unsatt. .§.\A 3a yaw“ fiw . ‘ fisurfl u. . $4 ‘ _ ”:9 5.3.x z. . u . : parka: 1 “mm". 35%. 3 In. {1}.) :95...“ ti. .Jflnhlhhiciu 1 .. 6.. x. 3.6!...thaflhfll 222....nahz 3.51.. 5.... a; x .. ‘5: 81!}: ‘1‘ :‘arfiwgzvit! A. .13).;- . v! .5 . 1.1.0.2“:5 4% fig. a mommy “rm-:31; Michigan State M Unuversuty This is to certify that the dissertation entitled UNDERSTANDING THE ORIGIN AND FUNCTION OF ORGANELLAR METABOLITE TRANSPORT PROTEINS IN PHOTOSYNTHETIC EUKARYOTES: GALDIERIA SULPHURARIA AND ARABIDOPSIS THALIANA AS MODEL SYSTEMS presented by MARC LINKA has been accepted towards fulfillment irements for the Ph.D. Genetics Major Professor’s Signature 12/05/2008 Date MSU is an Affirmative Action/Equal Opportunity Employer I.-o-n-OOI-o-I-----o-u-o-u-—---o--n--.—-—.----o-.-o-c-. l ¢I 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 K:IProj/Acc&Pres/CIRC/DateDue.indd UNDERSTANDING THE ORIGIN AND FUNCTION OF ORGANELLAR METABOLITE TRANSPORT PROTEINS IN PHOTOSYNTHETIC EUKARYOTES: GALDIERIA SULPHURARIA AND ARABIDOPSIS THALIANA AS MODEL SYSTEMS By Marc Linka A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Genetics 2008 ABSTRACT UNDERSTANDING THE ORIGIN AND FUNCTION OF ORGANELLAR METABOLITE TRANSPORT PROTEINS IN PHOTOSYNTHETIC EUKARYOTES: GALDIERIA SULPHURARIA AND ARABIDOPSIS IHALIANA AS MODEL SYSTEMS By Marc Linka Membrane-bound compartments, especially the organelles mitochondria and plastids, are a hallmark of eukaryotic cells. Organellar metabolite transport proteins facilitate the exchange of metabolites across membranes in a cell and are crucial for connecting biochemical pathways that operate in separate compartments. Of particular interest to plant scientists is the plastid organelle of photosynthetic eukaryotes. Plastids synthesize and deliver major biological molecule classes such as carbohydrates, fatty acids, amino acids, and nucleic acids to the rest of the cell and thus the plastid has to be extensively connected to the cytosol. Plastids originated fi'om an endosymbiotic cyanobacteria-like ancestor about 1.6 billion years ago and in this thesis the differences between plastid metabolite transporters of the eukaryotic model plant Arabidopsis thaliana and prokaryotic cyanobacteria were investigated. A phylogenomic analysis of 83 predicted plastid metabolite transporters from Arabidopsis thaliana has been conducted in collaboration with Dr. Debashish Bhattacharya. These studies allowed the conclusion that the majority of the transport proteins in extant plastids are absent fi'om free-living cyanobacteria and originated from eukaryotic host genes. They represent true innovations associated with organelle evolution. Transporters became likely targeted to the endosymbiont via the endoplasmic reticulum of the host early in its evolution. Furthermore, the results suggest that export of photosynthates from the plastid in form of sugar-phosphates has been a selective advantage to set-up a permanent endosymbiosis between the host and the endosymbiont. While these sugar-phosphate transport proteins are conserved in all photosynthetic eukaryotes, their biochemical properties co-evolved to meet the specific metabolic requirements in the distinct groups of the eukaryotic kingdom. As reported in Chapter 3, further studies have shown that in contrast to higher plants, the red alga Galdieria sulphuraria has a high affinity export system for tn'ose- phosphates and lacks hexose-phosphates transport across the plastid envelope membrane. This reflects an adaptation for an efficient export of photosynthates from the organelle due to an absence of a plastidic starch pool in red algae. Metabolite carriers facilitate also the transport of compounds in a single, highly compartmentalized cellular pathway. A prime example is the photorespiratory pathway. 2-phosphoglycolate is produced by the oxygenase reaction of the enzyme ribulose- 1,5- bisphosphate Carboxylase/Oxygenase (RubisCO) and subsequently recycled to 3- phosphoglycerate in the compartments chloroplast, cytosol, peroxisome, and mitochondrion. A reverse genetic approach was used to identify the transport proteins involved in photorespiration and 32 candidates have been designated. Five of these were genetically analyzed to test their role in the recycling of 2-phosphoglycolate, leading to the discovery of a novel transporter required for a functional photorespiratory pathway. This protein is localized to the inner envelope membrane of mitochondria and the transporter most likely imports a cofactor from the cytosol, which is required for the mitochondrial glycine decarboxylase enzyme complex. ACKNOWLEDGEMENTS I would like to sincerely thank Dr. Andreas Weber for this one-time opportunity to join his research group and complete my graduate studies here at Michigan State University. Under his supervision I developed my confidence and capabilities to inquire into the phenomena on my daily scientific endeavor. He allowed me to be a student, an instructor, a teacher and an independent thinker. He always encouraged me to broaden my scientific knowledge and to enjoy research. I would also like to thank the Director Dr. Barbara Sears and Associate Director Dr. David Amosti, the faculty and Jeannine Lee of the Genetics Graduate Program for organizing such a great educational program at Michigan State University. The fulfillment of the requirements during the first three years of my graduate studies was of paramount importance for my scientific progress and will be for my future goals. My committee members Dr. Robert Larkin, Dr. John Ohlrogge, Dr. Steve van Nocker, and Dr. Richard Allison deserve a special note. They contributed by discussion and opinions to my progress and allowed the transfer to Germany without any concern during my last year. I would also like to thank all graduate students, postdocs and faculty members for their interaction and discussion at seminars, in the laboratories, and off—campus. In addition, these acknowledgements would not be complete if I did not mention the former and recent members of “The Weber Lab”. Thanks for your support and your tremendous scientific and non-scientific help. Most importantly, I would like to thank my wife Nicole. You are the sunshine of my life. I thank my parents, for their support to leave "Old Europe Into The West". But I have to confess that there are places on this Earth, which are even farther away than Michigan. iv TABLE OF CONTENTS LIST OF TABLES .................................................................................. vii LIST OF FIGURES .................................................................................. viii KEY TO ABBREVIATIONS ........................................................................ x INTRODUCTION PLASITD EVOLUTION FROM THE PERSPECTIVE OF METABOLITE TRANSPORT PROTEINS .......................................................................... 1 References ........................................................................................ 25 CHAPTER 1 PHYLOGENY OF THE PLASTID PHOSPHATE TRANSLOCATOR FAMILY ....... 34 Abstract ............................................................................................. 35 Introduction ..................................................................................... 35 Material and Methods .......................................................................... 36 Results and Discussion ........................................................................... 37 References ....................................................................................... 45 CHAPTER 2 PHYLOGENY AND ORIGIN OF PLASTID METABOLITE TRANSPORT PROTEINS FROM ARABIDOPSIS THALIANA ................................................. 47 Abstract .......................................................................................... 48 Introduction ....................................................................................... 49 Material and Methods ............................................................................ 52 Results and Discussion ........................................................................... 56 References ......................................................................................... 81 CHAPTER 3 ANALYSIS OF THE PLASTID PHOSPHATE TRANSLOCATOR FAMILY FROM THE RED ALGA GALDIERIA SULPHURARIA ................................................ 87 Abstract .......................................................................................... 88 Introduction ..................................................................................... 89 Material and Methods .......................................................................... 92 Results ............................................................................................ 95 Discussion ..................................................................................... 117 References ..................................................................................... 127 CHAPTER 4 PROPOSED METABOLITE TRANSPORT PROCESSES FOR THE RE-ASSIMILATION OF AMMONIA DURING PHOTORESPIRATION ............... 133 Abstract ....................................................................................... .134 Results .......................................................................................... 134 References ...................................................................................... 146 CHAPTER 5 A REVERSE GENETIC APPROACH TO IDENTIFY THE PROPOSED METABOLITE TRANSPORT PROTEINS IN THE PHOTORESPIRATORY PATHWAY.. .................................................................................................................. 149 Abstract ......................................................................................... 150 Introduction .................................................................................... 151 Material and Methods ........................................................................ 155 Results ........................................................................................... 168 Discussion ...................................................................................... 201 References ..................................................................................... 208 CHAPTER 6 CONCLUSION AND PERSPECTIVES ........................................................ 215 References ..................................................................................... 225 LIST OF TABLES Table 2.1. List ofArabidopsis thaliana chloroplast solute transporters analyzed in this study and their putative evolutionary origins ................................. 58 Table 3.1. Oligonucleotide primers designed in this study ................................... 97 Table 3.2. Molecular characteristics of the plastidic phosphate transporter GsTPT, GsPPT and GSGPT from Galdieria sulphuraria ..................................... 100 Table 3.3. Comparison of kinetic constants from Galdieria sulphuraria and higher plant pPT homologues ...................................................................... 1 13 Table 5.1. Primer sequences used in Chapter 5 ............................................... 158 Table 5.2. Co-expressed photorespiratory enzymes from Arabidopsis thaliana... . . . 170 Table 5.3. Putative transporter genes, which are transcriptionally co-regulated with the photorespiratory pathway ............................................................... 171 Table 5.4. Systematic search for the BOU substrate ......................................... 196 Table 5.5. Presence of a membrane potential does not stimulate glycine of serine transport by BOU .................................................................................. 197 Table 5.6. Several cofactors must be provided for the mitochondrial glycine decarboxylase (GDC) .................................................................... 198 vii LIST OF FIGURES Figure l. The photorespiratory pathway ......................................................... 24 Figure 1.1. Maximum likelihood phylogeny of endomembrane and plastid translocators .......................................................................................... 42 Figure 1.2. Model for plastid and translocator origin in the common ancestor of the Plantae during primary endosymbiosis ........................................................... 43 Figure 1.3. Model for plastid and translocator origin in the common ancestor of the chromalveolates during secondary endosymbiosis .............................................. 44 Figure 2.1. Origin of plastid targeted solute transporters in Plantae ........................ 61 Figure 2.2. Plastid targeted solute transporters of putative 'Host' origin in Plantae. . 64 Figure 2.3. Plastid targeted solute transporters of putative ’cyanobacterial'l‘plastid endosymbiont' origin in Plantae ................................................................... 67 Figure 2.4. Plastid targeted solute transporters of putative 'Chlamydia-like' origin in Plantae ...................................................................................... 74 Figure 2.5. Plastid targeted solute transporters of 'Other’ or 'PIantae-specific' origin in Plantae ..................................................................................... 79 Figure 3.1. Protein alignment of plastidic phosphate transporters from Arabidopsis thaliana and Galdieria sulphuraria ............................................... 102 Figure 3.2. Expression of plastid phosphate transporter (pPT) homologues from Galdieria sulphurarz'a in yeast ............................................................... 105 Figure 3.3. Time-kinetics of pPT homologues from Galdieria sulphuraria .............. 108 Figure 3.4. Internal substrate dependence on the transport activities of GsTPT, GsPPT, and GSGPT ................................................................... 112 Figure 3.5. Transcript abundance and protein activity of GsTPT, GsPPT, and GSGPT in Galdieria sulphuraria cells ............................................ 1 16 Figure 3.6. Flexible transport of triose-phosphate across the plastid membrane is central for the cellular carbon metabolism in the red alga Galdieria sulphuraria. . .. 125 Figure 4.1. Simplified scheme of the current textbook photorespiratory cycle. . . . . .138 viii Figure 4.2. A citrulline/ornithine shuttle between mitochondria and plastids ............. 142 Figure 4.3. A glutamate/glutamine Shuttle between mitochondria and plastids .......... 145 Figure 5.1. Isolation of A. thaliana T-DNA insertion lines for At1g54350, At2g42770, At2g48070, At3g26710, and At5g46800 ......................................... 176 Figure 5.2. Growth phenotype of T-DNA insertion lines from the candidate genes Atlg54350, At2g42770, At2g48070, and At3g26710 ......................................... 178 Figure 5.3. Growth phenotype of T-DNA insertion line bouI ............................... 180 Figure 5.4. Concentrations of photorespiratory marker metabolites glyoxylate, ammonium ions, glycine, and serine in bouI leaves ............................................ 183 Figure 5.5. Carbon dioxide-dependent effect on additional metabolite levels in b0ul...185 Figure 5.6. Molecular complementation of bouI .............................................. 188 Figure 5.7. Complementation of the bard mutant restores glycine concentrations to wild type levels ................................................................................. 189 Figure 5.8. Mitochondrial glycine oxidation is impaired in bad ........................... 190 Figure 5.9. Irnmunoblot analysis of mitochondrial enzymes ................................. 1.91 Figure 5.10. Heterologous expression and purification of BOU protein in Escherichia coli and Saccharomyces cerevisiae ............................................................... 195 Figure 5.11. BOU relevant branch from the maximum-likelihood unrooted tree of the mitochondrial carrier family (MFC) 199 Figure 5.12. Reconstituted BOU protein exhibits ion channel properties .................. 200 ix KEY TO ABBREVIATIONS 2-PG: 2-phospoglycolate 3-PGA: 3-phosphoglycerate or-KG: alpha-ketoglutarate ADP: adenine diphosphate AMP: adenine monophosphate ATP: adenine triphosphate At: Arabidopsis thaliana A.thaliana: Arabidopsis thaliana CaClZ: calcium chloride CCM: carbon concentrating mechanism cDNA: complementatry DNA cMDH: chloroplastic malate dehydrogenase C02: carbon dioxide DCT: dicarboxylate transporter DHAP: dihydroxyacetone phosphate DiT: dicarboxylate transporter DNA: deoxyribonucleic acid EGT: endosymbiont gene transfer E. coli: Escherichia coli ER: endoplasmic Reticulum EST: expressed sequence tag E-value: expectation-value FAD: flavin adenine dinucleotide FBPase: fructose-1,6-bisphosphatase Fd: ferredoxin Fd-GOGAT: glutamate synthase Frcl,6BP: fructose-1,6-bisphosphate Frc6P: fi'uctose-6-phosphate Gal: galactose GallP: galactose-l-phospbate GAPDH: glyceraldehydes-B-phosphate dehydrogenase GDC: glycine decarboxylase gDNA: genomic DNA GFP: green fluorescent protein GGT: glutamatezglyoxylate aminotransferase Glc: glucose Glc 1 P: glucose- 1 -phosphate Glc6P: glucose-6-phosphate Glu: glutamate Gluc6P: gluconate-6-phosphate Gly3P: glycerol-3-phosphate GLYK‘ glycerate kinase GOGAT: glutamate synthase GOX: glycolate oxidase GPT : glucose-6-phosphate/phosphate translocator xi Gs: Galdieria sulphuraria G. sulphuraria: Galdieria sulphuraria GS: glutamine synthetase H+: proton H20: water H202: hydrogen peroxide HGT: horizontal gene transfer HPR: hydoxypyruvate reductase IPTG: isopropyl B-D-l-thiogalactopyranoside Ki: inhibitory constant KM: Michaelis-Menten constant Li: lithium LiAc: Lithium acetic acid Mal: malate MCF: mitochondrial carrier family MgCl2: Magnesium chloride mMDH: mitochondrial malate dehydrogenase MP: maximum parsimony NAD(P): nicotinamide adenine dinucleotide (phosphate) NAD(P)H: reduced nicotinamide adenine dinucleotide (phosphate) NJ: neighbor joining NDP: nucleotide diphosphate NH3: ammonia xii NH4+: ammonium ion 02: oxygen 0AA: oxaloacetic acid OPPP: oxidative pentose phosphate pathway pABA: para-aminobenzoic acid PCR: polymerase chain reaction PGK: 3-phosphoglycerate kinase PEP: phosphoenolpyruvate PGP: phosphoglycolate phosphatase Pi: orthophosphoric acid PLP: pyridoxal-phosphate pMDH: peroxisomal malate dehydrogenase pPT: plastid phosphate translocator PPT: phosphoenolpyruvate/phosphate translocator PR: photorespiration Pyr: pyruvic acid RibSP: ribose-S-phosphate RNA: ribonucleic acid RubisCO: ribulose- l ,5-bisphosphate carboxylase/oxygenase RuBP: ribulose- 1 ,5-bisphosphate S. cerevisiae: Saccharomyces cerevisiae SGT: serinezglyoxylate aminotransferase SHMT: serine hydoxymethyltransferase xiii TCA: tricarboxylic acid THF: tetrahydrofolate Tic: translocon at the inner chloroplast envelope Toc: translocon at the outer chloroplast envelope TP: triose-phosphate TPT: triose-phosphate/phosphate translocator UDP: uridine diphosphate UMP: uridine monophosphate XPT xylulose-S-phosphate/phosphate translocator xiv INTRODUCTION PLASTID EVOLUTION FROM THE PERSPECTIVE OF METABOLIT E TRANSPORT PROTEINS INTRODUCTION Approximately three billion years ago, an ancestor of the extant free-living cyanobacteria introduced oxygenic photosynthesis to a literally oxygen-free atmosphere (Blankenship and Hartman, 1998). Descendants of these early oxygenic photosynthetic organisms are sustaining the oxygen concentration in the present Earth’s atmosphere (Blankenship, 1992; Cavalier-Smith, 2006; Falkowski and Godfrey, 2008). Sunlight is absorbed and utilized by the photosynthetic apparatus for withdrawing electrons fi'om water, thereby generating a proton gradient and oxygen (02) as a byproduct. The electrons and protons are stored as usable chemical energy to assimilate inorganic nutrients (i.e., C, N, P and S) into organic matter. Thus, autotrophic organisms are of central importance for the Earth, providing the essential 02 for cellular respiration of aerobic organisms, and are the primary producers at the base of almost every food chain (Field et al., 1998). More than one billion years ago it is likely that a phagotrophic eukaryote gained the ability to conduct photosynthesis by engulfing a cyanobacterium-like prokaryote, thus changing fiom a heterotrophic to a mixotrophic and subsequently autotrophic lifestyle (Yoon et al., 2004; Cavalier-Smith, 2006). This permitted the invasion of and thriving in new habitats. Over millions of years the endosymbiotic cyanobacterium evolved into the organelle plastid (Reyes-Prieto et al., 2007). Today’s photoautotrophic organisms exist ubiquitously on the Earth with a remarkable diversity of species and appearance (Bhattacharya et al., 2004). The smallest and largest photosynthetic eukaryotic organisms described to date are the green micro-alga Ostreococcus tauri (Lanier et al., 2008) and the Giant Sequoia tree, respectively. The red alga Galdieria sulphuraria and green alga Chlamydomonas species adapted to the upper (56°C) (Gross, 2000) and lower temperature limits (Morgan-Kiss etal., 2006) of eukaryotic life. Phototrophic organisms recently attracted much attention as renewable and sustainable sources for biomass and energy production and as alternative bioreactors to the established eubacteria, yeasts and human cell lines (Gordon and Polle, 2007; Field et al., 2008; Raja et al., 2008; Rubin, 2008). Plastids are the site of carbon dioxide, nitrite, and sulfur assimilation (Calvin, 1962; Crawford, 1995; Kopriva, 2006). They partially or fully conduct the biosynthesis of various metabolite classes, such as fatty acids, starch, amino acids, nucleic acids, isoprenoides (chlorophyll, carotenoids, gibberelline) or phenylpropanoids (cell wall components, pesticides, anthocyane) (Weber et al., 2005). To understand and fully utilize the potential of photosynthetic eukaryotes, it is helpful to reconstruct the events that led to plastid origin and decipher the common features and specific adaptations in extant, distantly related photosynthetic organisms. Plastid origin from a cyanobacterial—like prokaryote was postulated already at the turn of the 19th and 20th century by microscopists Sachs, Altrnann, and Schimper (reviewed in Kutschera and Niklas, 2005) and led in 1971 to the “Theory of Endosymbiosis” by Lynn Margulis (Margulis and Nealson, 1989). Based on numerous single and multi-gene phylogenetic analyses fiom diverse plastid genomes and nuclear- encoded plastid-targeted genes (Martin et al., 2002; Reyes-Prieto and Bhattacharya, 2007a), it is now commonly accepted that the primary endosymbiosis had been a single event in eukaryotic evolution and all photosynthetic species containing primary plastids have been recently summarized in the supergroup Archaeplastida (Adl, 2005). Within approximately 150 million years, the unicellular primordial alga (i.e., the protoalga) had been established, and it gave rise to the three major lineages of the Archaeplastida: Glaucophyta (glaucophyte algae), Rhodophyceae (red algae) and Chloroplastida (green alga/plants) (Y oon et al., 2004;Ad1, 2005; Reyes-Prieto and Bhattacharya, 2007b). Multicellularity evolved independently in the Rhodophyceae and in the Chloroplastida, whereas only the Chloroplastida mastered the colonization of the land. This was accompanied by an extraordinary radiation of the existing species (Bhattacharya et al., 2004). Heterotrophic eukaryotes later captured green or red algae, respectively, transforming them to secondary plastids of green algal or red algal origin (Gould et al., 2008). The process of secondary endosymbiosis gave rise to a plethora of new species mainly micro- and macro-algae, dominating aquatic environments as major primary producers (Field et al., 1998). Serial secondary endosymbiosis by replacing a secondary plastid by another one and even tertiary endosymbiosis via an engulfinent of algae harboring a secondary plastid by another heterotrophic eukaryote are evident (Gould et al., 2008). Also of interest is the complete or partial loss of plastid functions in several species (Fast et al., 2001). Noteworthy is also continuing evolution of novel photosynthetic organisms of primary endosymbiotic origin, as it lms been recently reported for the host species Paulinella (Nowack et al., 2008). The antiquity of the last common ancestor, the recurrent acquisition of the plastid and species radiation has scrambled the evolutionary signal of plastid origin. In the last decade, plastid-, EST- and complete genome sequence projects fi'om a diverse group of photosynthetic organisms were initiated by the research community to clarify the complex host-endosymbiont partnership. The green lineage has been followed from the unicellular and multi-cellular green algae to mosses (Cove, 2005; Merchant et al., 2007), gymnosperms and several mono- and dicotyledonous plants (Arabidopsis Genome Initiative, 2000; Goff et al., 2002; Tuskan et al., 2006; Pavy et al., 2007). A catalog of sequences fi'om diverse micro- and macro-a1 gae with primary, secondary and even tertiary plastids was initiated to compile red algae evolution (Matsuzaki et al., 2004). The genome of the glaucophyte alga Cyanophora paradoxa is currently sequenced to complement the available sequence resources with the third major lineage of the Archaeplastida (Reyes-Prieto et al., 2006; Loeffelhardt, 2007). Besides all the diversity, common themes and specialization in plastid evolution can be now validated and addressed: Endosymbiont gene transfer (EGT) to the host nucleus massively reduced the plastid genome size (Tirnmis et al., 2004). The model-plant Arabidopsis thaliana has a chloroplast genome of ~155 kb with 87 protein-coding genes (Sato et al., 1999), which is just a small fraction of extant cyanobacteria! genomes. On the other hand, bioinformatic and proteomic predictions assume nearly 4,500 proteins in the organelle (Martin et al., 2002; Sun et al., 2008). Expression and re-import of these nucleus-encoded plastid targeted proteins has been predicted for a long time to be a hallmark for a successful organelle establishment (Dyall et al., 2004). An intriguing question is to what extent and for what purpose cyanobacteria] genes had been transferred and maintained in the host nucleus. According to Sato et a1. (2005) and Martin et al. (2002), the A. thaliana nuclear genome harbors between 1,200 — 4,500 genes of cyanobacteria] origin, respectively. The majority of these cyanobacteria] genes have non-plastid function (Martin et al., 2002), hence, the present plastid proteome evolved from multiple sources including cyanobacterial, host proteins and eubacterial genes acquired by horizontal gene transfer. Data from glaucophytes and red algae suggest a much smaller portion of transferred cyanobacterial genes and in contrary to A. thaliana, the majority of these candidates are plastid targeted (Reyes-Prieto et al., 2006). Although these findings are not in full agreement, they still demonstrate a Significant and complex contribution of EGT to the evolution of the host cell. Host genes and endosymbiont genes have been replaced by one another, duplicated, diverged and several times redirected to a new compartment or even existing homologous genes recombined with each other. A challenge will be to follow up these rearrangements in numerous model organisms and assemble the tree of photosynthetic eukaryotes. Prerequisites for the gene transfer to the nucleus were (i) the regaining transcriptional and translational activation and (ii) the retargeting of plastid-needed proteins (Martin, 2003). Recent studies demonstrated the transfer of plastid-encoded antibiotic resistance genes that were introduced by plastome transformation to the nucleus, including gene activation, within a short timefi‘ame (Huang et al., 2003; Stegemann et al., 2003). More elaborate must have been the evolution of the plastid protein import machinery. The sophisticated Toc (Translocon at the outer chloroplast envelope) and Tic (Translocon at the inner chloroplast envelope) system evolved not only to import the majority of the proteins across both envelope membranes, but also to distribute them to one of the six possible destinations (outer and inner envelope membrane, interrnembrane space, plastid stroma, thylakoid membrane and lumen) (Inaba and Schnell, 2008). Comparative analysis of the Toc and Tie apparatus across many photosynthetic eukaryotes points again to a monophyly of plastids and suggests its early establishment, likely already in the protoalga. Core parts of the cyanobacterial protein transport machinery have been recruited and were functionally replaced or improved with new elements, which are absent in the cyanobacterial membrane (Gould et al., 2008). Another major control point in plastid biogenesis was the establishment of coordinated division of endosymbiont and host cell (Miyagishima, 2005). Unrestricted growth of the endosymbiont would harm the host and inhibiting its division would ultimately lead to its loss. The plastid division apparatus follows the same evolutionary path described earlier for