation (Christie 2007). Amongst these types of modification, the phosphorylation and dephosphorylation of serine residues is a common theme amongst proteins involved in signal-transduction (Braun and Walker 1996, Michniewicz, Zago, et al. 2007, Sokolovski, Hills, et al. 2005). Chloroplast division in plants and algae is mediated by two ring-like structures, both of which are composed of polymer-forming GTPases: (1) the FtsZ ring (Z-ring) on the stromal face of the inner envelope (McAndrew, Froehlich, et al. 2001, Vitha, McAndrew, et al. 2001) and (2) the ARC5 ring on the cytosolic face of the outer envelope membrane (Gao, Kadirjan—Kalbach, et al. 2003). The Z-ring is made up of two non-redundant FtsZ proteins, Ftle and FtsZ2 (McAndrew, F roehlich, et al. 2001, Schmitz, Glynn, et al. 2009, Vitha, McAndrew, et al. 2001). ARC5 is a dynamin-like protein involved in the final squeeze that separates the two daughter organelles during plastid division (Gao, Kadirjan-Kalbach, et al. 2003, Miyagishima, Nishida, et al. 2003a, Yoshida, Kuroiwa, et al. 2006). In land plants, ARC5 is recruited to the surface of the plastid by the outer envelope proteins PDVl and PDV2. While either PDV] or PDV2 112 can recruit ARC5 to the surface of the plastid, both PDVl and PDV2 are required for full ARC5 contractile activity in vivo (Miyagishima, Froehlich, et al. 2006). ARC6 is a bitopic chloroplast inner membrane protein (V itha, Froehlich, et al. 2003) that coordinates operation of the internal and external plastid division complexes (Glynn, Froehlich, et al. 2008). In the stroma, the N~terminus of ARC6 facilitates FtsZ ring (Z-ring) assembly (Vitha, Froehlich, et al. 2003); the Z-ring is a critical component of the division apparatus, acting as a scaffold for other division proteins and providing nominal contractile force during division (Erickson 2009, Osawa, Anderson, et al. 2008). The intermembrane space (IMS)-localized portion of ARC6 binds PDV2 directly and positions PDV2 at the division site; PDV2 is required for full dynamin (ARC5) activity during the final steps of chloroplast division (Glynn, Froehlich, et al. 2008, Miyagishima, Froehlich, et al. 2006). Here, we characterize the PDV2 binding domain of ARC6 (ARC6pBD) and identify a predicted serine phosphoacceptor within this domain that may modulate ARC6 activity in response to serine modification. We show that ARC6pBD occupies amino acids 636-759 and that serine 744 might play a critical role in both mediating ARC6- PDV2 interaction within the IMS and in regulating Z-ring assembly within the stroma. ARC6S744 and the surrounding motif is found only in ARC6 orthologs from embryophytes, the lineage in which PDV2 first emerged (Glynn, Froehlich, et al. 2008, Miyagishima, Froehlich, et al. 2006). Phosphomirnetic mutation of serine 744 (S744E) 113 weakens ARC6-PDV2 interaction, leads to arc5-like plastid division defects, and perturbs stromal Z-ring structure in vivo — but strangely, this mutation does not affect localization of PDVl, PDV2, or ARC5 at the division site. Our results suggest that persistent phosphorylation of ARC6S744 slows recruitment of PDV2, but leads to profound upregulation of Z-ring assembly within the stroma. We propose that phosphorylation of ARC6S744 is a regulatory feature that retards the recruitment of PDV2 until the Z-ring is fully established and ready for envelope constriction to occur. Results The PD VZ-Binding Domain of ARC 6 (ARC 61031)) Occupies Amino Acids 636- 759. We previously showed that the C-terrninal IMS-localized regions of ARC6 and PDV2 interact; this interaction serves to position PDV2 within the outer envelope membrane during division (Glynn, F roehlich, et al. 2008). While we showed that this interaction required the conserved terminal glycine residue of PDV2, the discrete domain within ARC6 that binds PDV2 was not defined. To gain further understanding of ARC6 function and its constitutive domains, we generated a series of two-hybrid deletion constructs for ARC61MS (Figure 4.1). Following yeast transformation and selection on appropriate medium, we assayed for growth on dropout medium containing or lacking histidine; growth on medium lacking histidine is consistent with the presence of a protein-protein interaction between bait and prey leading to activation of the HIS3 reporter gene. Based on the growth ratios, we observed that full interaction 114 INTERACTION PAIR ARC6 AA636-801 VS- PDV2 AA233-307 S—D/-ULT S___D/-ULTH ._. ARC6 AA636-759 VS- PDV2 AA233-307 N ARC6 AA636-728 VS- PDV2 AA233-307 U.) ARC6 AA648-801 VS- PDV2 AA233-307 A ARC6 AA661-801 VS- PDV2 AA233-307 \1 kn ARC6 AA636-801 VS. EMPTY ON ARC6 AA636-759 vs. EMPTY ARC6 AA636-728 VS. EMPTY 0° ARC6 AA648-801 vs. EMPTY \O ARC6 AA66l-801 VS. EMPTY O EMPTY vs. PDV2 AA233-307 ._r ._n EMPTY vs. EMPTY l2 1 1/10 1/100 1 1/10 1/100 Figure 4.1. The PDV2 binding domain of ARC6 (ARC6pBD) occupies AA 636-759. Yeast two-hybrid HIS reporter assays showing interaction strength between the IMS region of PDV2 (AA 233-307) and fragments of the IMS region of ARC6. Left hand column shows growth on synthetic dropout (SD) medium supplemented with histidine and right hand column shows growth on medium lacking histidine. For a schematic of the ARC6 protein, refer to Figure 1.4. Dilutions from a starting culture of OD600 = 1.0 are indicated at the bottom of panel. EMPTY = Empty vector control. AA = amino acids. 115 between ARC6 and PDV2 only requires ARC6 A A636-759 (Figure 4.1). We call this region of ARC6 the PDV2-binding _d_omain (ARC6pBD). The reporter activation of ARC6pBD is comparable to that observed for the full length C-terminus of ARC6, ARC6 A A636-801 (Figure 4.1). PDV2 does not interact with ARC6AA661-301 nor ARC6AA536-723, but weakly interacts with ARC6 A A648-801 (Figure 4.1), suggesting that the residues adjacent to the transmembrane domain of ARC6 are involved, but not critical for ARC6-PDV2 interaction. Structural analysis of ARC6 1931) and PD VZIMS- To gain additional insights into the interaction between ARC6-PDV2 and guide site-directed mutagenesis, we attempted to generate a model of ARC6pBD. ARC6pBD was best-matched to the known structure of the Tim44 C-terminal domain (2CW9) using the meta-server at http://bioinfo.pl (Ginalski, Elofsson, et al. 2003). However, attempts to model ARC6pDB onto the structure of 2CW9 were unsuccessful, yielding several gaps between residues in the main chain of ARC6pBD (data not shown), suggesting that this is probably not a valid structural comparison. From this, we conclude that ARC6pBD probably has a structure unique from any of the structures currently present within the Protein Data Bank (http://wwwpdbcrgD. 116 We also attempted to generate a homology model for PDVZIMS, hoping that it might reveal insights into both PDV2 and ARC6 function. Using the meta-server at http://bioinfo.pl/, we determined that the best structural match for PDVZIMS is 2J6Z (1 W53), the N-terminal domain of a phosphoserine phosphatase that is involved in stress response in Bacillus subtilis and other gram-positive bacteria (Delumeau, Dutta, et al. 2004, Hardwick, Pane-Farre, et al. 2007). We made a pairwise comparison between RSbUNT and PDVZIMS (Figure 4.2) and used this alignment to generate a homology model of PDV21MS, using identical residues as anchor points for the PDVZIMS model. The PDVZIMS sequence was highly similar to RSbUNT and was well-fit to 2J6Z (Figure 4.3, inset), suggesting that these two proteins might have similar functions. The N-terrninus of Rst is required for binding the Rst stress response serine kinase (Dutta and Lewis 2003, Kang, Vijay, et al. 1998, Yang, Kang, et al. 1996), but RSbUNT does not appear to have kinase or phosphatase activity itself— this activity is thought to be exclusive to the C-terminal domain of Rst (Delumeau, Dutta, et al. 2004, Kang,_Vijay, et al. 1998). In this case, the C-terminus of ARC6 might be analogous to Rst— perhaps even possessing a kinase or phosphatase activity itself. From this rudimentary structural analysis, we hypothesize that PDV2 may have a structure akin to a known protein phosphatase or may influence the activity of its binding partner through a simple protein-protein interaction, analogous to the Rst-Rst signal transduction cascade in 117 RstNT I HQL S YI . ETS QAQ PDV2IMS S GPE R ‘ LF PHR RstNT KFS TIE HRK PDV2IMS ATR RTP Euvl EDG RstNT YPSLP D DFLI MI PDV2IMS VKER I VAK T Figure 4.2. Pairwise alignment between the N-terminal domain of Rst and the IMS region of PDV2. The N-terminal domain of Rst used for crystal structure of 2J6Z (RstNT, amino acids 7-81) and IMS region of PDV2 (PDVZIMS, amino acids 233-307) are aligned using a BLOSUM62 scoring matrix. Boxed positions indicate sequence similarity and shaded positions highlight sequence identity. 118 Figure 4.3. Structural model of PDVZIMS. The IMS region of PDV2 (amino acids 233-307, main panel) was modeled using the known structure of RSbUNT (2J6Z/1 W53, inset panel). Shown above are cartoon representations ofthe two structures. The side chains of conserved residues within the C-terminal extension of PDV2 proteins (see Figure 3.6) are highlighted in the PDVZIMS structural model. 119 some gram positive bacteria (Delumeau, Dutta, et al. 2004, Kang, Vijay, et al. 1998, Yang, Kang, et al. 1996). This IMS-localized interaction between ARC6 and PDV2 might modulate posttranslational modification of ARC6, thereby altering the activity of ARC6 within the IMS or chloroplast stroma. A Conserved Serine within ARC 6 pm) Influences ARC6-PD V2 Interaction. Because of the possible connection to protein kinase/phosphatase activity, we closely examined alignments between tracheophyte ARC6pBD and the corresponding region of cyanobacterial F tn2 proteins for embryophyte-conserved residues that might act as phosphoacceptors (Figure 4.4), as PDV proteins are found only in land plants. We identified a plant-specific motif (Figure 4.4, red bar) within the ARC6pBD that was conserved amongst higher plants, but missing from or poorly aligned with algal ARC6 and cyanobacterial F tn2 sequences; all ARC6 orthologs fi'om higher plants possessed a serine near the middle of this motif (Figure 4.4, arrowhead indicates S744 in AtARC6); serine residues are common sites of posttranslational modification in eukaryotes. Because algae and cyanobacteria do not encode orthologs of PDV2, this embryophyte- specific serine residue (ARC63744), or other residues within the surrounding motif, might be targets of ARC6 posttranslational modification that have coevolved with emergence of PDV2 in land plants. 120 * . SVTVSADGT AtARC6 L . S OsARC6 . SITISLDGR T. A PtARC6 . SVTVSVDGLS V . S AlARC6 . SVTVSADGT L . S BnARC6 . SVTVSADGT L . S RsARC6 . SVTVSADGT L . S GrARC6 . SVTLSLDGQ V . S lnARC6 . SVTVSVDGR I . S CiARC6 . SVTISLDGR V . S CcARC6 . SVTVSVDGQ I . L CsARC6 . SVTLSQEGR W . S OlARC6 .SVQVIG.TE F. V CrARC6 LKRVTHKGA F S GthnZ . QVADRR . PD A . V NthnZ . VEKIGLFAD A . V SeFtn2 .VQLSDG .DQ V. V CsFth . FKPDSNNPN V. V NsFth .FDQK. . .SD V. V LsFth .VQIDSSNPN F . T Figure 4.4. Identification of a plant-specific motif within the C-termini of ARC6/F m2 family members. A segment of a CLUSTALW alignment comparing the C-termini of ARC6 and Ftn2 family members is shown above and corresponds to Arabidopsis thaliana ARC6 amino acids 729-759. The red bar above the sequences indicates an amino acid motif unique to higher plants that may be involved in ARC6- PDV2 interaction, based on two hybrid results (Figure 4.1). The black arrowhead points to a serine (S744) predicted to be a phosphoacceptor (see text). Arabidopsis thaliana ARC6 (AtARC6); Oryza sativa cv. japonica (OsARC6); Populus trichocarpa ARC6 (PtARC6); Arabidopsis lyrata (AlARC6); Brassica napus ARC6 (BnARC6); Raphanus sativus (RsARC6); Gossypium raimondii (GrARC6); Ipomoea nil (InARC6); Cichorium intybus (CiARC6); Coffea canephora (CcARC6); Citrus sinensis (CsARC6); Ostreococcus lucimarinus CCE9901 (OlARC6); Chlamydomonas reinhardtii (CrARC6); Gloeobacter violaceus PCC7421 (Gthn2); Nostoc punctiforme PCC 73102 (NpF m2); Synechococcus elongatus PCC 6301 (SeFtn2); Cyanothece sp. CCYOl 10 (CsFtn2); Nostoc sp. PCC 7120 (NsFtn2); and Lyngbya sp. PCC 8106 (LsFtn2). Label colors in the lefthand column indicate groups of organisms: red = land plants, green = green algae, and blue = cyanobacteria. Boxed regions indicate sequence similarity in >70% of the lineages shown. Shaded regions indicate conserved residues. 121 To determine if ARC6S744 was a potential phosphoacceptor, we queried two phosphorylation-prediction programs, PhosPhAt (Heazlewood, Durek, et al. 2008) and NetPhos (Blom, Garnmeltoft, et al. 1999) to see if ARC6S744 might be a target for phosphorylation. Both of these programs gave significant scores to ARC6S744 (PhosPhAt, 0.496; NetPhos, 0.932), suggesting that ARC6S744 may be phosphorylated in vivo. Notably, both S740 and T742 also yielded significant scores using one, but not both, of these prediction algorithms — suggesting that other residues within the plant- specific motif of ARC6pBD (Figure 4.4) might also be targets for posttranslational modification. To determine if phosphorylation of ARC6S744 might affect binding of PDV2, we generated site-directed mutants of ARC61MS, targeting the serine at position 744 in two- hybrid vectors. We named these mutant proteins ARC61MS ($744 A) and ARC61MS (S744E), which mimic the unphosphorylatable (S744A) and constituitively phosphorylated (S744E) forms of the protein, respectively. Our results showed binding strengths consistent with PDV2 affinities as ARC6IMS > ARC63744 A > ARC6S744E (Figure 4.5A), indicating that PDV2 has low affinity for the phosphomimetic form of ARC6, ARC6S744E. We validated this result using pulldown assays with recombinant Hiss-ARC61MS and GST-PDV2IMS proteins, which confirmed that ARC61MS (874415) 122 A INTERACTION PAIR ARC61MS VS. PDVZIMS SD/-ULT SD/-ULTH ._‘ EMPTY vs. PovleS N ARC6IMS vs. EMPTY DJ ARC61MS (S744A) VS. PDVZIMS 4:- ARC6IMS (S744A) VS. EMPTY ARC6IMs (S744E) VS- PDVZIMs O\ ARC61MS (S744E) VS. EMPTY \l EMPTY vs. EMPTY (I! 00 1 1/10 1/100 1 1/10 1/100 1 2 3 4 l... .- - Ia-HIS I" ‘. .: lot-GST Figure 4.5. A phosphomimetic mutation in the IMS region of ARC6 (ARC6S744E) decreases its affinity for PDV2. (A) Site-directed mutations in ARC61MS were tested for their ability to interact with PDV21MS by yeast two-hybrid. Interaction strengths are as follows: ARC6-PDV2 (0.64); ARC61MS (S744A)-PDV2 (0.54); ARC61MS (57445)- PDV2 (0.14); all negative controls (5 0.01). Relative interaction strength was determined by ratio of growth on —HIS/+HIS synthetic dropout (SD) media. (B) Confirmation of two hybrid results by pulldown assay using immobilized ARC6 proteins treated with equal amounts of GST-PDVZIMS. Immunoblots of eluates from pulldowns are shown: Hiss- ARC61MS (lane 1), H153-ARC6IMS (S744A)a H188-ARC61MS (37445), or uncoated beads treated with GST-PDV21MS (lane 4). 123 has significantly lower affinity for PDVZIMS than the wild type (ARC6IMS) protein (Figure 4.58). ARC6IMS (S744A) had slightly decreased affinity for PDVZIMS in two- hybrid assays and slightly increased affinity for PDVZIMS in pulldown assays, suggesting this mutation may not perturb ARC6 function. A Phosphomimetic Mutation in ARC 61231) Causes Chloroplast Division Defects in vivo. Because of the profound effect of our phoshphomimetic mutation (ARC6S744E) on ARC6-PDV2 interaction, we wanted to know if this mutation was detrimental to the division process in vivo. Therefore, we generated a series of transgenes encoding wild type (ARC6), unphosphorylatable (ARC6S744A), and phosphomimetic (ARC6S744E) versions of ARC6; all of these were driven by the native ARC6 promoter sequence. We transformed wild type (Col-0) and arc6 (SAIL_693_G04) plants with each of these transgenes using Agrobacterium (Clough and Bent 1998) and selected T1 individuals using hygromycin. We examined chloroplast morphology in independent T1 lines that were expressing the transgene within the range that should complement the arc6 mutant, based on the expression of the control (wild type) transgene (Figure 4.6). Both the wild type (ARC6) protein and the unphosphorylatable (ARC6S744A) proteins were able to complement the arc6 mutant when expressed at similar levels. However, ARC6S744E was unable to fully complement arc6 mutants (Figure 4.6) and imposed a dominant- negative effect upon Col-0 transformants (data not shown), suggesting that this transgene 124 A l 2 3 4 5 6 7 8 9 10 “7“»:- * tag—..-s. i 97kD>“? [i a-ARCO 117kD> =nlah'xv.:.s‘...'.'..$‘ ninth,» B 97kD>Ei _ ' j HI Ponceau 70 60 l = Q O In 0 - Q- m H m N — Q: G i- O _ i: U Col-0 arc6 arc6 + arc6 + arc6 + ARC6 ARC6S744A ARC65744E Figure 4.6. ARC6S744E is dysfunctional. (A) Sample of immunoblot data used to quantitate ARC6 protein levels in transformed arc6 plants. Marker (lane 1); arc6 (lane 2); Col-0 (lane 3); four independent arc6 T1 transformants complemented by ARC6 transgene (lanes 4-7); 3 independent arc6 T1 ARC6S744E transformants (lanes 8-10). Protein levels in ARC6S744 A lines were quantitated on a separate immunoblot (not shown). (B) Analysis of chloroplast number in expanded leaves of Col-0, arc6, and transgenic lines from 10 cells with average areas of 3109 d: 443 umz. Lines used for quantitative analysis in panel B are marked with asterisks (*) in panel A. Error bars = standard deviation. 125 was not functional in vivo, consistent with its low affinity for PDV2 in pulldowns and two-hybrid assays (Figure 4.5). Interestingly, many of the lines expressing ARC 657441.; contained elongated and partially-constricted chloroplasts, indicative of a defect in PDV or ARC5 protein function (Figure 4.7E, arrowheads) (Gao, Kadirjan-Kalbach, et al. 2003, Miyagishima, Froehlich, et al. 2006). Taken together, we conclude that ARC6S744 is probably important for the ARC6-PDV2 interaction in vivo, but cannot yet conclude that this residue is actually subject to post-translational modification in vivo. PD V2, PD VI, ARC5 and F tsZ localize to the division site in ARC 657441.; mutants. To determine the effect of our phosphomimetic mutation upon the characteristics of other division components, we examined our transgenic lines and controls for localization of PDV2, PDVl , ARC5, and F tsZ proteins. Surprsingly, despite their lower affinity for PDV2 and perturbation in chloroplast morphology, ARC6S744E mutants were still able to localize YFP-PDV2 to rings at sites of chloroplast constriction (Figure 4.8C). This conflicts with the reduced affinity of ARC6S744E for PDVZIMS in binding assays (Figure 4.5), but might reflect the ability of ARC6S744E to bind PDV2, albeit with less stringency than the wild type protein (Figure 4.5). 126 ...- Figure 4.7. Phenotypes observed in lines expressing ARC6S744A and ARC6S744E. Mesophyll cells from expanding leaves are shown for Col-0 (A); arc6 (B); arc6 expressing transgenic ARC6 (C); arc6 expressing transgenic ARC 65744 A (D); and arc6 expressing transgenic ARC6S744E (E). Arrowheads in (E) indicate elongated and partially constricted chloroplasts. Scale bars = 10 pm. 127 PDV2 ARC5 PDVl FmZZ Figure 4.8. Localization of PDV2, ARC5, PDV], and FtsZ in ARC6S744E-expressing transgenics. All fluorescently-labeled proteins were expressed using their respective native promoters. Merged images showing chlorophyll autofluoresence (red) and the indicated fluorescent marker (green) are shown. YFP-PDV2 localization in (fol-l) (A); arc6 (B): and ARC657445 (C). GFP-ARC5 in Col—0 (D); arc6 (E); and ARC6S744E (F). GFP-PDVl in Col—0 (G); arc6 (H); and ARC7657445(1). Immunolabeling of FtsZ2-l in Col-U (J); arc6 (K); and .4RC6S7445 (L). Scale bars = 5 pm. 128 To determine if the division defect observed in ARC6S744E transgenics was a result of an indirect effect upon ARC5 or PDVl localization, we examined GFP-ARC5 and GFP-PDVl localization in lines ARC6S744E transgenics. Both GFP—ARC5 (Figure 4.8F) and GFP-PDV] (Figure 4.81) were localized to the division site, though GFP-PDVl was not observed in a ring in lines expressing ARC6S744E as it is in wild-type plants (Miyagishima, Froehlich, et al. 2006). Because Z-ring morphology is affected in the dumbbell-shaped plastids observed in pdvl , pdv2, and arc5 mutants, we examined Z-ring morphology in ARC6S744E mutants using an FtsZ2-specific antibody (Figure 4.8L). In mutants with defects in dynamin activity, we typically observe multiple Z-rings at the division site, presumably as a feedback response to the defect in dynamin activity at the outer envelope (Gao, Kadirjan- Kalbach, et al. 2003, Miyagishima, Froehlich, et al. 2006, Yang, Glynn, et al. 2008). Consistent with a defect in dynamin activity (Gao, Kadirjan-Kalbach, et al. 2003, Miyagishima, Froehlich, et al. 2006), ARC6S744E mutants have multiple Z-rings at the division site (Figure 4.8L). Taken together, we conclude that ARC6S744 is important for ARC6-PDV2 interaction, indirectly affecting dynamin (ARC5) pinchase activity at the outer envelope through PDV2, as ARC6S744E mutants have defects in chloroplast and FtsZ filament morphology similar to pdvl , pdv2, and arc5 mutants (Gao, Kadirjan- Kalbach, et al. 2003, Miyagishima, Froehlich, et al. 2006). 129 Discussion Here, we identify and characterize the PDV2-binding domain of ARC6 (ARC6pBD). ARC6pBD occupies only a portion (AA 636-759) of the IMS-localized C- terminal region of ARC6. While our structural analysis of ARC6pBD was not insightful, the IMS-localized C-terminus of PDV2 (PDVZIMS) modeled onto the known structure of Rst (Delumeau, Dutta, et al. 2004, Hardwick, Pane-Farre, et al. 2007), a bacterial phosphoserine phosphatase (Delumeau, Dutta, et al. 2004, Hardwick, Pane-Farre, et al. 2007). Because of the structural similarity of PDV2IMS to Rst and the prospect of posttranslational modification of ARC6, we examined alignments (Figure 4.4) within ARC6pBD for plant-specific motifs containing predicted phosphoacceptor sites. This analysis revealed that serine 744 (ARC6S744) as a high-probability phosphoacceptor within a plant-specific motif of ARC6pBD, based on NetPhos and PhosPhAt prediction programs. Our site-directed mutagenesis of this residue revealed that a phosphomimetic mutation (ARC6S7445) causes reduced affinity for PDV21MS; this same mutation leads to chloroplast division defects (Figures 4.6-4.8) similar to those seen in mutants with defects in dynamin (ARC5) activity (Gao, Kadirjan-Kalbach, et al. 2003, Miyagishima, Froehlich, et al. 2006). However, despite the reduced affinity for ARC6S744E for PDV2, arc6 mutants expressing an ARC6S744E transgene still recruit YFP-PDV2 and other division factors to the division site (Figure 4.8). 130 Our identification of the rough boundaries of the PDV2-binding domain is a first step in understanding how ARC6 and PDV2 interact with each other and how this interaction might be regulated. Further work is required to identify the precise boundaries of ARC6pBD, which might allow ftu'ther refinement of the current structure- function model for this domain. While structural similarity to Rst (2J6Z/1W53) is apparent, it is not yet clear if PDVZIMS possesses phosphatase or kinase activity and if ARC6pBD is truly a substrate for any posttranslational modification activity in vivo. While ARC6S744 might act as a phosphoacceptor in vivo, PDV2 itself probably does not modify this residue and likely uses ARC6pBD as a binding site, as the structural similarity to the Rst phosphatase is limited to the Rst-binding domain of Rst. However, modification of the plant-specific motif within ARC6pBD clearly reduces affinity for PDV2, suggesting that a posttranslational modification at serine 744 might be a mechanism by which PDV2 recruitment to the division site is modulated. Alternatively, serine 744 of ARC6 might be required for structural intergrity of ARC6pBD and our mutant proteins are simply misfolded, leading to reduced affinity for PDV2. Either of these scenarios might explain the prevalence of dumbbell-shaped chloroplasts in lines expressing ARC6S744E (Figures 4.7-4.8). Curiously, we occasionally observed chloroplasts within petiole cells that possessed several constrictions along their length in ARC6S7445-expressing lines (Figure 4.9), a phenotype 131 A R C 657445 l’etiole Figure 4.9. An unusual phenotype observed in ARC6S744E petioles phenocopies plastidic Min system defects. Multiple constrictions were observed along the length of elongated chloroplasts within large petiole cells of ARC6S744E transgenics. This phenotype has not been observed in arc5 mutants and may reflect ARC6-mediated downregulation of AtMinD (Colletti, Tattersall, et al. 2000, Fujiwara, Nakamura, et al. 2004) and/or upregulation of AtMinE activity within the chloroplast (Fujiwara, Hashimoto, et al. 2008, Maple, Chua, et al. 2002). Scale bar = 10 pm. 132 that resembles both atminD/arcI 1 loss-of-function mutants (Colletti, Tattersall, et al. 2000, Fujiwara, Nakamura, et al. 2004) and AtMinE overexpressors (Fujiwara, Hashimoto, et al. 2008, Maple, Chua, et al. 2002). The significance of this finding is not yet clear, nor is it understood why these phenotypes appear prominently in petiole cells, but the connection between ARC6 activity and the operation of the plastidic Min system is intriguing. The high proportion of elongated dumbbell-shaped plastids in the ARC6S744E transgenic lines caused us to look the localization of ARC5, PDVl, PDV2, and F tsZ. Surprisingly, the localization of each of these proteins to the mid-plastid was mostly unperturbed. However, PDV] was present exclusively as single foci associated with the chloroplast (Figure 4.81) in ARC6S744E transgenics, but never as a complete punctate ring, as is occasionally observed in wild type GFP-PDVl expressing lines (Miyagishima, Froehlich, et al. 2006, Yang, Glynn, et al. 2008). The reason for this is unclear, but the GFP-PDVl punctate ring is the most uncommon PDV] localization pattern observed; PDVl is typically observed as a single spot in wild type cells and more observation of these transgenic lines may be required. Additionally, while FtsZZ was localized to the mid-plastid in ARC6S744E lines, several adjacent Z-rings were observed (Figure 4.8L), reminiscent of the FtsZ2 localization pattern in pdvl, pdv2, and arc5 (Gao, Kadirjan- Kalbach, et al. 2003, Miyagishima, Froehlich, et al. 2006); all these mutants have defects in dynamin activity, which presumably generates a signal leading to the formation of several central Z-rings (Yang, Glynn, et al. 2008). Based on the Z-ring morphology in 133 our ARC6S744E transgenics, we propose that this signal is transduced through ARC6. Because of this arc5-like Z-ring phenotype in ARC6S744E lines, we propose that the dumbbell-shape chloroplasts in these lines result from a decrease of dynamin pinchase activity at the outer envelope, perhaps resulting from a decrease in ARC6-PDV2 affinity due to the phosphomimetic mutation in ARC6. Alternatively, upregulation of stromal Z- ring assembly in ARC65744 E mutants (Figure 4.8L) might impede dynamin-mediated envelope constriction, as the Z-ring must presumably either constrict or disassemble to allow the division process to conclude. How and why would ARC6S744 phosphorylation within the IMS affect Z-ring assembly and/or dynamin activity? One highly speculative possibility is that S744 phosphorylation occurs to slow or regulate the division process. Perhaps ARC6S744 is phosphorylated as the Z-ring is being assembled— this leads to controlled recruitment of PDV2 until the Z-ring is fully assembled within the stroma and ready to constrict. At this point, ARC6 molecules would then be progressively dephosphorylated at position 744, causing increased ARC6-PDV2 affinity and hence increased dynamin (ARC5) activity at the outer envelope with concurrent disassembly of the Z-ring in the stroma. If this is the case, it is unclear why we did not observe smaller chloroplasts in ARC6S744 A lines; perhaps this phosphoregulatory mechanism may only be relevant when PDV2 is overexpressed, functions only under a particular set of growth conditions, or the modification of additional residues within the plant-specific motif of ARC6pBD (such as 134 S740 and/or T742) might also contribute to regulating ARC6-PDV2 interaction. Regardless, this hypothesis is in-line with observations that the rate of plastid division is affected by the amount of PDV2 protein: less PDV2 leads to a slower rate of division and more PDV2 leads to a faster rate of division (Okazaki, Kabeya, et al. 2009). ARC6 is absolutely required for PDV2 localization at the division site (Figure 4.8B) and ARC6 is absolutely required for PDV2-mediated division activity, as overexpression of PDV2 in the arc6 background does not modify the number of chloroplasts observed within arc6 mutants (Glynn, Froehlich, et al. 2008). Presumably, the amount of PDV2 present at the division site is a rate-limiting step, with regard to organelle constriction —— and phosphorylation of ARC65744 might be a mechanism by which individual chloroplasts regulate PDV2 recruitment to the division site, providing a checkpoint for proper Z-ring assembly prior to dynamin-driven constriction of the organelle. Further work on this aspect of the division process might lead to the discovery of a novel auto- kinase/phosphatase activity for ARC6, or a separate novel kinase/phosphatase, that aids fine-tuning of the division process by controlling the recruitment of PDV2 to the division site. 135 Materials and Methods Yeast two-hybrid analysis. ARC6 A A 637-801 was cloned into pGADT7 using Ndel-Xmal and PDV2 A A233- 307 was cloned into pGBKT7 using Ndel-Xmal sites. Site-directed alleles of ARC6 were made by SOE-PCR (Warrens, Jones, et al. 1997) and cloned into pGBKT7 using Ndel- XmaI sites. Yeast strain AH109 (Clontech) was cultured and transformed as recommended by the manufacturer using standard Synthetic Dropout (SD) Media (Clontech) as indicated. Growth assays were conducted using previously-established protocols. Pulldown assays. All fusion proteins were expressed in E. coli BL21 (DE3) Codon Plus cells (Stratagene) induced at OD600 = 0.8 with 2 mM IPTG for 2 hours at 37°C. 750 ug of protein from induced cell extract was used for each bait/prey combination. Pulldowns between Hi83-ARC61MS and GST-PDVZIMs, ARC61MS (3744 A) and GST-PDVZIMs, or ARC61MS (874413) and GST-PDVZIMS were performed as in previous experiments (Glynn, Froehlich, et al. 2008), except Triton X-100 was present at 0.1% in all wash buffers to prevent clumping of the sepharose beads. 136 Structural modeling. A structural model of PDVZIMS was generated by comparison of PDV2 A A23 3- 307 to all current PDB structures (http://wwwpdborg) using the MetaServer at http://bioinfo.pl. The best-matched protein (in terms of its secondary structure) was aligned with the relevant query protein using CLUSTALW (Larkin, Blackshields, et al. 2007) and a PDB file containing data for a three dimensional model was generated using the homology modeling server at http://prOteinsmsuedu. The final images used in figures herein were rendered using PyMol 0.99rc6 (Delano Scientific). Generation of transgenic lines. ARC6S744 A and ARC6S744E transgenes were generated using SOE—PCR and placed into a derivative of pCAMBIA-l 302 containing the native ARC6 promoter using AvrII and BstEII restriction sites. Clones were sequence-verified and transformed into Agrobacterium tumefaciens GV3101 prior to transformation of Arabidopsis (Clough and Bent 1998). Trangenic lines were selected on Linsmaier-Skoog medium containing hygromycin (25 ug/mL). Only hygromycin-resistant individuals were transplanted to soil and used for further analysis. Immunoblotting. Plant extracts for determination of ARC6 protein level were taken from floral bud tissue. Flower buds were ground in liquid nitrogen and prepared for 10% SDS-PAGE using 6X sample buffer according to previously-established protocols (Wiegel and 137 Glazebrook 2002). 2.5 mg of homogenized tissue was loaded per lane. Immunoblotting of plant extracts was performed using an ARC6-specific antibody at 1:2500 in blocking buffer containing 5% nonfat dry milk (McAndrew, Olson, et al. 2008). The blot was washed several times in TBS-T before applying the secondary antibody. The anti-rabbit HRP-conjugated secondary was used at 1:5000 in TBS-T containing 5% nonfat dry milk and the blot was washed several times in TBS-T before applying the HRP chemilluminescent substrate (Therrno Scientific, Inc.) and exposing to film. Quantitative analysis of chloroplast number. Tissue preparation was carried out as described previously (Miyagishima, Froehlich, et al. 2006, Pyke and Leech 1991, Vitha, Froehlich, et al. 2003). Cell size and chloroplast number were measured using ImageJ v1.37 (NIH) (Bearer 2003). Quantitative analysis of phenotypes was performed using 10 cells of similar size. Microscopy. Light micrographs depicting chloroplast morphology in expanded leaf cells were taken using DIC Optics on a Leica DMI3000B Inverted Microscope outfitted with a Leica DF C320 Camera. Samples for chloroplast morphology and quantitation were prepared and analyzed using established protocols (Pyke and Leech 1991). Fluorescence micrographs were taken using a Leica DMRA2 using Q-Capture Camera Control Software (Q-Imaging) and the filter sets indicated (Leica) as previously described (V itha, Froehlich, et al. 2003). Image analysis and RGB composites were made using ImageJ v1.37 (NIH) (Bearer 2003). 138 Acknowledgements I thank Aaron Schmitz, Yue Yang, and Katherine Osteryoung for helpful discussions during the development and progression of this project. 139 Chapter 5 PARC6 Influences FtsZ Assembly and PDVl Recruitment in Arabidopsis. 140 Abstract Chloroplast division in plant cells is accomplished through the coordinated action of the tubulin-like F tsZ ring inside the organelle and the dynamin-like ARC5 ring outside the organelle. This coordination is facilitated by ARC6, an inner envelope protein required for both FtsZ assembly and ARC5 recruitment. Recently, we showed that ARC6 specifies the mid-plastid positioning of the outer envelope proteins PDVl and PDV2, which have parallel functions in dynamin recruitment. PDV2 positioning involves direct ARC6-PDV2 interaction but PDVl and ARC6 do not interact, indicating an additional factor functions downstream of ARC6 to position PDVl. Here, we show that PARC6 (Paralogue of ARC6), an ARC6-like protein unique to vascular plants, fulfills this role. Like ARC6, PARC6 is an inner envelope protein with its N-terminus exposed to the stroma and Arabidopsis parc6 mutants exhibit chloroplast and FtsZ filament morphology defects. However, whereas ARC6 promotes FtsZ assembly, PARC6 appears to inhibit FtsZ assembly, suggesting ARC6 and PARC6 function as antagonistic regulators of FtsZ dynamics. The FtsZ inhibitory activity of PARC6 may involve its interaction with the FtsZ-positioning factor ARC3. A PARC6-GFP fusion protein localizes both to the mid- plastid and to a single spot at one pole, reminiscent of the localization of ARC3, PDVl and ARC5. We show that PARC6 positions and binds PDVl, but PARC6 is not required for localization of PDV2 or ARC5. Our findings indicate that PARC6, like ARC6, plays a role in coordinating the internal and external components of the chloroplast division complex, but that PARC6 has evolved distinct functions in the division process. 