the protein import apparatus (Miyagishima, 2005). The organelle division is based on ancient bacterial binary fission and several genes with cyanobacterial and other eubacterial origin are still conserved in extant plastids. The nuclear genome took control over the expression of most of the genes (Glynn et aL, 2007). In land plants, the transfer is completed and none of the proteins are encoded on the organelle genomes sequenced to date, but some algal genomes sporadically still encode division factors (Miyagishima, 2005; Glynn et al., 2007). Analogous genes replaced or had been added as division components originated from eukaryotic and eubacterial sources. Even after the split of the Archaeplastida, lineage-specific remodeling occurred. This makes it difficult to provide a universally valid plastid division model and in particular challenging to find new division related genes via sequence-based comparative analysis (Glynn et al., 2007). From the engulfinent of the plastid ancestor to the split into the Glaucophyta, red algae and green algae/plants, a period of 150 million years has been estimated (Y oon et al., 2004). All above-mentioned major events in plastid origin are results of a complex multi-gene acquisition and build the foundation of all descendants of the proto-alga. To date, it is not satisfactorily answered how the two organisms managed to coexist long enough to provide sufficient time for the transformation of the endosymbiont into an organelle. In the following paragraph and in chapters one and two I will outline the potential of metabolite exchange and integrating metabolisms to set-up a strong mutual symbiosis, which ultimately led to an enslavement of the cyanobacterimn by the protist. This initial metabolic interaction might have provided the time needed for the emergence of a full- fledged organelle, including large scale EGT and establishment of effective protein import apparatus and division machinery. My aim was to use comparative genomics to unravel the differences of the metabolite transport protein composition between the plastid inner envelope membrane and the plasma membrane of free-living cyanobacteria. In this work, I could Show that recruitment of host-derived metabolite transporters to the plastid envelope membrane was very likely a driving force in the transition of a mutualistic endosymbiosis to the proto-alga. Chapters one and two had been prepared in collaboration with the laboratory of Debashish Bhattacharya (University of Iowa) and were published in two different scientific journals. In In the proposed scenario, a complete cyanobacterium was engulfed by a phagotrophic eukaryote and escaped digestion in the food vacuole. Eventually, the membrane surrounding the cyanobacterium was lost or merged with the cyanobacterial outer membrane (Cavalier-Smith, 2000), resulting in two membranes marking the barrier to the new environment: the cytosol of the host. The composition of the outer membrane changed considerably over evolutionary time scales (Block et al., 2007). Eubacterial lipoproteins and lipopolysaccharides are absent fi'om higher plant chloroplasts and the phospholipid phosphatidylcholine was specifically introduced into the outer envelope membrane from the endomembrane system of the host (Douce and Joyard, 1990; Cavalier-Smith, 2000). Extant Archaeplastida display an intimate connection between the Endoplasmic Reticulum (ER) and the envelopes of the plastid. Lipids are synthesized in the plastid stroma and a sophisticated traffic system has been established between plastid and ER in both directions for the biogenesis of plastidial and extra-plastidial membranes (Benning, 2008). As a conserved feature of extant primary plastids, the Tic/Toe protein import apparatus recognizes most nuclear-encoded genes via a N—terminal leader sequence of 25 - 125 amino acids (Soil and Schleifi’, 2004). Still, some plastid-targeted proteins do not exhibit an obvious leader sequence and an alternative protein translocation route via the ER and Golgi vesicle sorting system has been recently discovered (Villarejo et al., 2005). The host-plastid interaction was repeatedly established through endosymbiosis in secondary plastid evolution. Three to four envelope membranes surround secondary plastids (Bhattacharya et al., 2004) and nuclear-encoded proteins are delivered by a multi-partite leader sequence through the concerted action of the ER sorting system and the Tic/Toc related import machinery of the organelle (Kilian and Kroth, 2003 ). These direct contacts between ER and plastid support the idea that vesicle transport to and from the plastid was initiated at an early stage of endosymbiosis. ER or Golgi-derived vesicles initially might have fused randomly with the outer membrane of the cyanobacterium, releasing their contents into the periplasmatic space and the proteins were inserted into the cyanobacterial plasma membrane or taken up by the cyanobiont. Thus, even if insertion of host metabolite transporters into the cyanobiont’s plasma membrane were initially not very efiicient, it would have given the host the opportunity to connect the cyanobiont with its cytosol and tap into photosynthates fi'om the cyanobacterium. This could have become a selective advantage for the phagotrophic eukaryote and the endosymbiont and favored the establishment of a permanent partnership. In chapter one we address the phylogenetic origin of the plastid-targeted triose- phosphate/phosphate transporter (TPT). As a starting point, I used the available EST database from the unicellular red alga G. sulphuraria to annotate and assemble the homologous plastid phosphate translocator (pPT) genes. This work was part of the publication: “EST analysis of the thermo-acidophilic red microalga Galdieria sulphuraria reveals potential for lipid A biosynthesis and unveils the pathway of carbon export from rhodoplasts” (Weber et al., 2004). The results led us to hypothesize that the ancestor protein of the TPT has evolved fi'om a host-derived endomembrane transporter, which was “mistargeted” to the cyanobacterial membrane and thus enabled a metabolic connection between the endosymbiont and the host. The TPT as a member of the pPT family was the first solute transport protein identified on a molecular basis in the inner envelope membrane in higher plants (F liigge, 1999). This transporter is of particular interest, because it is the major route for export of assimilated carbon from the plastid into the cytosol dining the day in seed plants. Our analysis in chapter one revealed that homologous TPT genes are widespread in the red algal lineage including chromalveolate algae with secondary plastids, but related genes are absent from cyanobacteria and other prokaryotes. Export of triose-phosphate is truly an innovation of plastid evolution. The pPT family originated from nucleotide-sugar transporter present in the endomembrane system of extant eukaryotes. However, none of the extant pPT members 10 do transport nucleotide sugars, thus leading to the question how a nucleotide sugar transporter might have fimctioned in the initial establishment of a metabolic connection between cyanobiont and host. Nucleotide-sugars are the substrate for literally all glycosylation reactions in a cell. They are incorporated into disaccharides, oligo- and polysaccharides, which are utilized as compounds for transport, storage, structural integrity and post-translational modifications in the cell. Recently, (Deschamps et al., 2008) highlighted the importance of the nucleotide-sugar transporters for merging the storage polysaccharide metabolism in host and cyanobiont. According to their hypothesis, both organisms initially harbored complete and independent sets of enzymes to synthesize, store and mobilize carbon in their respective compartments. The insertion of an NDP-sugar transporter would have allowed the host to withdraw the cyanobacterial starch precursor ADP-glucose fiom the cyanobiont. The key assumption of this hypothesis is that ADP-glucose represents an activated form of organic carbOn that has already been committed to storage, thus it can be withdrawn from the evolving plastid without disturbing its primary carbon metabolism. However, ADP-glucose is a poor substrate for the soluble starch synthase in heterotrophic eukaryotes. To counter this argument, it has been argued that the gene encoding soluble starch synthase of cyanobacterial origin has been transferred to the host nucleus at an early stage and was subsequently expressed in the cytosol of the host. Importantly, this would only require the transfer of a single gene to the host. It is not necessary for the gene product to be retargeted to the cyanobacterium and proteins for mobilization of the host a-glucans are already present. Thus, the proposed scenario is a very simple one in evolutionary terms — it does require only a single gene transfer and ‘mistargeting’ of a single endomembrane- 11 derived transporter. The scenario would be even simpler if the ancestral host starch or glycogen synthase would have accepted ADP-glucose as precursor, even with low affinity. In this case, the initial connection would not have required EGT or evolutionary innovations. Nucleotide-sugar transport proteins operate in a strict counter-exchange mode and the transport direction depends on the concentration of the substrates on either sides of the membrane. The export of NDP—sugar is compensated with an import of nucleotide- monophosphates, such as AMP or UMP (Handford et al., 2006; Rollwitz et al., 2006). In the endosymbiosis context, this counter-exchange mode of transport would have prevented the depletion of the cyanobacterial nucleotide pool. A relatively higher concentration of ADP—glucose at its side of biosynthesis inside the endosymbiont would favor its eXport and immediate consumption in the cytosol. As outlined above, the transporter was likely inefficiently inserted and only carbon committed to storage was drained, thus it did not affect the viability of the symbiont (Deschamps et al., 2008). However, to date, a paralogous NDP-sugar export system has not been identified in the inner envelope membrane of any extant photosynthetic eukaryote. Identification of such a transporter would represent the missing link that would lend strong support to the “metabolic symbiosis and birth of the Plant kingdom” hypothesis put forward by Deschamps et a1. (2008). According to my work presented in chapter one, the pPT gene family with at least three distinct members already evolved in the last common ancestor of the red algae and the green algae and plants, shortly after the initial endosymbiosis was established (Weber et al., 2006). They all are equipped with an N—terminal leader sequence for an efficient retargeting and insertion by the Toc and Tic system. As a result, 12 the major flux of sugar-phosphates across the membrane of plastid in higher plants was reorganized fi'om an NDP-sugar transporter towards a triose-phosphate (TP)/phosphate, phosphoenolpyruvate (PEP)/phosphate and glucose-6-phosphate (Glc6P)/phosphate shuttle system (Flfigge, 1999). The transport systems fit perfectly into metabolic compartrnentalization of higher plants: triose-phosphate is exported to the cytosol and is mainly used for sucrose or cell wall biosynthesis. PEP is taken up from the cytosol to drive fatty acid biosynthesis and synthesis of compounds by the shikimate pathway (i.e. aromatic amino acids and a heterogeneous group of phenolics). The glucose-6-phosphate transporter (GPT) provides glucose-6-phosphate (G6P) for starch synthesis and the oxidative pentose phosphate pathway (OPPP) in heterotrophic tissues, or the Ribulose- 1,5-bisphosphate Carboxylase/Oxygenase (RubisCO) bypass reaction in mixotrophic tissues. In summary, the most likely scenario in evolutionary terms is that the NDP-sugar transporter represented the starting point for connecting the previously independent carbon metabolisms of host and cyanobiont. Gene duplications and further radiation of the gene family delivered the genetic material for the evolution of functional diversification, eventually promoting the evolution of a more efficient targeting system to the plasma membrane of the endosymbiont. The findings about the phylogenetic origin of the plastid phosphate transporter family encouraged me to extend the phylogenomic analysis of the repertoire of plastid metabolite transporters to a broad range of different photosynthetic organisms. In present- day plastids of seed plants, photosynthesis provides the energy for the Calvin-Benson- Cycle and several other anabolic pathways, including the synthesis of starch, fatty acids, 13 several amino acids, nucleic acids, and the reductive assimilation of inorganic ions like nitrate and sulfate. For each pathway, a number of precursors and intermediates have to be transported across the envelope membrane (Weber et al., 2005). Plastids must be extensively connected to the cytoplasm to distrrbute the numerous products within the cell. In contrast, the cyanobacterial membranes generally prevent leakage of their primary metabolites to the surrounding The plasma membrane is mainly configured for the uptake of inorganic and organic nutrients and the excretion of waste products or components for the periplasmic space, the outer membrane, and cell wall (Kaneko et al., 1996). In chapter two we addressed two questions concerning the transporter evolution of the plastid inner envelope membrane. First, we sought to estimate the contribution of host and symbiont on the transporter repertoire in photosynthetic eukaryotes. Is the pPT the rule or the exception for host proteins, which had been inserted into the cyanobacterial plasma membrane? Second, we determined the differences between the A. thaliana transporter set and photosynthetically active prokaryotic and eukaryotic organisms. This would give us for the first time an insight into the conserved and lineage- specific metabolite “permeome” of higher plant plastids. To address these hypotheses, we analyzed the phylogenetic origin and distribution of well-annotated transporter proteins from A. thaliana. Transport proteins have characteristic hydrophobic transmembrane spans to reside in the membrane as well as hydrophilic domains to allow interaction and transport of their substrates (Schwacke et al., 2003). In silica searches for genes in the A. thaliana genome with a predicted plastid target sequence and hydrophobic domains provided an initial catalogue of candidate genes (Weber et al. 2005). Several plastid- l4 Specific proteomic studies confirmed the presence and extended the list of proteins targeted to inner envelope membranes of plastids to approximately ~130 putative metabolite transporter proteins (Sun et al., 2008). This number does not include genes from the Toc, Tic and thylakoid-related protein import machinery and prominent thylakoid-localized electron-transporting proteins of the photosynthetic core complex, such as D1, D2, cytochrome b6, cytochrome f, or proton-coupled ATP—synthase subunits. Our results for the first time revealed that the host has been the major source for metabolite transporter acquisition by the plastid. We also found an tmexpectedly high proportion of Chlamydiae-derived genes encoding plastid metabolite transporters. The group of Debashish Bhattacharya examined this result in more detail and identified a total of 55 genes in Plantae with chlamydial origin (Moustafa et al., 2008). A similar result was demonstrated for the red alga Cyanidioschyzon meralae (Huang and Gogarten, 2007). This provides a minor but crucial third genome source for plastid evolution by horizontal gene transfer or even an “endosymbiotic” gene transfer from a parasitic Chlamydiae bacterium. Intriguingly, most of the metabolite transport proteins were distributed in both green and red algae and to a much lesser extent present in only one of them. This suggests that the initial phase of the proto-alga, before the split into the Archaeplastida lineages, was accompanied by a massive invasion of the envelope membrane with host-proteins. As mentioned earlier, the extant pPT family evolved from endomembrane NDP- sugar transporters. The three pPT subfamilies do not simply encode differentially expressed or targeted gene products with identical functions. They evolved distinct substrate specificities and define the major flux of carbon across higher plant plastid 15 envelope membranes (F lfigge, 1999). The phylogenetic tree shown in chapter one argues for highly conserved orthologous sugar-phosphate transporters with the preferred substrates Pi, TP, PEP, and Glc6P in the red alga G. sulphuraria. This is in contrast to the enormous differences in how starch and soluble sugars are synthesized in the green and the red lineages (Viola et al., 2001). Red algae harbor an UDP-glucose specific starch synthase of ancient eukaryotic origin, which is localized in the cytosol. The soluble starch synthase fiom green algae and plants is closely related to cyanobacterial genes and incorporates ADP-glucose into the polysaccharide chain in the stroma of the organelle (Deschamps et al., 2008). Probably, UDP- and ADP-glucose dependent starch synthesis was still manifested in the genome of the proto-alga. By the time the common ancestor split up into the Rhodophyceae and Chloroplastida, the carbon metabolism was far from being established as we find it in extant species. The pathway has been relocated to the plastid stroma in green plants and algae (Deschamps et al., 2008) or stayed and gradually developed in the cytosol of Rhodophyceae. In addition, red algae produce the predominant soluble carbohydrate floridoside, composed of a glycerol and galactose moieties (Viola et al., 2001) and plants synthesize the disaccharide sucrose from UDP- glucose and fi'uctose-6-phosphate (Huber and Huber, 1996). Based on phylogenomic analysis, I developed the hypothesis that transport of glycolytic intermediates, i.e. PEP, TP, and Glc6P, was a successfirl invention in plastid biogenesis and is conserved in primary photosynthetic eukaryotes. Hence, if the hypothesis is correct, the homologues in G. sulphuraria should represent truly orthologous genes with identical substrate specificities as their counterparts in the green lineage. However, the major differences in carbon metabolism between photosynthetic 16 eukaryotes challenge the hypothesis. Possibly, the ancient NDP—sugar transporters continued to evolve divergently in red algae from the TPT, PPT and GPT of plants. To find an answer to these contradictory hypotheses, I functionally tested the pPT genes from G. sulphuraria by recombinant expression and biochemical characterization, and comparison to their counterparts in higher plants. The results are presented in detail in chapter three. I discussed alternative mechanisms in green plants and red algae to organize their primary carbon partitioning between rhodoplasts and cytosol. Prior to this study, carbon assimilation and especially the metabolic fate of the photosynthates in Rhodophyceae did not receive great attention compared to green algae and land plants. This does not match with the major role of red algae in the global carbon cycle, since they dominate carbon fixation in aquatic environment, which approximately accounts for half of the global photoautrophic C02 assimilation (Field et al., 1998). In chapter three, 1 demonstrate that the red alga G. sulphuraria, in comparison to the Plantae, possesses an export system for triose-phosphates with very high specificity and afiinity. In red algae, TPT controls the flux of carbon during autotrophic and heterotrophic growth conditions and, in contrast to plants, glucose-6-phosphate transport activity does not exist in the red alga as a major import route for sugar-phosphates. The lower affinity TPT in green plants might represent an adaptation to maintain a relatively high concentration of TP that is required for driving plastidial starch biosynthesis. The TPT protein fiom G. sulphuraria is a good example for the biochemical diversity of known proteins in distantly related photoautotrophic organisms and implies that this gene pool is an underestimated toolbox to engineer plants for applied sciences. 17 Besides understanding of the origin and contribution of metabolite carriers for plastid evolution, a major challenge is linking the catalogue of uncharacterized transport proteins to their biological function. A defining feature of eukaryotes is their cellular compartrnentation. The cytosol is separated from the internal space of mitochondria, plastids, vacuole, peroxisomes, nucleus, and ER by biological membranes. Lipid bilayers create a diffusion barrier for hydrophilic and charged molecules and define separate compartments for enzymatic reactions. In a single cell, pathways can be confined in a specific or even partitioned in more than one compartment. In multi-cellular organisms, several pathways are connected between different cells, tissues or even organs and often separated by more than one membrane. Thus, metabolite transport proteins are required for nutrient and metabolite uptake, secretion of toxic by-products, to sustain and even regulate the metabolic flux through pathways and distribute a single metabolite to distinct compartments at the same time (Kunze et al., 2002). One quarter of the A. thaliana genome encodes for proteins with at least one transmembrane span that are predicted to be embedded in one of the cellular membranes (Schwacke et al., 2003; Komatsu et al., 2007). This nrunber includes numerous receptor kinases, electron transport chain or protein import components fi'om mitochondria and chloroplasts. Nevertheless, it is estimated that up to 40% of these membrane-localized proteins could represent metabolite transporters (Ward, 2001). Most of the predicted transporters lack homology to known proteins from other species and less than 20% of the putative carriers are characterized at the biochemical and molecular level (Barbier-Brygoo et al., 2001). During the day, the major flux across plastid membranes is the export of triose- phosphate, which is catalyzed by the TPT (Flitgge, 1999). TP is the product of the 18 reductive part of the Calvin-Benson-cycle (Calvin, 1962). Carboxylation of ribulose-l,5- bisphosphate (RuBP) by the RubisCO results in two molecules of 3-phosphoglycerate (3- PGA) and the following two enzymatic reactions generate triose-phosphate, consuming ATP and NADPH provided by the photosynthetic light reactions. Atmospheric oxygen (02) competes with C02 for the acceptor molecule RuBP, yielding significant amounts Of 2-phosphoglycolate (2-PG) and 3-phosphoglycerate (3- PGA) (Ogren, 1984). In contrast to 3-PGA, 2-PG cannot be metabolized further in the Calvin-Benson-cycle. It is hydrolyzed in the chloroplast to glycolate and exported to the cytosol to be recycled by the photorespiratory pathway (F igrue l) (Douce and Neuburger, 1999). Measurements on chloroplast membranes demonstrated the presence of a transport system, which is able to export glycolate with rates required to cope with the flux through the photorespiratory pathway (Howitz and McCarty, 1991). Two molecules of the 02 carbon glycolate are converted to one 03 molecule glycerate in a complex cycle organized in peroxisomes, mitochondria and cytosol. Glycerate is taken up by the chloroplast, phosphorylated and reduced to triose-phosphate (Howitz and McCarty, 1991; Schwarte and Bauwe, 2007). Numerous studies identified the enzymes, which hydrolyze 2-PG to glycolate (Somerville and Ogren, 1979), oxidize and transaminate glycolate to glycine in the peroxisomes (Kendall et al., 1983; Liepman and Olsen, 2001; Igarashi et al., 2003; Queval et al., 2007). Two molecules of glycine are converted to serine in the mitochondria (Somerville and Ogren, 1981; Vol] et al., 2006; Engel et al., 2007), and serine is deaminated to hydroxypyruvate and reduced to glycerate in the peroxisomes (Liepman and Olsen, 2001; Igarashi et al., 2003) (Murray et al., 1989). During the 1970s and 19803, forward genetics screens with mutants of A. thaliana and Hordeum vulgare 19 has been a powerful tool to identify many of the photorespiratory genes (Ogren, 1984). Isolated mutants grew normally under elevated carbon dioxide levels, but the enzymes were essential under photorespiratory conditions (Ogren, 1984). Actually, genetic dissection of photorespiration using Arabidopsis mutants paved the way for A. thaliana as a model organism in plant biochemical genetics and plant physiology (Somerville, 2001). Surprisingly, however, map-based cloning of the photorespiratory A. thaliana genes did not reveal a single transporter of the mandatory shuttle systems needed for the described conversion of glycolate to glycerate between chloroplast, peroxisomes, and mitochondria. The only isolated transport protein linked to the photorespiratory pathway and identified by forward genetic screen is the dicarboxylate transporter DCT 1/DiT2.l in the inner envelope membrane of chloroplasts (Figrn'e 1) (Somerville and Ogren, 1983; Renne et al., 2003). During glycine oxidation in the mitochondria, two molecules of glycine are converted to one molecule of serine, NADH, C02 and ammonia. The DCT protein is required in the so-called nitrogen cycle of the photorespiratory pathway to re-assimilate the released ammonia (Keys, 2006). The concerted activity of glutamine synthetase (GS) and glutamate synthase (GOGAT) incorporates ammonia into 2-oxoglutarate in the chloroplast yielding glutamate. Export of glutamate depends on the activity of the DCT/DiT2.l protein. An A. thaliana mutant defective in this carrier cannot export glutamate to the cytosol and the amino-donor for the peroxisomal aminotransferase reaction of glyoxylate to glycine is depleted fast (Renne et al., 2003). Similarly, impaired GS or GOGAT enzymes short-circuit the re-assimilation of 2-PG, demonstrating that the 20 nitrogen cycle is essential for the growth under photorespiratory conditions (W allsgrove et al., 1987; Coschigano et al., 1998). With a concentration of 380 ppm C02 and 210,000 ppm of 02 in the air, the carboxylation to oxygenation ratio of RubisCO is approximately 3:1 in C3 plants without a carbon concentrating mechanism (CCM) (Sharkey, 1988). Hence, a massive flux through the carbon and nitrogen cycle of photorespiration results (Figure l). The protein- mediated transport of glycerate/glycolate (Howitz and McCarty, 1991), glycolate/glycine/serine/glycerate/glutamate/Z-oxoglutarate (Reumann et al., 1998) and glycine/serine (Walker et al., 1982; Yu et al., 1983) is documented on isolated membranes from chloroplast, peroxisomes and mitochondria, respectively. However the genes encoding the corresponding transporters have not yet been identified at the molecular level. Several reasons could explain why the forward genetic screens were unsuccessful to identify additional metabolite transporters. So far, only plants that show a severe decrease in chlorophyll content within four days after shifting fiom high C02 to ambient air have been further analyzed (Somerville, 2001). Prolonging the period of photorespiratory conditions could serve to identify weaker alleles for the pale green phenotype. Overlapping substrate specificity of metabolite transporters could mask the phenotype of a single gene mutant. The mitochondrial glycine decarboxylase complex for glycine oxidation is absolutely essential for the fitness of the plant. The gdc knockout mutant is lethal at ambient and elevated carbon dioxide and thus demonstrate that not every gene in the photorespiratory pathway is dispensable under elevated C02 conditions (Engel et al., 2007). It is tempting to speculate that this could be true for other proteins, such as the porin-like channel of the peroxisomes, which facilitates the transport of a 21 broad range of mono- and dicarboxylate anions (Reumann et al., 1998). A loss of this activity would presumably interfere with the peroxisomal glyoxylate cycle to mobilize fatty acids (Graham, 2008) and the oxaloacetate/malate Shuttle to exchange reduction equivalents with the cytosol (Reumann and Weber, 2006). In chapter four and five I revisited the highly compartmentalized photorespiratory pathway and initiated a search for the long outstanding metabolite transporters. In chapter four I summarize putative alternative pathways leading to ammonia re-assimilation in the organelles mitochondria and chloroplast. I emphasize the required, but not yet identified transport systems. In chapter five I took advantage of publicly available bioinforrnatics databases to screen for genes with two or more predicted transmembrane spans, which are highly co-expressed with known photorespiratory genes (Schwacke et al., 2003; Steinhauser et al., 2004). I initiated a reverse genetics candidate approach for five of the best 32 co-expressed putative metabolite transporters and report on the discovery of a novel transporter involved in the photorespiratory pathway. The long-term goal will be to identify the transporters involved in the photorespiratory pathway in higher plants. 