141 Introduction The chloroplasts of vascular plants replicate by binary fission, similar to their cyanobacterial relatives, but utilize a machinery that is a composite of cyanobacterial and host-derived components (Aldridge, Maple, at al. 2005, Gao, Kadirj an-Kalbach, et al. 2003, Kuroiwa, Kuroiwa, et al. 1998, Maple, Vojta, et al. 2007, Miyagishima, Takahara, et al. 2001, Miyagishima, Wolk, et al. 2005, Osteryoung 2001, Osteryoung and Nunnari 2003, Osteryoung and Vierling 1995). The proteins that make up this dynamic complex, all of which are nuclear—encoded, occupy the stroma, intermembrane space (IMS), and cytosol, with several key factors acting at or across the envelope membranes (Gao, Kadirjan-Kalbach, et al. 2003, Miyagishima, Froehlich, et al. 2006, Miyagishima, Nishida, et al. 2003a, Vitha, Froehlich, et al. 2003, Yang, Glynn, et al. 2008). Within the stroma, two tubulin-like proteins, F tle and FtsZ2, both derived from cyanobacterial FtsZ via endosymbiosis (Osteryoung, Stokes, et al. 1998, Osteryoung and Vierling 1995, Stokes and Osteryoung 2003), are central players in the division process, comprising a mid-plastid contractile ring (Z-ring) (McAndrew, F roehlich, et al. 2001, Stokes, McAndrew, et al. 2000, Vitha, McAndrew, et al. 2001). In bacteria, the Z-ring marks the division site, acts as a scaffold for other division components, and may provide contractile force driving membrane constriction (Lan, Wolgemuth, et al. 2007, Li, Trimble, et al. 2007, Margolin 2005, Osawa, Anderson, et al. 2008). A third FtsZ- derived chloroplast division protein, ARC3, is a chimera of FtsZ and other host-derived functional domains (Shimada, Koizumi, et al. 2004). ARC3 has likely replaced MinC, an 142 inhibitor of FtsZ polymerization and a component of the Min system, which positions the Z-ring in bacteria (Glynn, Miyagishima, et al. 2007, Maple, Vojta, et al. 2007, Rothfield, Taghbalout, et al. 2005). On the cytosolic surface of the chloroplast, factors of host origin mediate organelle constriction and separation (Gao, Kadirjan-Kalbach, et al. 2003, Glynn, Miyagishima, et al. 2007, Miyagishima, Froehlich, et al. 2006, Miyagishima, Nishida, et al. 2003a, Yang, Glynn, et al. 2008). ARC5 is a dynamin-like protein required for the late stages of division, acting as a pinchase (Gao, Kadirjan-Kalbach, et al. 2003, Miyagishima, Nishida, et al. 2003a, Yoshida, Kuroiwa, et al. 2006). Dynamins are best known for their roles in vesicle-budding and endocytosis in eukaryotes (Hinshaw 2000, Hinshaw and Schmid 1995, Praefcke and McMahon 2004, Wiejak and Wyroba 2002), though the origins of dynamins can be traced to prokaryotes (Low and Lowe 2006). ARC5 is recruited to the division site from cytosolic patches by PDVl and PDV2, two coiled-coil transmembrane outer envelope proteins of host origin that are unique to land plants (Glynn, Froehlich, et al. 2008, Miyagishima, Froehlich, et al. 2006). The finding that ARC5 localizes to the division site in pdvl and pdv2 mutants but not in the double mutant indicates that PDVl and PDV2 function independently in recruiting ARC5 to the chloroplast, but both PDV proteins are required for full ARC5 contractile activity (Miyagishima, Froehlich, et al. 2006). The stromal F tsZ and cytosolic dynamin complexes that drive division of the organelle must be coordinated across the two envelope membranes to ensure their 143 synchronous operation. A key factor coordinating these complexes is ARC6, a transmembrane protein of the inner envelope that evolved from the cyanobacterial cell division protein Ftn2 (Koksharova and Wolk 2002, Vitha, Froehlich, et al. 2003). The stromal region of ARC6 binds FtsZ2 (Maple, Aldridge, et al. 2005) and mediates Z-ring assembly (V itha, Froehlich, et al. 2003), while the IMS region interacts with and positions PDV2 at the division site to facilitate ARC5 recruitment (Glynn, Froehlich, et al. 2008). ARC6 is also required for PDVl positioning at the division site, but our finding that ARC6 and PDVl do not interact suggested that an additional factor might function between ARC6 and PDVl to localize PDVl (Glynn, F roehlich, et al. 2008). Here we show that PARC6, an ARC6-like protein unique to vascular plants, performs this function. Mutant analysis reveals that PARC6 is required for wild-type chloroplast replication and accumulation. Like ARC6, PARC6 resides in the inner envelope membrane (IEM) at the division site with its N-terminus facing the stroma. However, PARC6 acts downstream of ARC6 to mediate PDVl localization, consistent with a previous model (Glynn, Froehlich, et al. 2008, Yang, Glynn, et al. 2008) and show that PARC6 interacts with the cytosolic domain of PDV]. In further contrast to ARC6, PARC6 localizes at one pole of some chloroplasts, similar to the localization patterns of PDVl and ARC3 (Maple, Vojta, et al. 2007, Miyagishima, Froehlich, et al. 2006). Analysis of FtsZ filament morphology in parc6 suggests that PARC6 inhibits FtsZ assembly, in contrast with ARC6, which promotes F tsZ assembly (V itha, Froehlich, et al. 2003). We further show that PARC6 interacts with ARC3, and hypothesize that PARC6 inhibits FtsZ assembly via this interaction. Collectively, these findings suggest that 144 PARC6 and ARC6 may function as antagonistic regulators of Z-ring dynamics, and that PARC6 plays a role in coordinating Z-ring activity in the stroma with dynamin activity at the outer envelope. Additionally, the interaction of PARC6 with a component of the Min system and its localization pattern hint at the possibility that PARC6 may play a role in establishing polarity within newly divided chloroplasts. Results PARC6 Family Members are Distinct fi'om ARC6 and are Unique to Vascular Plants. We identified PARC6 by BLAST (Altschul, Madden, et al. 1997), shortly after the identification of ARC6 by a map-based approach (V itha, Froehlich, et al. 2003). The PARC6 locus (At3g19180) encodes a predicted protein product that shares ~21% identity with Arabidopsis ARC6 (V itha, Froehlich, et al. 2003) and is predicted by ChloroP (Emanuelsson, Nielsen, et al. 1999) to bear a chloroplast transit peptide at its N-terminus (Figure 5.1). Consistent with this prediction, the N-terminus of PARC6 (amino acids (AA) 1-76) targets a YFP fusion to the chloroplast stroma in tobacco (Glynn, Yang, et al. 2009) In contrast to ARC6, which contains one transmembrane helix that anchors it in the IEM (V itha, Froehlich, et al. 2003), Arabidopsis PARC6 has two predicted transmembrane domains (Figure 5.1) based on output from Aramemnon (Schwacke, Schneider, et al. 2003). Multiple sequence alignment between PARC6, ARC6, and Ftn2 proteins from several organisms revealed two major regions of similarity, one near the N- 145 TP 1-76 TMl 357-377 TM2 574-596 / / / AtPARC6 E] [1 | 819AA TP 1-67 TM 615-635 / / AtARC6 It. I in ] 801AA TM 535-557 / |720AA Figure 5.1. Summary of PARC6 protein features and similarity to related proteins. Lines under each diagram indicate regions of high similarity between all three proteins. A multiple sequence alignment comparing sequences of PARC6, ARC6, and Ftn2 from several species is shown elsewhere (Glynn, Yang, et al. 2009). Transit peptide (TP); transmembrane domain (TM); and amino acids (AA). 146 terminus and the other near the C-terminus (Figure 5.1), suggesting that these regions of PARC6, ARC6, and Ftn2 may have related functions. Several small regions conserved within, and unique to, PARC6 family members were also evident upon close examination of the alignment (Glynn, Yang et al. 2009); these regions may harbor functions that are specific to PARC6 proteins. Notably, the conserved proline that resides within the central motif of the predicted J -domain of most ARC6/Ftn2 family members (V itha, Froehlich, et al. 2003) is not evident in PARC6 sequences (Figure 5.1), suggesting that PARC6 does not function as a DnaJ-like co-chaperone. While ARC6-like sequences are found in cyanobacteria, algae, and moss, PARC6 sequences were only detected in vascular plants (tracheophytes) suggesting that PARC6 emerged as a chloroplast division factor in this group of organisms. Consistent with this observation, phylogenetic analysis shows that PARC6 and ARC6 sequences from tracheophytes cluster into distinct clades and that both of these proteins may have diverged from a common ARC6-like ancestor (Figure 5.2). As postulated previously (Vitha, Froehlich, et al. 2003), ARC6 is more closely related to cyanobacterial Ftn2 than is PARC6 (Figure 5.2), suggesting that ARC6 and Ftn2 share greater functional similarity than PARC6 and Ftn2. These results, in combination with the sequence similarities and differences highlighted above, suggest that PARC6 arose as a result of gene duplication and divergence in primitive vascular plants and that it probably has evolved functions distinct from those of ARC6 within the tracheophyte lineage. 147 '1 64 ~——VvPARC6 54_[ PtPARC6 98 MtPARC6 '00 AtPARC6 OsPARC6 46 1 oo ZmARC6 __‘22{ OsARC6 AtARC6 ‘ °° PpARC6 OlARC6 TeFtn2 ‘00 . SeFtn2 41L NsFth 93 —---CwFt112 100 - CsFtn2 |-—-—l 0.1 Figure 5.2. Phylogenetic analysis of PARC6, ARC6, and Ftn2 family members. Neighbor-joining tree showing the relationship between PARC6, ARC6, and Ftn2 family members from various species. Numbers at the nodes represent bootstrap values and scale bar corresponds to the number of amino acid substitutions per site. Arabidopsis thaliana (At); Medicago truncatula (Mt); Oryza sativa cv. japonica (Os); Populus trichocarpa (Pt); Vitis vinifera (V v); Ostreococcus lucimarinus (Ol); Physcomitrella patens (Pp); Zea mays (Zm); Thermosynechococcus elongatus BP-l (Te); Synechococcus elongatus PCC 7942 (Se); Nostoc sp. PCC 7120 (N s); Crocosphaera watsonii WH 8501 (Cw); and Cyanothece sp. ATCC 51142 (Cs). 148 Mutations in PARC6 cause aberrations in chloroplast morphology and F tsZ filament morphology. The similarity of PARC6 to ARC6 led us to investigate its role in plastid division. We identified a T-DNA insertion (SALK_100009) in the first exon of PARC6 (parc6-1, Figure 5.3A) and characterized homozygous lines for chloroplast morphology and number (Figure 5.3C). parc6-I mutants are less drastically impaired in chloroplast division than arc6 mutants (compare Figure 5.3C and Figure 5.3F), but mesophyll cells in parc6-I mutants contain nearly 10-fold fewer chloroplasts (6.7 i 2.4 chloroplasts per cell, 11 = 15 cells) than those in wild type (59.7 :t 11.2 chloroplasts per cell, n = 15 cells). parc6-I mesophyll cell chloroplasts possess characteristics of chloroplasts in both arc3 and arc5 mutants (Pyke and Leech 1994). Like arc3 mutants, parc6-I mutants exhibit a heterogeneous mixture of chloroplast sizes within individual cells (Marrison, Rutherford, et al. 1999, Pyke and Leech 1992, Pyke and Leech 1994) (Figure 5.3C). Like arc5 mutants, parc6-1 mutants have some chloroplasts with prevalent constrictions (Figure 5.3C, arrowheads), suggesting a block in dynamin (ARC5) function (Gao, Kadirjan- Kalbach, et al. 2003, Miyagishima, Froehlich, et al. 2006). We also identified two additional alleles of PA RC6, parc6-2 and parc6-3, in a forward genetic screen (Figure 5.3D-E). These alleles confer phenotypes similar to those in parc6-I . parc6-2 behaves as a dominant-negative mutation while parc6-I and parc6-3 are recessive to the wild- type allele. These phenotypes lead us to conclude that PARC6 is a bona fide plastid 149 (SALK_100009) (Q526‘) parc6-I parc6-3 TC'iA At3g1 91 80 0 0 ATG parc6~2 4549 bP (D112N) Figure 5.3. Diagram of the PARC6 locus and phenotypic analysis of parc6 mutants. (A) Gene structure and locations of characterized mutations in the PARC 6 locus. White boxes indicate untranslated regions. Black boxes indicate exons. Introns are denoted by thin grey lines. The transcribed region of At3gl9180 is 4549 base pairs (bp). ATG, start codon; TGA, stop codon. (B-F) Chloroplast phenotypes in mesophyll cells of: Col-0 (B); parc6-1 (C); parc6-2 (D); parc6-3 (E); and arc6-1 (F). Arrowheads denote sites of chloroplast constriction in the parc6-I mutant. Scale bar = 10 pm. 150 division protein and suggest that PARC6 could influence operation of the division machinery through ARC3, ARC5, or the ARC5 recruitment factors PDVl and PDV2. PARC6 is an inner envelope protein with localization similar to PD VI. To examine the subcellular localization of PARC6, we generated a transgene encoding PARC6 fused to GFP expressed under control of the CaMV 35S promoter (35SprosPARC6-GFP). This transgene was capable of complementing the division defect in parc6-I mutants, indicating that it generates a functional protein product (Glynn, Yang et al. 2009). Following selection of T2 individuals, we examined epidermal cells in young leaves for expression of the fusion protein. PARC6-GFP localized to the middle of ovoid, partially constricted, and deeply constricted plastids (Figure 5.4), though whether it forms a complete ring during the early stages of division was not clear, as the GFP signal in these plastids was very weak. The GFP signal was most evident in deeply constricted plastids as single foci (Figure 5.4B), suggesting concentration of the fusion protein at the isthmus connecting plastids just prior to separation. In some plastids, PARC6-GFP was localized at the organelle surface in a single spot at one pole (Figure 5.4C), perhaps representing the persistence of the fusion protein at the pole following separation of the daughter plastids. A similar pattern of localization has been observed in Arabidopsis for both PDVl and ARC5 (Miyagishima, Froehlich, et al. 2006), but not for ARC6 or PDV2 (Glynn, F roehlich, et al. 2008, Vitha, Froehlich, et al. 2003). In addition, we observed both polar and medial localization of PARC6-YFP at the 151 Figure 5.4. PARC6 is a plastid protein that localizes to the division site and to polar spots. PARC6-GFP localizes to mid—plastid puncta (A); mid-plastid spots (B); or polar spots (C) in Arabidopsis. Scale bar 5 pm. 152 chloroplast periphery in transiently transfected tobacco leaf cells (Glynn, Yang et al. 2009), reminiscent of ARC3 localization under similar experimental conditions (Maple, Vojta, et al. 2007). In tobacco, the dual localization in the same plastid may be a consequence of overexpression as we did not observe both medial and polar localization in the same plastid in Arabidopsis parc6-I mutants complemented with the PARC6-GE P transgene (Figure 5.4). In both Arabidopsis and tobacco, PARC6-GFP was associated with the chloroplast periphery, suggesting it is localized in the envelope membrane. To verify that PARC6 is an envelope membrane protein, we examined the fractionation of native PARC6 protein in isolated pea chloroplasts using an antibody generated against the region of Arabidopsis PARC6 residing between the transit peptide and the first predicted transmembrane domain (AA 77-356, Figure 5.1). The antibody detected a protein migrating at ~1 16 kD in whole-cell extracts from pea (Figure 5.5, lane 1) and a competition assay in which the antibody was preincubated with the immunizing antigen confirmed that this protein is native pea PARC6 (Glynn, Yang, et al. 2009). This protein was enriched in fractions containing isolated intact chloroplasts (Figure 5.5, lane 2) and chloroplast membrane fractions, (Figure 5.5, lane 3), suggesting that PARC6 is a chloroplast envelope protein. Preliminary analysis of PARC6 topology from pea chloroplast fractions indicates that the N-terminus of the protein resides in the stroma, but these experiments lacked controls and the orientation of the C-terminus of the protein remains unclear (Glynn, Yang. et al. 2009). Further work is needed to fully verify the topology of PARC6 in vivo. 153 1234 5678 <116kD> <54kD> OL-PARCG a-Ticl 10 Figure 5.5. PARC6 is a chloroplast membrane protein. Whole-cell extract (lanes 1 and 5); isolated chloroplasts (lanes 2 and 6); chloroplast membrane fractions (lanes 3 and 7); and chloroplast soluble; fractions (lanes 4 and 8) were probed with an anti-PARC6 antibody (left) or an anti-Tic110 antibody (right) (Davila-Aponte, Inoue, et al. 2003). SDS-PAGE migration of PARC6 is slower than would be predicted based on its molecular weight, but we observed similar behavior with ARC6 (McAndrew, Olson, et al. 2008). Asterisk (*) indicates an unknown non-specific anti-PARC6 cross-reacting protein. Kilodaltons (kD). 154 PARC6 inhibits F tsZ assembly and interacts with ARC3. Because ARC6 acts as a positive regulator of FtsZ assembly (V itha, Froehlich, et al. 2003), we examined FtsZ organization in homozygous parc6-I mutants by immunostaining to determine if PARC6 also affects F tsZ assembly in vivo. In contrast to the short FtsZ filaments observed in the enlarged chloroplasts of arc6 mutants (Vitha, F roehlich, et al. 2003), we observed long Ftle filaments that appeared to be multiple rings or spirals within the larger chloroplasts of parc6~1 (Figure 5.6B). Additionally, we observed tubular chloroplasts containing clusters of FtsZ rings (Figure 5.6C, arrowheads), reminiscent of the multiple Z-rings observed in arc3 mutants (Glynn, Miyagishima, et al. 2007). The elongated FtsZ filaments observed in parc6-I were not due to a change in Fle or FtsZZ levels, which were similar to those in wild type (Figure 5.6D). These results suggest that PARC6 might act as an antagonist of ARC6, with PARC6 inhibiting FtsZ assembly, and also implicate PARC6 in the fimctioning of the plastidic Min system (Fujiwara, Hashimoto, et al. 2008, Fujiwara, Nakamura, et al. 2004, Itoh, Fujiwara, et al. 2001, Maple and Moller 2007, Maple, Vojta, et al. 2007, Nakanishi, Suzuki, et al. 2009). Arabidopsis ARC6 has been shown to interact with both FtsZ2 family members, but not Ftle (Maple, Aldridge, et al. 2005, Schmitz, Glynn, et al. 2009). The profound effect of PARC6 depletion on Z-ring morphology led us to test for interaction between the putative N-terrninal stromal region of PARC6 (AA 77-357) and both families of Arabidopsis FtsZ using two-hybrid assays. However, neither Ftle nor FtsZ2 interacted 155 123123 I. Ponceau a-Ftle a—FtsZZ 50kD > 37kD > Figure 5.6. Analysis of FtsZ morphology and FtsZ protein levels in parc6-1 . Immunolocalization of Ftle in Col-0 (A); and parc6-1 mutants (B, C). Arrowheads point to structures formed by F tsZ. Chlorophyll autofluorescence (red) and Ftle localization (green) is indicated. FtsZ2 filament morphology is also perturbed in parc6-1 mutants, as both FtsZ2-1 and FtsZ2-2 do not form regular rings, but instead form highly elongated filaments (Glynn, Yang et al. 2009) that are unlike the short FtsZ filaments observed within the chloroplasts of arc6 mutants (Vitha, Froehlich, et al. 2003). Immunoblots showing FtsZ levels are shown in panel (D). Ponceau stain of the membrane used for immunoblotting with Ftle and FtsZ2-1 specific antibodies (left); F tle immunoblot (center); and FtsZ2—1 immunoblot (right). After Ponceau staining, blot was probed for Ftle, stripped, and then reprobed for FtsZ2-1. Molecular weight marker (lane 1); Col-0 whole cell extract (lane 2); and parc6-1 whole cell extract (lane 3). Scale bars = 5 pm. 156 with PARC6 (Figure 5.7). This result was somewhat surprising given the high affinity of ARC6 for FtsZ2 (Maple, Aldridge, et al. 2005). Based on the chloroplast size heterogeneity in parc6-1 (Figure 5.3) and arc3 (Marrison, Rutherford, et al. 1999, Pyke and Leech 1992, Pyke and Leech 1994), the affect of PARC6 on FtsZ ring and filament morphology (Figure 5.6) and the ARC3-like localization of PARC6 in tobacco (Glynn, Yang et al. 2009, Maple, Vojta, et al. 2007), we decided to test for interaction between PARC6 and ARC3 by two-hybrid. In contrast to ARC6 (Maple, Aldridge, et al. 2005), the N-terminus of PARC6 interacted strongly and specifically with the predicted mature ARC3 protein (Figure 5.8); these results were confirmed by another group (Zhang, Hu, et al. 2009b). Interestingly, the PARC6-ARC3 interaction requires the Membrane Occupation and Recognition Nexus (MORN) domain (Shimada, Koizumi, et al. 2004) (Figure 5.8), a region of ARC3 that was shown previously to inhibit a specific interaction between ARC3 and Ftle (Maple, Vojta, et al. 2007) PA RC6 acts downstream of ARC6. ‘ To investigate the functional relationship between ARC6 and PARC6, we generated arc6 parc6 double mutants and compared their phenotypes with those of both parental mutants and wild-type plants (Figure 5.9A-D). arc6 parc6 plants (Figure 5.9D) possess mesophyll chloroplast phenotypes that are more like those of arc6 mutants (Figure 5.9C) than parc6 mutants (Figure 5.9B), suggesting that PARC6 acts downstream of ARC6 during chloroplast division. 157 INTERACTION PAIR SI)/-ULT SD/-ULTI'I PARC6AA77_ 357 vs. Ftle- 1 -- PARC6AA77_ 357 vs. FtsZ2- l - EMPTY vs. Ftle—l -_ EMPTsz. 14.2.2-1 -- EMPTY vs. EMPTY Figure 5.7. PARC6 does not interact with an FtsZ protein. Unlike ARC6, PARC6 does not interact with an FtsZ protein, as growth was not observed on synthetic dropout (SD) media lacking histidine (right column). Only constructs representing processed stromal portions of each protein were used to test for interaction. Dilutions at bottom of each column refer to dilution from a starting culture of OD600 = 1.0. 158 TP 1-76 TM] 357—377 TM2 574-596 / / / PARC6 El El 1 TP 141 MORN 598-741 / / ARC3 - B INTERACTION PAIR SD/-ULT SD/-ULTH PARC6AA77_357 vs. ARC3 PARC6AA77-357 vs. ARC3-M PARC6AA77_357 vs. EMPTY EMPTY vs. ARC3 EMPTY vs. ARC3_M EMPTY vs. EMPTY 1 1/10 1/100 1 1/10 1/100 Figure 5.8. PARC6 interacts with ARC3. A schematic of the proteins used for interaction assays is shown in (A). The underline indicates the portion of the protein used for assay. Results of the two-hybrid assay are shown in (B). Dilutions at bottom of each column refer to dilution from a synthetic dropout (SD) starting culture of OD600 = 1.0. For interaction test between PARC6 A A77-3 57 and ARC3, a growth ratio of 0.67 was observed (growth on SD/-ULTH: growth on SD/-ULT) ~48h after spotting; all other interaction pairs had growth ratios of 0.07 or less. Transit peptide (TP); transmembrane domain (TM); ARC3-MORN (ARC3_M). 159 I 7 lipttrC6~l .xncnxirv ’ \Rt‘tntil-‘l’ ' Figure 5.9. PARC6 acts downstream of ARC6. Chloroplast phenotypes of Col-0 (A); parc6-1 (B); arc6-1 (C); and a parc6-I arc6-1 double mutant are shown. ARC6-GFP localizes to mid-plastid rings in wild type Col-0 (E) and parc6-I mutants (F), indicating that PARC6 is not required for ARC6 localization in Arabidopsis. Scale bar = 5 pm. 160 If PARC6 acts downstream of ARC6, we expected that ARC6-GFP should localize normally to chloroplast constriction sites in parc6 mutants. To test this, we examined ARC6-GFP localization in Col-0 and parc6-l mutants using an ARC6prosARC6-GF P transgene that was used previously to examine ARC6 localization in Arabidopsis (Glynn, Froehlich, et al. 2008). Consistent with ARC6 acting upstream of PARC6, we observed localization of ARC6-GFP to sites of constriction in young leaf cells of both wild-type (Figure 5.9E) and parc6 mutants (Figure 5.9F). From these data, we conclude that PARC6 acts downstream of ARC6 and that PARC6 is not required for localization of ARC6 to sites of chloroplast constriction. PARC6 is required for PD VI localization, but not for PD V2 or ARC5 localization. We previously reported that ARC6 interacts directly with PDV2 and positions PDV2 at the division site (Glynn, Froehlich, et al. 2008). We also showed that ARC6 is required for equatorial positioning of GFP-PDV] , though direct interaction between ARC6 and PDVl could not be confirmed by two-hybrid. At that time, we hypothesized the presence of at least one factor that acts as an intermediary between ARC6 and PDVl. Because PARC6 acts downstream of ARC6 and because parc6 mutants (Figure 5.9B) possess a subpopulation of chloroplasts that resemble the partially-constructed chloroplasts observed in pdvl mutants (Miyagishima, Froehlich, et al. 2006), we hypothesized that PARC6 might be the factor that acts between ARC6 and PDVl. To test whether PARC6 is required for PDVl localization, we examined GFP-PDVl signals (Miyagishima, Froehlich, et al. 2006) in Col-0 and parc6-1 mutants (Figure 5.10). We 161 1. A . GFP-PDVI purctfil purc'O— I ( 'ol—(l purco-I Figure 5.10. Localization of PDVl, PDV2, and ARC5 in parc6 mutants. GFP-PDVI localizes to mid-plastid constrictions in Col-0 (A), but is unable to localize to the division site in parc6-I mutants (B). YFP-PDV2 (C, D) and GFP-ARC5 (E, F) localize to sites of constriction in Col-0 (C. E) and parc6-1 mutants (D, F). Scale bar = 5 pm. 162 were unable to observe any GFP-PDV] signal at sites of chloroplast constriction in parc6-I mutants (Figure 5.103), though GFP-PDV] localized properly to dividing chloroplasts in wild-type (Figure 5.10A), was expressed in both wild type and parc6-1 backgrounds based on immunoblotting (Glynn, Yang et al. 2009), and can functionally complement the pdvl mutant (Miyagishima, Froehlich, et al. 2006). Taken together, the PDVl-like localization of PARC6 in Arabidopsis (Figure 5.4) and the loss of PDVl localization in parc6 mutants (Figure 5.10B) indicate a major role for PARC6 in positioning PDV] at the division site. Consistent with PARC6 acting downstream of ARC6 (Figure 5.9) and ARC6- dependent positioning of PDV2 (Glynn, Froehlich, et al. 2008), we observed YFP-PDV2 at sites of chloroplast constriction in parc6-I mutants as well as in wild type (Figure 5.10C-D). Further, in agreement with previous fmdingings that either PDV] or PDV2 are capable of recruiting ARC5 to the division site (Miyagishima, Froehlich, et al. 2006), we also observed GFP-ARC5 positioned correctly in parc6 mutants despite loss of PDVl positioning (Figure 5.10E-F). PARC6 binds the cytosolic domain of PD VI in two—hybrid assays. Because PARC6 is required for positioning PDVl at the division site, we aimed to determine if PARC6 bound PDVl in two hybrid assays. The C-terminus of PDVl was previously shown to be localized to the IMS (Miyagishima, F roehlich, et al. 2006). We initially tested for interaction between the C-termini of PARC6 and PDVl , assuming that the C-terrninus of PARC6 (AA 597-819, Figure 5.1) might reside within the 163 intermembrane space, with PARC6 having a topology similar to ARC6 (V itha, Froehlich, et al. 2003). However, no interaction was detected between these two C-terminal domains, nor did the middle region of PARC6 (AA 378-573, Figure l) interact with the C-terminus of PDVl (M. Hemmes, J. Glynn, and K. Osteryoung, unpublished). Because PARC6 possesses two predicted transmembrane domains (Figure 1), possibly allowing it to span both envelope membranes (thereby placing the C-terminus (AA 597-819, Figure 5.1) in the cytosol), we also tested for interaction between the C-terminus of PARC6 and the cytosolic N—terminus of PDVl. To our surprise, a weak interaction between these two proteins was detected in yeast (Figure 5.11). This interaction is consistent with PARC6-dependent positioning of PDVl , but the precise topology of PARC6 remains to be determined. However, it must be seriously considered that PARC6 might have a more dynamic or complex topology relative to other division proteins. PD VI and PD V2 Independently Localize to the Division Site. If PARC6 and ARC6 are responsible for positioning PDVl and PDV2, at the division site, respectively, we predicted that PDVl and PDV2 should localize independently of each other. To confirm this, we expressed GFP-PDVl in pdv2-1 mutants and YFP-PDV2 in pdvl-1 mutants. GFP-PDV] localized to the mid-plastid in the pdv2-I background (Figure 5.12B), indicating that PDV2 is not required for PDVl localization. Similarly, YFP-PDV2 localized to the mid-plastid in the pdvl-1 background (Figure 5.12D). We conclude that PDVl and PDV2 localize to the division site independently of one another — PDVl through PARC6 and PDV2 through ARC6 — though PARC6 localization and/or activity probably depends upon ARC6 (Figure 5.9). 164 TP/1-76 TM 1/357-377 Th/AZ 574-596 PARC6 I TM 207-225 PDV] I B INTERACTION PAIR SD/—ULT SDl-ULTH PARC6m-vs. EMPTY EMPTY vs. PDV] N, EMPTY vs. EMPTY 1 1/10 1/100 1 1/10 1/100 Figure 5.11. The C-terminus of PARC6 binds the cytosolic domain of PDVl. Schematic showing proteins tested in yeast two-hybrid interaction assays (A). The underline indicates the segment of each protein used in the interaction assays. Results of two-hybrid reporter assays on synthetic dropout (SD) media are shown in (B), showing growth on media supplemented with histidine (left panels) and media lacking histidine (right panels). Growth ratio (growth on SD/-ULTH: growth on SDl-ULT) for PARC6CT and PDVlN—r was 0.37, while all other interactions tested were 0.1 or less. Dilutions at the bottom of each column indicate dilution from a starting culture of OD600 = 1.0. 165 (ilil:‘l’-l’l)Vl (bho YFP—PDVZ q pdvl - I Figure 5.12. PDVl and PDV2 localize independently of each other. GFP-PDV] localization (A, B) and YFP-PDV2 localization (C, D) in Col-0 (A, C), pdv2-1 (B), and pdvI-I (D) mutants. PDV2 is not required for localization of PDVl (B) nor is PDVl required for localization of PDV2 (D). Scale bar = 5 pm. 166 Discussion Here we characterize the new chloroplast division protein PARC6, a protein of tracheophyte origin that, like ARC6, is a multi-functional integral inner envelope protein that aids the coordination of the FtsZ and dynamin rings across the envelope membranes. However, our results show that PARC6 acts downstream of ARC6 in the division process and that its function differs significantly from that of ARC6: (1) PARC6 and ARC6 have distinct localization patterns in Arabidopsis. In incompletely constricted chloroplasts, ARC6 appears as a continuous ring at the division site (Glynn, Froehlich, et al. 2008, Vitha, F roehlich, et al. 2003, Yang, Glynn, et al. 2008) whereas PARC6 localizes to puncta. PARC6 also localizes to foci at the poles of some plastids, which has not been observed for ARC6. (2) PARC6 acts as an inhibitor of FtsZ assembly whereas ARC6 promotes FtsZ assembly. (3) ARC6 is required for recruitment of PDVl , PDV2, and ARC5 to the division site whereas PARC6 is required only for recruitment of PDVl. A working model depicting the roles of ARC6 and PARC6 in the coordination and activity of the FtsZ and dynamin rings at the division site is shown in a later chapter. Despite the overall sequence identity between ARC6 and PARC6 (~21%), there are many differences that may contribute to their functional divergence. PARC6 lacks the J -like domain present in ARC6, suggesting it also lacks the co-chaperone activity hypothesized for ARC6 (V itha, Froehlich, et al. 2003). Additionally in contrast with ARC6, which has one transmembrane domain, multiple PARC6 sequences bear two predicted transmembrane domains (Glynn, Yang et al. 2009), one of which (WI; 167 Figure 5.1) interrupts a segment conserved in ARC6 family members that resides in the stroma (Glynn, Yang. et al. 2009, Vitha, Froehlich, et al. 2003); the divergence in ARC6- PARC6 similarity in this region coincides with absence of the FtsZ2-binding domain (ARC6 A A 3 51-503) in PARC6 proteins (Figure 5 .1). Preliminary protease protection assays in pea support the TM] prediction (Glynn, Yang et al. 2009) and, along with the interaction between PARC6 A A77-3 57 and ARC3 (Figure 5.8), are consistent with a topology placing the N-terminus of PARC6 in the stroma. Though the TM2 prediction remains to be confirmed, it is predicted in several PARC6 sequences and is aligned closely with the ARC6 transmembrane domain (Figure 5.1) (Glynn, Yang, et al. 2009). Moreover, we show here that the C-terminus of PARC6 interacts with the cytosolic domain of PDVl (Figure 5.1 1), suggesting that PARC6 might traverse both envelope membranes. While preliminary protease protection assays indicate that PARC6 does not traverse the outer envelope (Glynn, Yang, et al. 2009), PARC6 may have a more complicated or dynamic topology in vivo. If such an orientation is validated, it is curious as to why the C-termini of ARC6 and PARC6 would be compartmentally separated, given the sequence similarity near their C-termini. Full topological analysis will be critical to confirm these results and for further dissection of PARC6 ftmctional domains. The long FtsZ filaments observed in the enlarged chloroplasts of parc6-I mutants (Figure 5.6) suggest that, in vivo, PARC6 inhibits FtsZ assembly. This activity may be indirect because PARC6 does not interact with Ftle or FtsZ2 in two-hybrid assays (Figure 5.7). The inhibitory effect of PARC6 on FtsZ assembly may rather be a consequence of its interaction with ARC3 (Figure 5.8) (Zhang, Hu, et al. 2009b), because 168 ARC3 has been proposed as a functional replacement for MinC (Maple, Vojta, et al. 2007), a prokaryotic protein known to promote disassembly of bacterial F tsZ (Hu, Mukherjee, et al. 1999, Margolin 2003, Pichoff and Lutkenhaus 2001). The similar chloroplast and FtsZ morphologies in parc6 (Figure 5.3 and Figure 5.6) and arc3 mutants (Glynn, Miyagishima, et al. 2007 , Marrison, Rutherford, et al. 1999, Pyke and Leech 1992, Pyke and Leech 1994), are consistent this hypothesis. The opposing activities of ARC6 and PARC6 in promoting and inhibiting FtsZ assembly, respectively, suggest that these two proteins may fimction as antagonistic regulators of Z-ring dynamics within the stroma. We do not yet know whether this or other activities of PARC6 involve direct interaction with ARC6, though preliminary data suggests that they may interact through their stromal domains (Y. Yang and K. Osteryoung, unpublished). The requirement of the ARC3 MORN domain for PARC6-ARC3 interaction (Figure 5.8) contrasts with the finding that the MORN domain inhibits interaction between ARC3 and Ftle (Maple, Vojta, et al. 2007). These observations suggest that FtsZ disassembly may be controlled, at least in part, by the availability of the ARC3 MORN domain. Interaction of ARC3 with PARC6 could sequester the MORN domain, allowing ARC3 to interact with Ftle to promote F tsZ filament disassembly, perhaps through disruption of the Ftle -FtsZZ heteropolymer. Dynamic interaction between ARC3 and PARC6 could thereby regulate FtsZ dynamics in vivo, though other factors, including MinD, MinE and MCDl, certainly contribute (Aldridge and Moller 2005, Colletti, Tattersall, et al. 2000, Fujiwara, Hashimoto, et al. 2008, Fujiwara, Nakamura, et al. 2004, Itoh, Fujiwara, et al. 2001, Maple, Chua, et al. 2002, Maple and Moller 2007, 169 Nakanishi, Suzuki, et al. 2009, Reddy, Dinkins, et al. 2002). Because MORN domains have been shown to be important for membrane association (Ma, Lou, et al. 2006), an alternative possibility for the MORN-mediated PARC6-ARC3 interaction is that it only occurs near the surface of a membrane. Removal of the MORN sequence from ARC3 could prevent its association with the IEM, thereby preventing interaction with PARC6. PDVl and PDV2 localization at the division site, though independent of one another (Figure 5.12), both require ARC6 (Glynn, Froehlich, et al. 2008). PDV2 localization is established by direct interaction with ARC6, but it was previously unclear how ARC6 directs PDVl localization, as we did not detect interaction between ARC6 and PDVl using two hybrid assays (Glynn, Froehlich, et al. 2008). Our current results show that PARC6 acts as an intermediary between ARC6 and PDVl, functioning to organize PDVl at the division site (Figure 5.9 and Figure 5.10). PDVl positioning by PARC6 may involve direct interaction with PDVl within the cytosol, based on our two hybrid experiments (Figure 5.11). Beyond this question, the mid-plastid localization of either PDVl or PDV2 is clearly sufficient for ARC5 recruitment, but our collective data indicate that localization of both PDV] and PDV2 at the division site is required for full ARC5 contractile activity (Glynn, Froehlich, et al. 2008, Miyagishima, Froehlich, et al. 2006). This aspect of chloroplast division remains to be explored. During division, PDVl initially localizes to a medial punctate ring that becomes reduced to a single spot in deeply constricted chloroplasts. The PDVl spot persists at the pole of one of the two daughter organelles following their separation. ARC5 behaves 170 similarly (Miyagishima, F roehlich, et al. 2006). Intriguingly, the localization patterns we observed for PARC6 (Figure 5.4) suggest that it also follows a PDVl-like progression during division, possibly being retained at the pole of a single daughter chloroplast. We speculate that the lingering PARC6 spot could play a role in establishing organelle polarity within the stroma, which might be important for defining the new division site prior to the next round of division. In E. coli, cell polarity and division-site positioning are established by the dynamic behavior of the Min system, which restricts FtsZ ring assembly to the midcell (Lutkenhaus 2007). The fact that ARC3 interacts with PARC6 and that ARC3 and PARC6 exhibit partial polar localization raises the possibility that PARC6, perhaps via interaction with ARC3 and/or other components of the plastidic Min system (Fujiwara, Hashimoto, et al. 2008, Fujiwara, Nakamura, et al. 2004, Itoh, Fujiwara, et al. 2001, Maple, Chua, et al. 2002, Maple and Moller 2007, Maple, Vojta, et al. 2007, Nakanishi, Suzuki, et al. 2009), could play a role in directing the orientation of the Min system within the stroma and hence in placing the new division site. Analysis of parc6 arc3 double mutants, ARC3 localization in parc6, and PARC6 localization in arc3 mutants will be critical to ordering these factors in a pathway and understanding how each of them impact division site selection. The emergence of PARC6 in vascular plants follows the emergence of ARC3 in green algae (Yang, Glynn, et al. 2008) and PDVl/PDV2 in mosses (Miyagishima, F roehlich, et al. 2006), perhaps allowing for greater control of FtsZ dynamics and chloroplast morphology. Presumably, the dissimilar regions of ARC6 and PARC6 confer protein-specific functions and will provide clues to the individual domains responsible 171 for their unique functions. For example, the N-terminal stromal domain of PARC6 does not interact with Arabidopsis FtsZ2 (Figure 5.7), in contrast to that of ARC6 (Maple, Aldridge, et al. 2005), but has apparently evolved to interact with ARC3, perhaps via the ARC3 MORN domain. Similarly, while the C-termini of PARC6 and ARC6 share a high degree of sequence similarity (Figure 5.1) (Glynn, Yang, et al. 2009), subtle differences within this conserved region might mediate specific interactions with other proteins, though the precise location of the C-terminus of PARC6 remains to be determined. The activities of the F tsZ and dynamin rings are presumably coordinated with those of the inner and outer plastid-dividing (PD) rings, electron-dense components of the plastid division machinery whose compositions are not yet known (Kuroiwa, Kuroiwa, et al. 1998, Miyagishima, Nishida, et al. 2003b, Yoshida, Kuroiwa, et al. 2006). It will be interesting to learn whether PARC6 and ARC6 interact with or otherwise influence the operation of the PD rings. Further studies to elucidate the functional and evolutionary relationship between PARC6 and other plastid division components should deepen our understanding of the mechanisms contributing to coordination of the internal and external division complexes. 