22 Figure 1: The photorespiratory pathway The enzymes involved in photorespiratory pathway are compartmentalized between chloroplasts, peroxisomes, and mitochondria (cMDH, chloroplastic malate dehydrogenase; GDC, glycine decarboxylase; GGT, glutamatezglyoxylate aminotransferase; GLYK, glycerate kinase; GOX, glycolate oxidase; GS/GOGAT, glutamine synthetase/ glutamate synthase; HPR, hydoxypyruvate reductase; mMDH, mitochondrial malate dehydrogenase; PGP, phosphoglycolate phosphatase; pMDH, peroxisomal malate dehydrogenase; RubisCO, ribulose-l,5-bisphosphate carboxylase/oxygenase; SGT, serine:glyoxylate aminotransferase; SHMT, serine hydoxymethyl transferase). Photorespiratory metabolites are abbreviated as follows: 3- PGA, 3-phosphoglycerate; Glu, glutamate; KG, or-ketoglutarate; Mal, malic acid; 0AA, oxaloacetic acid; RuBP, ribulose-bisphosphate; THF, tetrahydofolate. The transport proteins represented in this scheme have been characterized biochemically (white box) or the corresponding genes have been cloned (black box). The scheme was adapted from Reumann and Weber (2006). The figure in this dissertation is presented in color. 23 If Peroxisome Glycerate 2x Glycolate NAD a\“ 02 -MDHX J HPR GOX ‘ -CAT H20 NADH V H202 —-> + 2x Glyoxylate Glu Hydroxy- 0.5 02 pyruvate 2x Glycine Figure 1: The photorespiratory pathway 24 REFERENCES Adl S.M., Farmer M.A., Andersen R.A., Anderson O.R., Barta J.R., Bowser S.S., Brugerolle G., Feusome R.A., Fredericq S., James T.Y., Karpov 8., Kugrens P., Krug J., Lane C.E., Lewis L.A., Lodge J., Lynn D.H., Mann D.G., McCourt RM., Mendoza L., Moestrup 0., Mozley-Staudridge S.E., Nerad T.A., Shearer C.A., Smiruov A.V., Spiegel F.W., and Taylor MF. (2005) The New Higher Level Classification of Eukaryotes with Emphasis on the Taxonomy of Protists. Journal of Eukaryotic Microbiology 52:399-451. Arabidopsis Genome Initiative (2000) Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature 408:796-815. Barbier-Brygoo H., Gaymard F., Rolland N., and Joyard J. (2001) Strategies to identify transport systems in plants. Trends in Plant Science 6: 577-585. Benning C. (2008) A role for lipid trafficking in chloroplast biogenesis. Progress in Lipid Research 47:381-389. Bhattacharya D., Yoon 11.5., and Hackett J .D. (2004) Photosynthetic eukaryotes unite: endosymbiosis connects the dots. BioEssays 26:50-60. Blankenship RE. (1992) Origin and early evolution of photosynthesis. Photosynthesis Research 33:91-111. Blankenship RE, and Hartman H. (1998) The origin and evolution of oxygenic photosynthesis. Trends in Biochemical Sciences 23:94-97. Block M.A., Douce R., Joyard J., and Rolland N. (2007) Chloroplast envelope membranes: a dynamic interface between plastids and the cytosol. Photosynthesis Research 92:225-244. Calvin M. (1962) Path of Carbon in Photosynthesis. Science 135:879-889. Cavalier-Smith T. (2000) Membrane heredity and early chloroplast evolution. Trends in Plant Science. 5:174—182. Cavalier-Smith T. (2006) Cell evolution and Earth history: stasis and revolution. Philosophical Transactions of the Royal Society B: Biological Sciences 361:969-1006. Coschigano K.T., Melo-Oliveira R., Lim J., and Coruzzi G.M. ( 1998) Arabidopsis gls mutants and distinct Fd-GOGAT genes: Implications for photorespiration and primary nitrogen assimilation. Plant Cell 10:741-752. Cove D. (2005) The moss Physcomitrella patens. Annual Review of Genetics 39:339- 358. 25 Crawford N.M. (1995) Nitrate—nutrient and Signal for plant growth. Plant Cell 7:859- 868. Desclramps P., Colleoni C., Nakamura Y., Suzuki E., Putaux J.L., Buleon A., Haebel S., Ritte G., Steup M., Falcon L.l., Moreira D., Loffelhardt W., Raj J.N., Plancke C., d'Hulst C., Dauvillee D., and Ball S. (2008) Metabolic Symbiosis and the Birth of the Plant Kingdom. Mol Biol Ev0125:536-548. Deschamps P., Moreau H., Worden A.Z., Dauvillee D., and Ball SC. (2008) Early Gene Duplication Within Chloroplastida and Its Correspondence With Relocation of Starch Metabolism to Chloroplasts. Genetics 178:2373-2387. Douce R., and Joyard J. (1990) Biochemistry and Function of the Plastid Envelope. Annual Review of Cell Biology 6:173-216. Douce R., and Neuburger M. (1999) Biochemical dissection of photorespiration. Curr Opinion in Plant Biology 2:214-222. Dyall S.D., Brown M.T., and Johnson PJ. (2004) Ancient invasions: From endosymbionts to organelles. Science 304:253-257. Engel N., van den Daele K., Kolukisaoglu U., Morgenthal K., Weckwerth W., Parnik T., Keerberg 0., and Bauwe H. (2007) Deletion of glycine decarboxylase in Arabidopsis is lethal under nonphotorespiratory conditions. Plant Physiology 144: 1 328- 1335. Falkowski P.G., and Godfrey L.V. (2008) Electrons, life and the evolution of Earth's oxygen cycle. Philosophical Transactions of the Royal Society B: Biological Sciences 363:2705-2716. Fast N.M., Kissinger J.C., Roos D.S., and Keeling PJ. (2001) Nuclear-encoded, plastid-targeted genes suggest a single common origin for apicomplexan and dinoflagellate plastids. Molecular Biology and Evolution 18:418. Field C.B., Behrenfeld M.J., Randerson J.T., and Falkowski P. (1998) Primary production of the biosphere: Integrating terrestrial and oceanic components. Science 281:237-240. Field C.B., Campbell J.E., and Lobell D.B. (2008) Biomass energy: the scale of the potential resource. Trends in Ecology & Evolution 23:65-72. Fltlgge U.I. (1999) Phosphate translocator in plastids. Annual Review of Plant Physiology and Plant Molecular Biology 50:27-45. 26 Glynn J.M., Miyagisbima S., Yoder D.W., Osteryoung K.W., and Vitha S. (2007) Chloroplast division. Traffic 8:451-461. Goff S.A., Ricke D., Lan T.H., Presting G., Wang R., Dunn M., Glazebrook J., Session A., Oeller P., Varma H., Hadley D., Hutcbison D., Martin C., Katagiri F., Lange B.M. et a1. (2002) A drafi sequence of the rice genome (Oryza sativa L. ssp. japom'ca). Science 296:92-100. Gordon J., and Polle J. (2007) Ultrahigh bioproductivity from algae. Applied Microbiology and Biotechnology 76:969-975. Gould S.B., Waller RE, and McFadden GJ. (2008) Plastid Evolution. Annual Review of Plant Biology 59:491-517. Graham LA. (2008) Seed storage oil mobilization. Annual Review of Plant Biology 59:115-142. Gross W. (2000) Ecophysiology of algae living in highly acidic environments. Hydrobiologia 433:3 l -3 7. Handford M., Rodriguez-Furln C., and Orellana A. (2006) Nucleotide-sugar transporters: structure, function and roles in vivo. Brazilian Journal of Medical and Biological Research 39:1 149-1 158. Howitz K.T., and McCarty RE. (1991) Solubilization, Partial-Purification, and Reconstitution of the Glycolate Glycerate Transporter fi'om Chloroplast Inner Envelope Membranes. Plant Physiology 96: 1060-1069. Huang C.Y., Aylifl’e M.A., and Timmis J.N. (2003) Direct measurement of the transfer rate of chloroplast DNA into the nucleus. Nature 422:72-76. Huang J., and Gogarten J. (2007) Did an ancient chlamydial endosymbiosis facilitate the establishment of primary plastids? Genome Biology 8: R99. Huber S.C., and Huber J.L. (1996) Role and regulation of sucrose-phosphate synthase in higher plants. Annual Review of Plant Physiology and Plant Molecular Biology 47:431-444. Igarashi D, Miwa T, Seki M, Kobayashi M, Kato T, Tabata S, Shinozaki K, and Ohsumi C (2003) Identification of photorespiratory glutamate : glyoxylate aminotransferase (GGAT) gene in Arabidopsis. Plant Journal 33: 975-987 Inaba T., and Schnell DJ. (2008) Protein trafficking to plastids: one theme, many variations. Biochemical Journal 413: 15-28. 27 Kaneko T., Sato S., Kotani H., Tanaka A., Asamizu E., N akamura Y., Miyajima N., Hirosawa M., Sugiura M., Sasamoto S., Kimura T., Hosoucbi T., Matsuno A., et a1. (1996) Sequence Analysis of the Genome of the Unicellular Cyanobacterium Synechocystis sp. Strain PCC6803. 11. Sequence Determination of the Entire Genome and Assignment of Potential Protein-coding Regions. DNA Res 3: 109-136. Kendall A.C., Keys A.J., Turner J.C., Lea PJ., and Miflin B.J. (1983) The Isolation and Characterization of a Catalase-Deficient Mutant of Barley (Hordeum vulgare L). Planta 159:505-511. Keys A.J. (2006) The re-assimilation of ammonia produced by photorespiration and the nitrogen economy of C-3 higher plants. Photosynthesis Research 87: 165-175. Kilian 0., and Kroth P.G. (2003) Evolution of protein targeting into "complex'" plastids: The "secretory transport hypothesis'". Plant Biology 5:350-358. Komatsu S., Konisbi H., and Hashimoto M. (2007) The proteomics of plant cell membranes. J. Exp. Bot. 58:103-112. Kopriva S. (2006) Regulation of sulfate assimilation in Arabidopsis and beyond Annals of Botany 97:479-495. Kunze R., Frommer W.B., andl Fltlgge U.I. (2002) Metabolic Engineering of Plants: The Role of Membrane Transport. Metabolic Engineering 4: 57-66. Kutschera U., and Niklas KJ. (2005) Endosymbiosis, cell evolution, and speciation. Theory in Biosciences 124:1-24. Lanier W., Moustafa A., Bhattacharya D., and Comeron J .M. (2008) EST Analysis of Ostreococcus lucimarinus, the Most Compact Eukaryotic Genome, Shows an Excess of Introns in Highly Expressed Genes. Public Library of Science (PLOS) ONE 3:E217l. Liepman AH, and Olsen LJ. (2001) Peroxisomal alanine: glyoxylate aminotransferase (AGT 1) is a photorespiratory enzyme with multiple substrates in Arabidopsis thaliana. Plant Journal 25:487-498. Loeffelhardt W. (2007) The cyanelle of Cyanophora paradoxa: Missing link of plastid evolution. Journal of Phycology 43:28. Margulis L., and Nealson KH. (1989) Symbiosis as the Source of Evolutionary Innovation. Endocytobiosis and Cell Research 6:235-239. Martin W. (2003) Gene transfer from organelles to the nucleus: fiequent and in big chunks. Proceedings of the National Academy of Sciences of the USA 100:8612-8614. 28 Martin W., Rujan T., Richly E., Hansen A., and Cornelsen S. (2002) Evolutionary analysis of Arabidopsis, cyanobacterial, and chloroplast genomes reveals plastid phylogeny and thousands of cyanobacterial genes in the nucleus. Proceedings of the National Academy of Sciences of the USA 99:12246-12251. Matsuzaki M., Misumi 0., Shin I.T., Maruyama S., and Takahara M. (2004) Genome sequence of the ultrasmall unicellular red alga Cyanidioschyzon merolae 10D. Nature 428:653. Merchant S.S., Prochnik S.B., Vallon 0., Harris E.H., Karpowicz 8.1., Witrnan C.B., Terry A., Salamov A., Fritz-Laylin L.K., Marechal-Drouard L., Marshall W.F., Qu L.H., Nelson D.R., Sanderfoot A.A., Spalding M.H., Kapitonov V.V., Ren Q., Ferris P., Lindquist E., Shapiro H., Lucas S.M., et a1. (2007) The Chlamydomonas Genome Reveals the Evolution of Key Animal and Plant F rmctions. Science 318:245- 250. Miyagishima S.Y. (2005 ) Origin and evolution of the chloroplast division machinery. Journal of Plant Research 118:295-306. Morgan-Kiss R.M., Priscu J.C., Pocock T., Gudynaite-Savitch L., and Huner N.P.A. (2006) Adaptation and Acclirnation of Photosynthetic Microorganisms to Permanently Cold Environments. Microbiology and Molecular Biology Reviews 70:222-252. Moustafa A., Reyes-Prieto A., and Bhattacharya D. (2008) Chlamydiae Has Contributed at Least 55 Genes to Plantae with Predominantly Plastid Functions. Public Library of Science (PLOS) ONE 3:E2205. Murray A.J.S., Blackwell R.D., and Lea PJ. (1989) Metabolism of Hydroxypyruvate in a mutant of barley lacking NADH-dependent hydroxypyruvate reductase, an important photorespiratory enzyme activity. Plant Physiology 91:395-400. Nowack E.C.M., Melkonian M., and Gltlckner G. (2008) Chromatophore Genome Sequence of Paulinella Sheds Light on Acquisition of Photosynthesis by Eukaryotes. Current Biology 18:410-418. Ogren W.L. (1984) Photorespiration: Pathways, Regulation, and Modification. Annual Review of Plant Physiology 35:415-442. Pavy N., Johnson J.J., Crow J.A., Paule C., Kunau T., MacKay J., and Retzel E.F. (2007) ForestTreeDB: a database dedicated to the mining of tree transcriptomes. Nucleic Acids Research 35:D888-D894. Queval G., Issakidis-Bourguet E., Hoebericbts F.A., Vandorpe M., Gakiere B., Vanacker H., Miginiac-Maslow M., Van Breusegem F., and N octor G. (2007) Conditional oxidative stress responses in the Arabidopsis photorespiratory mutant cat2 demonstrate that redox state is a key modulator of daylength-dependent gene expression, 29 and define photoperiod as a crucial factor in the regulation of H202-induced cell death. Plant Journal 52:640-657. Raja R., Hemaiswarya S., Kumar N.A., Sridhar S., and Rengasamy R. (2008) A perspective on the biotechnological potential of microalgae. Critical Reviews in Microbiology 34:77-88. Renne P., Dressen U., Hebbeker U., Hille D., Flugge U.l., Westhofi' P., and Weber A.P.M. (2003) The Arabidopsis mutant dct is deficient in the plastidic glutamate/malate translocator DiT2. Plant Journal 35:3 16—33 1. Reumann S., Maier E., Heldt H.W., and Benz R. (1998) Permeability properties of the porin of spinach leaf peroxisomes. European Journal of Biochemistry 251:359-366. Reumann S., and Weber A.P.M. (2006) Plant peroxisomes respire in the light: Some gaps of the photorespiratory C-2 cycle have become filled - Others remain. Biochimica et Biophysica Acta-Molecular Cell Research 1763: 1496-1510. Reyes-Prieto A., and Bhattacharya D. (2007a) Phylogeny of Calvin cycle enzymes supports Plantae monophyly. Molecular Phylogenetics and Evolution 45:384-391. Reyes-Prieto A., and Bhattacharya D. (2007b) Phylogeny of Nuclear-Encoded Plastid- Targeted Proteins Supports an Early Divergence of Glaucophytes within Plantae. Molecular Biology and Evolution 24:2358-2361. Reyes-Prieto A., Hackett J.D., Soares M.B., Bonaldo ME, and Bhattacharya D. (2006) Cyanobacterial contribution to algal nuclear genomes a primarily limited to plastid functions. Current Biology 16:2320—2325. Reyes-Prieto A., Weber A.P.M., and Bhattacharya D. (2007) The Origin and Establishment of the Plastid in Algae and Plants. Annual Review of Genetics 41: 147-168. Rollwitz 1., Santaella M., Hille D., Flugge U.l., and Fischer K. (2006) Characterization of AtNST-KTl, a novel UDP-galactose transporter fi'om Arabidopsis thaliana. Federation of European Biochemical Societies (FEBS) Letters 580:4246-4251. Rubin E.M. (2008) Genomics of cellulosic biofuels. Nature 454:841-845. Sato N ., Ishikawa M., Fujiwara M., and Sonoike K. (2005) Mass identification of chloroplast proteins of endosymbiont origin by phylogenetic profiling based on organism-optimized homologous protein groups. Genome Informatics 16: 56-68. Sato S., N akamura Y., Kaneko T., Asamizu E., and Tabata S. (1999) Complete Structure of the Chloroplast Genome ofArabidopsis thaliana. DNA Res 6:283-290. 30 Schwacke R., Schneider A., van der Graafi' E., Fischer K., Catoni E., Desimone M., Frommer W.B., Flugge U.l., and Kunze R. (2003) ARAMEMNON, a novel database for Arabidopsis integral membrane proteins. Plant Physiology 131:16-26. Schwarte S., and Bauwe H. (2007) Identification of the photorespiratory 2- phosphoglycolate phosphatase, PGLPI , in Arabidopsis. Plant Physiology 144:1580-1586. Sharkey TD. (1988) Estimating the Rate of Photorespiration in Leaves. Physiologia Plantarum 73:147-152. Sol] J ., and Schleiff E. (2004) Protein import into chloroplasts. Nature Reviews Molecular Cell Biology 5:198. Somerville CR. (2001) An early Arabidopsis demonstration. Resolving a few issues concerning photorespiration. Plant Physiology 125:20-24. Somerville CR, and Ogren W.L. (1979) Phosphoglycolate Phosphatase-Deficient Mutant ofArabidopsis. Nature 280:833-836. Somerville CR, and Ogren W.L. (1981) Photorespiration deficient mutants of Arabidopsis thaliana lacking mitochondrial serine transhydroxymethylase activity. Plant Physiology 67:666-671. Somerville S.C., and Ogren W.L. (1983) An Arabidopsis thaliana mutant defective in chloroplast dicarboxylate transport. Proceedings of the National Academy of Sciences of the United States of America-Biological Sciences 80:1290-1294. Stegemann S., Hartmann S., Ruf S., and Bock R. (2003) High-frequency gene transfer from the chloroplast genome to the nucleus. Proceedings of the National Academy of Sciences of the USA 100:8828. Steinhauser D., Usadel B., Luedemann A., Thimm 0., Kopka J. (2004) CSB.DB: a comprehensive systems-biology database. Bioinformatics 20:3647-3651. Sun 0., Zybailov B., Majeran W., Friso G., Olinares P.D.B., and van Wijk KJ. (2008) PPDB, the Plant Proteomics Database at Cornell. Nucleic Acids Research Published online on October 2nd 2008. Timmis J.N., Ayliffe M.A., Huang C.Y., Martin W. (2004) Endosymbiotic gene transfer: Organelle genomes forge eukaryotic chromosomes. Nature Reviews Genetics 5: 123-U l l 6. Tuskan G.A., DiFazio S., Jansson S., Bohlmann J., Grigoriev 1., Hellsten U., Putnam N ., Ralph S., Rombauts S., Salamov A., Scbein J., Sterck L., Aerts A., Bhalerao R.R., Bhalerao R.P., Blaudez D., Boerj an W., Brun A., Brunner A., Busov V., 31 Campbell M., Carlson J., Chalot M, Chapman J., et al. (2006) The genome of black cottonwood, Populus trichocarpa (Torr. & Gray). Science 313: 1 596-1604. Villarejo A., Buren S., Larsson S., Dejardin A., Monne M., Rudhe C., Karlsson J ., Jansson S., Lerouge P., Rolland N., von Heijne C., Grebe M., Bako L., and Samuelsson G. (2005) Evidence for a protein transported through the secretory pathway en route to the higher plant chloroplast. Nature Cell Biology 7: 1224-1231. Viola R., N yvall P., and Pedersen M. (2001) The unique features of starch metabolism in red algae. Proceedings of the Royal Society of London Series B-Biological Sciences 268:1417-1422. Voll L.M., J amai A., Renne P., Vol] H., McClung CR, and Weber A.P.M. (2006) The photorespiratory Arabidopsis shmI mutant is deficient in SHMT 1. Plant Physiology 140:59-66. Walker G.H., Saroj ini G., and Oliver DJ. (1982) Identification of a glycine Transporter from pea leaf mitochondria. Biochemical and Biophysical Research Communications 107 :856-861. Wallsgrove R.M., Turner J .C., Hall N.P., Kendall A.C., and Bright S.W.J. (1987) Barley mutants lacking chloroplast glutamine synthetase - Biochemical and genetic analysis. Plant Physiology 83: 155-158. Ward J .M. (2001) Identification of novel families of membrane proteins fi'om the model plant Arabidopsis thaliana. Bioinformatics 17:560-563. Weber A., Oesterhelt C., Gross W., Briutigam A., lmboden L., Krassovskaya 1., Linka N ., Truchina J., Schneidereit J., Vol] H., Voll L., Zimmermann M., Jamai A., Riekhof W., Yu B., Garavito R., and Benning C. (2004) EST-analysis of the thermo- acidophilic red microalga Galdieria sulphuraria reveals potential for lipid A biosynthesis and unveils the pathway of carbon export from rhodoplasts. Plant Molecular Biology 55:17-32. Weber A.P., Linka M., and Bhattacharya D. (2006) Single, ancient origin of a plastid metabolite translocator family in Plantae from an endomembrane-derived ancestor. Eukaryotic Cell 5:609-612. Weber A.P.M., Schwacke R., and Flugge U.l. (2005) Solute transporters of the plastid envelope membrane. Annual Review of Plant Biology 56:133-164. Yoon H.S., Hackett J.D., Ciniglia C., Pinto G., and Bhattacharya D. (2004) A molecular timeline for the origin of photosynthetic eukaryotes. Molecular Biology and Evolution 21:809-818. 32 Yu C., Claybrook D.L., and Huang A.H.C. (1983) Transport of glycine, serine, and proline into Spinach leaf mitochondria. Archives of Biochemistry and Biophysics 227:180-187. 33 CHAPTER 1 PHYLOGENY OF THE PLASTID PHOSPHATE TRANSLOCATOR FAMILYl I This work has been published in Weber A.P.M., Linka M., and Bhattacharya D. (2006) Single, ancient origin of a plastid metabolite translocator family in Plantae from a endomembrane-derived ancestor. Eukaryotic Cell 5:609-612. 34 ABSTRACT Phylogenetic analyses Show the single origin of a plastid metabolite translocator family in the Plantae from a gene encoding an existing endomembrane-derived protein. Red algal secondary endosymbiosis has spread a translocator gene into the ancestor of the “chromalveolate” protists, where it has diversified into a novel clade of proteins. INTRODUCTION The photosynthetic organelle (plastid) in red, green, and glaucophyte algae (Plantae) likely had a single origin fiom a cyanobacterial endosymbiosis (Cavalier-Smith, 1992; Palmer, 2003). This ancient event (ca. 1 — 1.5 billion years ago) (Yoon et al., 2004; Douzery et al., 2004) was followed by the transfer of genetic material from the endosymbiont to the nuclear genome of the host, the evolution of a protein import apparatus for the plastid, and targeting sequences for nucleus-encoded plastid-targeted proteins, and the establishment of genome-plastome intracellular communication and regulation (Brown et al., 2001; Fey et al., 2005; Strand, 2004). Current hypotheses regarding plastid origin and evolution provide plausible explanations for the later stages of organelle establishment (e.g., gene transfer to the nucleus), but do not specifically address the initial formation of the endosymbiosis. We hypothesize that the insertion of a metabolite antiporter into the ancestral plastid membrane was essential for establishing the primary endosymbiosis, allowing the ancestor of the Plantae to profit immediately from cyanobacterial carbon fixation. This antiporter likely evolved from an existing host metabolite translocator with either 35 mitochondrial function (Bullerwell and Gray, 2004; Gray et al., 2004) or the endomembrane system. In contrast, autotrophic free-living cyanobacteria have no obvious need for such antiporters (i.e., bacterial translocators are not detectable with BLAST searches). Members of the nucleotide-sugar/triose phosphate translocator gene family (Knappe et al., 2003) are likely ancestors to plastid translocators because (i) genes encoding these proteins are found in all sequenced eukaryotic genomes, (ii) orthologous proteins are absent from prokaryotes, and (iii) some gene family members are targeted to plant plastids. MATERIAL AND METHODS To address plastid translocator origin, we gathered available sequences (genome and expressed sequence tag data) from the NCBI WWW, DOE Joint Genome Institute (hgp'J/wwwjgidoegovb, Michigan State University Galdieria . ,z r m ,, , Porphyra yezoensis expressed sequence tag index thgp://www.kamsa.or.jp/en/plant/mrphjga/ESTO, and the Cjzam‘dioschyzon merolae Genome Project (http://merolae.biol.s.u-tokyo.ac.jp/). Homologs were identified using BLAST searches W with an E-value cut-off of 5 0.0001. A translocator phylogeny was inferred using protein maximum-likelihood (ML) method (PHYML V2.4.3 WWW, neighbor joining (NJ) (PHYLIP V3.63, , m ), maximum parsimony (MP) (PAUP*V4.0blO, hgpzl/paupesitfsueduh, and Bayesian inference (MrBayes V3.0b4, 36 WEI—928121112) (for details of phylogenetic approach, see Hackett et al., 2004). A total of 250 amino acids from 65 plastid phosphate translocators from red and “chromalveolate” (i.e., chlorophyll c—containing) algae and land plants and their non- plastid homologs were included in these analyses. RESULTS AND DISCUSSION The resulting tree (Figure 1.1) shows that plastid translocators are monophyletic and had a single origin from an existing endomembrane translocator (bootstrap support: ML, 100%; NJ, 98%; MP, 93%), consistent with our model. Each major group of plant plastid translocators (i.e., TPT, triose-phosphate/phosphate translocator; GPT/XPT, glucose-6-phosphate and xylulose-S-phosphate/phosphate translocator, respectively; PPT, phosphoenolpyruvate/phosphate translocator, shown in green text in Figure 1.1) form robustly supported lineages (ML and MP bootstrap values > 90%) that are sister to homologs in the red algae (Figure 1.1, shown in red text). This result supports the monophyly of red and green algae and land plants (e.g., Rodriguez-Ezpeleta et al., 2005) and is consistent with the origin of the different plastid translocators in their common ancestor; i.e., all plastid translocators are'monophyletic and each is divided into sister groups comprised of the red and green clades. Presumably, these Plantae translocators diversified because of the selective advantage they offer; i.e., to harvest the different fixed carbon products of the cyanobacterial primary endosymbiont. This phylogeny provides strong evidence for a Single primary endosymbiosis in the studied Plantae, reflecting a critical and early step in plastid evolution. The monophyly of GPTS and 37 XPTS has previously been reported (Knappe et al., 2003) and likely reflects a more recent plant-specific gene duplication event. The chromalveolates are a taxonomically diverse assemblage of protists comprised of alveolates (dinoflagellate algae, ciliates, and parasitic apicomplexans) and chromists (cryptophytes, haptophytes, and stramenopiles) that are thought to have ancestrally contained a plastid of red algal origin (Cavalier-Smith, 1999). This group is yet to be substantiated in a global eukaryotic phylogeny. It is therefore of interest that the available (i.e., apicomplexan, haptophyte, and stramenopile) chromalveolate sequences are monophyletic in our tree (Figure 1.1) and solidly associated (ML, 80%, NJ, 95%) with one clade of red algal proteins (i.e., C. merolae CMK114C, G. sulphuraria HET‘39C12). The addition of cryptophyte and dinoflagellate plastid translocators is needed to verify this result. Interestingly, preliminary biochemical analysis of the G. sulphuraria HET39C12 translocator suggests that it is a TPT (Marc Linka unpublished data); therefore, the ancestral chromalveolate plastid translocator was likely of this type. The topology of the chromalveolate subtree supports a single origin of the plastid translocator gene in the common ancestor of these species (i.e., supporting their monophyly) from the nucleus of the red algal endosymbiont. It is notWorthy that PPT and GPT/XPT fi'om Plantae are absent from the chromalveolates. In organisms containing plastids of secondary endosymbiotic origin with three or four plastid envelope membranes, transfer of a plastid phosphate translocator-related gene to the host nucleus and retargeting to the inner plastid membrane alone are not sufficient for the export of photosynthates to the host. Additional translocators with identical transport properties would be required in the third, the remnant of the endosymbiont plasma membrane, and 38 the fourth membranes, the plastid endoplasmic reticulum, to connect the metabolism of the host and the endosymbiont. This essential metabolic connection has been accomplished in the ancestor of the chromalveolates in our tree with the red algal gene Imdergoing diversification through duplication and divergence giving rise to the complex branching pattern shown in Figure 1.1. The specific functions and the detailed membrane localizations of the different chromalveolate translocators remain to be determined. We analyzed chromalveolate translocators with significant N-terminal extensions to determine if they are putatively plastid (or apicoplast) targeted. Using PATS . no) and PlasmoAP ' r m), apicoplast targeting is supported for the apicomplexan Plasmodium spp. (e.g., P. falciparum NP_703 643; PATS probability = 0.954, PlasmoAP = 5/5 tests positive). Similarly, the N-terminal extension in the diatom (stramenopile) Thalassiosira pseudonana Sol 5 (gene located on Scaffold 15) sequence contains a typical bipartite plastid-targeting sequence (Lang et al., 1998) comprised of a signal sequence using SignalP WW probability = 0.997; predicted length is 18 amino acids) followed by a transit peptide (using TargetP, probability = 0.740). These results suggest that the chromalveolate clade in our tree includes plastid-targeted translocators, although it is possible that others in this group may have lost this function. For example, the translocator from Iheileria annulata apparently does not encode a N-terrninal extension, but is clearly related to the other apicomplexan sequences. Although we do not have glaucophyte translocators in the tree, the signature activity of plastidic phosphate translocators has been detected in isolated cyanelles of Cyanophora paradoxa, indicating the presence of plastidic phosphate 39 translocators (Schlichting and Bothe, 1993). We therefore hypothesize that the insertion of a phosphate translocator into the plasma membrane of the endosymbiont occurred before the split of the Plantae and was probably a critical step in rendering the association between the cyanobacterium and the mitochondriate eukaryote irreversible. The proposed sequence of events for plastid establishment (Figure 1.2 and Figure 1.3) does not involve major evolutionary innovations. The basic components of a plastid envelope protein import apparatus were present in the cyanobacterium (Reumann et al., 2005) and it is reasonable to assume that this machinery was capable of importing proteins from the host cell cytosol, albeit with low efficiency; i.e., the orrter leaflet of the plastid outer envelope membrane consists mainly of ER—derived phospholipids (Douce and J oyard, 1981; Cavalier—Smith, 2000), indicating a close interaction between plastid envelope membranes and the host endomembrane system. It is likely that the cyanobacterium was engulfed as a prey item (as often occurs today) through phagocytosis, an event that likely occurred countless times and in some of these cells, the prey was not digested in the food vacuole but rather maintained as an endosymbiont. A descendant of these cells gave rise to the Plantae. The origin (or replacement) of plastids through cell (or organelle) engulfinent has occurred several times in evolution and can be found in taxa such as the filose amoeba Paulinella chromatophora (Bhattacharya et al., 1995) and in several dinoflagellate lineages (Schnepf and Elbrachter, 1999; Saldarriaga et al., 2001; Hackett et al., 2004). The position of the Plantae in the tree of life (Baldauf, 2003; Rodriguez-Ezpeleta et al., 2005) suggests that its common ancestor was a highly developed flagellate (or had flagellated stages) that most likely had the capacity for phagocytosis and endomembrane formation (Cavalier-Smith, 2000). 40 Figure 1.1. Maximum-likelihood phylogeny of endomembrane and plastid translocators ML bootstrap values (200 replications) are shown above the branches on the left of the slash mark, whereas the values to the right are from a NJ analysis (100 replications). Only bootstrap values 2 60% are shown. The different plastid translocators are GPT (glucose-6-phosphate/phosphate translocator), PPT (phosphoenolpyruvate/phosphate translocator), TPT (triose-phosphate/phosphate translocator), and XPT (xylulose-S- phosphate translocator). Apicomplexans: Iheileria annulata, Plasmodium falciparum, Toxoplasma gondii; HaptophyteszEmiliana huxleyi; Stramenopiles: Thallassiosira pseudonana. The numbers in the circles indicate translocators to the right that resulted from primary ( l) and secondary (2) endosymbiosis. T. pseudonana: Ihallassiosira pseudonana; A. thaliana: Arabidopsis thaliana; C. merolae: Cymidioschyzon merolae 41 Theileria annulata Plasmodiumfalciparwn NP_703428 Taxoplasma gondii Plasmodium falciparum NP_703643 100 r—Emiliam'a huxleyi L_Thalassiosira pseudonana Sc153 T. pseudonana Sc5 T. pseudonana Sc9 Thalassiosira pseudonana Sc15 / Sc44 Galdieria sulphuraria HET39C12 Physcomitrella patens 4787 Arabidopsis thaliana AAC83815 Zea mays P491 33 Galdieria sulphuraria A414H8 7 Physcomitrella patens 824 100 100 Arabidopsis thaliana AAL153 10 GPT ‘ Arabidopsis thaliana AAC28500 Arabidopsis thaliana K10A8 (XPT) Arabidopsis thaliana AAF01540 Arabidopsis thaliana AAM91743 Chlamydomonas reinhardtii AA020101 Galdieria sulphuraria A416F5 Homo sapiens KIAA044 100 C. meralae CMS488C 72 Karena brevis C01596 Karena brevis C00075 A. thaliana AtGONST4 85 _lllQ C. merolae CMOSSC 64 Cyanidioschyzon merolae CMT467C 199 Phytophtora ramorum 8384 Galdieria sulphuraria A413D9 Phytophtora ramorum 7748 Homo sapiens BAB5530 ‘_1m_l—Neurospora crassa CAB9145 Saccharomyces cerevisiae Sly4l TPT 88 84 Figure 1.1. Maximum likelihood phylogeny of endomembrane and plastid translocators 42 Acquisition of a cyanobacterium by a mitochondriate protist. Endomembrane-derived translocator of fixed carbon originates (triose-phosphate, TPT). Origin of plastid targeting system and ATP/ADP transporter to supply plastid with energy for e.g., fatty acid biosynthesis. Organellar genome reduction to the nucleus. Origin of additional translocators; e.g., dicarboxylate, hexose through gene duplication and divergence by a mitochondriate protist. Glaucophytes Red algae Radiation of the Plantae. Green algae Land plants Figure 1.2. Model for plastid and translocator origin in the common ancestor of the Plantae during primary endosymbiosis CB: cyanobacterium; ATP: Adenine triphosphate; TPT: Tr' translocator. The figure in this dissertation is presented in color 43 Acquisition of a red alga by the mitochondriate ancestor of the chromalveolates Transfer Of the red algal plastidic translocator (and other) genes(s) to the chromalveolate nucleus. Duplication and divergence of the chromalveolates translocator gene family 4- Radiation of the chromalveolates Cryptophytes Haptophytes Alveolates Stramenopiles Chromalveolates Figure 1.3. Model for plastid and translocator origin in the common ancestor of the chromalveolates during secondary endosymbiosis RA is red alga. The chromalveolate tree reflects the present understanding of the phylogeny of this group (e.g., Bhattacharya et al., 2004; Harper et al., 2005). The figure in this dissertation is presented in color. 44 REFERENCES Baldauf 8.1.. (2003) The deep roots of eukaryotes. Science 300: 1703-1706. Bhattacharya D., Yoon H.S., and Hackett JD. (2004) Photosynthetic eukaryotes unite: endosymbiosis connects the dots. Bioessays 26:50-60. Bhattacharya D., Helmchen T., Bibeau C., and Melkonian M. (1995) Comparisons of nuclear-encoded small-subunit ribosomal RNAS reveal the evolutionary position of the Glaucocystophyta. Molecular Biology and Evolution 12:415-420. Brown E.C., Somanchi A., and Mayfield SP. (2001) Interorganellar crosstalk: new perspectives on signaling from the chloroplast to the nucleus. Genome Biology 2: 1021- 1024. Bullerwell C.E., and Gray MW. (2004) Evolution of the mitochondrial genome: protist connections to animals, fungi and plants. Current Opinion in Microbiology 7:528-534. Cavalier-Smith T. (2000) Membrane heredity and early chloroplast evolution. Trends Plant Science. 5: 174-1 82. Cavalier-Smith T. (1999) Principles of protein and lipid targeting in secondary symbiogenesis: euglenoid, dinoflagellate, and sporozoan plastid origins and the eukaryotic family tree. The Journal of Eukaryotic Microbiology 46:347-366. Cavalier-Smith T. (1992) The number of symbiotic origins of organelles. Biosystems 28:91-106. Douce R., and Joyard J. (1981) Does the plastid envelope derive from the endoplasmic reticulum? Trends in Biochemical Sciences 6:237-239. Douzery B.J., Snell E.A., Bapteste E., Delsuc F., and Philippe H. (2004) The timing of eukaryotic evolution: does a relaxed molecular clock reconcile proteins and fossils? Proceedings of the National Academy of Sciences of the USA 101: 15386-15391. Fey V., Wagner R., Brautigam K., and Pfannschmidt T. (2005) Photosynthetic redox control of nuclear gene expression. Journal of Experimental Botany 56:1491-1498. Gray M.W., Lang B.F., and Burger G. (2004) Mitochondria ofprotists. Annual Review of Genetics 38:477-524. Hackett J.D., Yoon H.S., Soares M.B., Bonaldo M.F., Casavant T.L., Scheetz T.E., Nosenko T., and Bhattacharya D. (2004) Migration of the plastid genome to the nucleus in a peridinin dinoflagellate. Current Biology 14:213-218. 45 Harper J.T., Waanders E., and Keeling PJ. (2005) On the monophyly of chromalveolates using a six protein phylogeny of eukaryotes. International Journal of Systematic and Evolutionary Microbiology 55:487-496. Knappe S., Fltigge U.l., and Fischer K. (2003) Analysis of the plastidic phosphate translocator gene family in Arabidopsis and identification of new phosphate translocator- homologous transporters, classified by their putative substrate-binding site. Plant Physiology 131:1178-1190. Lang M., Apt KE., and Kroth P.G. (1998) Protein transport into "complex" diatom plastids utilizes two different targeting signals. Journal of Biological Chemistry 273:30973-30978. Palmer J.D. (2003) The symbiotic birth and spread of plastids: How many times and whodunit? Journal Of Phycology 39:4—1 1. Reumann S., Inoue K., and Keegstra K. (2005) Evolution of the general protein import pathway of plastids. Molecular Membrane Biology 22:73-86. Rodriguez-Ezpeleta, N., H. Brinkmann, Burey S.C., Roure B., Burger G., Lotfelhardt W., Bohnert H.J., Philippe H., and Lang B.F. (2005) Monophyly of primary photosynthetic eukaryotes: green plants, red algae, and glaucophytes. Current. Biology 15:1325-1330. Saldarriaga J.F., Taylor F.J., Keeling P.J., and Cavalier-Smith T. (2001) Dinoflagellate nuclear SSU rRNA phylogeny suggests multiple plastid losses and replacements. Journal of Molecular Evolution 53:204-213. Schlichting R., and Bothe H. (1993) The cyanelles (organelles of a low evolutionary scale) possess a phosphate-translocator and a glucose-carrier in Cyanophora paradoxa. Botanica Acta 106:428-434. Schnepf E., and Elbrachter M. (1999) Dinophyte chloroplasts and phlyogeny - A review. Grana 38:81-97. Strand A. (2004) Plastid-to-nucleus Signaling. Current Opinion Plant Biology 7 :621-625. Yoon H.S., Hackett J.D., Ciniglia C., Pinto G., and Bhattacharya D. (2004) A molecular timeline for the origin of photosynthetic eukaryotes. Molecular Biology and Evolution 21:809-818. 46 CHAPTER 2 PHYLOGENY AND ORIGIN OF PLASTID METABOLITE TRANSPORT PROTEINS FROM ARABIDOPSIS THALIANA 2 2 This work has been published in Tyra B.M., Linka M., Weber A.P.M., and Bhattacharya D., (2007) Host origin of plastid solute transporters in the first photosynthetic eukaryotes. Genome Biology 8:R212-224. Taira HM. and I contributed equally to this work. Taira H.M. wrote the initial draft of the manuscript. Dr. D. Bhattacharya was responsible for the final phylogenetic trees. 47 ABSTRACT Background: It is generally accepted that a single primary endosymbiosis in the Plantae (red, green (including land plants), and glaucophyte algae) a common ancestor gave rise to the ancestral photosynthetic organelle (plastid). Plastid establishment necessitated many steps, including the transfer and activation of endosymbiont genes that were relocated to the nuclear genome of the 'host' followed by import of the encoded proteins into the organelle. These innovations are, however, highly complex and could not have driven the initial formation of the endosymbiosis. We postulate that the retargeting of existing host solute transporters to the plastid fore-runner was critical for the early success of the primary endosymbiosis, allowing the host to harvest endosymbiont primary production. Results: We tested this model of transporter evolution by conducting a comprehensive analysis of the plastid permeome in Arabidopsis thaliana. Of 137 well-annotated transporter proteins that were initially considered, 83 that are broadly distributed in Plantae were submitted to phylogenetic analysis. Consistent with our hypothesis, we find that 58% of A. thaliana transporters, including all carbohydrate transporters, are of host origin, whereas only 12% arose fiom the cyanobacterial endosymbiont. Four transporter genes are derived from a Chlamydia-like source, suggesting that establishment of the primary plastid likely involved contributions fi'om at least two prokaryotic sources. Conclusion: Our results indicate that the existing plastid solute transport system shared by Plantae is derived primarily from host genes. Important contributions also came from the cyanobacterial endosymbiont and Chlamydia-like bacteria likely co-resident in the first algae. 48 INTRODUCTION Plastids in eukaryotes that contain chlorophyll are capable of carrying out photosynthesis, a process that converts light energy, carbon dioxide, and water into organic compounds. The evolutionary history of this organelle unfolded over a billion years ago when a previously non-photosynthetic protist engulfed and maintained a fi'ee-living cyanobacterium in its cytoplasm (Y oon et aL, 2004). It is hard to over-state the importance of this ancient and extraordinarily rare primary endosymbiosis because plastids allowed the evolution of algae and the plants that form the base of the food chain for many ecosystems on Earth. Current data suggest that the primary endosymbiosis occurred once in the common ancestor of the red, green (including land plants), glauCOphyte algae, the Plantae (Rodriguez-Ezpelata et al., 2005; Weber et al., 2006; Reyes-Prieto and Bhattacharya, 2007), with the Original plastid and the nuclear-encoded machinery for running the organelle spreading in subsequent cell captures to other branches of the eukaryotic tree (Cavalier-Smith et al., 1994; Mo Fadden, 1999; Bhattacharya et al., 2004). The only other known case of a potential bonafide cyanobacterial primary endosymbiosis occurred relatively recently in the thecate amoeba Paulinella chromatophora (Marin et al., 2005; Yoon et al., 2006). The gradualist view of evolution through mutation-selection suggests that it would have taken millions of years for the captured prokaryote to become fully integrated into the 'host' eukaryote, ultimately becoming the site not only for carbon fixation but also for other complex functions, such as lipid, isoprenoid, and amino acid biosynthesis (Weber et al., 2005). These processes were associated with the migration of much of the cyanobacterial genome to the host nucleus'and development of the complex protein 49 import system that are key shared features among all canonical plastids (Timmis et al., 2004; Weber et al., 2006; Gutensohn et al., 2006). A remarkable exception to the view that endosymbiosis was a gradual process of integration is offered by the katablepharid protist 'Hatena', which undergoes large-scale morphological changes following the engulfment of a green alga (Okamato and Inouye, 2005). Regardless of whether the ancient primary endosymbiosis fostered an accelerated rate of morphological evolution in the Plantae ancestor or whether general cell morphology was unchanged as in the Paulinella example (Johnson et al., 1919), one thing is clear - in the absence of rapid benefits to the host it is unlikely that the endosymbiosis would long have been sustained. Given the need for short-term survival, a key feature of early success for the endosymbiosis must have been the integration of the metabolism of the two cells. A solution for this process would have been solute transporters that regulate the flux of metabolites (for example, ATP, phosphate, sugars and sugar phosphates, metal ions, and other important ions) across the organelle membranes. Controlled exchange in response to environmental factors such as changes in light intensity and trace metal availability (Raven et al., 1999; Renne et al., 2003; Reiser et al., 2004) is decisive because the unregulated flux of metabolites would have had detrimental effects and, thereby, lowered the evolutionary fitness of the endosymbiosis. A complex system of solute transporters is in place today in extant plastids that provides the link between this organelle and the surrounding cytosol (F erro et al., 2003; Weber, 2004; Weber and Fischer, 2007). Here we focus on the evolutionary history of these plastid metabolite transporters to infer early events in plastid evolution. 50 We make two assumptions in this study. First, a system of metabolite transporters was a critical and early development in plastid evolution to supply the endosymbiont with essential nutrients and to enable the host to reap immediate benefit from photosynthetic primary production. It is rmclear why the cyanobacterium that was destined to become the plastid escaped digestion in the host but this scenario has also played out in 'Hatena' and in Paulinella. Second, whereas the genome of the previously flee-living cyanobacterium encoded all the transport systems required for the uptake of essential inorganic nutrients, it most likely did not harbor genes encoding transporters for the export of organic solutes to the host - this would have served no obvious pre-existing purpose in the prokaryote. Precisely how the plastid solute transport system was established is unknown. One possible model involves a primarily cyanobacterial origin, in which the plastid continued to utilize its own original cyanobacterial solute transporters with their evolution over time into proteins that perform most or all currently known plastid permeome functions. An alternative model involves a host—driven solute transport system, likely derived from the vacuolar envelope that initially surrounded the endosymbiont after its engulfinent (Weber et al., 2006). And finally, both of the new partners could have contributed proteins equally to this machinery, resulting in a chimeric system composed of the most beneficial combination possible of prokaryotic and eukaryotic transporters. To determine which of these competing hypotheses best explains plastid transporter evolution, we undertook an initial bioinformatics analysis of 137 A. thaliana solute transporters and then a detailed phylogenetic analysis of a subset of 83 conserved proteins that included available data from other Plantae. The A. thaliana transporters are either predicted or have been shown to be chloroplast targeted and are 51 ideal for tracking plastid permeome evolution. Using these data we demonstrate that over one-half of Plantae plastid targeted transporters are putatively of host origin whereas less than a quarter arose from the cyanobacterial endosymbiont. This suggests that primarily host genes made the lasting contribution to the Plantae host-endosymbiont relationship with regard to the plastid solute transport system. We also find evidence for the origin of four transporter genes or gene families from a Chlamydia-like source. This latter result raises the possibility that establishment of the ancient primary plastid may have involved contributions from at least two prokaryotic sources, perhaps explaining its singular nature. This hypothesis received substantial support from the recent finding of at least 21 genes of Chlamydia-like origin in the nuclear genome of the extremophilic red alga Cyanidioschyzon merolae (Huang and Gogarten et al., 2007). MATERIAL AND METHODS Initial transporter analyses As a starting point for the compilation of a conservative set of predicted or confirmed plastid envelope membrane transporters, we used a previously published list of 137 plastid-targeted membrane proteins that was based on predicted plastid localization and classification by the transporter classification system (Weber et al., 2005). This list was manually curated to remove proteins from the list if published evidence indicated that they were localized to a cellular location other than chloroplasts, if they represented membrane-bound enzymes, or if they were annotated as components of the Tic/Toe protein import apparatus, the photosynthetic machinery of the thylakoid membrane, or 52 the Sec or Tat protein targeting pathways. This curated list of candidate genes was updated and amended with recently published chloroplast envelope membrane transporters, such as AtFOLTl , a plastid localized transporter belonging to the mitochondrial carrier family that does not contain a plastid targeting signal (Bedhomrne et al., 2005) and was thus not included in previous lists. The final list contained 83 A. thaliana predicted or confirmed chloroplast solute transporters. The sequence for each protein was obtained fi'om The Arabidopsis Information Resorn'ce website (11W). These protein sequences were used as queries in the BLAST program “blastp” and “tblastn” searches of the NCBI Database WWW), the plant and algal genomes available through the Joint Genome Institute Eukaryotic Genomics Website (hmflgmejgi; W), the Cyanidioschyzon merolae Genome Project website W the Galdieria sulphuraria Genome Project website (m://genomics.msu.edu/ga_ldieria_/),_and Dragonblast V2.1 (Ruemmele SE, unpublished data), a web based database in Bhattacharya’s lab that contains EST datasets for several chromalveolates, Plantae, excavates, Rhizaria, and Amoebozoa. We used the predicted protein sequences for the following species for our analysis whenever available: Arabidopsis thaliana, Oryza sativa, Physcomitrella patens, Chlamydomonas reinhardti i, Ostreococcus tauri, Ostreococcus Iucimarinus, Cyanidioschyzon merolae, Galdieria sulphuraria, cyanophora paradoxa, Dictyostelium discoideum, Strongylocentrotus purpuratus, Xenopus Iaevis, Danio rerio, Mus musculus, Canisfamiliaris, and Homo sapiens. In addition, we included at least one insect, three fungal species, and a broad range of Bacteria and Archaea in our analysis. The BLAST searches used an E-value cut- 53 off < 0.00001. If a translated EST sequence was not available, the nucleotide sequence was translated over Six frames using the EXPASY translate tool (hm://ca.expa§y.grgtools/dna.htrnl). The resulting protein sequences were used in a BLAST search against the NCBI protein database to ensure the correct translation was obtained. We used the ClustalW feature included with BioEdit V7.0.5.3 to generate protein alignments (Hall, 1999). Alignments were visually inspected and manually corrected if necessary. Trees were generated under maximum likelihood using PHYML V2.4.4 utilizing the WAG model of amino acid substitution and estimating both the proportion of invariable sites and the alpha parameter (that is, WAG +1 + I') (Guindon and Gascuel, 2003). We performed non-parametric bootstrap analysis with 100 replicates for each PHYML analysis. The resulting trees were analyzed to determine the origin of the _ transporter in A. thaliana and other Plantae. The designation of gene origin was done as follows. When the Plantae solute transporter formed a well-supported (usually > 70% bootstrap support) monophyletic group with homologs in opisthokonts (that is, animals and fungi) and secondarily with other eukaryotes such as excavates and chromalveolates (if present), then it was classified as having a 'Host' origin. Under this scheme, no bacterial sequences interrupted the eukaryotic domain. 'Cyanobacterial' or 'Chlamydia-like' origin was inferred if the Plantae sequence formed a monophyletic group with protein sequences finm either of these lineages with strong bootstrap support. Other bacterial or eukaryotic sequences could (not necessarily) be in these trees but there had to be a robust separation of the Plantae and cyanobacteria or Chlamydia-like clade from all other homologs. We had two other 54 categories of gene origin that likely reflected a lack of phylogenetic resolution or pervasive horizontal gene transfer (HGT) among taxa that defied a clear inference of origin. The first was the 'Other' category in which the Plantae transporter formed a well- supported monophyletic clade but its position relative to available bacterial data was Imresolved, thereby not allowing us to identify the donor taxon. The second, 'PIantae— specific', was for transporters that had no significant bits to sequences in GenBank or other databases and appeared to be limited to the Plantae. The protein alignments are available in the download section of the Bhattacharya Lab website Detailed phylogenetic analyses For eight representative transporters fiom the five categories described above we inferred a maximum likelihood phylogeny using RAxML (RAxML-VI-HPC, v2.2.1) (Stamatakis et al., 2005) and the WAG + F evolutionary model. The specific transporters were: 'Host' - At4g39460 (SAMT, S-adenosylmethionine transporter, 236 amino acids), At5g663 80 (AtFOLTl, Arabidopsis thaliana folate transporter l, 268 amino acids); 'Cyanobacterial‘ - Atl g19800 (T GDl, trigalactosyldiacylglycerol l, permease-like protein for lipid transport, 234 amino acids), At5g64940 (ABC transporter protein, 487 amino acids); 'Chlamydia-like' - Atlg15500 (plastidic ATP/ADP transporter 2 (AtNT'T 2), 436 amino acids), At4g37270(HMA1, copper exporting ATPase, 444 amino acids); 'Other' - Atl g32080 (putative membrane protein, 211 amino acids); 'Plantae-specific' - At5g24690 (hypothetical expressed protein, 274 amino acids). These detailed analyses used a random starting tree (one round of taxon addition) and the rapid hill-clirnbing 55 algorithm (that is, option -f d in RAxML). To generate bootstrap values for these phylogenies, we used RAxML with the same settings and 100 replications. In addition, we used Bayesian inference (MrBayes V3.0b4) (Huelsenbeck and Ronquist, 2001) with each of the eight data sets using the WAG +1 + I‘ model to calculate posterior probabilities for nodes in the RAxML trees. Metropolis-coupled Markov chain Monte Carlo from a random starting tree was used in this analysis with two independent runs and 1 cold and 3 heated chains. The Bayesian analyses were run for two million generations each with trees sampled every 100th generation. To increase the probability of chain convergence, we sampled trees after the standard deviation values of the two runs were < 0.01 to calculate the posterior probabilities. We also ran the Bayesian analysis for the remaining two putative 'Chlamydia-like' genes in Plantae (dicarboxylate translocators DiTl, DiT2.1, and DiT2.2, and the phosphate transporter PHT2;1) to assess the topologies. We incorporated a representative diversity of available sequences in all of these trees. RESULTS AND DISCUSSION Distribution of transporters within Plantae Phylogenetic analysis of the best-annotated transporter data that are currently available from Arabidopsis was used to identify and putatively annotate homologs fiom other Plantae. Of 137 transporter proteins that were initially considered, BLAST and phylogenetic analyses and manual curation of recently available data led to the identification of 83 proteins that were of sufficient conservation and broad distribution 56 among Plantae to be used for firrther analyses. Each of these 83 proteins that included gene families (that is, representing 63 distinct, ancestral genes; Table 2.1) was used as input in BLAST and PHYML bootstrap analyses to infer the trees. This approach identified 41 proteins that are present in both red and green algae (including land plants) and, therefore, were likely found in the Plantae ancestor (glaucophyte homologs were found for some of these genes; for example, ADP/ATP translocase, hypothetical protein At3g45890). Eleven proteins were restricted to green algae and land plants, seven were plant-specific, and two were limited to red algae and land plants. The distribution of these proteins with respect to their putative origin in Plantae is shown in Figure 2.1(A). Given the lack of evidence for widespread horizontal gene transfer in extant Plantae, which most likely lost the capacity for phagotrophy early in its evolution (Archibald et al., 2003; Reyes-Prieto et al., 2007), we postulate that the patchy distribution for many plastid targeted transporters primarily reflects differential gene loss over the greater than one billion years that has passed since the primary endosymbiosis (Y oon et al., 2004). Under this interpretation, the large set of shared transporters among Plantae lineages provides resounding support for the monophyly of this supergroup (Hackett et al., 2007). 57 Table 2.1. List of Arabidopsis thaliana chloroplast solute transporters analyzed in this study and their putative evolutionary origins Arabidopsis thaliana solute transporters Putative evolutionary origin: Host AGI code Atl g05580 At1g54320 Atl g59870 Atl g61 800 Atl g64150 At1g66950 Atl g70610 Atl g79450 At2g04620 At2g13 100 At2g27810 At2g28070 At2g29650 At2g3 8060 At2g38330 At2g40420 At3g01550 At3g12740 At3g17690 At3g17700 At3 g45 890 AB g523 10 At4g00370 At4g l 3590 At4g 1 7 340 At4g25 750 At4g32400 At4g32650 At4g383 80 At4g39460 Gene Annotation Cation/hydrogen exchanger Ligand-effect modulator 3 (LEM3) family ABC transporter Glucose-6-phosphate/phosphate translocator 2 (GPT2) Expressed protein ABC transporter Transporter associated with antigen processing protein 1 (AtTAPl) Ligand-effect modulator 3 (LEM3) family Cation efilux family protein Glycerol-3-phosphate transporter Xanthine/uracil perrnease ABC transporter Na+-dependent inorganic phosphate cotransporter Na+-dependent inorganic phosphate cotransporter Multi antimicrobial extrusion (MATE Efflux) protein Amino acid transporter Phosphoenolpyruvate/phosphate translocator 2 (PPT2) Ligand-effect modulator 3 LEM3 family Cyclic nucleotide-binding transporter 2 Cyclic nucleotide-binding transporter l Expressed protein ABC transporter Anion transporter 2 (ANTR2) Expressed protein Major intrinsic family protein ABC transporter Adenine nucleotide uniporter Arabidopsis thaliana K+ rectifying channel 1 (ATKCl) Multi antimicrobial extrusion (MATE Efflux) protein S-adenosylmethionine carrier 2 (SAMT) 58 Table 2.1. (cont’d). Arabidopsis thaliana solute transporters Putative evolutionary origin: Host AGI code At5g04770 At5g05630 At5g13550 At5g14040 At5gl6150 At5gl7630 At5g19410 At5g19600 At5g22830 At5g26820 At5g33320 At5g42130 At5g45450 At5g46l 10 At5g52860 At5g54800 At5g59250 At5g66380 Gene Annotation Amino acid permease Amino acid permease Sulfate transporter Mitochondrial phosphate transporter Hexose transporter Xyluose-S-phosphate/phosphate transporter l (XPT) ABC transporter (White) Sulfate transporter CorA-like magnesium transporter F erroportin-related protein Phosphoenolpyruvate/phosphate translocator (PPTl) Mitochondrial substrate carrier family Iron transporter-related Triose phosphate/phosphate translocator (TPT) ABC transporter (White) Glucose-6-phosphate/phosphate transporter 1 (GPTl) Sugar transporter Folate transporter l (AtFOLTl) Putative evolutionary origin: Cyanobacteria AGI code Atl g045 70 Atl g08640 At1g19800 Atl g78620 At2g32040 At3g51 140 At3g60590 At4g33520 At5g12470 At5g64940 Gene Annotation Integral membrane family protein Expressed protein Trigalactosyldiacylglycerol 1, permease-like protein (TGDl) Integral membrane family protein Folate monoglutamate transporter, FT Expressed protein Expressed protein Metal-transporting P-type ATPase (PAAl) Expressed protein ABC] -family protein 59 Table 2.1. (cont’d). Arabidopsis thaliana solute transporters Putative evolutionary origin: 'ChIamydia-like' AGI code At1g15500 Atlg80300 At3g26570 At4g37270 At5g12860 At5g64280 At5g64290 Gene Annotation Adenine nucleotide translocase 2 (AtNTT 2) Adenine nucleotide translocase 1 (AtNTT 1) Low affinity phosphate transporter (PHT2;1) Heavy metal ATPase HMAl Dicarboxylate translocator 1 (DiTl) Dicarboxylate translocator 2.2 (DiT2.2) Dicarboxylate translocator 2.1 (DiT2. l) Putative evolutionary origin: ‘Other’ AGI code Atlg01790 Atlg32080 Atlg44920 At1g54350 At1g78560 At2g02590 At2g21340 At2g26900 At3g25410 At4g30580 At5g03555 At5gl3720 At5g52540 At5g62720 Gene Annotation Potassium transporter Membrane protein Expressed protein ABC transporter Bile acid:sodium symporter Expressed protein (Putative small multi-drug export) Enhanced disease susceptibility protein Bile acid:sodium symporter Bile acid:sodium symporter l-Acylglycerol-3 -phosphate O-acyltransferase (ATS2) Cytosine/purines, uracil, thiamine, allantoin family permease Expressed protein Expressed protein Integral membrane HPP family protein Putative evolutionary origin: 'Plantae-specific' AGI code At2g38550 At3g57280 At5gl7520 At5g24690 Gene Annotation Expressed protein Expressed protein Maltose transporter (MEXl) Expressed protein 60 30 25 I Host I Chlam dia—like 20 I Cyano acteria I Other + Plantae-specific Number of Genes 6i 0 Red and Green Green Red and Plant Plantae-specific B 12% 8% 5% I Host 17% El Chlamydia-like 1 5% I Cyanobacterial 7% I Plantae-specific D Other 50% 7% 20% Figure 2.]