172 Materials and Methods Sequence and Phylogenetic Analysis. These sequences were used for alignments and phylogeny: Arabidopsis thaliana (AtPARC6, NP_188549); Medicago truncatula (MtPARC6, IMGAG-annotated pseudomolecule AC157350_28.4); Oryza sativa cv. japonica (OsPARC6, NP_001054252); Populus trichocarpa (PtPARC6, a GENSCAN-based prediction (Burge and Karlin 1997) of Populus trichocarpa genome Scaffold_122 204390-215000); Vitis vinifera (VvPARC6, CAO48483); Arabidopsis thaliana (AtARC6, NP_199063); Oryza sativa cv. japonica (OsARC6, NP_001045726); Ostreococcus lucimarinus (OlARC6, XP_001421185); Physcomitrella patens (PpARC6, XP_001778770); Zea mays (ZmARC6, ACG29776); Thermosynechococcus elongatus BP-l (TeFth, BAC08309); Synechococcus elongatus PCC 7942 (SeFtn2, ABB57973); Nostoc sp. PCC 7120 (N sFtn2, BAB74406); Crocosphaera watsonii WH 8501 (CwFtn2, EAM48783); and Cyanothece sp. ATCC 51142 (CsFtn2, ACB49642). Evolutionary history was inferred by Neighbor-Joining (Saitou and Nei 1987) following multiple sequence alignment with MEGA4 (Tamura, Dudley, et al. 2007) using an identity matrix. The bootstrap consensus-tree inferred from 1000 replicates represents the evolutionary history of the taxa analyzed (Felsenstein 1985). Branches corresponding to partitions reproduced in less than 50% bootstrap replicates are collapsed. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (1000 replicates) are shown next to the branches (Felsenstein 1985). Tree is drawn to scale, with branch lengths in the same units as evolutionary distances used to infer the phylogenetic tree. 173 Evolutionary distances were computed using the Poisson correction method (Zuckerkandl and Pauling 1965) and are given as the number of amino acid substitutions per site. All positions containing gaps and missing data were eliminated from the dataset (Complete deletion option). There were 441 positions in the final dataset. Analysis of Chloroplast Morphology. Tissue preparation was carried out as described previously (Miyagishima, Froehlich, et al. 2006, Pyke and Leech 1991, Vitha, Froehlich, et al. 2003). Quantitative analysis of Col-0 and parc6-1 phenotypes was performed using the distal 3-4 mm of the largest leaf present on plants 30 days after sowing. Analysis of Subcellular Localization of PARC 6 35Spro-PARC6 Tp— YF P and 35Spm-PARC6-YF P constructs were made using the Gateway System (Invitrogen, http://www.invitrogen.com/). A PARC6 cDNA was used as template for PCR. Analysis of PARC6 localization in tobacco and in Arabidopsis was performed as described (Glynn, Yang, et al. 2009). Production of PARC 6 antibody. The coding sequence for the N-terminus of mature PARC6 (PARC6 A A 77-357) was cloned into pHIS9 (Yang, Xu, et al. 2008) and His-tagged protein was expressed in E. coli Rosetta DE3 (N ovagen). The insoluble recombinant protein was solubilized with 6M urea and purified using Ni-agarose. Subsequent SDS-PAGE removed urea and purified the antigen. Afier brief coomassie-blue staining, antigen was excised from the 174 gel and rabbit antibodies for PARC6 were generated by Covance Research Products, Inc. Exsaguination serum was used at 1:5,000 for immunoblotting. Analysis of PARC 6 Fractionation and Topology. Pea chloroplast isolation and fiactionation into membrane and soluble fractions was performed as described previously (Bruce, Perry, et al. 1994); additional details can be found elsewhere (Glynn, Yang, et al. 2009). Analysis of F tsZ Localization. Tissue preparation, fixation, and immunofluorescence analysis were carried out as described previously (Miyagishima, Froehlich, et al. 2006, Vitha, Froehlich, et al. 2003, Vitha, McAndrew, et al. 2001). Two-hybrid Analysis. Clones for two hybrid analysis were made using methods as described (Glynn, Yang, et al. 2009). HIS3 reporter assays were performed as described previously (Glynn, Froehlich, et al. 2008, Maple, Aldridge, et al. 2005). Tests of Epistasis. arc6-I was crossed to parc6-1 and arco parc6 double mutants were identified by genotype. SALK genotyping primers were used as described (Alonso, Stepanova, et al. 2003) for identifying homozygous parc6-I plants. An NlaIII-containing dCAPS marker (N eff, Neff, et al. 1998) was used to identify the arc6-1 mutation; the product generated 175 from these two primers harbors an additional NlaIII site when arc6-1 genomic DNA is amplified. Lines homozygous for both mutations were used for double mutant analysis. All primers used for genotyping are described elsewhere (Glynn, Yang, et al. 2009). Analysis of ARC 6, PDV], PDV2, and ARC5 Localization. Transgenes for ARC6pm-ARC6-GF P, PD VIpm-GFP-PD VI, PD V2pm- YFP- PD V2, and ARC5 pro-CF P-A RC 5 were introduced by Agrobacterium-mediated transformation (Clough and Bent 1998). Following selection of T1 individuals, young leaves of transgenic plants were examined by fluorescence microscopy using a Leica DMRA2 microscope outfitted with a Retiga EXi camera. Extract Preparation and Immunoblotting. Whole-cell extracts for anti-Ftle-l and anti-FtsZ2-1 immunoblots were prepared by grinding tissue from 11 day-old plate-grown plants in liquid nitrogen. The resulting powder was homogenized in 5 volumes of Extraction Buffer (25mM Tris, pH 7.4 containing 4mM dithiothreitol, 0.1% Triton X-100, and protease inhibitors). An equal amount of loading buffer was added to this homogenate before boiling. Four microliters of each sample was loaded for SDS-PAGE and transferred to nitrocellulose. The blot was probed with Ftle -1 or FtsZ2-1 antibodies at 1:15000 in 5% blocking buffer. 176 Acknowledgements I thank Yue Yang for developing the PARC6 antibodies, determining preliminary topology of PARC6, providing localization of PARC6 in tobacco, and sharing data on ARC6-PARC6 interaction. 1 thank Stan Vitha for examining FtsZ localization in parc6 mutants and determining localization of PARC6-GFP in Arabidopsis. I thank Aaron Schmitz (ARC3) and Mia Hemmes (PARC6) for generating the two hybrid clones used here. Finally, I thank Shin-ya Miyagishima for providing the parc6-2 and parc6-3 alleles. Most of this chapter and the figures therein are reproduced from (Glynn, Yang. et al. 2009) (httn://www.wilev.com/bw/iournflrsp?ref=0960-74l2/) and is Copyright Blackwell Publishing. Blackwell Publishing allows the author to re-use your own article in another publication providing you are editor or co-editor (author or co-author) of the new publication. 177 Chapter 6 A Conserved Aspartate within ARC6 Influences Plastid Size and Z-ring Position. 178 Abstract The systems controlling positioning and assembly of the complexes that mediate plastid division are likely to be integrated with each other to allow for efficient regulation of the division process. The site of bulk assembly of the FtsZ polymers within the stroma is restricted to the mid-plastid by the Min system, composed of ARC3, MinD, MinE, MCDl , and PARC6. However, ARC6 is absolutely required for Z-ring formation in vivo. It is unclear if or how the activities of ARC6 and the Min system are coordinated to ensure efficient assembly and operation of the plastidic Z-ring. Here, we perform preliminary characterization of a novel hypomorphic allele of ARC6, arc6 D20 5 N- arc6 D 20 5N mutants possess subtle alterations in chloroplast size and morphology, relative to wild type, perhaps due to decreased amounts of ARC6 protein. Most intriguingly, arc6 D 20 SN mutants harbor misplaced and miniaturized Z-rings within the stroma, implicating ARC6 in the operation of the plastidic Min system. The conservation of ARC6D205 amongst land plants, algae, and cyanobacteria suggests that this mechanism of crosstalk between factors that position and stabilize the Z-ring probably occurs in all ARC6- and F tn2-encoding organisms. 179 Introduction The FtsZ ring (Z-ring) is a central player in bacterial cell division and plastid fission, providing some small amount of contractile force and serving as a scaffold for other division proteins (Erickson 2009, Osawa, Anderson, et al. 2008, Rothfield, Taghbalout, et al. 2005). In most bacteria, a single FtsZ gene encodes a polymer-forming GTPase with structural similarity to eukaryotic tubulins (Erickson, Taylor, et al. 1996, Mukherjee, Dai, et al. 1993). In bacteria, positioning of the Z-ring is critical to ensuring an equal distribution of cytoplasm and genetic material between new daughter cells. At least two systems provide input into the positioning of the Z-ring in E. coli: the nucleoid occlusion (Noe) system and the minicell (Min) system (Rothfield, Taghbalout, et al. 2005). The Noc system prevents chromosomal scission during division by inhibiting Z- ring formation around the bacterial chromosome (Bernhardt and de Boer 2005). Uniquely, both plastids and cyanobacteria carry multiple chromosomal copies (Birky and Walsh 1992, Doolittle 1979, Falkow, Dworkin, et al. 2006, Schneider, Fuhrmann, et al. 2007) and do not utilize proteins with sequences similar to Noc proteins (Glynn, Miyagishima, et al. 2007). Moreover, it has been shown that Z-rings can form around the chromosome in the cyanobacterium Synechococcus elongatus (Miyagishima, Wolk, et al. 2005). Based on these and other observations, nucleoid occlusion is probably not a mechanism employed in these lineages; rather it is likely that chromosomal segregation 180 in the cyanobacteria is largely managed by maintaining multiple chromosomal copies (Bin, Guohua, et al. 2007). Because of its apparent lack of relevance to plastid division, we will limit our review of the nucleoid occlusion system and focus on the other Z-ring positioning input, the Min system. The Min system in gram-negative bacteria is composed of three primary factors: MinC, MinD, and MinE (Lutkenhaus 2007). MinC binds FtsZ and inhibits Z-ring assembly by preventing lateral associations (Hu, Mukherjee, et al. 1999, Scheffers 2008, Shen and Lutkenhaus 2009). However, the FtsZ assembly-inhibiting activity of MinC is regulated by MinD, which is tethered to the membrane and promotes MinC activity only in the polar zones of the cell (Hu and Lutkenhaus 1999, Raskin and de Boer 1999a, Szeto, Rowland, et al. 2002). MinD is. regulated by MinE through a mechanism in which MinE binds to MinD and causes it to be released from the membrane (Hu and Lutkenhaus 2001). The maximum concentration of MinE occurs near the midcell adjacent to the membrane, thereby causing the concentration of active MinC and MinD to be highest at the poles (Hale, Meinhardt, et al. 2001). While the determinants for MinE localization are unknown, the midcell zone created by MinE activity allows for FtsZ protofilament assembly along the inner leaflet of the plasma membrane at the midcell (Rothfield, Taghbalout, et al. 2005). In some bacteria that lack MinE, DivIVA tethers MinD at the cell poles and inhibits FtsZ polymerization within the polar zone by maintaining a higher concentration of active MinCD at the poles (Marston and Errington 1999). 181 The components of the plastidic Min system are the only known factors that are responsible for positioning the Z-ring within chloroplasts. The Min system is now known to be made up of at least five components in Arabidopsis: AtMinD (Colletti, Tattersall, et al. 2000), AtMinE (Maple, Chua, et al. 2002), ARC3 (Glynn, Miyagishima, et al. 2007, Maple, Vojta, et al. 2007), MCDl (Nakanishi, Suzuki, et al. 2009), and PARC6 (Glynn, Yang, et al. 2009, Zhang, Hu, et al. 2009a). There is no MinC ortholog in any of the vascular plants sequenced to date (Yang, Glynn, et al. 2008), but ARC3 seems to fulfill a similar functional role in vivo (Glynn, Miyagishima, et al. 2007, Maple, Vojta, et al. 2007). On the basis of their sequence and phenotypic similarity, it was hypothesized that AtMinD and AtMinE might work similarly to E. coli MinD and MinE (Colletti, Tattersall, et al. 2000, Fujiwara, Nakamura, et al. 2004, Maple, Chua, et al. 2002). AtMinD exhibits both polar and equatorial localization in chloroplasts (Fujiwara, Nakamura, et al. 2004, Nakanishi, Suzuki, et al. 2009), suggesting that it may relocate or oscillate, similar to E. coli MinD (Raskin and de Boer 1999b), to regulate Z-ring position. AtMinE localizes to poles in tobacco chloroplasts (Maple, Chua, et al. 2002) in contrast to its mostly mid-zone localization in bacteria (Hale, Meinhardt, et al. 2001, Sun and Margolin 2001), though the localization pattern of functional AtMinE under native expression conditions in Arabidopsis has not yet been demonstrated. MCDl is required for AtMinD function and binds AtMinD, but the inputs into this plant-specific division protein are not yet clear (Nakanishi, Suzuki, et al. 2009). Similarly, PARC6 appears to inhibit Z-ring assembly through its interaction with ARC3, but further work is required in order to firlly determine how this protein is integrated into the plastidic Min system (Glynn, Yang, et al. 2009, Zhang, Hu, et al. 2009b). 182 ARC6 is a bitopic inner envelope protein of cyanobacterial origin and was shown to aid assembly and/or stabilization of the Z-ring (Vitha, Froehlich, et al. 2003). However, it was not clear if ARC6/F m2 has any role in positioning the Z-ring within chloroplasts or cyanobacteria. Here we characterize a novel allele of ARC6, arc6Dzo5N, that exhibits min-like defects in chloroplast morphology and Z-ring positioning. Consistent with its profound effect upon Z-ring placement, we show that ARC69205 is a conserved residue in all ARC6-like proteins. The effect of arc6 020 5N upon Z-ring placement probably does not occur through AtMinE, but we show here ARC6 ring formation requires both F tsZ and AtMinE. We conclude that ARC6 plays a role in influencing Z-ring placement and propose that ARC6 might affect division-site selection through PARC6 or another Min-system component. Results arc6 D20 5N is a Hypomorphic Allele of ARC6 that is Associated with Defects in Chloroplast Morphology and Number. To identify important functional domains within ARC6, we used TILLING (McCallum, Comai, et al. 2000) to identify novel alleles of ARC6 that are associated with defects in chloroplast division. A major advantage of TILLING is that in vivo relevance for any new allele can be quickly determined by the same established assay(s) used to identify mutants in forward genetic screens (Henikoff, Till, et al. 2004). We targeted the two large conserved regions within ARC6 (see Figure 1.4) for TILLING (V itha, 183 Froehlich, et al. 2003) and identified 14 novel mutations within the ARC6 locus that lead to amino acid changes within the ARC6 protein: P858, P878, D205N, SZSON, G293D, L309F, A328V, A592T, A621V, R677K, E705K, D708N, S763F, and T7731. We were able to isolate homozygous lines for each of these polymorphisms, with one exception (A328V), by screening with molecular markers (Neff, Neff, et al. 1998). Of these, only arc6 D 20 5N exhibited defects in chloroplast morphology and number. arc6 D 205N mutants have slightly enlarged chloroplasts and possess slightly fewer chloroplasts than the corresponding wild type TILLING parental line, Col-er105 (Figure 6.1); the phenotype associated with arc6 D 20 5 N is recessive to the wild type allele. In addition to the enlarged chloroplasts present within arc6 D 20 5 N, we occasionally observed mini-chloroplasts within this line (see later sections). These mini-chloroplasts could arise from asymmetric positioning of the Z-ring, in a process similar to the asymmetric division events that generate chloroplasts of varying sizes in Arabidopsis min mutants (Colletti, Tattersall, et al. 2000); however, the number and incidence of mini-chloroplasts in arc6Dzo5N was highly variable. From these results, we conclude that ARC6DZO5 is a functionally- irnportant residue in vivo that could be involved in the operation of the plastidic Min system. 184 O Chloroplasts per cell Col-er105 arc6Dzo5N Figure 6.1. arc6D205N is associated with a defect in chloroplast size and number. Chloroplast phenotypes are shown from expanded leaf cells of Col-er105 (A) and arc6 D20 5N (B). Scale bars = 10 um. Quantitative analysis of chloroplast number is shown in panel (C). Error bars represent standard deviation from the mean; 25 cells were counted for each line; the average cell plan area was 3279 :t 467 umz. 185 ARC 6 D 20 5 is conserved in land plants, algae, and cyanobacteria. To examine the evolutionary conservation of ARC6DZO5 and gain insight into its function, we generated multiple sequence alignments between several ARC6, PARC6, and Ftn2 proteins (Figure 6.2). The alignment revealed that ARC6DZO5 is a conserved residue (Figure 6.2, arrowhead) that resides within a 9 amino acid motif (Figure 6.2, line). From this analysis, we conclude that ARC6Dzo5 is likely to be a functionally- important amino acid that impacts the operation of the divisomes of chloroplasts and cyanobacteria by a similar mechanism. arc6 D 20 5 N is Associated with Miniaturized and Misplaced Plastidic Z-rings, as well as a Decrease in ARC 6 Protein Levels . To gain insight into the mechanism by which arc6 D 20 5N might impact chloroplast division, we examined F tsZ localization in arc6Dzo5N mutants. In wild type chloroplasts, FtsZ proteins colocalize to a central ring (Figure 6.3A) (V itha, McAndrew, et al. 2001). In contrast, arc6 D 20 5 N mutants possess a miniature FtsZ ring that is localized near the chloroplast periphery (Figure 6.3B). This F tsZ localization pattern suggests that ARC6 not only regulates FtsZ assembly within plastids (V itha, Froehlich, et al. 2003), but may also influence the size and position of the Z-ring. Moreover, this peripheral Z-ring localization might be connected to the unusual mini-chloroplast phenotype observed in arc6Dzo 5N plants (Figure 6.3C-D). 186 YLE YLE VVPARCG PAAVAD I T L PtPARCG ..... AtARC6 ..FKQDVVLV FLDVS OsARC6 ..FKQDVVLA YVDLS PpARC6 ..LHRDVVFA YMELS OlARC6 RRHRRDVALT LCEYG ZmARC6 ..FKQDVVLA YVDIS NsFtn2 KPYIHDIFLS ECAIA SeFtn2 RPYVHDVLLA ECSIA CwFtn2 KPYADDLVLS ECTVA TeFtn2 KPYIHDLILS ECAIA CsFtn2 KPYVHDVLLS ECAIA AtPARcs ..... T SLELG OsPARCG ..... A QQSLG MtPARCG ..... s YLEIS T LG s ;s Figure 6.2. Multiple sequence alignment showing conservation of ARC6D205. Boxed regions indicate sequence similarity and shaded regions indicate sequence identity. This portion of the alignment has been cropped from an alignment of the full- length proteins and is published elsewhere (Glynn, Yang, et al. 2009). Black line (top) indicates 9-mer amino acid motif and aspartate 205 of Arabidopsis ARC6 is indicated with a red arrowhead. Arabidopsis thaliana (At); Medicago truncatula (Mt); Oryza sativa cv. japonica (Os); Populus trichocarpa (Pt); Vitis vinifera (Vv); Ostreococcus lucimarinus (OI); Physcomitrella patens (Pp); Zea mays (Zm); T hermosynechococcus elongatus BP-l (Te); Synechococcus elongatus PCC 7942 (Se); Nostoc sp. PCC 7120 (N s); Crocosphaera watsonii WH 8501 (Cw); and Cyanothece sp. ATCC 51142 (Cs). 187 Figure 6.3. FtsZ localization and mini-chloroplasts in arc6Dzo5N. Immunolocalization of FtsZZ-l in young leaves of Col-er105 (A) and arc6 D 2 501V (B). Chlorophyll autofluorescence is indicated in red and FtsZ protein is shown in green; arrowheads point to Z-rings. Mini-chloroplast phenotype observed in arc6 D 20 SN is shown in (C), boxed region is magnified and shown in (D). Arrowheads point to mini- chloroplasts in (D). Scale bars in (A-C) are 10 um. Scale bar in (D) is 2 pm. 188 To determine if ARC6 levels are affected in the arc6 D 20 SN background, we examined the relative amounts of ARC6 and FtsZ protein in Col-er105 and arc6 D 20 5N flower buds using ARC6- and FtsZ-specific antibodies (McAndrew, Olson, et al. 2008, Vitha, McAndrew, et al. 2001). Consistent with a defect in ARC6 function, we observed less ARC6 protein in arc6Dzo5N lines than in wild type (Figure 6.4). While no ARC6 protein was detected in this background by immunoblotting, there must be a small amount of ARC6 protein present, as arc6 null mutants have only one or two chloroplasts per cell (Glynn, Froehlich, et al. 2008, Vitha, Froehlich, et al. 2003); arc6Dzo5N likely has a reduced amount of ARC6 protein that is below the threshold of detection for our ARC6 antibody (McAndrew, Olson, et al. 2008). However, it is unclear if this decrease in ARC6 protein level is the basis for the defect observed in Z-ring morphology and position (Figure 6.3B). Unlike the arc6-1 mutant (V itha, Froehlich, et al. 2003), the levels of Ftle and F tsZ2 in arc6 D20 5 N were mostly unchanged relative to wild type plants (Figure 6.4). AtMinE- YF P has a unique localization pattern in Arabidopsis and ARC6 D 20 5 mutants do not interact with AtMinE. It was previously noted that both arc6-1 mutants and arc12 (atminE) mutants have very similar phenotypes with respect to chloroplast morphology, as the leaf cells within both of these mutants typically contain only one or two large chloroplasts (Glynn, Miyagishima, et al. 2007, Pyke, Rutherford, et al. 1994, Rutherford 1996). Interestingly, 189 l 2 Félhflfl‘m-fiflfhfl a' L- l' I -‘ .‘n'-“ .H I “HMR'E'IJDIZI ”HIT-r .9 . 1 v . p i; a-ARC6 a-FtsZI a-FtsZZ RuBisCo Figure 6.4. Preliminary analysis of ARC6, Ftle, and FtsZZ levels in arc6pzo5N. ARC6 protein levels (top panel) in arc6 020 5N (lane 2) are diminished relative to wild type (lane 1), while Ftle and FtsZ2-1 levels (center panels) are mostly unchanged. An estimate of relative loading is indicated by the Coomassie stain showing RuBisCo levels in each lane; based on this, there is slightly more protein in the lane carrying extract from an arc6 020 5N plant, probably explaining the higher Ftle and FtsZ2-l signals in the mutant. 190 not only are their chloroplast morphologies similar, but F tsZ immunolocalization patterns observed within arc6 and arc12 are indistinguishable from one another (Glynn, Miyagishima, et al. 2007), suggesting that the functions of the two proteins might be connected. Because previous studies have shown that ARC6 and AtMinE have similar phenotypes, we aimed to determine if AtMinE localizes to equatorial rings in Arabidopsis, similar to ARC6 (V itha, Froehlich, et al. 2003). Curiously, previous studies have shown that AtMinE localizes to polar spots within plastids of tobacco leaf cells (Maple, Chua, et al. 2002), in contrast to its largely mid-zone localization in bacteria (Hale, Meinhardt, et al. 2001, Sun and Margolin 2001). However, these localization results for AtMinE are disputed, as they are observed in tenninally-differentiated leaf cells with very high levels of 35S-AtMinE expression — AtMinE expression in Arabidopsis is highest in regions surrounding and including the apical meristem in vegetative tissues (Itoh, Fujiwara, et al. 2001). Other studies have shown that polar localization of Arabidopsis Min proteins within tobacco may not be representative of their native localization in Arabidopsis, as AtMinD localizes to a polar spot when transiently overexpressed in tobacco (Fujiwara, Nakamura, et al. 2004), but also localizes to a ring-like structure in Arabidopsis chloroplasts (N akanishi, Suzuki, et al. 2009). To clarify the precise localization pattern for AtMinE in Arabidopsis, we introduced an AtMinEpro-MinE-YFP transgene into C 01-0 and arc12 mutants; arc12 is an AtMinE loss- of—function allele (Glynn, Miyagishima, et al. 2007). We examined emerging leaves for a YFP signal using epifluorescence microscopy. To our surprise, we saw neither AtMinE-YFP spots nor rings in these samples; AtMinE-YFP was localized to a diffuse pattern within very small plastids (Figure 6.5D), suggesting that AtMinE may only have a 191 Figure 6.5. Localization of AtMinE-YFP in Arabidopsis. Micrographs from a young emerging leaf expressing AtMinE-YFP in the arc12 background is shown in the upper panels: bightfield (A); chlorophyll autofluorescence (B); AtMinE-YFP fluorescence (C); and a merged image of chlorophyll and YFP fluorescence (D). In the lower panels, verification of functional complementation the plastid division defect in the arc12 (atminE) mutant using an AtMinEpro-AtMinE- YFP transgene is shown: Col-0 (E); arc12 (F); arc12 rescued with the AtMinEpro-AtMinE- YFP transgene. The boxed region in (D) is magnified and shown in (H). Scale bars in (A-G) = 10 um. Scale bar in (H) = 1 pm. 192 significant impact upon proplastid division rather than chloroplast division, consistent with its high level of expression within the apical meristem (Itoh, Fujiwara, et al. 2001). We did not observe any AtMinE-YFP signal within plastids of older, expanded leaves (not shown). Our AtMinEpm-MinE- YFP transgene was fully ftmctional, as most arc12 lines carrying this transgene possessed wild type chloroplasts in fully expanded leaf cells (Figure 6.5G). Further work is required to determine if AtMinE-YFP exhibits aberrant localization in the arc6 020 5N background. Because of their similar phenotypes and expression patterns, we aimed to determine if a stromal portion of ARC6 (AA154-340) interacts with processed AtMinE (AA34-229) using yeast two hybrid assays; this region of ARC6 (AA154-340) has no defined firnction, but is conserved in all ARC6-like proteins. Previous work showed no interaction between these two proteins, but because of the defect in ARC6 function observed in arc6 D 20 5 N mutants and the prospect of aspartate phosphorylation in cyanobacteria and chloroplast response regulators (Jacobs, Connell, et al. 1999, Li and Kehoe 2005, Maeda, Sugita, et al. 2006, Ruiz, Salinas, et al. 2008), we hypothesized that perhaps ARC6 might conditionally interact with AtMinE as a result of posttranslational modification. We generated site-directed mutants of ARC6 A A] 54-340, coding for D205A and D205E amino acid changes. In no instance did ARC6 interact with AtMinEAA34- 229 using two hybrid assays (not shown), suggesting that these proteins probably do not interact in Arabidopsis, despite their similar mutant chloroplast and FtsZ morphology 193 phenotypes (Glynn, Miyagishima, et al. 2007, Itoh, Fujiwara, et al. 2001, Pyke, Rutherford, et al. 1994, Rutherford 1996, Vitha, Froehlich, et al. 2003). Taken together, we conclude that the mispositioned Z-rings observed in the arc6 D 20 5 N background occur through a MinE-independent mechanism, but the localization of all of the Min system components needs to be examined in the arc6 D 20 5 N background. ARC 6 probably fimctions downstream of F tsZI/F tsZZ assembly and downstream of AtMinE. Despite its lack of ARC6 interaction with AtMinE, our FtsZ immunolocalization data for arc6 D 20 5 N still suggest that ARC6 might be somehow connected to the plastidic Min system, as arc6 D 20 SN mutants clearly possess mispositioned Z-rings (Figure 6.3 B). To confirm where ARC6 functions within the pathway of plastidic Z-ring formation, we introduced an ARC6-GFP transgene (Glynn, F roehlich, et al. 2008) into ftsZZ mutants (Schmitz, Glynn, et al. 2009) and arc12/atminE (Glynn, Miyagishima, et al. 2007) to examine its localization in these backgrounds; we cannot use traditional tests of epistasis in this case because the arc6, arc12, and ftsZ2 phenotypes are not easily distinguished from one another. In both cases, we were unable to observe ARC6-GP P localization to organized structures (Figure 6.6), though we were able to confirm expression of the fusion protein in all cases by immunoblotting (data not shown). The absence of ARC6- GFP rings was probably not a consequence of plastid size, as the enlarged plastids 194 Figure 6.6. ARC6-GFP does not localize to rings in ftsZZ or arc12 mutants. Shown is preliminary ARC6-GFP localization in Col-0 (A); an fisZZ mutant; and (C) an arc12 (atminE) mutant. The background of the mutants shown above is Col-0. The ftsZ2 mutant carries homozygous T-DNA insertions in the loci for F tsZZ-I and F tsZZ-Z (Schmitz, Glynn, et al. 2009). Scale bar = 5 pm. 195 of pdv2 mutants still form ARC6-GF P rings (Glynn, Froehlich, et al. 2008). These results indicate that at least some degree of FtsZ assembly is required for localization of ARC6 to mid-plastid rings in vivo; when FtsZ is either depleted (as in ftsZZ mutants) or filament assembly is constituitively inhibited (as in atminE/arcIZ mutants), ARC6 does not form a ring. From these results, we conclude that the plastidic Min system, or at least AtMinE, functions upstream of ARC6 and that some degree of F tsZ assembly is required for ARC6 localization to a ring in vivo. Discussion Here we perform preliminary characterization of a novel allele of ARC6, arc6 D 20 5N: that possesses defects in chloroplast morphology and Z-ring placement that appear somewhat similar to Arabidopsis min mutants. The precise connection to the Min system was initially unclear, so we looked for interaction with a Min system component, AtMinE, whose mutant phenotype is strinkingly similar to arc6 mutants. While we were unable to detect any interaction between ARC6 and AtMinE, nor provide further indication of the basis for the arc6 D 20 5N phenotype, our preliminary results do shed light on a frequently-encountered question relating to plastid division: does FtsZ ring formation precede ARC6 localization or does ARC6 localize independently of the Z- ring? Our observations of ARC6-GFP in an fisZZ depletion mutant show that FtsZ is absolutely required for ARC6 localization to the division site. Further, our examination of ARC6-GFP localization in arc12 (an AtMinE loss-of-fimction allele) shows that the presence of FtsZ protein within the plastid is insufficient for ARC6 localization; 196 chloroplast polarity and probably some degree of FtsZ polymerization are required for ARC6 localization to the mid-plastid. While the connection of ARC6 to the plastidic Min system does not appear to occur through AtMinE, it is possible that another Min system component interfaces ARC6 with the Min system. In three very recent studies, novel components or novel localization patterns of known components have come to light. Similar to ARC6, native AtMinD protein was shown to localize to mid-plastid rings in Arabidopsis (N akanishi, Suzuki, et al. 2009), in contrast to the polar localization of AtMinD when overexpressed in mature tobacco leaf cells (Maple, Chua, et al. 2002). It is possible that ARC6 conditionally interacts with AtMinD, but previous work suggests that it does not (Maple, Aldridge, et al. 2005) and the plastid phenotypes of arc6 and atminD loss-of—function mutants are quite different (Fujiwara, Hashimoto, et al. 2008, Fujiwara, Nakamura, et al. 2004, Vitha, Froehlich, et al. 2003). A novel division protein, MCDl, was recently shown to be a modifier of the plastidic Min-system and possesses partial equatorial localization within the plastid (Nakanishi, Suzuki, et al. 2009), similar to ARC6 (V itha, Froehlich, et al. 2003). However, the precise topology and function of MCDl is still unclear, making its involvement with the Min system (and possibly ARC6) difficult to address. Most recently, PARC6 (a Paralog of ARC6) was shown to play a role in inhibiting Z-ring assembly and positioning, possibly through another Min system component, ARC3 (Glynn, Yang, et al. 2009, Zhang, Hu, et al. 2009b). Like ARC6, PARC6 localizes to a ring during plastid division (Glynn, Yang, et al. 2009) and PARC6AA77-355 interacts with ARC6 A A1 54-340, possibly through the conserved motifs 197 (Figure 6.2) harboring PARC6D211/ARC6DZO5 (Y. Yang and K. Osteryoung, unpublished). Consistent with this, both ARC6 (Maple, Aldridge, et al. 2005) and PARC6 (Zhang, Hu, et al. 2009b) have been shown to be capable of self-interaction. Furthermore, a recent study has shown that Cdv3, a cyanobacterial protein with some sequence similarity to a portion of the stromal domain of PARC6, binds the cytosolic region of Ftn2/ARC6 in bacterial two hybrid assays (Marbouty, Saguez, et al. 2009). These observations indicate that the arc6 D 20 5 N mutant may lead to a change in affinity between ARC6 and PARC6 (a component of the plastidic Min system), possibly explaining the mislocalization of the Z-ring observed in arc6 D 20 5N mutants (Figure 6.3). Alternatively, the irregular position of the Z-ring in arc6 D 20 5N mutants may somehow be a side-effect of low ARC6 protein levels (Figure 6.4). Regardless, further work that analyzes the ARC6-PARC6 interaction and the impact of site-directed mutations at or near ARC6Dzo5 might provide insights into the intriguing division defects observed in the arc6 D 20 SN mutant. In the course of this work, we also analyzed the localization of AtMinE-YFP in Arabidopsis. The AtMinEpm-AtMinE- YFP transgene was able to complement arc12, an AtMinE loss-of-function mutant (Glynn, Miyagishima, et al. 2007), suggesting that the transgene is functional and the localization pattern observed is relevant. Unlike the polar spot localization of AtMinE-YFP observed in mature tobacco leaf cells (Fujiwara, Nakamura, et al. 2004), we observed no YFP signals that colocalized with chlorophyll 198 autofluorescence and no obvious expression of AtMinE-YFP in expanded Arabidopsis leaf tissue. Instead, we observed a diffuse AtMinE-YFP signal associated with very small plastids within juvenile leaves of Arabidopsis, consistent with the expression of AtMinE within and proximal to the shoot apical meristem (Itoh, Fujiwara, et al. 2001). From our analysis (Figure 6.5), it was not possible to determine if this localization pattern was associated with the inner enve10pe membrane or if it was randomly distributed throughout the stroma. From these results, we conclude that the native localization pattern of AtMinE is diffuse within the plastid, and not polar, being more similar to the localization of E. coli MinE to a broad zone within the cell (Hale, Meinhardt, et al. 2001, Sun and Margolin 2001) rather than to polar spots observed during transient overexpression assays in tobacco leaf cells (Itoh, Fujiwara, et al. 2001). Regardless, analysis of the localization of Min system components (ARC3, MinD, MinE, MCDl , PARC6, and ARC6) in the arc6 D20 5N background is still required. In a related line of investigation, we showed that ARC6-GFP fails to localize to a ring in ftsZZ or arc12 (atminE) mutants, suggesting some basal level of FtsZ assembly must occur for ARC6 to localize to a ring; this presumably means that ARC6 acts downstream of the plastidic Min system. Further analysis of the interactions and localization patterns of proteins that make up the plastidic Min system will be critical to understanding the how the site for Z- ring assembly is selected and what conditions trigger bulk FtsZ assembly at any particular site. 199 Materials and Methods Identification of EMS-induced polymorphisms in ARC6 and genotypic analysis. Single nucleotide polymorphisms (SNPs) that cause missense mutations were identified as described previously through the Arabidopsis TILLING Project: hgp://tilling.fl1crc.org[ (Till, Colbert, et al. 2006). Progeny of mutagenized lines that harbored the mutation indicated were sown on Linsmaier-Skoog medium and screened by CAPS or dCAPS-based genotyping (Neff, Neff, et al. 1998). Primers used for screening the mutant pool for mutations in the ARC6 amino terminus were: CCTCCGATTCCTCC TCCTCCTCCT (left) and GACAACCCCTGCCAACCAGGTTTC (right). Primers used for screening the mutant pool for mutations in the ARC6 amino terminus were: GGCAGGGGTTGTCTTTCCTAGGTTCAG (left) and GATCAAGGAAAAGGGTG TGCCAAGAAC (right). The polymorphisms in each line are as follows (nucleotide positions are indicated relative to the primer start position for each TILLING screen: P858 (C96T); P87L (C103T); D205N (G538R); SZSON (G761A); G293D (G890R); L309F (C937T); A328V (C995Y); A592T (G377A); A621V (C465T); R677K (G922A); E705K (G1005A); D708N (G1105A); S763F (C1271T); and T7731 (C1301T). None of the seeds from the line harboring the A328V mutation could be germinated, but this may be due to a background mutation. Phenotypic analysis of plants carrying TILLING-derived missense alleles of ARC6. Individuals that were shown to be homozygous for the indicated mutation were examined for chloroplast morphology and compared to the parental TILLING line (C01- 200 er105), using established fixation and analysis protocols (Miyagishima, Froehlich, et al. 2006, Pyke and Leech 1991, Vitha, Froehlich, et al. 2003). Multiple sequence alignment. These sequences were used for multiple sequence alignment: Arabidopsis thaliana (AtPARC6, NP_188549); Medicago truncatula (MtPARC6, IMGAG-annotated pseudomolecule AC157350_28.4); Oryza sativa cv. japonica (OsPARC6, NP_001054252); Populus trichocarpa (PtPARC6, a GENSCAN-based prediction (Burge and Karlin 1997) of Populus trichocarpa genome Scaffold_122 204390-215000); Vitis vinifera (V vPARC6, CAO48483); Arabidopsis thaliana (AtARC6, NP_199063); Oryza sativa cv. japonica (OsARC6, NP_001045726); Ostreococcus lucimarinus (OlARC6, XP_001421185); Physcomitrella patens (PpARC6, XP_001778770); Zea mays (ZmARC6, ACG29776); Thermosynechococcus elongatus BP-l (TeFtn2, BAC08309); Synechococcus elongatus PCC 7942 (SeFtn2, ABB57973); Nostoc sp. PCC 7120 (N sFtn2, BAB74406); Crocosphaera watsonii WH 8501 (CwFtn2, EAM48783); and Cyanothece sp. ATCC 51142 (CsFtn2, ACB49642). Multiple sequence alignment was performed with MEGA4 (Tamura, Dudley, et al. 2007) using an identity matrix. F tsZ immunolocalization. Tissue preparation, fixation, and immunofluorescence analysis were carried out as described previously (Miyagishima, Froehlich, et al. 2006, Vitha, Froehlich, et al. 2003, Vitha, McAndrew, et al. 2001). 