. Origin of plastid targeted solute transporters in Plantae (A) Gene distribution among Plantae and gene origin for 63 distinct transporters considered in this study. (B) Summary pie charts showing the origin of all the 83 transporters (top chart) and the 63 distinct genes (lower chart) considered in this study. The figure in this dissertation is presented in color. Most proteins of the plastid envelope permeome are host-derived Analysis of the phylogenetic data supports the notion that the host drove the integration of plastid and host metabolism. We find that the majority (58%, when considering all 83 genes; Figure 21(3) and Table 2. l) of the plastid solute transporters were most likely derived from existing host membrane proteins. These 48 proteins are diverse in nature, including several ATP binding cassette (ABC) transporters, nucleotide and amino acid permeases, sulfate, potassium, magnesium, and iron transporters, and cation efflux proteins (see Figure 2.2 (A) for S-adenosylmethionine transporter (SAMT) and Figure 2.2 (B) for A. thaliana folate transporter 1 (AtFOLTl) as representative trees). Of particular interest is the finding that in addition to the members of the nucleotide-sugar/triose phosphate translocator gene family previously reported to be of host origin (Weber et al., 2006), all other carbohydrate transporters included in our analysis were derived from existing host proteins. This result strongly suggests that the host utilized existing eukaryotic transport proteins pre-adapted to this function to 'tap' into the photosynthates produced by the captured cyanobacterium. In addition, the Plantae host also provided transporters to facilitate the movement of valuable nutrients such as magnesium, potassium, iron, and phosphate into the captured prokaryote. The replacement of pro-existing cyanobacterial anion and cation transporters with host derived proteins again suggests that there was strong selection to rapidly establish control over and utilize the endosymbiont. This process was most likely accomplished by using transporters derived fi'om the host vacuolar envelope (Weber et al., 2006). 62 Figure 2.2. Plastid targeted solute transporters of putative host’ origin in Plantae These are RAxML trees with the nrunbers above the branches inferred fi'om a RAxML bootstrap analysis and the thick branches showing significant (P > 0.95) support fi'om a Bayesian phylogenetic inference. Only bootstrap values 2 60% are shown. Branch lengths are proportional to the number of substitutions per site (see scale bars). The filled magenta circle shows the node that unites the Plantae taxa within the eukaryotic domain. The different algal groups are shown in different text colors: red for red algae, green for green algae and land plants, and brown for chromalveolates. The inclusion of chromalveolates within the Plantae is believed to reflect horizontal or endosymbiotic gene transfer events (for example, Li et al., 2006). The two transporters are: (A) SAMT, S-adenosylmethionine carrier transporter; and (B) AtFOLTl, A. thaliana folate transporter 1. The name of the A. thaliana solute transporter used for the query is indicated for both trees shown in this figure. The figure in this dissertation is presented in color. 63 Danio rerio Xenopus laevis Mus musculus Homo sapiens . Strongylocentrotus purpuratus Oplsthokonts Drosophila melanogaster C oprinopsis cinerea 87 Schizosaccharomyces pombe Edmowia lipolytica Candida albicans 98 Piraeodacty/um tricornutum T/ralassiosira pseudonana .-'l lexrmdrium tamarense 98 PhyIO/J/rt/zuru ramorum Plrr'mphllzora sofas ' 100 Cyanidimchvzon mero/ue Galdieria sulphuraria 88 Plasmodimn fir/Cipurum _'1 Theileria (mu/ma C h/amr'dmmmas rein/IuI'c/Iii ()strcuc'occus quri Ostreat/occur lucimurimrs P/II'xcomiIrc/lu pawns ()rr':a saliva Papa/us Irit'lzoc'w‘pa A-"icn/iunu hen/lumziuna Capsicum ummum Arabidopsis thaliana At4g39460 98 — 0.1 substitutions/Site B Milago maydis Cryptococcus neoformans _ 97 Aedes aegypti 100 Tribolium castaneum r—Strongylocentrotus purpuratus . Xenopus laevis OplSthOKOIltS 100 Gallus gallus Danio rerio T etraodon nigroviridis 100 Homo sapiens Mus musculus Cyanidirrs‘c/ryzr)n INC/‘0](lc’ Phaeou'actr'lum Iricormrmm P/zvtop/rt/rora rumarum (71/chI'dmmmas reinlzm'dlii Oslrem'nt't'us quri ()sIrcuc‘oc't'us lrlt'imurinus Papa/us Ir'ir'lmt'urpu Arabidopsis t/Iu/iuml r‘\15}.16038() ()rr':u saliva 100 100 —— 0.1 substitutions/site Figure 2.2. Plastid targeted solute transporters of putative 'host' origin in Plantae 64 The cyanobacterial contribution The cyanobacterial endosymbiont putatively contributed ten solute transporters to the plastid transport system (Table 2.1). These proteins include permease-like protein TGDl (trigalactosyldiacylglycerol 1, Figure 2.3), which is required for integrating the prokaryotic (that is, cyanobacterial) with the eukaryotic (that is, endoplasmic reticulum) pathway for lipid biosynthesis (Xu et al., 2003; Awai et al., 2006; Benning et al., 2006), the metal-transporting P-type ATPase PAAl (Shikanai et al., Abdel-Ghany et al., 2005), and a transporter required for folate/biopterin biosynthesis (Klaus et al., 2006). The remaining seven proteins of unknown function that are localized to the chloroplast inner membrane were included in the cyanobacterial group. Whereas the predicted secondary structure of most of these proteins indicates they represent transporters (that is, they contain at least four transmembrane domains that are connected by short loops), some, such as the ABCl-farnily protein At5g64940 (F igure 2.3) contain only one or two predicted transmembrane domains and may thus have functions other than metabolite transport. It is also intriguing that with the exception of the PAAl copper transporter the only cyanobacterial transport proteins apparently retained by Arabidopsis are those for which the host lacked a suitable replacement. For example, the initial steps of folic acid biosynthesis in plants are confined to the chloroplast; the final steps are localized in the cytosol and in mitochondria (Sahr etal., 2006; Basset et al., 2004a; Basset et al., 2004b). Plastids thus depend on an external folate supply and require an uptake system for this important metabolite. Interestingly, redundant systems for folate uptake exist in Arabidopsis chloroplasts, consisting of the cyanobacterial-derived folate transporter FT (Klaus et al., 2005) and the host-derived transporter AtFOLTl (Bedhomrne et al., 2005). 65 Figure 2.3. Plastid targeted solute transporters of putative 'cyanobacterial'l'plastid endosymbiont' origin in Plantae For details of tree building see Figure 2.2. The filled magenta circle shows the node that unites the Plantae taxa as sister to cyanobacteria. The different photosynthetic groups are shown in different text colors: blue for cyanobacteria, red for red algae, green for green algae and land plants, and brown for chromalveolates. The inclusion of chromalveolates or Euglenozoa (Eugl.) within the Plantae is believed to reflect horizontal or endosymbiotic gene transfer events (for example, Li et al., 2006). The two transporters are: (A) TGDl , trigalactosyldiacylglycerol l, lipid transporter; and (B) ABCl-farnily transporter protein The name of the A. thaliana solute transporter used for the query is indicated for both trees shown in this figure. The figure in this dissertation is presented in color. 66 T hermosinus carboxydivorans Norl * Solibacter usitatus Ellin6076 Mari rtzfimdus ferrooxydans PV-l ic ettsia bellii RML369-C 100 Rickettsia prowazekii str. Madrid E Non- Parvibaculum lavamentivorans DS-l Cyano- Acidi hilium crv tum JF-S bacteria Iasto ire/ ula marina DSM 3645 Syntrophus acidilrop icus SB Wolinella succinogenes DSM 1740 Myxococcus xanthus DK 1622 Thermos}'nec'llacocc'us e/(mgatus BP—l _{——— Sl'llé’t'lhu:l‘SliS sp. PC'C 6803 (2) Naxa‘wpza'zc'Ii/armu PC‘C 73 l 02 (2) — Crocmplzacra walsanii WH 8501 S)'nec-lmuu'cus elongatus PCC 6801 Prat/z/arar'ac'r'u.\' marinas 511'. 921 l Si'HUL‘IIUCUL'L'IIA' Sp. \‘Vll 57Ul Nos/ac- Hindi/army PCC 73 l ()2 ( l ) Nada/aria spamigcna CC Y9—1 I 4 .V'asmt' sp. P('(‘ 7120 Ana/menu variabi/ix' ATCC 294 l 3 Tric/zm{osmium uni/Human N S l 0| Lyng/n‘a sp. P('(' 8 I 06 G/(M'U/‘lclt'ft’l' via/accus PCC 743| Sl'llt’L'/2(I(.’()('(‘ll.\' Sp. JA-3-3 Ab Sl'm’c'llm'm't'ux S . JA-2-3B‘a(2- l 3) .Sl‘fl(’(‘/l()(:l'.\‘fi.9 Sp. PC‘C 68 )3 ( l ) 83 100 Ci'anidiasC/ntan mcrala Grad/aria Iwzuislipilala var. liui Porphyra purpurca Parp/rvra _1*e:()ensis O.s'lrca(-au'u.s lat'imarimzs ()xlrcacm'(‘Im‘ {am'i PratoI/wca u'ic'lx'cr/zamii Cit/am}'damanas I't'in/zardlii Ph}'sc'amilrclla palms ()rym saliva Pa m/us Iridmcarapa A cc/ic'aga a'mzc'aIu/a Arabidopsis thaliana AI l g [9800 100 Nuclear 79 — 0.1 substitutions/site Figure 2.3. Plastid targeted solute transporters of putative 'cyanobacterial'l'plastid endosymbiont' origin in Plantae Figure 2.3 continues on the following page. 67 Sl‘llcé“/2()(‘()('L‘ll.8' (”la/lgalus PCC630| lerznast'nec‘lzac'aa'us' elanglus BP-l Trir'lzmlcsmium (’l'l‘l/ll‘rlt’lllll [MS 1 0| Ana/mwza \‘al'ia/vilis ATCC 29413 Nr‘mlm' Sp. PCC 7 120 Noslac' pzmcli/armc PCC 7310?. Mala/aria .s‘pzmzigcna (CY 94 I 4 Cracmp/zaura l'l'c'lIS'Olll-l' WH 850i Sl'llct'lllu‘lNUS 5]). PC‘C 6303 Glue/ulc‘lw' Via/accux PCC 7421 Sl'llc’t‘llrlt'lu‘t‘l(.\' Sp. .1 A-JBB‘aQ-U) Lyngbm 3;). PCC 8106 95 fi 100 I Galdieria sulplna'al'ia (I) Galdieria .S'lllp/llll'ul‘ic’l (2) .8_5‘ 7 Cyanidioscllyzon memlae (I) Cyanidi().s'('/1_1.‘:an mam/ac) (2) Porphyrc'l I)’L’:()c'l7.8‘i5 Plzaeoa’acl)‘lum lricormllum T l?c7/d.§'.$'i05il‘('l pseua’mzana Euglena gracil is Chlam_\‘d(mmnas rcin/zardlii 100 Oslrc'ac'ac'cuslac'imarinus ()slreac'ac'c'as lam'i Pl?):s‘comilrc/la palms Papa/us Il‘it'llUCtll'pa Arabidopsis l/zaliana A15g649—10 Orin: saliva — 0.1 substitutions/site Figure 2.3. (cont’d). 68 'Chlamydia-like' transporters In addition to the host and cyanobacteria, a third significant contributor to the Plantae plastid solute transport system is the Chlamydiae. A surprisingly high number (four) of plastid envelope membrane transporters have been contributed by these prokaryotes. The presence of plant-like genes in Chlamydia has been noted in the past, sparking debate over whether their presence indicated a transfer fiom the ancestral plant to Chlamydia, an evolutionary relationship between cyanobacteria and Chlamydia, or a horizontal gene transfer (HGT) from a chlamydial parasite to the plant ancestor (Brinkman et al., 2002; Wolf et al., 1999; Schmitz-Esser et al., 2004). Phylogenetic analysis of plastid, Chlamydiae, and Rickettsiae ADP/ATP translocases (Schmitz-Esser et al., 2004) supports an ancient Chlamydia-to-Plantae direction of transfer. This explanation for the origin of the ADP/ATP translocase gene (and other ChlamydiaI-like genes) in Plantae was strongly supported by the phylogenomic analysis of Huang and Gogarten (2007). We found a monophyletic relationship between the AtNTTl and AtNTT2 (the A. thaliana plastid ADP/ATP translocases l and 2) and Chlamydiae ATP/ADP translocases (Figure 2.4) (Greub and Raoult, 2003; Linka et aL, 2003). In addition, the copper transporter heavy metal ATPase 1 (HMAl; Figure 2.4), the dicarboxylate translocators (DiT) DiTl, DiT2. l, and DiT2.2 (data not shown), and the low affinity phosphate transporter PHT2;] (data not shown) apparently has a chlamydial origin in Plantae (Huang and Gogarten, 2007). All of these trees provide bootstrap (except for the DiT tree) support for the monophyly of the 'ChIamydia-like' and plastid transporters. In the case of HMAl there are two ancient paralogs in plants, one of cyanobacterial likely endosymbiotic origin and one from a Chlamydia-like source that is shared with red and green algae. The DiTs 69 (Weber and Flugge, 2002) are present only in green algae, plants, and bacteria (that is, not in red algae). Whereas genomic data for glaucophytes are not yet available, transport experiments using isolated Cyanophora cyanelles showed that this glaucophyte uses a transport system for glutamine and 2-oxog1utarate that is distinct from green plant DiTs (Kloos et al., 1993). Taken together, these data indicate that 'Chlamydia-like' dicarboxylate translocators have likely been lost from red algae and glaucophytes. An alternative explanation is that the gene was acquired by the green lineage after the split of Chlorophyta and Rhodophyta. A DiT2 gene was also found in the dinoflagellates Amphidim'um carterae and Heterocapsa triquetra, which likely originated from an independent horizontal gene transfer (HGT). Several 'green' genes have been found in dinoflagellates and other chromalveolates that could have either originated from multiple independent HGTs or an ancient green algal endosymbiosis (for discussion, see Nosenko et al., 2006). In summary, it is surprising that bacteria not putatively involved in the endosymbiosis contributed 8% of the transporters that we have identified. When one considers the functions of these transporters, the chlamydial contribution becomes more important. HMAl increases copper and/or zinc transport into the plastid under conditions of high light, facilitating the production of copper/zinc superoxide dismutase (CuZnSOD), which protects the plant from superoxide radicals produced under high light conditions (Seigneurin-Berny et al., 2006, Lee et al., 2007). PHT2;], a phosphate transporter, controls phosphate allocation under conditions of phosphate-starvation (Versaw and Harrison, 2002). The DiT transporters are involved in assimilating nitrogen and recovering carbon lost to photorespiration, a process that is initiated by the 70 oxygenation reaction of RubisCO that primarily occurs under conditions when a high 022C02 ratio is present in the vicinity of RubisCO. Mutants lacking these transporters are tmable to survive in ambient C02 concentrations (Renne et al., 2003; Reumann and Weber, 2006; Schneidereit et al., 2006). Finally, the AtNTTl and AtNTT2 transporters are required for ATP import into the plastid during the dark (that is, in the absence of photosynthetic ATP production), particularly during lipid and chlorophyll biosynthesis. Although AtNTT2 mutants are still capable of producing lipids, indicating that the plastid has an alternative method for generating the ATP required for lipid biosynthesis, the production is significantly reduced and mutant plants have a sharply reduced growth rate Reiser et al., 2004). Arabidopsis mutants deficient in both AtNTTl and AtNTT 2 develop necrotic lesions when grown under short days, accumulate H202, and, strikingly, show constitutive expression of CuZnSOD2 and ascorbate peroxidase (Reinhold et al., 2007). The phenotype of the mutant was linked to reduced magnesium chelatase activity and it was concluded that ATP import into plastids in the dark is required for chlorophyll biosynthesis and for preventing photooxidative damage (Reinhold et al., 2007). The import of ATP into plastids in the dark is thus clearly a case in which the endosymbiont benefits from host metabolism. The ancient origin of these transporters in the tree of photosynthetic eukaryotes (Figure 2.4) is indicative of an essential role of this uptake system in the formation of the endosymbiosis. With the exception of the DiT translocators, each of these transporters appear to perform somewhat redundant functions (that is, copper and phosphate transport) but in a way that permits the plant to adapt to stresses involved in life on the land (that is, high light and 02 levels or low phosphate 71 availability). This may explain why the genes encoding these four plastid transporters have been retained in the Arabidopsis genome. How the 'ChIamydia-like' genes entered into the Plantae ancestor is unclear but it is possible that both the cyanobacterial endosymbiont and chlamydial parasites may have co-existed in the cell. Many environmental Chlamydia are known today that are broadly distributed in animals and protists (Molmeret et al., 2005). The co-existence of these two distinct prokaryotes may have provided the genetic 'toolkit' to make permanent the endosymbiosis with gene transfer from each cell providing essential functions for endosymbiont utilization. An alternative explanation is that the cyanobacterial endosymbiont was itself highly chimeric (that is, the 'fluid chromosome model') (Embley and Martin, 2006) and contained genes of chlamydial origin that had been gathered through HGT. Although possible, this scenario seems less plausible because it invokes, for example, the presence of an ATP/ADP translocator (a gene typical for 'energy parasites' such as Rickettsiae) in the genome of an oxygenic photosynthetic cell that is unlikely to encounter high concentrations of ATP in the surrounding environment; To this end, a 'Chlamydia-like' ATP/ADP translocator is absent from all studied cyanobacteria. 72 Figure 2.4. Plastid targeted solute transporters of putative 'Chlamydia-like’ origin in Plantae For details of tree building see Figure 2.2. The filled magenta circle shows the node that unites chlamydial taxa with plastid targeted Plantae transporters. The different photosynthetic groups are shown in different text colors: blue for cyanobacteria, red for red algae, green for green algae and land plants, magenta for glaucophytes, and brown for chromalveolates. The inclusion of chromalveolates within the Plantae is believed to reflect horizontal or endosymbiotic gene transfer events (for example, Li et al., 2006). The two transporters are: (A) ATP/ADP translocator (NTT); and (B) Heavy metal ATPase (HMAl) copper transporter. The name of the A. thaliana solute transporter used for the query is indicated for both trees shown in this figure. The figure in this dissertation is presented in color. 73 A 100 Rickettsia prowazekii Rickellsia bellii RML363-C Rickeltsia prowazekii C AA 1 4826 Rickellsia sibirica 246 Caedibacler caryophilus 100 Rickettsia belli RML369-C Ricketlsia montanensis 82 Lawsom'a inlracellularis PHE/MNl -00 Candidalus Paracaedibacter symbiosis Candidalus Protochlamydia amoephila 78 Candidalus Protochlamydia amoephila UWE25 93 Neochlamydia hartmannellae 100 Parachlamydia sp. Hall’s coccus 100 Chlamydia lrachomalis D/UW-3/CX Chlamydophila pneumoniae 1138 Glaucoc'vslis noslochincarum 96 100 Fragilariapsis (iv/indl'us‘ 97 T/ta/assiasil'a pseuc/(mana Plzamclac'lt‘lum ll'lt‘Ol'llllllHN Parp/n‘ra t'ciacnsis 8. Galdieria su/p/mraria Ct’anic/imc‘lzt'mn mam/ac Pralalheca wicker/lama 98 66 Dlllla/ic/la s‘a/illa C/I/cllllt‘a’amanas rcin/zardlil' Torin/a rural/'5‘ 93 Plans law/a 84 Onta saliva Salalmm lulwraxum " :‘ Cilrus Irt'hrid cullit‘ar .~ll'(ll)l(.l()p.\‘l.\' l/Ia/iana Al I g80300 lllawn/72')‘anllzcmum ('lft‘xlal/inmn 100 — 0.1 substitutions/site Figure 2.4. Plastid targeted solute transporters of putative 'Chlamydia-like' origin in Plantae Figure 2.4 continues on the following page. 74 lauri .. remhardm ( 2) Host ramorum. hansevm purpuratus JA-3-3 Ab walsonii 1 CC93Il Cyanobacterial reinhardlii amoebophila , sp. PCC 8l06 Chlamydia pneumoniae trachomatis . . . At4g37270 — 0.1 substrtutrons/srte merolae Figure 2.4. (cont’d). 75 'Other' and 'Plantae—specific' transporters We were unable to conclusively determine the origin of 18 transport proteins. Fourteen of these data sets resulted in PHYML trees in which the Plantae transporters were rooted within prokaryotes but without bootstrap support for a specific afi'rliation. An excellent example is provided by At] g32080 (Figure 2.5), which is a putative membrane protein conserved among Plantae, chromalveolates, and a diverse set of Eubacteria and Archaea (that is, the Thermococcus and Pyrococcus clade). Although the prokaryotic source of this gene in Plantae is unclear with the available data, the eukaryotic clade is clearly monophyletic, which is consistent with a single gene origin in the Plantae ancestor and, thereafter, transfer to chromalveolates (for example, diatoms in this tree) via secondary endosymbiotic gene transfer (Li et al., 2006). The unresolved provenance of At1g32080 and the 'Other' set of transporters in Plantae can be explained by pervasive HGT followed by full or partial gene replacement or differential gene loss among prokaryotes that has erased the ancient phylogenetic signal. Alternatively, these results may indicate erratic rates of sequence divergence that make it impossible to model protein evolution for these sequences. Given the growing evidence, however, for recurring HGT among bacteria (Dagan and Martin, 2007), it is likely that genes in the 'Other' category have reticulate evolutionary histories. In this regard it is noteworthy that the likely frequent HGTs seen in Figure 2.5 (A) among prokaryotes and other genes in this category contrasts starkly with the apparent single origin and vertical inheritance in Plantae. This result suggests a clear difference in rates of HGT for these genes with elevated rates in prokaryotes relative to eukaryotes. 76 Of the remaining transporters, four fell in the 'Plantae-specific' category because they lacked identifiable homologs outside of this supergroup and may simply be too divergent to determine their origin. This includes At5g24690 (Figure 2.5 (B), a hypothetical expressed protein) and the plastidic maltose exporter l (MEXl; data not shown). The latter is required for export of maltose resulting from starch breakdown from plastids at night in green plants. Storage of starch inside the chloroplast is exclusively found in the green linage. Therefore, MEXl has likely co-evolved with plastid-based starch biosynthesis and breakdown since it can be detected only in members of the Viridiplantae with one gene found in the dinoflagellate Karloa'inium micrum, which, as described above for DiT2, likely has resulted from a HGT. 77 Figure 2.5. Plastid targeted solute transporters of 'Other' or Plantae-specific' origin in Plantae For details of tree building see Figure 2.2. The filled magenta circle shows the node that unites the Plantae taxa. The different algal groups are shown in different text colors: red for red algae, green for green algae and land plants, magenta for glaucophytes, and brown for chromalveolates. The inclusion of chromalveolates within the Plantae is believed to reflect horizontal or endosymbiotic gene transfer events (Li et al., 2006). (A) A. thaliana solute transporter in the 'Other' category: putative membrane protein (Atl g32080) (B) A. thaliana solute transporter in the 'PIantae-specific' category: hypothetical expressed protein (At5g24690). The figure in this dissertation is presented in color. 78 Del lia acidovorans SPH-l Bur holderia multivorans ATCC176 l 6 Ralstom'a eutropha H16 77 Ralstonia eulropha JMP134 Ralslonia pickettli l2] Ralstonia solanacearum UW551 Polynucleobacler if. gLW-Pl DMWA-l Dechloromonas aromatica C Paracoccus denitrificans PD 1 222 Chloroflexus aglgregans DSM 9485 _ Roseiflexus sp. RS- 100 Geobacler bemidjiensis Bem Geobacler metallireducens GS-15 100 T hermococcus kodakarensis KODl '—EE Pyrococcus horikoshii 0T3 Pyrococcusfuriosus DSM 3638 Desulfitobaclerium hafm'ense Y5] 100 r— Pseudomonas pulida GB-l l-Pseudomonas enlomophila L48 Clostridium cellulol licum H10 Enlerococcus faeca is V583 __ Sfiphylococcus haemolyticus JC SC 1435 100 aemophilus in uenzae Rd KW20 -—-1_‘:- Vibriofischeri ES] 1 Photobaclerium s . SKA34 G aa’iel'ia su/ 7hza'aria Ct'anidiosr' titan mero/ac Pllc‘lé’0(l'a('l_l‘lllm T/m/assiasil‘a (ll/am}'danmnas rein/zardlii ()slrcac'm'("us lat'imarinus O.s'll'ca('()('c'zl.s' lam'i Pit)ascomilrc/la palms 98 Papa/us ll'lt'llUt'tl/‘(lpa Ol't'za .s'alit‘al I) ()ljt'3a sail/val?) LAl'ahidapxis llza/iana At 1 g32080 100 100 — 0.1 substitutions/site Figure 2.5. Plastid targeted solute transporters of 'Other' or 'PIantae-specific' origin in Plantae Figure 2.5 continues on the following page. 79 C t‘anap/mra paradm'a k Ct'alzidimc/irtan mam/ac Glac/iw'ia Slilp/llll't'll'la Cli/(lllll't/UNIUHHS i'cin/larc/lii 62‘ 100 O.\'ll't’(l('()('(’ll.\' luvinmrinus Oslrcat'mt'us lauri P/zit'sc'amilrt'l/a palms 99 013:4: salim Zea mars Medicaga lrum'ala/a Papa/us ll'lt'llfK'tll‘U/Hl Arabidopsis thaliana A15g24690 - 0.1 substitutions/site Figure 2.5. Plastid targeted solute transporters of 'Other' or 'Plamae-specific' origin in Plantae 80 Conclusion Here we determined the phylogeny of 83 Arabidopsis plastid solute transporters to determine whether they are of endosymbiotic origin from the captured cyanobacterium, of host origin, or of a ‘mixed' origin fi'om both of these sources. Our analysis has afforded a rare look at early, critical events in primary plastid evolution and support the notion that integration of plastid-host metabolism was primarily driven by host-derived transporters with important contributions coming from the cyanobacterial endosymbiont and Chlamydia-like bacteria. Another class of proteins of currently unknown origin included plant specific transporters such as MEX] . Despite the power of our comparative approach, our work has some important limitations. One is that because we used the Arabidopsis transporter set, we most certainly have missed a number of lertae transporters that are specific to red orgreen algae and have been lost from the Arabidopsis genome. In addition, we lack significant data fi‘om glaucophytes, but the upcoming Cyanophora paradoxa (glaucophyte) nuclear genome sequence WWW will allow us to incorporate this lineage into future inferences about transporter evolution. It is reasonable to assume, however, given the wealth of data supporting Plantae monophyly (McFadden, 1999; Rodriguez- Ezpeleta et al., 2005; Weber et al., 2006; Reyes-Prieto and Bhattacharya, 2007), that our inferences regarding the red and green lineages also apply to their glaucophyte sisters. Despite these limitations and the fact that phylogenetic signal is imperfectly maintained over a billion years of evolution, our comprehensive analysis of the chloroplast solute transport system will likely hold up and can be further tested as other genome sequences become available. 81 REFERENCES Abdel-Ghany S.E., Miiller-Moulé P., N iyogi K.K., Pilon M., and Shikanai T. (2005) Two P-type ATPases are required for copper delivery in Arabidopsis thaliana chloroplasts. Plant Cell 17:1233-1251. Archibald J.M., Rogers M.B., Toop M., Ishida K., and Keeling PJ. (2003) Lateral gene transfer and the evolution of plastid-targeted proteins in the secondary plastid- containing alga Bigelowiella natans. Proceedings of the National Academy of Sciences USA 10027678-7683. Awai K., Xu C., Lu B., and Benning C. (2006) Lipid trafiicking between the endoplasmic reticulum and the chloroplast Biochemical Society Transactions 34:395- 398. Basset GJ., Quinlivan E.P., Ravanel S., Rébeillé F., Nichols B.F., Shinozaki K., Seki M., Adams-Phillips L.C., Giovannoni J.J., and Gregory J.F. 3rd, Hanson A.D. (2004a) F olate synthesis in plants: the p—aminobenzoate branch is initiated by a bifunctional PabA-PabB protein that is targeted to plastids. Proceedings of the National Academy of Sciences USA 101:1496-1501. Basset GJ., Ravanel S, Quinlivan E.P., White R., Giovannoni J.J., Rébeillé F., Nichols R.P., Shinozaki K., Seki M., Gregory J.F. 3rd, and Hanson A.D. (2004b) F olate synthesis in plants: the last step of the p-aminobenzoate branch is catalyzed by a plastidial aminodeoxychorismate lyase. Plant Journal 40:453-461. Bedhomme M., Hoffmann M., McCarthy E.A., Gambonnet B., Moran R.G., Rébeillé F., and Ravanel S. (2005) F olate metabolism in plants: an Arabidopsis homolog of the mammalian mitochondrial folate transporter mediates folate import into chloroplasts. Journal of Biological Chemistry 280:34823-34831. Benning C., Xu C., and Awai K. (2006) Non-vesicular and vesicular lipid trafficking involving plastids. Current Opinion Plant Biology 9:241-247. Bhattacharya D., Yoon H.S., and Hackett J.D. (2004) Photosynthetic eukaryotes unite: endosymbiosis connects the dots. Bioessays 26:50-60. Brinkman F.S., Blanchard J.L., Cherkasov A., Av-Gay Y., Brunham R.C., Fernandez R.C., Finlay B.B., Otto S.P., Ouellette B.F., Keeling P.J., Rose A.M., Hancock R.E., Jones SJ., and Greberg H. (2002) Evidence that plant-like genes in Chlamydia species reflect an ancestral relationship between Chlamydiaceae, cyanobacteria, and the chloroplast. Genome Research 12:1159-1167. Cavalier-Smith T., Allsopp M.T., and Chao BE. (1994) Chimeric conundra: are nucleomorphs and chromists monophyletic or polyphyletic? Proceedings of the National Academy of Sciences USA 91:11368-11372. 