201 Immunoblotting. Plant extracts for determination of F tsZ and ARC6 protein levels were taken from floral bud tissue. Flower buds were ground in liquid nitrogen and prepared for 10% SDS- PAGE using 6X sample buffer according to previously-established protocols (Wiegel and Glazebrook 2002). Immunoblotting for Ftle and FtsZZ-l protein was performed as described previously (Stokes, McAndrew, et al. 2000). Immunoblotting of plant extracts was performed using an ARC6-specific antibody at 1:2500 in TBS-T containing 5% nonfat dry milk. The blot was washed several times in TBS-T before applying the secondary antibody. The anti-rabbit HRP-conjugated secondary was used at 1:5000 in TBS-T containing 5% nonfat dry milk and the blot was washed several times in TBS-T before applying the HRP chemilluminescent substrate (Therrno Scientific, Inc.) and exposing to film. Yeast two hybrid analysis. ARC6 and AtMinE clones for two hybrid analysis were made by PCR and cloned into pGADT7 and pGBKT7, respectively. Mutations causing the missense mutations D205A and D205E in ARC6 were generated by SOE-PCR and cloned into pGADT7. Growth assays (HIS3 reporter assays) were performed according to methods as described previously (Glynn, Froehlich, et al. 2008, Maple, Aldridge, et al. 2005). 202 Generation of AtMinEpm-AtMinE—YFP, transformation, selection, and microscopy. Analysis of ARC 6—GFP localization. The AtMinEpm-A tMinE- YFP transgene was generated by amplifying the promoter and coding regions of the AtMinE locus with primers TTTT'TTTCTGCAGACTTGTTT CAAAACGACTGTGTTTTTTG and CCAGAGAGATCTCTCTGGAACATAAAAATC GAACCTGACATC by PCR. This fragment was cloned into a derivative of pCAMBIA- 1302 (Hajdukiewicz, Svab, et al. 1994) carrying C-terrninal EYFP coding sequence by digesting the vector and PCR products with Pstl and BglII; digestion with Pstl and BglII removes the 35S promoter from the plant transformation vector. Clone was sequenced and shown to be free of coding errors prior to transformation of Agrobacterium and Arabidopsis. Plants were transformed as described previously (Clough and Bent 1998). Selection of transgenic individuals was performed on Linsmaier-Skoog media containing hygromycin (25 ug/mL). Hygromycin-resistant lines were transplanted to soil and allowed to recover for several days prior to analysis. Analysis of ARC6-GFP localization in arc12 and ftsZZ-I ftsZZ-Z double mutants was conducted using GFP fusion vectors as previously described (Glynn, Froehlich, et al. 2008, Vitha, Froehlich, et al. 2003) and selecting on Linsmaier-Skoog media containing hygromycin (25 ug/mL). Hygromycin- resistant lines were transplanted to soil and allowed to recover for several days prior to microscopic analysis. 203 Acknowledgements I thank Mia Hemmes for backcrossing the arc6 020 5N TILLING mutant, Joyce Bower for assisting with quantitative analysis of the arc6 D 20 5 N phenotype, and Deena Kadirjan—Kalbach for continuing technical assistance with this project. 204 Chapter 7 Conclusions and Future Directions 205 Summary of ARC6 and PARC6 Functional Analysis and Future Directions. In Figure 7.1, I show a schematic detailing the functional domains we have characterized within ARC6 and PARC6, in addition to previously annotated features. ARC6-FtsZZ Interaction Our body of work here identifies a prospective domain within ARC6 that binds the C-terminus of FtsZ2 family members (Figure 7.1, ZBD) and might have structural similarity to E. coli ZipA. While our pulldown assays and in vivo analysis of ARC6F44ZD site-directed mutants negates the criticality of the core hydrophobic phenylalanine (F442) of ARC6ZBD we had hypothesized based on two-hybrid results, rigorous structural analysis is still required to fully refiite the proposed ZipA-like structure we show in Chapter 2. While an X-ray crystal structure or NMR-based solution structure of ARC6ZBD is highly preferred, a comparison between circular dichroism (CD) spectra from ARC6ZBD and ZipAZBD may be sufficient for a preliminary test of the model we show in Figure 2.4 (Greenfield 2006, Whitrnore and Wallace 2008). If ARC6ZBD and ZipAZBD have similar CD profiles, it is reasonable to assume that their structures are probably similar. However, if their CD profiles are different, it would suggest that ARC6ZBD has a structure distinct from that of ZipAZBD; this result would be consistent with secondary structure predictions of ARC6ZBD (Figure 2.5) and in vivo analysis of ARC6ZBD in Arabidopsis (Figure 2.9 and Figure 2.13). However, we still 206 ARC6 801 AA 1.... .1 m: \TM llimcrimtion‘.’ / M1 {TM2 1 I a I I — a a; aw; .2:- , in ARC3 Interaction l‘JEVI Interaction PARC6 819 AA Figure 7.1. Domain Architecture of ARC6 and PARC6 Proteins. ARC6 (top) has a transit peptide (TP, AA 1-67); a predicted J-domain (J, AA 89-153); a hypothesized dimerization domain (AA 154-340) containing a conserved aspartate at position 205 (red dot); an FtsZZ-binding domain (ZBD, AA 351-503); a transmembrane domain (TM, AA 615-635); and a PDV2-binding domain (PBD, AA 636—759) containing a serine residue at position 744 that influences ARC6-PDV2 interaction and is predicted to be a phosphoacceptor site (asterisk, *). ARC6AA68_614 resides in the chloroplast stroma and ARC A A636-801 resides within the intermembrane space. PARC6 (bottom) has a transit peptide (TP, AA 1-76); an ARC3-interacting region (AA 77-357); a hypothesized dimerization domain (AA 169—335) containing a conserved aspartate at position 211 (red dot); two predicted transmembrane domains (TMl, AA 357-377; TM2, AA 574-596); and a PDVl-interacting region (AA 575-819). PARC6 A A77-3 56 is stromal, but the C- terminus of the protein has yet-undefined topology within the chloroplast. 207 emphasize that outside of its affinity for the C-terminus of FtsZ2 proteins, the precise function of ARC6ZBD remains undefined, but hypothesize that this domain is somehow involved in assembly or stabilization of FtsZ polymers within the chloroplast. To determine if ARC6ZBD has a direct effect upon the polymerization of Ftle and FtsZ2, analysis of FtsZ copolymer assembly in the presence of ARC6ZBD could be conducted. The effect of ARC6ZBD upon F tsZ assembly could be analde using light scattering assays and electron microscopy following treatment of FtsZ assembly reactions (containing Ftle and FtsZZ) with recombinant ARC6ZBD protein according to previously established protocols (Lu and Erickson 1998, Olson 2008). ARC 6-PD V2 Interaction We have shown that ARC6 is responsible for positioning PDV2 at the division site within dividing chloroplasts and that PDV2-positioning activity occurs within the intermembrane space of the chloroplast (Glynn, F roehlich, et al. 2008). In Chapters 3 and 4 we showed that the ARC6-PDV2 interaction only requires a small domain contained within the IMS-localized portion of the ARC6 protein (Figure 7.1, PBD). While we were unable to generate a contiguous structural model of this domain within ARC6, the structural similarity of the IMS-localized region of PDV2 to the Rst phosphoprotein phosphatase (Figure 4.3) (Delumeau, Dutta et al. 2004, Dutta and Lewis 2003, Hardwick, Pane-Farre, et al. 2007) led us to hypothesize that interaction between ARC6 and PDV2 might be controlled by differential phosphorylation of ARC6. Using multiple sequence comparison of ARC6/F m2 proteins and phosphorylation prediction 208 algorithms, we identified one high-confidence phosphoacceptor site (S744) and two lower-confidence phosphoacceptor sites (S740 and T742) within ARC6pBD. Phosphomimetic mutations at one of these sites (ARC6S744E) disrupts the ARC6-PDV2 interaction (Figure 4.5) and impedes chloroplast division (Figure 4.6 and Figure 4.7), but does not grossly alter PDV2 localization to the division site (Figure 4.8). The presence of dumbbell-shaped plastids and multiple Z-rings at the division site within chloroplasts of ARC6S744E-expressing lines are strikingly similar to those reported for both pdv2 and arc5 mutants (Gao, Kadirjan-Kalbach, et al. 2003, Miyagishima, Froehlich. et al. 2006). In other words, the ARC6S744 E phosphomimetic mutation generates two effects: (1) decreased dynamin activity at the outer envelope membrane and (2) an increase in Z-ring assembly within the stroma. While it is unclear if the ARC6S744 E mutation simply impedes dynamin activity by altering ARC6-PDV2 interaction affinity or by upregulating Z-ring assembly within the stroma (thus creating a rigid physical barrier that blocks constriction of the plastid membranes), the change of a single residue within the IMS- localized portion of ARC6 certainly leads to upregulation of Z-ring assembly within the stroma. However, it remains to be seen whether ARC6 residues S740, T742, or S744 are actually subject to posttranslational modification in vivo. There is no evidence of phosphorylation of any of these residues based on current phosphoproteomic data provided in the PhosPhAt database (hfip:flphosphatmpimp—golm.mpg.de[) (Heazlewood, Durek, et al. 2008), but this might reflect the transient nature of phosphorylation of the ARC6 protein in vivo — perhaps only a small fraction of ARC6 molecules are phosphorylated at these sites at any one time or are only phosphorylated under certain 209 environmental or developmental conditions. To address this potential mechanism further, it is suggested that in vitro assays are conducted in which radiolabeled phosphate is used to detect phosphorylation and/or dephosphorylation of ARC6S744 following treatment of recombinant ARC6pBD with plant cell or chloroplast extracts. If ARC6pBD is a substrate for a phosphatase and/or kinase, a change in the amount of radioactivity should be observable following treatment of recombinant proteins with plant cell extracts; the amount of radiolabel carried by S740A, T742A, and/or S744A site-directed mutants could be compared to that of the wild type recombinant protein to demonstrate that these residues are actually the target of a phosphatase and/or kinase enzyme contained within cell extracts. Similar methods have been utilized for other proteins (Ben-Nissan, Cui, et al. 2008, Xu, Wong, et al. 2008) and may provide insight into the validity of the hypothesis that ARC6 activity is controlled through phosphomodification of ARC6pBD. PARC 6-ARC3 Interaction. In Chapter 5, we introduced PARC6, a novel chloroplast division gene that probably arose as a result of ARC6 duplication in the tracheophyte lineage. While ARC6 and PARC6 do share some degree of sequence similarity, they vary significantly in function. Unlike ARC6, which promotes Z-ring assembly (V itha, Froehlich, et al. 2003), PARC6 appears to inhibit F tsZ assembly within the stroma (Figure 5.6) and does not directly interact with Ftle or F tsZZ (Figure 5.7). However, we showed that PARC6 interacts with ARC3, a protein with functional similarity to bacterial MinC and that this interaction requires the MORN region of ARC3 (Figure 5.8) — the MORN region of 210 ARC3 inhibits its interaction with F tle proteins (Maple, Vojta, et al. 2007). Based on these data, I hypothesize that PARC6 inhibits FtsZ assembly through ARC3: when PARC6 is present, it sequesters the MORN region of ARC3, allowing ARC3 to bind Ftle and disrupt the Ftle/FtsZ2 copolymer that makes up the Z-ring; in the absence of PARC6, the MORN region of ARC3 is exposed and inhibits ARC3 interaction with F tle, allowing for assembly of the Ftle/FtsZ2 copolymer. Furthermore, because of the . high PARC6-YFP signal intensity at the isthmus connecting the dividing plastid during the late stages of division (Figure 5.4) and because parc6 mutants commonly have chloroplasts that are dumbbell-shaped (Figure 5.3 and Figure 5.9), I hypothesize that PARC6 functions to aid disassembly of the Z-ring during the late stages of division. The plastid morphology phenotypes in parc6 mutants could result from disorganized or inefficient Z-ring disassembly in the absence of PARC6, causing hyperstabilized Z- ring(s) to block the final stages of plastid division. To determine whether this model is correct, in vitro assays could be conducted using purified recombinant proteins to monitor the assembly of the Ftle/FtsZ2 copolymer (Lu and Erickson 1998, Olson 2008) in the presence of various domains of ARC3 and PARC6. For this example, I hypothesize that recombinant ARC3 lacking its transit peptide and MORN region would be capable of inhibiting assembly of the Ftle/FtsZZ copolymer. In contrast, a recombinant form of ARC3 protein lacking only its transit peptide would have no effect on Ftle/FtsZ2 copolymer assembly due to the presence of the MORN region of ARC3, as this region of ARC3 inhibits interaction between ARC3 and Ftle (Maple, Vojta et al. 2007). The addition of the N-terminal stromal region of PARC6 to assembly assays containing Ftle/FtsZ2 and mature recombinant ARC3 protein would lead to 211 sequestration of the MORN region of ARC3, thereby allowing ARC3 to bind F tle and inhibit assembly (or perhaps promote disassembly) of the Ftle/FtsZ2 copolymer, as bulk FtsZ polymer assembly only occurs in the presence of both FtsZ] and FtsZZ protein (Olson 2008). The use of complex, multiprotein mixtures in assembly assays will be critical for determining how plastidic FtsZ assembly is regulated in vivo. PARC6-PDV] Interaction. . Another possible explanation for the common observation of dumbbell-shaped plastids in parc6 mutants might be a defect in dynamin (ARC5) activity at the outer envelope. Previously, PDVl was shown to be critical for full ARC5 pinchase activity at the outer envelope (Miyagishima, Froehlich, et al. 2006). In Chapter 5, we demonstrated that PARC6 is required for PDVl localization to sites of constriction (Figure 5.10) and that PARC6 interacts with PDVl in two-hybrid assays (Figure 5.11), suggesting that PARC6 might directly position PDVl during plastid division, thereby influencing ARC5 pinchase activity by directly recruiting PDVl to the division site. However, this hypothesis is somewhat cursory due to the lack of knowledge regarding the topology of PARC6. Our preliminary analysis indicated that PARC6 may only traverse the inner envelope membrane (Glynn, Yang, et al. 2009), but this experiment was preliminary and may not possess the sensitivity to detect multiple or transient topological orientations of PARC6 within the envelope membrane(s) in dividing chloroplasts; a complete diagnostic of PARC6 topology needs to be examined using protease protection assays to determine the validity and/or plausibility of an interaction between PARC6 and PDVl. 212 Furthermore, we also observed the persistence of a PARC6 spot at one pole of the chloroplast (Figure 5.4), presumably representing its localization following separation of the two daughter plastids. Similar localization patterns have been observed for PDVl and ARC5, but not for PDV2 or ARC6. We suspect that the similarity of PARC6, PDVl , and ARC5 at this stage represents a multiprotein complex that is retained on one pole of the chloroplast, possibly serving as a marker or reference point to aid division-site selection during the next round of division, perhaps analogous to the way some yeast species maintain cell polarity following cell division (Amberg, Zahner, et al. 1997, Drubin 1991, Glynn, Lustig, et al. 2001, Mine, Bratrnan, et al. 2009). Isolation of this putative polarity-determining complex from chloroplasts using affinity chromatography of affinity-tag-labeled PARC6 folowed by LC-MS/MS might allow for the identification of the factors composing this complex and provide insights into how polarity might be established within daughter plastids following division. A RC 6 is a Component of the Plastidic Min System. Historically, I had thought of the plastidic Min system and ARC6 as two separate entities with little or no connection between their apparent functional roles. However, as a result of two major observations, 1 now believe ARC6 function is integrated into the plastidic Min system that firnctions to position the Z-ring at the division site: (1) expression of ARC65744 E in vivo results in the production of elongated chloroplasts containing multiple sites of constriction within petiole cells (Figure 4.9), similar to phenotypes observed in petiole cells of Arabidopsis atminD/arc] I loss-of-function mutants (Fujiwara, Hashimoto, et al. 2008, Fujiwara, Nakamura, et al. 2004) and 35S- 213 AtMinE overexpressing plants (Reddy, Dinkins, et al. 2002); and (2) the minicelling-like phenotype and aberrant Z-ring position in arc6 D 20 5 N mutants (Figure 6.3) is consistent with a role for ARC6 in promoting equatorial positioning of the Z-ring prior to chloroplast division. To further examine this phenomenon, it is suggested that the localization patterns of other Min system proteins (AtMinD, AtMinE, ARC3, and PARC6) are examined in these mutant backgrounds to determine if the D205N or S744E mutations somehow impact function of the plastidic Min system. Additionally, preliminary data shows that a conserved segment of the stromal regions of ARC6 and PARC6 family members (Figure 5.1, grey underline) is important for homo- or hetero- dimerization between these proteins (Y. Yang and K. Osteryoung, unpublished). The conserved aspartate residues within these segments of ARC6 (ARC6D205) and PARC6 (PARC69211) might be critical for their ability to dimerize within the chloroplast stroma (Maple, Aldridge, et al. 2005, Zhang, Hu, et al. 2009a) and/or mediate their communication with other components of the plastidic Min system. Interestingly, the cyanobacterial Cdv3/DivIVA division protein (Marbouty, Saguez, et al. 2009, Miyagishima, Wolk, et al. 2005) also shares some degree of sequence similarity with PARC6 (J .M. Glynn and K. Osteryoung, unpublished) and like both parc6 and arc6 D 20 5 N mutants, Cdv3 -deficient cyanobacteria possess ectopic Z-rings (Marbouty, Saguez, et al. 2009), suggesting that dimerization of ARC6 with PARC6 (or Ftn2 with Cdv3 in the case of cyanobacteria) serves to interface ARC6/Ftn2 with the Min system. Consistent with this idea, the loss of Cdv3/DivIVA from the plastid divisome during the evolution of land plants (J .M. Glynn and K. Osteryoung, unpublished) loosely 214 corresponds with the acquisition of PARC6 in vascular plants (Glynn, Yang. et al. 2009). Further analysis of the interaction properties and localization of site-directed mutants of ARC6D205 and PARC69211 may provide key insights into the coordination of Z-ring placement with organized Z-ring assembly and disassembly in chloroplasts. Unanswered Questions, Additional Observations, and New Hypotheses. Does ARC 6 function as a DnaJ-like cochaperone? While shown to have sequence and probable structural similarity to DnaJ -1ike cochaperones (Vitha, Froehlich, et al. 2003), it remains to be seen whether the predicted J-domain of ARC6 actually stimulates the chaperone activity of a DnaK-like protein. Moreover, it is unclear as to what function this chaperone activity serves within the chloroplast. One possibility is that ARC6 is required for efficient import and/or refolding of FtsZ following import into the chloroplast stroma. Two points support this hypothesis: DnaJ/DnaK type chaperone systems have been shown to be important for protein import into eukaryotic organelles (Ivey and Bruce 2000, Zhang, Elofsson, et al. 1999); and intraplastidic FtsZ protein levels are significantly diminished in arc6 mutants (Figure 1.3) (V itha, F roehlich, et al. 2003). It is reasonable to suggest that comparative import experiments might highlight such an activity; perhaps higher efficiency FtsZ protein import would be observed in the presence of increasing amounts of ARC6 protein if ARC6 contributes to FtsZ import and/or folding. 