82 Dagan T., and Martin W. (2007) Ancestral genome sizes specify the minimum rate of lateral gene transfer during prokaryote evolution. Proceedings of the National Academy of Sciences of the USA 104:870-875. Embley T.M., and Martin W. (2006) Eukaryotic evolution, changes and challenges. Nature 440:623-630. Ferro M., Salvi D., Brugiere S., Miras S., Kowalski S., Louwagie M., Garin J., Joyard J., and Rolland N. (2003) Proteomics of the chloroplast envelope membranes from Arabidopsis thaliana. Molecular and Cellular Proteomics 2:325-345. Greub G., and Raoult D. (2003) History of the ADP/ATP-translocase-encoding gene, a parasitism gene transferred from a Chlamydiales ancestor to plants 1 billion years ago. Applied and Environmental Microbiology 69:5530-5535. Guindon S., and Gascuel 0. (2003) A simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood. Systematic Biology 52:696-704. Gutensohn M., Fan E., Frielingsdorf S., Harmer P., Hou B., Hust B., and Klosgen RB. (2006) Toc, Tic, Tat et al.: structure and function of protein transport machineries in chloroplasts Journal of Plant Physiology 163:333-347. Hackett J.D., Yoon H.S., Li S., Reyes-Prieto A., Ritrnmele S.B., and Bhattacharya D. (2007) Phylogenomic analysis supports the monophyly of cryptophytes and haptophytes and the association of 'Rhizaria' with Chromalveolates. Molecular Biology and Evolution 24:1702-1713. Hall T.A. (1999) BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symposium Series 41:95-98. Huang J., and Gogarten J.P. (2007) Did an ancient chlamydial endosymbiosis facilitate the establishment of primary plastids? Genome Biology 8:R99. Huelsenbeck J .P., and Ronquist F. (2001) MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics 17:754-755. Johnson P.W., Hargraves RE, and Sieburth J.M. (1919) Ultrastructure and ecology of Calycomonas ovalis Wulfi‘, (Chrysophyceae) and its redescription as a testate rhizopod, Paulinella ovalis n. comb. (F ilosea: Euglyphina). Journal of Protozoology 35:61 8-626. Klaus S.M., Kunji E.R., Bozzo G.G., Noiriel A., de la Garza R.D., Basset G.J., Ravanel S., Rébeillé F., Gregory J .F. 3rd, and Hanson AD. (2005) Higher plant plastids and cyanobacteria have folate carriers related to those of trypanosomatids. Journal of Biological Chemistry 280:38457-3 8463. 83 Kloos K., Schlichting R., Zimmer W., and Bothe H. (1993) Glutamine and glutamate transport in Cyanophora paradoxa. Botanica Acta 106:435-440. Lee Y.P., Kim S.B., Bang J.W., Lee H.S., Kwak S.S., and Kwon S.Y. (2007) Enhanced tolerance to oxidative stress in transgenic tobacco plants expressing three antioxidant enzymes in chloroplasts. Plant Cell Reports 26:591-598. Li S., Nosenko T., Hackett J.D., and Bhattacharya D. (2006) Phylogenomic analysis identifies red algal genes of endosymbiotic origin in the chromalveolates. Molecular Biology and Evolution 23:663-674. Linka N., Hurka H., Lang B.F., Burger G., Winkler E.H., Stamme C., Urbany C., Sell 1., Kusch J., and Neuhaus H.E. (2003) Phylogenetic relationships of non- mitochondrial nucleotide transport proteins in bacteria and eukaryotes. Gene 306:27-35. Marin B., Nowack E.C., and Melkonian M. (2005) A plastid in the making: evidence for a second primary endosymbiosis. Protist 156:425-432. McFadden GJ. (1999) Endosymbiosis and evolution of the plant cell. Current Opinion Plant Biology 2:513-519. Molmeret M., Horn M., Wagner M., Santic M., and Abu Kwaik Y. (2005) Amoebae as training grounds for intracellular bacterial pathogens. Applied and Environmental Microbiology 71:20-28. Nosenko T., Lidie K.L., Van Dolah E.M., Lindquist E., Cheng J.F., and Bhattacharya D. (2006) Chimeric plastid proteome in the Florida "red tide" dinoflagellate Karenia brevis. Molecular Biology and Evolution 23:2026-203 8. Okamoto N., and Inouye I. (2005) A secondary symbiosis in progress? Science 310:287. Raven J.A., Evans M.C.W., and Korb RE. (1999) The role of trace metals in photosynthetic electron transport in Oz-evolving organisms. Photosynthesis Research 60:111-150. Reinhold T., Alawady A., Grimm B., Beran K.C., J ahns P., Conrath U., Bauer J., Reiser J., Melzer M., Jeblick W., and Neuhaus H.E. (2007) Limitation of nocturnal import of ATP into Arabidopsis chloroplasts leads to photooxidative damage. The Plant Journal 50:293-304. Reiser J., Linka N., Lemke L., Jeblick W., and N euhaus H.E. (2004) Molecular physiological analysis of the two plastidic ATP/ADP transporters fiom Arabidopsis. Plant Physiology 136:3524-3536. Renne P., Dressen U., Hebbeker U., Hille D., Flugge U.l., Westhofl' P., and Weber 84 A.P. (2003) The Arabidopsis mutant dct is deficient in the plastidic glutamate/malate translocator DiT2. Plant Journal 35:316-331. Reumann S., and Weber A.P.M. (2006) Plant peroxisomes respire in the light: Some gaps of the photorespiratory C-2 cycle have become filled - Others remain. Biochimica et Biophysica Acta 1763:1496-1510. Reyes-Prieto A., and Bhattacharya D. (2007) Phylogeny of Calvin cycle enzymes supports Plantae monophyly. Molecular Phylogenetics and Evolution 45:384-391. Rodriguez-Ezpeleta, N., H. Brinkmann, Burey S.C., Roure B., Burger G., Loffelhardt W., Bohnert H.J., Philippe H., and Lang B.F. (2005) Monophyly of primary photosynthetic eukaryotes: green plants, red algae, and glaucophytes. Current Biology 15:1325-1330. Sahr T., Ravanel S., Basset G., Nichols R.P., Hanson A.D., and Rébeillé F. (2006) Folate synthesis in plants: purification, kinetic properties, and inhibition of aminodeoxychorismate synthase. Biochem Journal 396: 157-162. Schmitz-Esser S., Linka N., Collingro A., Beier C.L., Neuhaus H.E., Wagner M., and Horn M. (2004) ATP/ADP translocases: a common feature of obligate intracellular amoeba] symbionts related to Chlamydiae and Rickettsiae. Journal of Bacteriology 186:683-691. Schneidereit J ., Hausler R.E., Fiene G., Kaiser WM., and Weber A.P. (2006) Antisense repression reveals a crucial role of the plastidic 2-oxoglutarate/malate translocator DiTl at the interface between carbon and nitrogen metabolism. Plant Journal 45:206-224. Seigneurin-Bemy D., Gravot A., Auroy P., Mazard C., Kraut A., Finazzi G., Grunwald D., Rappaport F., Vavasseur A., and Joyard J ., Richaud P., and Rolland N. (2006) HMAl, a new Cu-ATPase of the chloroplast envelope, is essential for growth under adverse light conditions. Journal of Biological Chemistry 281:2882-2892. Shikanai T., Mitller-Moulé P., Munekage Y., Niyogi K.K., and Pilon M. (2003) PAA l , a P-type ATPase of Arabidopsis, firnctions in copper transport in chloroplasts. Plant Cell 15:1333-1346. Stamatakis A., Ludwig T., and Meier H. (2005) RAxML-1H: a fast program for maximum likelihood-based inference of large phylogenetic trees. Bioinformatics 21:456- 463. Timmis J.N., Ayliffe M.A., Huang C.Y., and Martin W. (2004) Endosymbiotic gene transfer: organelle genomes forge eukaryotic chromosomes. Nature Reviews Genetics 52123-135. Versaw W.K., and Harrison M.J. (2002) A chloroplast phosphate transporter, PHT2;1, 85 influences allocation of phosphate within the plant and phosphate-starvation responses. Plant Cell 14:1751-1766. Weber A.P.M., and Fischer K. (2007) Making the connections - The crucial role of metabolite transporters at the interface between chloroplast and cytosol. Federation of European Biochemical Societies (FEBS) Letters 581:2215-2222. Weber A.P., Linka M., and Bhattacharya D. (2006) Single, ancient origin of a plastid metabolite translocator family in Plantae from an endomembrane-derived ancestor. Eukaryotic Cell 5:609-612. Weber A.P.M., Schwacke R., and Flugge U.I. (2005) Solute transporters of the plastid envelope membrane. Annual Review of Plant Biology 56:133-164. Weber A.P.M. (2004) Solute transporters as connecting elements between cytosol and plastid stroma. Current Opinion Plant Biology 7:247-253. Weber A., and Fliigge U.I. (2002) Interaction of cytosolic and plastidic nitrogen metabolism in plants. Journal of Experimental Botany 53:865-874. Wolf Y.I., Aravind L., and Koonin EN. (1999) Rickettsiae and Chlamydiae: evidence of horizontal gene transfer and gene exchange. Trends in Genetics 15:173-175. Xu C., Fan J., Riekhof W., Froehlich J.E., and Benning C. (2003) A permease-like protein involved in ER to thylakoid lipid transfer in Arabidopsis. European Molecular Biology Organization (EMBO) Journal 22:2370-2379. Yoon H.S., Reyes-Prieto A., Melkonian M., and Bhattacharya D. (2006) Minimal plastid genome evolution in the Paulinella endosymbiont. Curr Biol 16:R670-672. Yoon H.S., Hackett J.D., Ciniglia C., Pinto G., and Bhattacharya D. (2004) A molecular timeline for the origin of photosynthetic eukaryotes. Molecular Biology and Evolution 21:809-818. 86 CHAPTER 3 ANALYSIS OF THE PLASTID PHOSPHATE TRAN SLOCATOR FAIVIILY FROM THE RED ALGA GALDIERIA SULPHUR/1M3 3 This work has been published in Linka M., Jamai A., Weber A.P.M. (2008) Functional characterization of the plastidic phosphate translocator gene family from the thermo-acidophilic red alga Galdieria sulphuraria reveals specific adaptations of primary carbon partitioning in green plants and red algae. Plant Physiology 148: 1487-1496. I performed all of the experiments shown here. 87 ABSTRACT In chloroplasts of green plants and algae, C02 is assimilated into triose-phosphates; a large part of these triose-phosphates is exported to the cytosol by a triose- phosphate/phosphate translocator (TPT) whereas some is stored in the plastid as starch. Plastidial phosphate translocators (pPT) have evolved from transport proteins of the host endomembrane system shortly after the origin of chloroplasts by endosymbiosis. The red microalga Galdieria sulphuraria shares three conserved putative orthologous transport proteins with the distantly related seed plants and green algae. However, red algae, in contrast to green plants, store starch in their cytosol, not inside plastids. Hence, due to the lack of a plastidic starch pool, a larger share of recently assimilated C02 needs to be exported to the cytosol. We thus hypothesized that red algal transporters have distinct substrate specificity in comparison to their green orthologs. This hypothesis was tested by expression of the red algal genes in yeast and assessment of their substrate specificities and kinetic constants. Indeed, two of the three red algal pPT candidate orthologs have clearly distinct substrate specificities when compared to their green homologs. GsTPT displays very narrow substrate specificity and high affinity; in contrast to green plant TPTs, 3-phosphoglyceric acid (3-PGA) is poorly transported and thus not able to serve as a triose phosphate/3-PGA redox shuttle in viva. Apparently, the specific features of red algal primary carbon metabolism promoted the evolution of a highly efficient export system with high affinities for its substrates. The low-affinity TPT of plants maintains triose phosphate levels sufficient for starch biosynthesis inside of chloroplasts, whereas the red algal TPT is optimized for efficient export of triose phosphate from the chloroplast. 88 INTRODUCTION In plants, the photosynthetic light reactions provide the energy for major plastid localized pathways, such as C02 assimilation, the synthesis of starch, fatty acids, several amino acids, nucleic acids, and the reductive assimilation of inorganic ions like nitrate and sulfate (Weber et al., 2005; Zrenner et al., 2006). To supply the cell and the organism with these primary metabolites, a large number of precursors, end products, and intermediates have to be transported across the organelle envelope membrane and therefore present-day plastids are extensively connected to the cytoplasm by metabolite transporters that reside in the envelope membranes (Weber and Fischer, 2007; Tegeder and Weber, 2006). Chloroplasts originated approximately 1.6 billion years ago through a single primary endosymbiosis between a non-photosynthetic primitive mitochondriate eukaryote and a cyanobacterium (Y can et al., 2004; Bhattacharya et al., 2007; Reyes- Prieto et al., 2007). Within a period of 0.15 billion years, establishment of the plastid and divergence of the three major lineages of the Archaeplastida (Adl et al., 2005), that is the red algae (Rhodophyceae), green a1 gae/ land plants (Chloroplastida), and glaucophytes (Glaucophyta), began (Bhattacharya et al., 2004; Yoon et al., 2004). Establishment of the chloroplast within the host cell required massive remodeling of its membrane proteome; novel transport proteins to connect its metabolism with the metabolic network of the host cell had to be acquired (Bhattacharya et al., 2007 ; Weber and Fischer, 2007). Phylogenetic and phylogenomics analyses recently revealed that a large portion of these plastid-resident transporters is host-derived, indicating that integration of the chloroplast with host metabolism was predominantly a host-driven process (Tyra et al., 2007). Genes 89 encoding these transporters are not present in extant cyanobacterial genomes but most are conserved throughout the Archaeplastida, indicating they have been established at an early stage during formation of endosymbiosis, likely already at the stage of the protoalga (Tyra et al., 2007). Of particular importance for establishment of chloroplasts was an efficient and controlled export of photoassimilates from the endosymbiont to its host cell (Weber et al., 2006). In the Chloroplastida, a triose-phosphate/phosphate transporter (TPT) exports a significant amount of the dihydroxyacetone phosphate (DHAP) that is generated by the Calvin-Benson cycle to the cytosol; there it mainly serves as precursor for sucrose and cell wall biosynthesis (Flugge, 1999; Reiter, 2002). However, part of the triose phosphate is retained in the plastid stroma and used to fuel starch biosynthesis and other intraplastidial biosynthetic pathways (Zeeman et al., 2007). TPT candidate orthologs are highly conserved throughout the photosynthetic eukaryotes (Weber et al., 2006) and transport experiments with isolated organelles or reconstituted organellar membranes demonstrated triose-phosphate transport activity in the red algae and glaucophytes lineages, respectively (Schlichting and Bothe, 1993; Weber et al., 2004). Apparently, triose-phosphate (TP) already served as the main export photoassimilate at the stage of the proto-alga (Weber et al., 2006). In the plant model organism Arabidopsis thaliana, additional sugar phosphate transporters with specialized functions have been characterized (Flugge, 1999). Import of phosphoenolpyruvic acid (PEP) fi'om the cytosol by the PEP/phosphate transporter (PPT) drives fatty acid biosynthesis and synthesis of compounds by the shikimate pathway (i.e. aromatic amino acids) (Fischer et al., 1997; Knappe et al., 2003a; Voll ct aL, 2003) and a glucose-6-phosphate/phosphate transporter 90 (GPT) provides g1ucose-6-phosphate (Glc6P) for starch synthesis in heterotrophic plastids (Kammerer et aL, 1998; Niewiadomski et al., 2005). Lastly, a protein closely related to the GPT, the pentose-phosphate/phosphate transporter (XPT) connects the oxidative pentose phosphate pathways (OPPP) in cytosol and plastid (Eicks et al., 2002; Fliigge and Gao, 2005). Phylogenetic analysis showed that candidate orthologs for the PPT and the GPT/XPT type of translocators are present in the genome of the ancient red microalga G. sulphuraria, which is separated from green plants by an evolutionary distance of at least one billion years (Y oon et aL, 2004). However, reconstituted membrane fractions from G. sulphuraria showed neither significant PEP nor Glc6P transport activity (Weber et al., 2004). This raises the intriguing question as to whether the genetic repertoire present in the last common ancestor of Rhodophyceae and Chloroplastida was functionally maintained or whether it diverged after separation of the two phyla To address this question, we used G. sulphuraria as a model for the Rhodophyceae. The draft genome of G. sulphuraria is publicly available and the unicellular organism shares core features of carbon metabolism with other red algae (Viola et al., 2001; Barbier et al., 20053). In contrast to higher plants, starch in red algae is produced in the cytosolic compartment, using UDP-Glucose (UDPGlc) as precursor (Coppin et al., 2005; Patron and Keeling, 2005). Cell wall polysaccharides as well as the major soluble carbohydrate floridoside (a-D-galactopyranosyl-l-2’-glycerol) are produced in the same compartment. Notably, in addition to photoautotrophic growth, G. sulphuraria is also able to grow mixo- or heterotrophically on more than 50 different carbon sources (Gross and Schnarrenberger, 1995; Barbier et al., 2005b). Carbon partitioning thus has to be coordinated not only 91 during alternating light and dark periods but also under continuous heterotrophic growth conditions. It is not known in detail how the carbon allocation between the plastid and the cytosol is accomplished in red algae. Based on phylogenetic data, the plastid phosphate translocators presumably are the major routes for metabolite exchange. They have to balance a high demand for photoassimilates in the cytosol with maintaining sufficient levels of triose-phosphates for rhodoplast—localized pathways. Here we report the heterologous expression and biochemical characterization of the plastid phosphate transporter family fi‘om G. sulphuraria. With regard to higher plants we will discuss an alternative strategy to fine tune carbon flux across the plastid membrane in photosynthetic eukaryotes. MATERIAL AND METHODS Growth and sampling of G. sulphuraria cells Galdieria sulphuraria strain 074W (Gross et al., 1999) was cultured autotrophically in a minimal salt medium at 37°C with a light intensity of 80 umol photons per square meter and second. The culture was grown at pH 2.0 in Erlenmeyer flasks under vigorous shaking at ambient air conditions. Heterotrophic growth of the algal cells was performed at 37°C in culture flasks in complete darkness in the identical salt medium containing 25 mM glucose as sole carbon source. Cells were harvested at the late logarithmic phase by centrifugation (3,000 g, 5 min, 4°C), washed with leE buffer (10 mM Tris-HCl pH 7.5, 1 mM EDTA), flow in liquid nitrogen and stored at -80°C or directly used for nucleic acid extraction (Barbier et al., 2005) or total membrane enrichment. 92 Isolation of genomic DNA and total RNA from G. sulphuraria Isolation of genomic DNA was performed as previously described (Barbier et al., 2005). Briefly, cells were ground in liquid nitrogen and proteins were removed by ‘ Phenol/Chloroform/Isoamylalcohol (PCI) extraction. Nucleic acids were precipitated by ethanol and RNA was removed by a DNAse-free RNAse treatment and purified by an additional PCI and ethanol precipitation step. Total RNA was isolated with an acid guanidium isothiocyanate-phenol/chloroform solution as described by (Chomczynski and Sacchi, 1987). Enrichment of total membrane fractions from G. sulphuraria Harvested cells were resuspended in breaking buffer (100 mM HEPES/KOH, pH 7.6, 1 mM EDTA, 5% glycerol, 5 mM ascorbic acid, 5 mM DTT, 1 mM PMSF) to yield an OD600 of ~100 (0.5 mL) and an equal volume of acid-washed glass beads (0.4-0.6 mm size, Sigma Aldrich) were added. Cells were lysed with a mixer mill (MM301, Retsch GmbH, Germany) for 4 minutes. Broken cells were diluted with 10 mL breaking buffer and centrifuged (2,000 g, l min, 4°C). The supernatant contained the total membrane fractions. Membranes were pelleted from the supernatant by ultracentrifugation (100,000 g, 50 min, 4°C), resuspended in 0.4 mL 10 mM I-IEPES/KOH (pH 7.6), 1 mM MgC12, and 0.05 mL aliquots were reconstituted into liposomes. Expression of GsPTs in S. cerevisiae The coding sequences of all three GsPT were amplified from G. sulphuraria cDNA by PCR (Platinum Pfic polymerase, Invitrogen, Carlsbad, CA) using the primer combinations 93 listed in Table 3.1. PCR products were subcloned into the pGEM-T Easy vector system (Promega, Madison, WI) and sequenced. Forward and reverse primers for the genes GsHET39C12 and GsAl6F5 were designed with BamHI and XhoI restriction recognition sites, respectively. The GsA14H8 specific forward primer had a KpnI and the reverse primer an XhoI restriction site. Each cDNA was ligated in frame with an N-terminal polyhistidine-tag into the yeast expression vector pYESZ/NT (Invitrogen). Standard molecular methods were applied for DNA restriction and cloning (Sambrook et al., 1995). The resulting constructs were transformed into competent Saccharomyces cerevisiae INVScl cells (Invitrogen). Selection, maintenance of the transformants, and galactose-inducible expression of the recombinant proteins was done according to the manufacturer’s instructions (pYESZ/NT expression system, Invitrogen). Preparation of yeast membranes containing heterologously expressed GsI-IET3912, GsAl6F 5, and GsA14H8 proteins, respectively, was done as described previously (Bouvier et al., 2006), except that the expression was induced at an OD600 of 0.6 in the presence of 2% galactose and cells had been cultured for eight hours at 30°C. Presence of recombinant proteins was verified by standard SDS-PAGE and immunoblot analysis (Sambrook et aL, 1995) using an anti-penta-His antibody (Qiagen, Hilden, Germany) and a secondary anti- mouse IgG antibody conjugated with an alkaline phosphatase (Promega). Reconstitution into liposomes and transport activity assays Acetone-washed bet-phosphatidylcholine (PC) (Sigma-Aldrich, St. Louis, MO) had been sonicated (5 min on ice, Branson Sonicator 250, output 2, duty cycle 30%) to a final concentration of 2% (w/v) in 120 mM Tricine-KOH (pH 7.5) and 30 mM internal 94 substrate. 50 uL yeast membrane fraction was reconstituted with 950 uL liposome suspension by the freeze-thaw procedure (Kasahara and Hinkle, 1977). After thawing, the proteoliposomes were pulsed 20 times on ice to yield unilamellar vesicles. PDlO- desalting columns (GE Healthcare) were pre-equilibrated with 150 mM Tricine-KOH (pH 7.5) and used to remove non-incorporated substrate from the external medium. If not stated otherwise, all transport studies were initiated with radiolabeled [32P]- orthophosphate at a final concentration of 0.25 mM. For each data point, 190 uL proteoliposomes were terminated with 10 uL 10x inhibitor stop solution (200 mM PLP, 20 mM DIDS, 100 mM Tricine-KOH, pH 8.0). Control experiments with membranes from yeast cells harboring the empty vector were performed in parallel. Adding of inhibitor mix before addition of radiolabeled substrate was used to monitor unspecific binding. External phosphate was removed by strong anion-exchange chromatography with AG-l X8 resin (Acetate form, pre-equilibrated with 150 mM sodium acetate, BioRad, Hercules, CA). The flow-through of the anion-exchange columns containing the proteoliposomes with the imported radiolabeled phosphate was quantified by a liquid scintillation counting. All substrates used for uptake studies were purchased from Si grna- Aldrich. Kinetic constants were determined by measuring the initial velocity of each experiment. Michaelis-Menten constant (KM) has been analyzed with at least six external Pi concentrations ranging between 0.05 mM — 10 mM and competitive inhibition of Pi transport was assessed by the inhibitor constant Ki as described by Dixon (Dixon, 1953). GraphPad-Prism Software was used for non-linear regression fitting of all enzyme kinetic data. 95 RT-PCR gene expression study DNAse-treated RNA was used for first strand cDNA synthesis (Superscript II F irst- Strand Synthesis System, Invitrogen). Oligonucleotide primers for expression studies of the G. sulphuraria genes are summarized in Table 3.1. Other methods Protein concentrations were determined with a standard Bradford assay. Proteins in the membrane fractions were delipidated prior to determination (Shultz et al., 2005). 96 Table 3.1. Oligonucleotide primers designed in this study Amplified Galdieria sulphuraria gene products were used for cloning heterologous expression (HE) constructs or Reverse-Transcriptase PCR (RT-PCR) expression analysis. Underlined nucleotides highlight restrictions sites for cloning into expression vector system pYES. Fwd, forward primer", Rev, Reverse primer. Primer Sequence (5'—3') Direction Purpose Gs39C12 ML58 TAWATGAACCTTCTAAAGAATCCA fwd HE TCA ML59 ATCIEQAQATGAACCCAGAGTCT ACT'IT AT rev HE TT ML 1 37 AGCACTTGAACCCI‘TTGTCG fwd RT-PCR ML 1 3 8 TGCATTGGCAGGAGTGATAG rev RT-PCR Gsl6F5 ML56 TAQGATCCCAAGGAAACTCACTATTCGTC fwd HE ML66 GTCT CGAGTCTATTCT ATI'I‘TCTCC'ITCI' TC rev HE TTGG ML135 CAACCATGGCTACCTITCGT fwd RT-PCR ML136 CGTCAAAAAGCCAATCCAAT rev RT-PCR Gsl4H8 ML94 TGGTACCCAAAGGAGAAAAGGATATCATA fwd HE AGAGC ML95 ATCLQQAGACTACTGC’ITI‘TCTCGCTTCG rev HE ML133 AGAGTGGTGGCTCTCCTCAA fwd RT—PCR ML134 GATGAATCCGGTCCAGGTAA rev RT-PCR GsActin ML132 CACAGCATCTGGCCAAAC'IT fwd RT-PCR ML13 1 GACCCACATI‘TGTI‘GGAAGG rev RT-PCR 97 RESULTS Molecular features of the plastidic phosphate translocator (pPT) homologues from G. sulphuraria Three genes encoding proteins with significant similarity to higher plant pPT have previously been identified in the genome of the red alga G. sulphuraria (Weber et aL, 2006). Based on detailed phylogenetic analysis, GsHET39C12 is the candidate ortholog to the plastid triose-phosphate/phosphate translocator (TPT) and will thus be called GsTPT. GsAl4H8 and GsAl6F5 cluster with the functionally characterized green plant plastid glucose-6-phosphate/phosphate transport proteins (GPT), and phosphoenolpyruvate/phosphate translocator proteins (PPT), respectively. They were thus assigned with the acronyms GSGPT and GsPPT. The coding sequences of GsTPT and GsPPT contain 1224 nucleotides, corresponding to 407 amino acid residues and computed molecular masses of 45.83 kDa (T PT) and 44.98 kDa (PPT). GsGPT has a calculated molecular mass of 45.45 kDa, consisting of 410 amino acid residues that are encoded by 1233 nucleotides. The N-terminal regions of all three proteins represent putative plastid target sequences that display only low sequence similarity with the pPTs from higher plants (Figure 3.1) and the red alga Cyanidioschyzon merolae. The G. sulphuraria proteins show an average sequence identity of 37% and a 55% similarity with their A. thaliana homologs and a slightly higher identity of 48% to the corresponding C. merolae proteins (Table 3.2). Six lysine and two arginine residues that are invariantly embedded in conserved motifs in all fimctionally characterized pPT proteins to date (Knappe et al., 2003a), are also found in all three G. sulphuraria homologues. Relative to the GsTPT sequence, the positively charged amino acids are at 98 positions K123, K199, R261, K266, K267, K361, R362, and K399 (Figure 3.1). All three putative transport proteins are highly hydrophobic proteins that are predicted by ConPred Il (Arai et al., 2004) to contain nine to ten membrane spanning alpha-helices. The GsTPT and GSGPT genes contain one intron that separates a short first exon from the residual coding sequence (Table 3.2). GsPPT has a total of four exons. The last exon covers approximately 50% of the CDS. All three sequences have been submitted to GeneBank and can be also found at the MSU Galdieria genome database (http://genomicsmsuedu/ galdieria). Accession numbers and annotations are listed in Table 3.2. 99 Table 3.2. Molecular characteristics of the plastidic phosphate transporter GsTPT, GsPPT and GSGPT from Galdieria sulphuraria. GsTPT GsPPT GSGPT Accession number EU853171 EU853172 EU853173 . . . Stig_15 Stig_53 Stig_45 Galdrerra Genome Browser Burld 3.0 Gs21 6 60 Gs53050 Gs48050 Amino acid statisitc report [%] Identity to AtTPT 33 30 32 Similarity to AtTPT 53 51 53 Identity to AtPPT 31 41 31 Similarity to AtPPT 52 57 52 Identity to AtGPT 33 33 37 Similarity to AtGPT 56 51 55 # of Exons 2 4 2 Exon length (nt) 119 244 1 12 l 105 174 l 121 191 615 # of Introns 1 3 1 Exon length (nt) 53 55 50 47 70 # of transmembrane spans 9 10 9 100 Figure 3.1. Protein alignment of plastidic phosphate transporters (pPT) from Arabidopsis thaliana (At) and Galdieria sulphuraria (Gs) Multiple sequence alignment program ClustalW2 (Larkin et al., 2007) at EMBL-EBI lined up the sequence similarities and the tool GeneDoc WWW WWI) visualized and edited the higher plant pPT proteins and their red alga] homologues Gs39C12 (GsTPT), Gsl6F5 (GsPPT) and Gsl4H8 (GSGPT). AtTPT, triose-phosphate/phosphate translocator, AGI genome code At5g46110; AtGPT l, glucose-6-phosphate/phosphate translocator, At5g54800; AtPPT] , phosphoenolpyruvate/phosphate translocator, At5 g33320. 