215 A second possibility is that the J -domain activity is important for chaperoning the assembly or disassembly of FtsZ proteins into ordered structures within the stroma. Bacterial dnaK and dnaJ mutations affect cell division (Clarke, Jacq, et al. 1996, McCarty and Walker 1994), FtsZ dynamics (Uehara, Matsuzawa, et al. 2001), and overexpression of FtsZ has been shown to complement the cell division defects observed in a dnaK mutant (Bukau and Walker 1989). Furthermore, a recent study has suggested that DnaK chaperone activity is important for proper localization of the Z-ring in E. coli (Sugimoto, Saruwatari, et al. 2008). Presumably, the role of chaperones in Z-ring assembly and cell division in bacteria is conserved in plastids. Unfortunately, the analysis of plastidic hsp 70 T-DNA mutants has revealed partial redundancy between the two Arabidopsis chloroplast Hsp70 proteins and implicated them in chloroplast biogenesis in vivo; mutants that lack both plastidic Hsp70 proteins are non-viable (Su and Li 2008) and as a result are probably uninformative with regard to the role of these proteins in FtsZ import or assembly in vivo. However, the analysis of J -domain mutants of ARC6 is currently underway (Y. Yang and K. Osteryoung, unpublished) and may provide some exciting insights into this aspect of ARC6 function. Is ARC 6 localized to thylakoid membranes? Interestingly, analysis of proteomic data from plastid membrane fractions suggests that the localization and topology of ARC6 might be more complex or more dynamic than import and fractionation assays suggest. The initial analysis of ARC6 fractionation and t0pology was performed using an in vitro approach with isolated pea 216 chloroplasts; this work placed ARC6 within the inner envelope membrane (V itha, Froehlich, et al. 2003). However, more recent work indicates that ARC6 is present within the inner envelope and also the thylakoid membrane system, as multiple ARC6 peptide fragments are found in isolated thylakoid membrane fractions (http://ppdb.tc.comell.edu/) (Sun, Zybailov, et al. 2009). While it is possible that these ARC6-derived peptides simply represent contamination of the thylakoid membrane fraction, this experimentally-determined thylakoid fractionation of ARC6 might represent a novel function for ARC6 during the early stages of division, or perhaps occurs as a result of the process of membrane breakage and resealing during the final stage of daughter plastid separation. It is unclear as to whether the thylakoid membrane is extricated from the division furrow prior to the final pinch, or if portions of the thylakoid membrane fuse with the inner envelope. Such a fusion event could lead to temporary relocation of ARC6 from the inner envelope to the thylakoid membrane, but the fate and/or purpose of these thylakoid-localized ARC6 molecules is unclear. How is aynamin (ARC5) recruited to the division site? While the roles of ARC6 and PARC6 in recruiting the PDV proteins to the division site are becoming more transparent, it is still unclear how the PDV proteins might recruit dynamin (ARC5) to the division site. Interestingly, pdvl pdv2 double mutants are blocked in their ability to recruit ARC5 from cytosolic patches, but pdvl and pdv2 mutants can localize ARC5 at the division site (Miyagishima, Froehlich, et al. 2006). Similarly, arc6 mutants are defective in their ability to recruit ARC5 patches from the cytosol (Glynn, Froehlich, et al. 2008) and phenocopy pdvl pdv2 mutants in this 217 regard (Miyagishima, Froehlich, et al. 2006). While PDVl and PDV2 localize to the chloroplast periphery in arc6 mutants, neither of the PDV proteins localize to a central ring in the arc6 background (Glynn, Froehlich, et al. 2008); this suggests that PDVl and/or PDV2 must be concentrated at a single site within the outer envelope to facilitate dynamin recruitment to the dividing chloroplast. Curiously, neither PDVl (Miyagishima, F roehlich, et al. 2006) nor PDV2 (M. Hemmes and J .M. Glynn, unpublished) interact with ARC5 in yeast two-hybrid assays. I have also constructed novel N-terminal BiF C vectors to test for identify interaction between PDVl-ARCS, PDV2-ARC5, PDVl- PDV2, PDVl-PDVl, PDV2-PDV2, and ARC5-ARC5 — but these experiments have yet to be performed. Collectively, these data hint at a factor (or factors) that bridge PDV proteins to ARC5, and aid ARC5 patch recruitment from the cytosol. The N-terminal domains of PDVl and PDV2 are both localized in the cytosol, exhibit some low-level sequence similarity with each other, and have coiled-coil regions, indicating that they might have related functions (Glynn, Froehlich, et al. 2008, Miyagishima, Froehlich, et al. 2006). Unfortunately, the PDV proteins of land plants share little sequence similarity to any other proteins (Miyagishima, Froehlich, et al. 2006). To gain further insight into the proteins or entities that might bridge the PDV proteins with ARC5 recruitment, I performed comparative structural analysis of the N- terrninal domains of PDVl (AA 1-206) and PDV2 (AA 1-213) to all extant PDB structures using the MetaServer at http://bioinfo.pl (Ginalski, Elofsson, et al. 2003). 218 The best match for PDVl A A1-206 was a crystal structure of the C-terminus of human E31 (1 YIB), a protein domain with homodimeric fold comprised of a coiled-coil and four-helix bundle (Slep, Rogers, et al. 2005). E31 belongs to a conserved protein family that localizes to microtubule plus ends and recruits cell polarity and signaling molecules to growing microtubule tips (Askham, Moncur, et al. 2000, Bu and Su 2003, Nakamura, Zhou, et al. 2001, Slep and Vale 2007, Wen, Eng et al. 2004). The N- terrninus of EB] is required for binding microtubles (Barth, Siemers, et al. 2002, Hayashi and Ikura 2003), while the C-terminus (i.e. the portion of EB] with structural similarity to PDVl A A1-206) has been shown to be a cargo-recognition domain. The C-terminus of EB] family proteins bind several cargo proteins that regulate microtubule dynamics, including APC in mammals (Su, Burrell, et al. 1995), Kar9p in yeast (Miller, Cheng, et al. 2000), and may be involved in crosslinking the microtubule and actin cytoskeletons through spectraplakin-like proteins (Slep, Rogers, et al. 2005). The best match for PDV2 A A1-213 was a crystal structure of the repetitive elements of fruit fly Spectrin (2SPC). Spectrin is a cytoskeletal protein descended from an a-actinin-like ancestor (Dixson, Forstner, et al. 2003). Spectrins enhance the mechanical stability of membranes and promote the assembly of specialized membrane domains (Broderick and Winder 2005, Kordeli 2000). Spectrin functions as a tetramer (Liu and Palek 1980) that cross-links transmembrane proteins, membrane lipids, and actin; this cross-linking activity occurs directly (Brenner and Korn 1979) or through 219 adaptor proteins such as ankyrin and 4.1 (Bennett and Baines 2001, Goodman, Krebs, et al. 1988). Dynamins do not generally interact directly with actin, but are typically associated with membrane structures (Wiejak and Wyroba 2002). My preliminary structural analysis of the PDV proteins point toward a mechanism where PDVl (and possibly PDV2) might reorganize the actin and/or microtubule cytoskeleton to facilitate trafficking of dynamin (ARC5) patches to the cytosolic side of the chloroplast outer envelope through microtubule and/or actin-anchored motor proteins. Consistent with this hypothesis, at least two kinesin-like proteins, At5g10470 and At4g38950, have PDVl- like mRN A expression profiles (J .M. Glynn, unpublished) based on analysis using Expression Angler (http://bbc.botany.utoronto.ca/) (Toufighi, Brady, et al. 2005). Furthermore, the membrane remodeling properties of Spectrin-like proteins, perhaps including PDV2, could create lipid microdomains that may be necessary (Lemmon 2004) for attracting the pleckstIin homology domain of ARC5 (Gao, Kadirjan-Kalbach, et al. 2003, Hong, Bednarek, et al. 2003), distributing ARC5 around the organelle surface, and/or facilitating stimulation of ARC5 pinchase activity at the division site (Yoshida, Kuroiwa, et al. 2006). 220 Working Model of the Coordination of the Chloroplast Divisome in Tracheophytes. In Figure 7.2, I show a rudimentary model that reflects our current understanding of the roles of ARC6 and PARC6 in coordinating the assembly and operation of the plastid divisome. The plastid PD rings (Kuroiwa, Kuroiwa, et al. 1998) are omitted from this model, as their constituents are yet unknown and uncharacterized. This schematic reiterates the complexity of the process of plastid division and hints at several new hypotheses regarding the operation and regulation of some of the components that make up the plastid divisome. Firstly, while ARC6 certainly plays a role in the assembly and/or stabilization of the Z-ring, the mechanism by which it does so remains enigmatic. Our attempt to identify the discrete region of ARC6 that binds FtsZ2 family members within plant chloroplasts was a necessary step, but indicates that this issue may be more complicated than previously thought. Beyond its role in regulating FtsZ assembly within the stroma, we showed that ARC6 also mediates the activity of division factors that act upon the outer envelope. ARC6 is required to position PDVl and PDV2 within the outer envelope membrane. ARC6 positions PDV2 by direct interaction, but only influences PDVl localization indirectly through its paralog, PARC6. Our genetic and cytological analyses suggest that ARC6 is probably required for PARC6 localization and/or activity during division. It remains to be seen if ARC6 interacts directly with PARC6 in vivo or if other proteins bridge these two division factors. Furthermore, the precise topology of PARC6 is still rather murky, making it difficult to determine the validity of PARC6- PDVl interaction data. The interaction between PARC6-ARC3 and the similarity of 221 ®® / Cfioskeletal reorganization, Motor Proteins, Membrane Reorganization? PDV2 ‘l" "it...“ u C 4; OEM C C IMS i IEM ARC6 , PARC 6 Stroma Z-ring Assembly Z-ring Disassembly \/ Z-ring Dynamics Figure 7.2. Coordination of Z-ring Dynamics and ARC5 Recruitment through ARC6 and PARC6. PD rings are omitted from this model for simplicity. Division is initiated by FtsZ assembly at the division site, which is determined by the Min system (not shown). ARC6 may influence Z-ring positioning and stabilize the Z-ring at the division site through its interaction with FtsZ2. PARC6 is a Min system component hypothesized to inhibit Z-ring assembly through ARC3. ARC6 and PARC6 form homodimers; ARC6 and PARC6 may (?) form heterodimers within the stroma. PDV2 is positioned in the outer envelope by interacting with ARC6 in the IMS. ARC6 indirectly positions PDVl in the outer envelope, but does so through PARC6. Preliminary data suggests that PARC6 binds PDVl in the cytosol. It is unknown (?) if PDVl and PDV2 form hetero- or homodimers in vivo. The recruitment of ARC5 may occur through cytoskeletal reorganization, kinesin-driven directed movement, and/or membrane lipid reorganization by the PDV proteins. Not all detail is shown. See text for details. 222 their respective mutant phenotypes indicate that these two proteins might work together to inhibit FtsZ assembly, perhaps contributing to Z-ring disassembly during the final stages of division and/or aiding division-site selection along with other components of the plastidic Min system. While previously thought to be distinct from the Min system, ARC6 may also contribute to positioning of the Z-ring, based on analysis of ARC6S744 E and arc6 020 5N alleles. Though the exact connection is unclear, ARC6 may interface with the Min system through heterodimerization with its paralog, PARC6. The recruitment and activity of dynamin (ARC5), both regulated by the PDV proteins, remains a complicated matter. While both PDVl and PDV2 are presumed to be required for full ARC5 contractile activity, either PDV protein is sufficient for ARC5 recruitment to the division site. ARC5 recruitment may occur through reorganization of the cytoskeleton, directing the dynamin-like protein to the division site through a vesicular transport mechanism that utilizes microtubule and/or actin-tracking motor proteins like kinesins (microtubules) or myosins (actin). Further work on this aspect of the division process may not only prove useful for improving our understanding of plastid replication, but may provide novel insights into general aspects of dynamin recruitment and activity within other types of eukaryotes as well. 223 APPENDIX 224 Appendix A NEXT and CIENA: Pipelines for Identification of Plastid Division Genes. 225 Summary Reverse genetic screens for plastid division factors have relied heavily upon gene identification and characterization of division factors in other systems; when a new division component is identified in E. coli or another prokaryote, a researcher typically performs a BLAST query against finished plant genomes to determine if similar factors might contribute to organelle division. Several plastid division genes have been identified using this approach, including: FtsZ (Osteryoung, Stokes, et al. 1998, Osteryoung and Vierling 1995), GCl (Maple, Fujiwara, et al. 2004), AtMinD (Colletti, Tattersall, et al. 2000), and AtMinE (Itoh, Fujiwara, et al. 2001).' However, this mode of gene identification is limiting and novel approaches are required to determine the complete inventory of genes involved in organelle division. Here we highlight two platforms that we tested in an attempt to identify new plastid division factors. These platforms utilize genomic context and/or transcriptional profiling to identify new candidate genes and assign them priority. NEXT: A Platform for Identification of Candidate Genes by Genome Neighborhood Analysis and Priority Ranking Based on Gene Expression and Intracellular Targeting. To start our search for new plastid division genes, we utilized The Seed gene neighborhood analysis tool (http://www.theseed.org/mltiMain_Pag§) (Overbeek, Begley, et al. 2005). In prokaryotes, conserved gene proximity can be indicative of 226 related firnctions, as genes linked to a similar cellular process are often clustered together on the bacterial chromosome. Prokaryotic gene operons are a popular example of gene clustering (Overbeek, Fonstein, et al. 1999), and gene-gene association approaches have been used successfully to define the function of several genes of unknown function (de Crecy-Lagard 2007, de Crecy-Lagard and Hanson 2007). We pinned known prokaryotic cell division genes (Miyagishima, Wolk, et al. 2005) and then examined the surrounding genomic sequence amongst several bacterial species for coding loci that frequently reside within proximity (< 5000 bases) to a query gene. We identified approximately 80 candidates that commonly co-occur with our set of query genes. Within this set, 55 of these had orthologous sequences in Arabidopsis using BLAST-based queries (Altschul, Madden, et al. 1997). We then used Genevestigator (hgpszflwwwgenevestigatorcomD (Zimmerrnann, Hirsch-Hoffmann, et al. 2004) to prioritize our panel of ~55 candidates. Genevestigator is a gene expression analysis tool that allows the user to compare the degree of correlative expression between any two genes in Arabidopsis; we used ARC 6 (At5g42480) as a reference gene, as it is likely to be a critical regulator of plastid division in vivo (Marrison, Rutherford, et al. 1999, Pyke, Rutherford, et al. 1994, Vitha, Froehlich, et al. 2003), and arranged our candidates in order, based on the degree of expression profile similarity to ARC6. To further prioritize candidates coming out of the pipeline described above, we used the TargetP (httgflwwwcbs.dtu.dl_