101 * 20 * 4O AtTPT -MESRVLLRATANVVGIPKLRRPIGAIHRQFSTASSSSFSVKPIG 44 Gs39C12 ------------------------------- MRKFDWITRQYLVA 14 AtPPTl MQSSAVFSLSPSLPLLKPRRLSLRHHPITTAASSSDLNVSPNVVS 45 GsAl6F5 MIAFATPLGNTSWKLGYRSLKSVIYFQSTTSSHPRKLTIRPSIQP 45 AtGPTl -MVLSVKQTLSP ------- KIGLFRRNP ----- SSSLGRSPVSLS 32 GsAl4H8 MMIEAAFVPLSCSLFSKQHRWTVVSRNKNKNSISSHMEGSKKLLG 45 * 6O * 80 * AtTPT GIGEGANLISGRQLR ----- PILLLDSSAINGGEKREILKPVKAA 84 Gs39C12 KARRLPQFYCLANTS ------ LDEPSK ------------------ 35 AtPPTl IPSLSRRSWRLASSD----SPLRAWSGVP--SPISHSLDTNRFRT 84 GsAl6F5 WLPFVQ---NQKTVD----HPSHLTSSFS--SPMMTLGDG----A 77 AtGPTl FPSTELPKRTVLAVS----KPLHLSSSLRAKSPVVRCEAYEADRS 73 GsAl4H8 TPRFTLSRSQFLNVSYLRTKYNNVASSSKGEKDIIRAACRRAE-S 89 100 * 120 * AtTPT AAEGGDTAGDAKVGFLAKYP TGFFFF F :. 129 Gs39C12 ESIKVTEASQPSQNTASWKR VASYFF I Ii- 80 AtPPTl AATAVPESAEEGDNSGKLTK ELGLLF L I EK' 129 GsAl6F5 ESSTGTSSSNVRQPVQSLQK LTFYIGC 1 LL 122 AtGPTl EPHPIGDDAAAAETKSEAAK IGIYFA :K' 118 GsAl4H8 GGSPQ~--KSSVGVSPTLVH VGFYFF F F E ‘ 131 140 * 160 * 180 AtTPT YNYFPYPYFVS HL YCLIS SV PKRAPIDSNLLK I 174 Gs39C12 LNAYPFPWTVA QLA FYVVP LLH KAPHIPLEDIK L 125 AtPPTl LKALHAPMTVT QFA SVLITI L YKRPKISGAQL L 174 GsAl6F5 LKVFPLFATVT QFL SLVGLA IS HRFQKASLEDLK 167 AtGPTl LNAYPYPWLTS SL SLMMLIS V ETPKTDFDFWK F 163 GsAl4H8 LNMWKYPWVLS QLG LYCTF L TKPNVSKKLI I 176 AtTPT 218 Gs39C12 169 AtPPTl 218 GsAl6F5 212 AtGPTl 207 GsAl4H8 220 Figure 3.1. Protein alignment of plastidic phosphate transporters from Arabidopsis thaliana (At) and Galdieria sulphuraria (Gs) The Figure 3.1 continues on the following page. 102 AtTPT GQSIPITLWLSA 1L Gs39C12 RSVFPZPVYLS} CG AtPPTl GEKPTPWVLG‘. f GsAl6F5 GTAYTLWVYLSn 1C AtGPTl GETFPTSVYLSC , GsAl4H8 GEFFEPLTYL vs 280 * AtTPT TD —————— MDSTN Gs39C12 Is DQTSYKHMSPAN AtPPTl KK--DDSLDNIT GsAl6F5 K KG----VQFDNLN AtGPTl K K—---GKSVSGMN GsAl4H8 F DFKNEKTLIAQNT 320 * 340 AtTPT ———PKLLNHGFADAIAKVGMTKFISDLFWVGMFYH Gs39C12 ———PKLYQGWILATSGKTTSMQLZTGLLTSGLFFY AtPPTl ————— IKFTPSYIQSAGVNVKQIYTKSLIAALCFH GsAl6F5 GRWREMASVATHIGSEGCT:PVLLLRIAIAGFLHF AtGPTl ———PQMWVDGWQTALATVG-PQFVWWVVAQSVFYH GsAl4H8 -——FPP;VS-———A:AGVSKAKLFGSIMFCSLFYH * 380 AtTPT Gs39C12 AtPPTl GsAl6F5 AtGPTl GsAl4H8 * 420 AtTPT v II KIEEEKRQGKKA 410 Gs39C12 v L YYYSQKIK ————— 362 AtPPTl v R IKPKPKTA---— 408 GsAl6F5 v Q ISTKKKEKIE-- 407 AtGPTl T Q L ———————————— 388 GsAl4H8 T L KLPSKREKQ——- 411 Figure 3.1. (cont’d). * 240 103 263 214 263 257 252 265 302 259 306 298 293 310 344 301 346 343 334 348 389 346 391 388 379 393 Heterologous expression of GsTPT, GsPPT, and GsGPT To assess the substrate specificities of the G. sulphuraria pPTs, we cloned the corresponding cDNAs into the yeast expression vector pYES/NT under control of the galactose-inducible GAL4 promoter. The regions of the cDNAs encoding for the putative target sequences of each protein were removed and instead fused to an N-terminal hexa- histidine tag. After transformation of the corresponding constructs into the yeast strain INVScl , all three pPT homologues could be successfully expressed and accumulated in the membrane fraction (Figure 3.2). Immunoblot analysis with an anti-polyhistidine tag antibody verified the galactose-inducible accumulation of the pPT proteins (Figure 3.2, lane 2-4) compared to controls, which maintained the empty expression vector (Figure 3.2, lane 1). The calculated molecular masses of the N-terminal His-tagged proteins were 41.6 kDa, 45.2 kDa and 41.2 kDa for GsTPT, GsPPT and GsGPT, respectively. The presence of recombinant protein was verified for all biological replicates by Western blot prior to reconstitution. 104 Figure 3.2. Expression of plastid phosphate transporter (pPT) homologues from Galdieria sulphuraria (Gs) in yeast Upper part: Membrane proteins from galactose-induced S. cerevisiae cells transformed either with an unmodified expression vector (Lane 1) or with constructs for GsTPT (lane 2), GsPPT (lane 3) and GsGPT (lane 4) were separated by SDS-PAGE and stained with Coomassie Blue R250. Lower part: Immunoblot analysis of an identical SDS- polyacrylamide-gel as described above. N-terminal poly-histidine tagged proteins were colormctrically verified with anti-His antibody and anti-mouse-IgG antibody linked to an alkaline phosphatase. TPT, triose-phosphate/phosphate translocator; GPT, Glucose-6- Phosphate-Translocator; PPT, L L ',,......;.."L L ‘ translocator; kDa, kilo r r K Dalton. 105 Functional analysis of three putative transport proteins from G. sulphuraria All functionally characterized plastidic phosphate translocators of higher plants catalyze, in addition to their characteristic substrates, also a strict homo-exchange of ortho- phosphate (F liigge, 1999). To test whether the yeast-expressed, recombinant GsPTs are functional, we thus first examined their ability to catalyze the signature Pi homo- exchange. To this end, they were reconstituted into liposomes that were preloaded with 30 mM Pi (i.e., the liposomes contained 30 mM Pi inside). Then radiolabeled [32P]- phosphate was added to the liposomes and the uptake kinetics were recorded. In GsTPT and GsPPT containing proteoliposomes, protein-mediated isotope equilibrium was reached within ten minutes, using an external Pi concentration of 0.25 mM (Figure 3.3). GsGPT mediated Pi uptake equilibrated after 15 minutes with a four times higher external [32P]-Pi concentration of 1 mM. Ortho-phosphate exchange followed a first- order kinetic with an equilibrium plateau (Vmax) of 400:13.74, 187.3133] and 237.1:702 nmol per mg protein and initial rates of 60.691248, 30.94:0.43 and 23.23:].27 nmol per mg protein and minute for GsTPT, GsPPT, and GSGPT, respectively. Control membranes from yeast cells transformed with an empty pYES/N T vector or boiled recombinant proteins showed negligible accumulation rates over time (data not shown). PLP and DIDS also effectively inhibited GspPT mediated import of [3 2P]-phosphate (data not shown). Marginal uptake of radiolabeled [3 2P] -phosphate compared to control experiments was measured when vesicles did not contain Pi as a counter substrate (Figure 3.3). In summary, all putative phosphate transport proteins from G. sulphuraria were frmctionally expressed in yeast and are able to catalyze the signature Pi homo-exchange activity, similar to their higher plants homologs. 106 Figure 3.3. Time-kinetics of pPT homologues fiom Galdieria sulphuraria (Gs) Liposomes were reconstituted with membrane proteins fi'om S. cerevisiae cells expressing GsTPT (panel A), GsPPT (panel B) or GsGPT (panel C). Uptake of ortho- Phosphate (Pi) was measured in the presence (0) or absence (0) of intraliposomal Pi (30 mM). Transport was initiated with a final concentration of 0.25 mM [32P]-Pi (panel A & B) or 1 mM [32P]-Pi (panel C). Diagrams represent a typical result of at least three independent studies. 107 400- 300- 2004 100- Uptake of Pi [nmol/mg protein] 0 2 4 6 s in Time [min] Uptake of Pr [nmol/mg protein] 3 f i 0 2 4 6 Time [min] A ‘1 I 8 Uptake of Pi [nmol/mg protein] Time [min] Figure 3.3. Time-kinetics of pPT homologues from Galdieria sulphuraria (Gs) 108 Transport properties of GsTPT, GsPPT and GsGPT To assess the substrate specificity of the recombinant GspPT proteins, vesicles were preloaded with saturating concentrations (i.e., 30 mM) of various counter-exchange substrates and initial rates of [32P]-Pi uptake were determined. GLUE Relative to the Pi/Pi homo-exchange experiment, the uptake rate of [32P]- Pi into proteoliposomes reconstituted with GsTPT was 86.7:13.3% when proteoliposomes were preloaded with the triose-phosphate DHAP. Uptake rates into liposomes preloaded with 3-phosphoglyceric acid (3-PGA), phosphoenolpyruvic acid (PEP), or glucose-6-phosphate (Glc6P), respectively, were low (Figure 3.4A). Uptake was also negligible when GsTPT liposomes were preloaded with glucose-l-phosphate (Glc l P), fructose-6-phosphate (Frc6P), galactose-1-phosphate(GallP), glycerol-3- phosphate (Gly3P) (Figure 3.4A), adenosine-monophosphate (AMP), uridine- monophosphate (UMP), gluconate-6—phosphate (Gluc6P), ribose-S-phosphate (Rib5P), ADP-glucose (ADP-Glc), UDP-glucose (UDP-Glc), UDP-galactose (UDP-Gal), alpha- ketoglutaric acid (or-KG), malic acid (Mal), pyruvic acid (Pyr), or oxaloacetic acid (0AA) (data not shown). GsTPT has an apparentKM value of 0.33:0.07 mM for Pi and DHAP competitively inhibited Pi uptake with a Ki value of 0520.04 mM (Table 3.3). 3- PGA and PEP inhibited [32P]-Pi uptake only at non-physiological concentrations of 8 and 9 mM, respectively, and Glc6P did not show significant inhibition of Pi transport at any of the tested concentrations. GsPPT: Reconstituted GsPPT protein efficiently used Pi and PEP as a counter substrate for the import of radiolabelled [32P]-phosphate (Figure 3.43). 18 additional substrates, as listed above for the GsTPT, showed marginal initial Pi uptake rates of less 109 than 30% of the Pi/Pi homo-exchange rate (Figure 3.4B and data not shown). GsPPT has an apparent KM value of 0.76:0.075 mM for Pi and a Ki value of 0.36:0.04 mM for PEP. Compared to PEP, ll-times higher 3-PGA and 22-times higher DHAP levels, respectively, were needed to inhibit Pi uptake by 50% (Table 3.3). Glc6P had no affinity for the Pi binding site. @911: Vesicles reconstituted with the GPT homologue from G. sulphuraria were preloaded with the same 20 substrates as given above; in all cases the [32P]-Pi uptake was much lower than for liposomes preloaded with 30 mM Pi (Figure 3.4C). GsGPT has raflier low affinity to ortho-phosphate (KM: 5.07:0.4 mM) (Table 3.3). Compared to non-preloaded proteoliposomes (l3.03:5.9%), approximately three- to fourfold higher uptake rates were measured when liposomes were preloaded with either 3-PGA or DHAP. Importantly, however, even high concentrations of 3-PGA and DHAP did not significantly inhibit Pi homo-exchange (Table 3.3). Glc6P and PEP are non-relevant substrates for the GsGPT protein. 110 Figure 3.4. Internal substrate dependence on the transport activities of GsTPT, GsPPT, and GsGPT Liposomes were preloaded with various substrates (30 mM) and reconstituted with membrane proteins from yeast cells expressing GsTPT (panel A), GsPPT (panel B) or GsGPT (panel C). Uptake experiments were initiated with 0.25 mM [32P]-Pi (panel A & B) or 1 mM [32P]-Pi (panel C) and terminated afier 1.5 min (GsTPT), 2 min (GsPPT) or 4 min (GsGPT). As negative control, proteoliposomes were preincubated (2 min) with inhibitor stop solution to calculate net uptake activity. Relative uptake activities were compared to the Pi/Pi counter-exchange experiment, which was set to 100%. Data are summarized as the arithmetic mean 1 SE of at least three independent experiments. Pi, ortho-phosphate; DHAP, dihydroxyacetone phosphate; 3-PGA, 3-phosphoglyceric acid; PEP, phosphoenolpyruvic acid; Glc6P, glucose-6-phosphate; GlclP, glucose-1- Phosphate; F rc6P, fructose-6-phosphate; GallP, galactose-l-phosphate, Gly3P, glycerol- 3-phosphate. 111 Internal substrate 0 20 40 60 80 100 Relative uptake [%] B Gly3P GallP Frc6P GlclP Glc6P PEP 3-PGA DHAP Non Pi 0 20 40 60 80 100 Relative uptake [%] Internal substrate Gly3P C GallP Frc6P GlclP Glc6P PEP 3-PGA DHAP Non Pi 0 20 40 60 80 100 Relative uptake [%] Internal substrate Figure 3.4. Internal substrate dependence on the transport activities of GsTPT, GsPPT and GsGPT 112 Table 3.3. Comparison of kinetic constants from G. sulphuraria and higher plant pPT homologues The Michaelis-Menten constant (KM) for Pi was determined using various external [32P]-Pi concentrations (0.05 mM - 10 mM). The competitive inhibition constant (Ki) of [32P]-Pi uptake was assayed at two different external Pi concentrations with increasing inhibitor concentrations (0.05 mM — 20 mM) (Dixon, 1953). All proteoliposomes were preloaded with 30 mM Pi. Data represent the arithmetic mean :1: SE of at least three independent experiments. Recombinant higher plant pPT data are taken from Flugge (1999). “None” refers to: No competitive inhibitory constant could be measured under the given experimental conditions. Abbreviations are listed in Figure 3.4. Kinetic TPT homologue PPT homologue GPT homologue Constants [mM] [mM] [HIM] Plants Gs Plants Gs Plants Gs KM (Pi) 1.0 0.33 2 0.07 0.8 0.76 :r: 0.08 1.] 5.07 a: 0.4 Ki (DHAP) 1.0 0.5 :t 0.04 8.0 7.93 :1: 0.23 0.6 None Ki (3-PGA) 1.0 7.71 a 0.47 4.6 3.95 :r: 0.1 1.8 None Ki (PEP) 3.3 8.75 = 0.40 0.3 0.36 a: 0.04 2.9 None Ki (Glc-6P) None None None None 1.1 None Transcript levels and protein activity of the plastidic phosphate translocator homologues in G. sulphuraria cells In land plants, TPT expression is confined to photosynthetic tissues whereas GPT expression is highest in heterotrophic tissues (Kammerer et al., 1998; Flugge, 1999; Niewiadomski et al., 2005). PPT-genes are expressed at low levels in various tissues (Knappe et al., 2003b). To determine the expression pattern of GspPTs, semi-quantitative RT-PCR was used. Figure 3.5 shows that Gs TPT, GsPPT, and GsGPT transcripts are clearly detectable and abundant under both, autotrophic and heterotrophic growth conditions. The steady state transcript levels of all three genes were higher than those of the reference gene actin (G515600; Galdieria Genome Browser Build 3.0). Liposomes 113 reconstituted with total membrane fractions exhibit high [32P]-Pi import using DHAP as a counter substrate under both culturing regimes (Figure 3.5). Exchange rates with PEP and 3-PGA as counter-exchange substrates were much lower as compared to DHAP. Exchange of glucose-6-phosphate with Pi was close to the background values of imloaded vesicles (Figure 3.5). 114 Figure 3.5. Transcript abundance and protein activity of GsTPT, GsPPT, and GsGPT in G. sulphuraria cells Upper part: Total RNA from auto- and heterotrophically cultured G. sulphuraria cells was isolated, treated with DNase and used for first strand cDNA synthesis. Expression of GsTPT, GsPPT, GsGPT and actin transcripts was amplified by PCR with 25, 30 or 35 elongation cycles and visualized on an ethidium bromide-stained 1.25% (w/v) agarose- gel. Lower part: Relative [32P]-Pi uptake rates into liposomes reconstituted with total G. sulphuraria membranes isolated from heterotrophically (black bar) or autotrophically (white bar) grown cells, respectively. Membrane fractions were reconstituted into liposomes preloaded with 30 mM various substrates. Experiment was performed and abbreviations were used as given in the legend to Figure 3.4. 115 Autotrophic Heterotrophic PCR m m GsTPT GsPPT GsGPT «II-I- GsActin 120- D Autotrophic — 100. - Heterotrophic W O I ch 6 1 Relative uptake [% N 65 e G l I ml— (““399 “0193 .5308 9"? (X6? Figure 3.5. Transcript abundance and protein activity of GsTPT, GsPPT, and GsGPT in G. sulphuraria cells 116 DISCUSSION In this study, we report the molecular and biochemical characteristics of the plastidic phosphate translocator family from the unicellular red alga G. sulphuraria. All three members of the GsPT gene family were heterologously expressed in S. cerevisiae, functionally reconstituted into liposomes (Figure 3.2 and Figure 3.3), and their substrate specificities and kinetic constants were assessed (Figure 3.4 and Table 3.3). The G. sulphuraria genome encodes a high-affinity TPT with narrow substrate specificity Recombinant G. sulphuraria TPT mediates a strict counter-exchange of the triose- phosphate DHAP with ortho-phosphate (Figure 3.4). Interestingly, the kinetic constants of GsTPT considerably differ from those of the green plant TPT (Table 3.3). The red algal transporter exhibits a three times higher affinity for Pi and, as indicated by its low Kr value, the substrate DHAP has a twofold higher affinity to the Pi binding site. Further, GsTPT is highly specific for DHAP. In contrast to the green plant ortholog, 3-PGA poorly acts as a counter substrate, even under saturating conditions (Figure 3.4A). In addition, 3-PGA is a weak competitive inhibitor of Pi import with an eightfold and fifteen-fold higher Ki value as compared to the TPT fi'om plants and the Ki (DHAP) constant from GsTPT, respectively (Table 3.3). These results argue strongly against 3- PGA as a physiological relevant substrate of GsTPT. PEP and Glc6P were even less efficient exchange substrates and inhibitors of Pi transport (Figure 3.4 and Table 3.3). Very likely, the distinct kinetic properties of GsTPT represent an important adaptation to red algal carbon metabolism. In contrast to the Chloroplastida, red algae store an 117 insoluble starch-like polymer called floridean starch and synthesize cell wall building blocks and the main soluble carbohydrate floridoside, composed of UDP-Gal and Gly3P moieties in their cytosol (Viola et al., 2001; Collen et al., 2004; Barbier et al., 2005a). Hence, primary carbon partitioning in red algae is predominantly organized within the cytosol and GsTPT enables solely the export of triose-phosphates, even at low stroma] concentrations, to cope with the massive cytosolic demand of recently fixed carbon (Figure 3.6). On the other hand, GsTPT is highly expressed and active in heterotrophically cultured G. sulphuraria cells (Figure 3.5). Under these conditions, GsTPT presumably mediates TP import into the plastid, thus mediating carbon flux between cytosol and non-photosynthetic rhodoplasts. In contrast, land plants export TP from photosynthetic chloroplasts via TPT and import Glc6P into heterotrophic plastids via GPT. The TPT is an integral part of the photoassimilate partitioning in plants (Hausler et al., 2000; Smith and Stitt, 2007). Accumulation of sucrose and hexose- phosphates in the cytosol sequesters cytosolic Pi and thus slows down the export of triose-phosphate. A resulting increased stroma] 3-PGA/Pi ratio allosterically activates ADP-glucose pyrophosphorylase, which is the committing step of plastidic starch biosynthesis (Ballicora et al., 2004). Plants have to maintain sufficiently high levels of TP in the plastid stroma to enable both the regeneration of the C02 acceptor ribulose-1,5- bisphosphate and starch biosynthesis during the light period (Zeeman et al., 2007). The apparent lower affinity of the plant transporter for Pi and TP thus ensures higher steady state levels of TP in the stroma to drive these reactions. 118 G. sulphuraria has a PPT with similar properties as its green plant ortholog PEP uptake from the cytosol is required for stroma localized fatty acid and shikimate biosynthesis (F lfigge, 1999). Previously published results have been ambiguous about the existence of a putative PEP/Pi transporter (PPT) in G. sulphuraria. Phylogenetic analysis revealed an orthologous candidate gene (Weber et al., 2006), although isolated membrane fractions did not exhibit pronounced PEP uptake activity (Weber et al., 2004). Here, we show that EST GsAl6F5 (Weber et al., 2006) represents a highly conserved PPT with almost identical kinetic constants in both the green and red lineages (Table 3.3). Both transporters are highly specific for Pi and PEP and poorly accept 3-PGA, DHAP, or any other of the 16 tested metabolites as a substrate (Figure 3.4B, data not shown). In addition, we now employed an improved reconstitution method for total membranes from autotrophic and heterotrophic cell cultures that yielded lower background and a weak but clearly detectable PEP/Pi counter-exchange activity (Figure 3.5). The low PEP transport capacity (Figure 3.5) mirrors published data from chloroplasts of C3 land plants (Flugge and Weber, 1994). TPT is by far the most abundant transport protein in C3 chloroplasts and accounts for a high DHAP exchange capacity (F lugge, 1999). Fatty acid biosynthesis and the shikimate pathway in red algae are predicted to be similar to that of seed plants and are presumably localized in the plastid, based on phylogenetic analyses and N- terminal target sequence predictions (Weber et al., 2004; Richards et al., 2006). A question that cannot be conclusively answered at the moment is whether rhodoplasts, analogous to chloroplasts of green plants, also have negligible activities of plastid phosphoglyceromutase and enolase to produce PEP from plastidial 3-PGA (Trimming and Emes, 1993). The G. sulphuraria genome encodes for two phosphoglyceromutase 119 isozymes (Gs04140 and Gs52680) and a single enolase (Gs21490), which are all three computationally predicted as cytosolic enzymes (Nielsen et al., 1997) without any N- terminal targeting sequences, when compared with their homologous genes from A. thaliana (phosphoglyceromutase: At4g09520, At3g08590, At1g09780; enolase: At2g36530, At2g29560, At1g74030) (Friso et al., 2004; Larkin et al., 2007). Alternatively, PEP could be generated fi'om pyruvic acid via the pyruvate orthophosphate dikinase (PPDK) reaction. The annotated PPDK gene Gs42070 is computationally predicted to be localized in the cytosol, exhibiting a high similarity with cyanobacterial PPDK genes, and is not phylogenetically related to the dual targeted enzyme from A. thaliana (At4g15530) (Parsley and Hibberd, 2006). While computational targeting predictions have to be takenwith a grain of salt, bioinformatics analysis indicates that conversion of TP to PEP is not possible in G. sulphuraria and rhodoplasts thus depend on PEP import fi'om the cytosol to drive PEP-depended reactions in the stroma (Figure 3.6). G. sulphuraria does not possess a plastidic hexose phosphate importer Reconstituted recombinant putative GsGPT mediated a strict Pi/Pi homo-exchange that followed a first order rate kinetics (Figure 3.3). An unusually high KM value of 5 mM for Pi (Table 3.3) indicates that GsGPT under physiological conditions is not able to compete with GsTPT and GsPPT for the common substrate Pi. While liposomes preloaded with high (i.e., saturating) concentrations of DHAP or 3-PGA (i.e., 30 mM internal substrate concentration) showed Pi uptake rates of approximately 50% of Pi/Pi homo-exchange, it is important to notice that neither 3-PGA nor DHAP significantly inhibited Pi/Pi homo- exchange either at physiological concentrations or at lOO-fold excess (Table 3.3). The 120 prime substrate candidate glucose-6-phosphate was poorly accepted under saturating conditions (Figure 3.4) and it did not inhibit Pi import in the Pi/Pi homo-exchange setup (Table 3.3). These results strongly indicate that 3-PGA, DHAP, and Glc6P are not physiologically relevant substrates of the putative GsGPT. We also tested various additional metabolites, such as nucleotide-sugars, mononucleotides, hexose-phosphates (GlclP, Frc6P, and Gall P), pentose-phosphates and a precursor of floridoside biosynthesis, glycerol-3-phosphate. All of these are poorly exchanged with [32P]-Pi and the physiological substrate of GsGPT thus remains unidentified. In non-green plastids of plants, Glc6P is the preferred precursor for starch synthesis and NADPH generation via the oxidative pentose phosphate pathway (OPPP) due to the absence of a fructose-1,6- bisphosphatase (FBPase) activity (Flugge, 1999). In photosynthetic tissues, plastidic F BPase is inactivated at night and thus hexose phosphates cannot be generated fi'om triose phosphates (and vice versa). In contrast, red algal starch biosynthesis is cytosolic and plastidic FBPase fiom G. sulphuraria is not subject to a strict redox regulation (Reichert et al., 2003; Oesterhelt et al., 2007). Importing triose-phosphates via the GsTPT could thus sustain the production of hexose-phosphate for carbon and NADPH supply in the rhodoplast during the night or prolonged heterotrophic growth conditions, thus bypassing the requirement for a plastidic hexose phosphate translocator (Figure 3.6). Further work will be required to pinpoint the catalytic activity of this enigmatic protein. An operative DHAP/3-PGA reduction shuttle is unlikely in G. sulphuraria 3-PGA is a major substrate for the TPT in higher plants (F lfigge, 1999). In spinach and other model plants, 3-PGA levels in the plastid and the cytosol frequently exceed DHAP 121 levels by several fold (Gerhardt et al., 1987). However, the affmity of the spinach TPT is identical for both metabolites. It has been proposed that in vivo a significant amount of 3- PGA transport occurs across the envelope membrane; the TPT could thus operate as a NAD(P)H reduction equivalent shuttle between stroma and cytosol (Heineke etal., 1991). The substrate specificity and the kinetic constants of recombinant proteins, as discussed in detail earlier, and total membrane fiactions from G. sulphumria presumably cannot sustain a physiological relevant exchange rate of 3-PGA across the rhodoplasts envelope membrane. Still, DHAP offers a possibility to export reduction equivalent to the cytosol. The non-phosphorylating glyceraldehyde-3 -phosphate dehydrogenase (NP- GAPDH) (Rius et al., 2006), which is also present in the draft genome sequence of the red alga (Gs00840) or the combined glycolytic reaction sequence of glyceraldehyde-3- phosphate dehydrogenase (GAPDH) and 3-PGA kinase (PGK) (Plaxton, 1996) generates NADPH, or NADH and ATP, respectively. However, in contrast to green plants, not 3- PGA but Pi must be postulated as the alternative counter-substrate for DHAP. Impact of subtle differences of TPT in Chloroplastida and Rhodophyta Triose-phosphate is the main entry and branching point in the glycolytic network in photosynthetically active cells (Plaxton, 1996). Metabolic flux has to be regulated between generating hexose-phosphates via gluconeogenesis and glycolytic “downstream” products, such as 3-PGA, PEP, and pyruvic acid. In higher plants, the redox transfer by a TP/3-PGA shuttle via (NP)-GAPDH and PGK is proposed to be marginal and GAPDH expression is mamly induced under stress conditions, such as anaerobiosis and heat shock or sucrose feeding (Yang et al., 1993; Bustos et al., 2008; Holtgrefe et al., 2008). On the 122 other hand, glycolysis has to provide adequate amounts of PEP to replenish TCA intermediates used for biosynthesis of, e.g., amino acids, to produce pyruvic acid for mitochondrial metabolism, or to fuel plastid fatty acid biosynthesis (Plaxton, 1996). Because 3-PGA transport is seemingly a specific adaptation in plants, we hypothesize that 3-PGA transport across the chloroplast envelope membrane might provide a bypass of the complex regulation of the cytosolic triose-phosphate pool that permits supporting cytosolic PEP biosynthesis and serves under stress conditions as an additional redox shuttle system. 123 Figure 3.6. Flexible transport of triose-phosphate across the plastid membrane is central for the cellular carbon metabolism in the red alga G. sulphuraria In photosynthetically active cells, triose-phosphate is exported via the GsTPT from rhodoplasts to sustain soluble and insoluble carbohydrate synthesis (floridoside and floridean starch) in the cytosol. Under heterotrophic growth conditions the same transporter supplies the rhodoplasts with triose-phosphate for, e.g., hexose-phosphate synthesis. Rhodoplasts are not capable to produce PEP via partial glycolysis and are thus dependent on the import of PEP from the cytosol via the PPT protein for fatty acid and aromatic amino acid biosynthesis. The red algal plastid envelope does not contain any 3- PGA and Glc-6—P transport activities, which is a specific trait of green plants. PPT, PEP/phosphate-translocator; TPT, triose—phosphate/phosphate-translocator; F rcl ,6BP, fructose-1,6-bisphosphate; Gal, galactose; Glc, glucose; Ru-l,SBP, ribulose-1,5- bisphosphate; other acronyms are listed in the legend to Figure 3.4. The figure in this dissertation is presented in color. 124 C I m” Floridoside KRhodoplast_ .. _____ —L UDP-Gal UDP- Glc Glc-6-P HGlc-1 -P GIc-6-P l Glucose UDP- Glc PEP T / Floridean Starch Figure 3.6. Flexible transport of triose-phosphate across the plastid membrane is central for the cellular carbon metabolism in the red alga G. sulphuraria 125 Conclusions Although the primary structures of plastidic phosphate translocators are highly conserved between green plants and red algae, these proteins have evolved quite distinct biochemical‘characteristics in the different lineages of the Archaeplastida. Comparative biochemical analysis of candidate orthologs fi'om distantly related organisms thus provides novel insights into alternative modes for regulating “ancient” metabolic pathways. While the triose-phosphate/phosphate antiport activity across the plastid envelope membrane is highly conserved across all photosynthetic eukaryotes analyzed to date, the red algal transporters catalyze neither 3-PGA nor Glc6P transport (Figure 3.6). G. sulphuraria must have evolved alternative mechanisms to distribute triose- phosphate into various pools ('Le. floridean starch, floridoside, fatty acid and amino acid synthesis), a less stringent redox control of Calvin cycle, oxidative pentose phosphate pathway, and glycolysis (Oesterhelt et al., 2007). Plastidial phosphate translocators thus represent crucial components of primary carbon partioning in higher plants and red algae that have evolved to the specific requirements of each lineage through modulation of substrate specificities and kinetic constants. 126 REFERENCES Ad] S.M., Farmer M.A., Andersen R.A., Anderson O.R., Barta J.R., Bowser S.S., Brugerolle G., Fensome R.A., Fredericq S., James T.Y., Karpov S., Kugrens P., Krug J., Lane C.E., Lewis L.A., Lodge J., Lynn D.H., Mann D.G., McCourt R.M., Mendoza L., Moestrup 0., Mozley-Standridge S.E., Nerad T.A., Shearer C.A., Smirnov A.V., Spiegel F.W., and Taylor M.F. (2005) The New Higher Level Classification of Eukaryotes with Emphasis on the Taxonomy of Protists. Journal of Eukaryotic Microbiology 52:399-451. Arai M., Mitsuke H., Ikeda M., Xia J-X., Kikuchi T., Satake M., and Shimizu T. (2004) ConPred II: a consensus prediction method for obtaining transmembrane topology models with high reliability. Nucleic Acids Research 32:W390-393. Ballicora M.A., Iglesias A.A., and Preiss J. (2004) ADP-glucose pyrophosphorylase: a regulatory enzyme for plant starch synthesis. Photosynthesis Research 79: 1-24. Barbier G., Oesterhelt C., Larson M.B., Halgren R.G., Wilkerson C., Garavito R.M., Benning C., and Weber A.P.M. (2005a) Comparative Genomics of Two Closely Related Unicellular Thermo-Acidophilic Red Algae, Galdieria sulphuraria and Cyanidioschyzon merolae, Reveals the Molecular Basis of the Metabolic Flexibility of Galdieria sulphuraria and Significant Differences in Carbohydrate Metabolism of Both Algae. Plant Physiology 137:460-474. Barbier G., Zimmermann M., and Weber A.P.M. (2005b) Genomics of the thermo- acidophilic red alga Galdieria sulphuraria. In BH Richard, VL Gilbert, YR Alexei, GR Gladstone, eds., SPIE, San Diego, USA, 5906:590-609. Bhattacharya D., Archibald J.M., Weber A.P.M., and Reyes-Prieto A. (2007) How do endosymbionts become organelles? Understanding early events in plastid evolution. BioEssays 29:]239-1246. Bhattacharya D., Yoon H.S., and Hackett J.D. (2004) Photosynthetic eukaryotes unite: endosymbiosis connects the dots. BioEssays 26:50-60. Bouvier F., Linka N., Isner J.C., Mutterer J., Weber A.P.M., and Camara B. (2006) Arabidopsis SAMT 1 Defines a Plastid Transporter Regulating Plastid Biogenesis and Plant Development. Plant Cell 18:3088-3105. Bustos B.M., Bustamante C.A., and Iglesias A.A. (2008) Involvement of non- phosphorylating glyceraldehyde-B-phosphate dehydrogenase in response to oxidative stress. Journal of Plant Physiology 165:456-461. Chomczynski P., and Sacchi N. (1987) Single-Step Method of RNA Isolation by Acid Guanidinium Thiocyanate Phenol Chloroforrn Extraction. Analytical Biochemistry 162: 1 56- 159. 127 Collen P.N., Camitz A., Hancock R.D., Viola R., and Pedersen M. (2004) Effect of nutrient deprivation and resupply of metabolites and enzymes related to carbon allocation in Gracilaria tenuistipitata (Rhodophyta). Journal of Phycology 40:305-314. Coppin A., Varré J.S., Lienard L., Dauvillee D., Guérardel Y., Soyer—Gobillard M.O., Buleon A., Ball S., and Tomavo S. (2005) Evolution of Plant-Like Crystalline Storage Polysaccharide in the Protozoan Parasite T oxoplasma gondii Argues for a Red Alga Ancestry. Journal of Molecular Evolution 60:257-267. Dixon M. (1953) The determination of enzyme inhibitor constants. Biochemical Journal 55: 1 70-1 71 . Eicks M., Maurino V., Knappe S., Fliigge U.l., and Fischer K. (2002) The Plastidic Pentose Phosphate Translocator Represents a Link between the Cytosolic and the Plastidic Pentose Phosphate Pathways in Plants. Plant Physiology 128:512-522. Fischer K., Kammerer B., Gutensohn M., Arbinger B., Weber A., Hausler R.E., and Fliigge U.I. (1997) A New Class of Plastidic Phosphate Translocators: A Putative Link between Primary and Secondary Metabolism by the phosphoenolpyruvate/ phosphate Antiporter. Plant Cell 9:453-462. Flugge U.l., and Gao W. (2005) Transport of Isoprenoid Intermediates Across Chloroplast Envelope Membranes. Plant Biology 7:91-97. Flugge DJ. (1999) Phosphate translocator in plastids. Annual Review of Plant Physiology and Plant Molecular Biology 50:27-45. Flugge U.l., and Weber A. (1994) A rapid method for measuring organelle-specific substrate transport in homogenates from plant tissues. Planta 194:181-185. Friso G., Giacomelli L., Ytterberg A.J., Peltier J .B., Rudella A., Sun Q., and van Wijk K.J. (2004) In depth analysis of the thylakoid membrane proteome of Arabidopsis thaliana chloroplasts: new proteins, new functions, and a plastid proteome database. Plant Cell 16:478-499. Gerhardt R., Stitt M., and Heldt H.W. (1987) Subcellular metabolite levels in spinach leaves: Regulation of sucrose synthesis during diurnal alterations in photosynthetic partitioning. Plant Physiology 83:399-407. Gross W., Lenze D., Nowitzki U., Weiske J., and Schnarrenberger C. (1999) Characterization, cloning, and evolutionary history of the chloroplast and cytosolic class I aldolases of the red alga Galdieria sulphuraria. Gene 230:7-14. Gross W., and Schnarrenberger C. (1995) Heterotrophic Growth of Two Strains of the Acido-Therrnophilic Red Alga Galdieria sulphuraria. Plant and Cell Physiology 36:633- 638. 128 Hiusler R.E., Schlieben N.H., Nicolay P., Fischer K., Fischer K.L., and Flugge U.I. (2000) Control of carbon partitioning and photosynthesis by the triose phosphate/ phosphate translocator in transgenic tobacco plants (Nicotiana tabacum L.). 1. Comparative physiological analysis of tobacco plants with antisense repression and overexpression of the triose phosphate/phosphate translocator. Planta 210:371-3 82. Heineke D., Riens B., Grosse H., Hoferichter P., Peter U., Flugge U.l., and Heldt H.W. (1991) Redox Transfer across the inner chloroplast envelope membrane. Plant Physiology 95:1131-1137. Holtgrefe S., Gohlke J., Starmann J., Druce S., Klocke S., Altmann B., Wojtera J., Lindermayr C., and Scheibe R. (2008) Regulation of plant cytosolic glyceraldehyde 3- phosphate dehydrogenase isoforms by thiol modifications. Physiologia Plantarum 133:2] 1-228. Kammerer B., Fischer K., Hilpert B., Schubert S., Gutensohn M., Weber A., and Fliigge U.l. (1998) Molecular Characterization of a carbon transporter in plastids from heterotrophic tissues: The glucose 6-phosphate/phosphate antiporter. Plant Cell 10: 105- 1 1 8. Kasahara M., and Hinkle P.C. (1977) Reconstitution and purification of the D-glucose transporter from human erythrocytes. Journal of Biological Chemistry 252:73 84-73 90. Knappe S., Fliigge 0.1., and Fischer K. (2003a) Analysis of the plastidic phosphate translocator gene family in Arabidopsis and identification of new phosphate translocator homologous transporters, classified by their putative substrate binding site. Plant Physiology 131:1178-1 190. Knappe S., Lottgert T., Schneider A., Vol] L., Fliigge U.l., and Fischer K. (2003b) Characterization of two functional phosphoenolpyruvate/phosphate translocator (PPT) genes in Arabidopsis (AtPPTl) may be involved in the provision of signals for correct mesophyll development. Plant Journal 36:411-420. Larkin M.A., Blackshields G., Brown N.P., Chenna R., McGettigan P.A., McWilliam H., Valentin F., Wallace I.M., Wilm A., Lopez R., Thompson J.D., Gibson T.J., and Higgins D.G. (2007) Clustal W and Clustal X version 2.0. Bioinformatics 23:2947-2948. Nielsen H., Engelbrecht J., Brunak S., and von Heijne G. (1997) Identification of prokaryotic and eukaryotic signal peptides and prediction of their cleavage sites. Protein Engineering 10:1-6. Niewiadomski P., Knappe S., Geimer S., Fischer K., Schulz B., Unte U.S., Rosso M.G., Ache P., Flugge U.l., and Schneider A. (2005) The Arabidopsis Plastidic glucose 6-phosphate/phosphate translocator (GPT 1) is Essential for pollen maturation and embryo sac development. Plant Cell 17:760-775. 129 Oesterhelt C., KIocke S., Holtgrefe S., Linka V., Weber A.P.M., and Scheibe R. (2007) Redox Regulation of chloroplast enzymes in Galdieria sulphuraria in view of eukaryotic evolution. Plant Cell Physiology 48: 1359-1373. Parsley K., and Hibberd J. (2006) The Arabidopsis PPDK gene is transcribed fiom two promoters to produce differentially expressed transcripts responsible for cytosolic and plastidic proteins. Plant Molecular Biology 62:339-349. Patron N.J., and Keeling P.J. (2005) Common evolutionary origin of starch biosynthetic enzymes in green and red algae. Journal of Phycology 41:1131-1141. Plaxton W.C. (1996) The organization and regulation of plant glycolysis. Annual Review of Plant Physiology and Plant Molecular Biology 47: 185-2 14. Reichert A., Dennes A., Vetter S., and Scheibe R. (2003) Chloroplast fi'uctose 1,6- bisphosphatase with changed redox modulation: comparison of the Galdieria enzyme with cysteine mutants fi'om spinach. Biochimica et Biophysica Acta 1645:212-217. Reiter W.D. (2002) Biosynthesis and properties of the plant cell wall. Current Opinion Plant Biology 5:536—542. Reyes-Prieto A., Weber A.P.M., and Bhattacharya D. (2007) The Origin and Establishment of the Plastid in Algae and Plants. Annual Review of Genetics 41: 147-168. Richards T., Dacks J.B., Campbell S.A., Blanchard J.L., Foster P.G., McLeod R., and Roberts C.W. (2006) Evolutionary origins of the eukaryotic shikimate pathway: gene fusions, horizontal gene transfer, and endosymbiotic replacements. Eukaryotic Cell 5:1517-1531. Rius S., Casati P., Iglesias A., and Gomez-Casati D. (2006) Characterization of an Arabidopsis thaliana mutant lacking a cytosolic non-phosphorylating glyceraldehyde-3- phosphate dehydrogenase. Plant Molecular Biology 61:945-957. Sambrook J., Fritsch E.F., and Maniatis T. (1995) Molecular Cloning: A Laboratory Manual 2nd Ed. Cold Spring Harbor Laboratory, Cold Spring Harbor, NY Schlichting R., and Bothe H. (1993) The cyanelles (organelles of a low evolutionary scale) possess a phosphate-translocator and a glucose-carrier in Cyanophora paradoxa. Botanica Acta 106:428-434. Shultz R.W., Settlage S.B., Hauley-Bowdoin L., and Thompson W.F. (2005) A Trichloroacetic acid-acetone method greatly reduces infiared autofluorescence of protein extracts from plant tissue. Plant Molecular Biology Report 23:405-409. 130 Smith A.M., and Stitt M. (2007) Coordination of carbon supply and plant growth. Plant, Cell and Environment 30:1126-1149. Tegeder M., and Weber A.P.M. (2006) Metabolite transporters in the control of plant primary metabolism. In WC Plaxton, MT McManus, eds, Control of Primary Metabolism in Plants. Blackwell Publishing Ltd, Oxford, UK, PP85-120. Trimming R.A., and Emes MJ. (1993) Glycolytic enzymes in non-photosynthetic plastids of pea (Pisum sativum L.) roots. Planta 190:439-445. Tyra H., Linka M., Weber A.P.M., and Bhattacharya D. (2007) Host origin of plastid solute transporters in the first photosynthetic eukaryotes. Genome Biology 8:R212. Viola R., Nyvall P., and Pedersen M. (2001) The unique features of starch metabolism in red algae. Proceedings of the Royal Society of London Series B-Biological Sciences 268:1417-1422. Vol] L., Hausler R.E., Hecker R., Weber A., Weissenbock G., Fiene G., Waffenschmidt S., and Fliigge U.I. (2003) The phenotype of the Arabidopsis cue] mutant is not simply caused by a general restriction of the shikimate pathway. Plant Journal 36:301-317. Weber A.P.M., and Fischer K. (2007) Making the connections - The crucial role of metabolite transporters at the interface between chloroplast and cytosol. Federation of European Biochemical Societies (FEBS) Letters 581:2215-2222. Weber A.P.M., Linka M., and Bhattacharya D. (2006) Single, ancient origin of a plastid metabolite translocator family in Plantae fi'om an endomembrane-derived ancestor. Eukaryotic Cell 5:609-612. Weber A.P.M., Schwacke R., and Flugge U.I. (2005) Solute transporters of the plastid envelope membrane. Annual Review of Plant Biology 56:133-164. Weber A., Oesterhelt C., Gross W., Brautigam A., Imboden L., Krassovskaya 1., Linka N., Truchina J., Schneidereit J., Vol] H., Vol] L., Zimmermann M., Jamai A., Riekhof W., Yu B., Garavito R., and Benning C. (2004a) EST-analysis of the thenno- acidophilic red microalga Galdieria sulphuraria reveals potential for lipid A biosynthesis and unveils the pathway of carbon export from rhodoplasts. Plant Molecular Biology 55:17-32. Yang Y., Kwon H.B., Peng H.P., and Shih MC. (1993) Stress responses and metabolic regulation of g1yceraldehyde-3-phosphate dehydrogenase genes in Arabidopsis. Plant Physiology 101:209-216. 13] Yoon H.S., Hackett J.D., Ciniglia C., Pinto G., and Bhattacharya D. (2004) A molecular timeline for the origin of photosynthetic eukaryotes. Molecular Biology and Evolution 21:809—818. Zeeman S.C., Smith S.M., and Smith A.M. (2007) The diurnal metabolism of leaf starch. Biochem Journal 401:13-28. Zrenner R., Stitt M., Sonnewald U., and Boldt R. (2006) Pyrimidine and purine biosynthesis and degradation in plants. Annual Review Plant Biology 57:805-836. 132 CHAPTER 4 PROPOSED METABOLITE TRANSPORT PROCESSES FOR THE RE- ASSIMILATION OF AMMONIA DURING PHOTORESPIRATION1 4 This work has been published in Linka M., and Weber A.P.M. (2005) Shufiling ammonia between mitochondria and plastids during photorespiration. Trends in Plant Science 10:46] -465. 133 ABSTRACT Surprisingly, it was recently found that glutamine synthetase is dual targeted to chloroplasts and mitochondria in Arabidopsis thaliana leaves. Mitochondrial glutamine synthetase likely assimilates ammonia, which is generated in large amounts in mitochondria during photorespiration. However, ammonia assimilation is a two-step process and the second step, catalyzed by glutamate synthase, is exclusively located in plastids. Hence, a shuttle for ammonia, possibly in the form of amino acids, is required between mitochondria and plastids. We discuss two alternative shuttles, an ornithine- citrulline shuttle and a glutamine-glutarnate shuttle. Both shuttles allow the safe transport of the toxic metabolite ammonium in the form of amino acids. The ornithine—citrulline shuttle also provides a means for the transport of carbon dioxide from mitochondria to plastids, but this shuttle requires more energy than the alternative glutamate-glutamine shuttle. RESULTS Photorespiration is a complex, multi-compartrnent biochemical pathway Ribulose-1,5-bisphosphate carboxlyase/oxygenase (RubisCO) is a bifimctional enzyme that catalyzes both the carboxylation of ribulose 1,5-bisphosphate (RuBP) and its oxygenation. Oxygenation of RuBP produces one molecule of 3-PGA and one molecule of 2-phosphoglycolate (2-PG). 2-PG cannot be directly metabolized inside chloroplasts; instead, two molecules of 2-PG are recycled to one molecule of 3-PGA by the complex, multi-compartrnent photosynthetic carbon oxidation cycle (Figure 4.1). Because this 134 pathway leads to consumption of oxygen and production of carbon dioxide in the light, it is also called photorespiration. The photorespiratory pathway involves chloroplasts, peroxisomes, and mitochondria. 2-PG is dephosphorylated by phosphoglycolate phosphatase (PGP) in the chloroplast and the resulting glycolate is shuttled to peroxisomes where it is oxidized to glyoxylate, which is subsequently transaminated to glycine. Glycine is transported to the mitochondria where two molecules of glycine are converted to one molecule each of serine, ammonia, and carbon dioxide by glycine decarboxylase (GDC) and serine hydroxymethyltransferase (SHMT). Serine is transported back to the peroxisomes, converted to glycerate by serine glyoxylate aminotransferase and hydroxypyruvate reductase (HPR) and glycerate is finally shuttled to the chloroplasts where it is phosphorylated to 3-PGA by glycerate kinase (GLYK). The reaction sequence converts two molecules of 2-PG into one molecule of each 3-PGA and carbon dioxide and thus leads to the net loss of one carbon (W ingler et al., 2000; Douce and Neuburger, 1999). The oxidation of glycine by GDC in mitochondria during photorespiration generates massive amounts of the cytotoxic metabolite ammonia, exceeding ammonia production fi'om nitrate via the concerted action of nitrate reductase and nitrite reductase by at least one order of magnitude (Ogren, 1984). It was assumed that photorespiratory ammonia would be either actively or passively transported to plastids to be assimilated into glutamate by the glutamine synthetase 2 (GS2) and Fd- dependent glutamate synthase (F d-GOGAT) system (Figure 4.1). 135 Glutamine synthetase is dual targeted to plastids and mitochondria The surprising discovery of dual targeting of GS2 to both plastids and mitochondria (T aira et al., 2004) has changed this view - the presence of G82 in mitochondria opens the possibility of ammonia assimilation by GS in the mitochondrial matrix, directly at the site of ammonia generation by GDC. However, to maintain the nitrogen stoichiometry of the photorespiratory pathway, glutamine (Gln) needs to be converted to glutamate (Glu), which serves as amino donor for the peroxisome-localized glyoxylate aminotransferase reaction. The conversion of glutamine and alpha-ketoglutarate (or-KG) to glutamate is catalyzed by F d-GOGAT and this enzyme is exclusively localized in the plastid stroma (Lancien et aL, 2002; Coschigano et al., 1998). Alternatively, it could be envisaged that ammonia would be directly assimilated into glutamate in mitochondria by glutamate dehydrogenase (GDH). However, GDH mutants do not show a photorespiratory phenotype in ambient air (Melo-Oliveira et al., 1996), whereas Fd-GOGAT mutants show a strong photorespiratory phenotype and are not viable in ambient air (Coschigano et al., 1998; Somerville and Ogren, 1980). Hence, Fd-GOGAT is essential and glutamine (or ammonia) needs to be shuttled fi'om mitochondria to chloroplasts to close the photorespiratory nitrogen cycle. 136 Figure 4.1. Simplified scheme of the current textbook photorespiratory cycle The oxygenation of ribulose-1,5-bisphosphate (RuBP) leads to the production of 2- phosphoglycolate (2-PG) which is dephosphorylated to glycolate (Glt). After transfer to peroxisomes, glycolate is oxidized to glyoxylate (Gox) and aminated to glycine (Gly), which is then transported to the mitochondria. Gly is oxidatively decarboxylated by glycine decarboxylase (GDC) in the mitochondrial matrix, yielding Methylen-THF, C02, and ammonia. Ammonia and C02 arrive back in the chloroplasts, presumably by diffusion, to be re-assimilated by glutamine synthetase (GS) and glutamate synthase (GOGAT) and RubisCO, respectively. Glutamate, the amino donor for the amination of Gox in the peroxisomes and the reaction product alpha-ketoglutarate (or-KG) are shuttled between plastids and peroxisomes by a plastidic malate-coupled two-translocator system (Weber and Flugge, 2002; Weber et al., 2004) and unknown transporters in the peroxisomal membrane. For sake of simplicity, the conversion of glycine to serine, and the downstream reaction sequence leading to the conversion of serine to 3- phosphoglycerate is omitted. (1) plastidic glycolate transporter; (2) peroxisomal glycolate transporter; (3) plastidic glutamate/malate translocator; (4) plastidic cr-KG/malate translocator; (5) peroxisomal glutamate transporter; (6) peroxisomal a-KG transporter; (7) peroxisomal glycine transporter", (8) mitochondrial glycine transporter. The figure in this dissertation is presented in color. 137 Peroxisomes 02 _ Glt H202 Figure 4.1. Simplified scheme of the current textbook photorespiratory cycle A citrulline-omithine shuttle between plastids and mitochondria To couple mitochondrial ammonia assimilation with plastidic glutamate biosynthesis, Taira et a]. (2004) have suggested a citrulline-omithine shuttle between chloroplasts and mitochondria (Figure 4.2) and have supported this concept by demonstrating that isolated leaf mitochondria are able to catalyze the glutamine- and ATP-dependent conversion of ornithine to citrulline. Ornithine-citrulline transporters are well characterized in yeast and mammals. For example, two closely related human omithine-citrulline transporters have been cloned and functionally characterized; they transport arginine, lysine, ornithine, and citrulline by exchange and unidirectional mechanisms (F iermonte et al., 2003). Less is 138 known about mitochondrial ornithine-citrulline transporters in plants. The closest relatives of mammalian ornithine-citrulline transporters in Arabidopsis are the basic amino acid carriers AthACl (At2g33820) and AthACZ (Atl g79900) and it was shown that AthACl and AthAC2 are able to relieve arginine-auxotrophy of a yeast mutant that is deficient in the mitochondrial ornithine/arginine transporter (Catoni et al., 2003; Hoyos et al., 2003). Recombinant reconstituted AthACl catalyzed the counter- exchange of arginine with lysine, ornithine, arginine, and histidine, but transport of citrulline was marginal (Hoyos et al., 2003). The substrate spectrum of AthAC2 has not yet been tested in a reconstituted system. Hence, the Arabidopsis ornithine-citrulline transporter has not yet been unequivocally identified. Robert Ludwig (1993) also previously showed that isolated Arabidopsis chloroplasts are able to convert externally supplied citrulline into ammonium, bicarbonate, ornithine, and ATP. This indicates the presence of a citrulline transporter in the chloroplast envelope membrane, although uptake of citrulline into chloroplasts has not been directly shown by, e.g., the silicon-oil filtration-centrifugation technique. In summary, evidence derived from experiments with isolated chloroplasts and mitochondria indicates that the components of the proposed citrulline-ornithine shuttle seem to be present and active in leaves but the corresponding proteins remain to be identified. Importantly, the ornithine-citrulline shuttle would not only (indirectly) transfer ammonia fi'om mitochondria to plastids, but also simultaneously carbon dioxide, and thus both products of the glycine decarboxylase reaction that need to be re-assimilated in the chloroplast stroma. Although this concept is appealing, it comes at a price: Whereas fixing one molecule of 02 by RubisCO and salvaging of the resulting 2-phosphoglycerate (2-PG) by a textbook C2-cycle (not including re-assimilation of C02 139 that is lost during the process) costs approximately 9.5 ATP (Canvin and Salon, 1997), two additional ATP would be required in the proposed citrulline-ornithine cycle due to mitochondrial carbamoylphosphate synthase activity (the additional ATP required by GS2 in mitochondria is offset by production of one ATP by carbamate kinase in plastids), thus raising the overall ATP-consurnption by 20% to 11.5 ATP. 140 Figure 4.2. A citrulline/ornithine shuttle between mitochondria and plastids Glycine (Gly), produced in peroxisomes , is decarboxylated and deaminated by GDC in the mitochondrial matrix. In contrast to the scheme shown in Figure 4.], GS assimilates ammonia into glutamate (Glu) directly within the mitochondrial matrix, yielding glutamine (Gln). Carbamoylphosphate (Caer) is synthesized from Gln and C02 by carbamoylphosphate synthetase (CPS), thereby regenerating the ammonia acceptor Glu. Citrulline (Cit) is generated from ornithine (Om) and Caer by Om carbamoyltransferase (OCT) and Cit is shuttled to plastids were it is converted back to Om and CarP by catabolic OCT. Carbamate kinase (CK) generates C02, ammonia, and ATP from Caer. For clarity, the generation of glycine, its conversion of glycine to serine, and the downstream reaction sequence leading to the conversion of serine to 3-PGA is omitted. (8) mitochondrial glycine transporter; (9) mitochondrial omithine/citrulline translocator; (10) plastidic omithine/citrulline translocator. The figure in this dissertation is presented in color. 141 Figure 4.2. A citrulline/ornithine shuttle between mitochondria and plastids A less costly alternative — the glutamate-glutamine shuttle Alternatively, a glutamate/glutamine shuttle between mitochondria and plastids can be envisaged that is compatible with ammonia assimilation by GS2 in mitochondria and that does not require more ATP than a textbook C2-cycle (Figures 4.1 and Figure 4.3). In this cycle, ammonia would be assimilated by GS2 in mitochondria, yielding glutamine. Glutamine is exported from mitochondria, imported into plastids, and the amido group from glutamine is transferred to or-KG by GOGAT, yielding two molecules of glutamate. One molecule of glutamate is shuttled back to mitochondria where it serves as substrate for the GS-reaction. The ATP expense for ammonia assimilation would thus be shified from plastids to mitochondria, but would remain otherwise unchanged. This hypothetical 142 glutamine/ glutamate shuttle between mitochondria and chloroplasts would require the presence of glutamine and glutamate transporters in both organelles. The activity of a glutamine-glutamate antiporter has been described in isolated chloroplasts (Y 11 and Woo, 1988); hence chloroplasts are able to take up glutamate in counter-exchange with export of glutamine. Relatively little is known about amino acid transporters of plant mitochondria. It was shown that glutamate enters beetroot and cauliflower bud mitochondria; the uptake required the presence of both phosphate and a permeable dicarboxylic anion (Day and Wiskich, 1977). The activity of a glutamate-aspartate exchanger was demonstrated in isolated pea mitochondria, the partially purified protein was constituted into liposomes, and capable to facilitate the exchange of glutamate for both aspartate and glutamate (Vivekananda and Oliver, 1989). Mammalian mitochondria also possess a mitochondrial glutamate-aspartate exchanger (Palmieri et al., 2001), and in addition related proteins that catalyze the co-transport of glutamate with H+ (or counter- exchange with a hydroxyl ion) (Fiermonte et al., 2002). The mitochondrial aspartate- glutamate carrier from Saccharomyces cerevisiae, unlike its human orthologs, catalyzes both aspartate-glutamate exchange and substrate uniport (Cavero et al., 2003). It is not known whether the plant glutamate-aspartate carrier is able to catalyze glutamate uniport or whether plant mitochondria possess a glutamate carrier. It is also not known whether glutamine can be transported across the membrane of plant mitochondria; however, in analogy to mammalian mitochondria, it is likely that plant mitochondria possess a transport system for glutamine. For example, rat liver mitochondria have a glutamine- transporter that is able to catalyze both the unidirectional flux of glutamine and glutamine-glutamine counter-transport (Indiveri et al., 1998). After reconstitution into 143 liposomes, this protein also catalyzes the counter-exchange of glutamine with glutamate or aspartate, albeit this activity was one order of magnitude lower than glutamine- glutamine counter-transport (Indiveri et al., 1998). In summary, with the exception of the mitochondrial glutamine transporter, all components of a glutamate/glutamine shuttle between mitochondria and plastids have been experimentally verified in leaf cells. In contrast to the above-described Orn/Cit shuttle, though, the Glu/Gln shuttle only provides a vehicle the transport of NHB between both organelles, whereas the transport of carbon dioxide would be dependent on passive diffusion. On the other hand, shuffling NH3 between mitochondria and chloroplasts by the Glu/Gln shuttle requires two ATP less than by the Om/Cit shuttle. Safety first - a bucket chain for intracellular ammonium transport Assimilation of photorespiratory ammonia in mitochondria by GS requires a shuttle system for ammonia transport between mitochondria and plastids. Both, an ornithine/citrulline shuttle (Figure 4.2) or a glutamate/glutamine shuttle (Figure 4.3) could fulfill this function. In both cases, no transport of free ammonia would occur; instead, this toxic metabolite would be transiently stored as amino acids and would be transported safely in this form. Future research will be required to examine whether either of the two alternative shuttles is physiologically relevant (if not both). Likely, the photorespiratory cycle just became more complex, involving even more metabolite transporters than previously assumed (Weber, 2005). 144 Peroxisomes. 02. G., H202 Figure 4.3. A glutamate/glutamine shuttle between mitochondria and plastids In contrast to scheme depicted in Figure 4.1, ammonia generated by GDC in mitochondria is not assimilated into Glu in the plastids, but by mitochondrial GS and the resulting Gln is shuttled to the plastid stroma were it is used for Glu biosynthesis fi'om or- KG by GOGAT. The ammonia acceptor Glu is returned to mitochondria by the same shuttle system. For clarity, the conversion of glycine to serine, and the downstream reaction sequence leading to the conversion of serine to 3-PGA is omitted. (1‘) plastidic glycolate transporter; (2) peroxisomal glycolate transporter; (3) plastidic glutamate/malate translocator; (4) plastidic or-KG/malate translocator; (5) peroxisomal glutamate transporter; (6) peroxisomal a-KG transporter; (7) peroxisomal glycine transporter; (8) mitochondrial glycine transporter; (l l) mitochondrial Glu/Gln translocator; (12) plastidic glutamate/glutamine translocator. The figure in this dissertation is presented in color. 145 REFERENCES Canvin D.T., and Salon, C. (1997) Photorespiration and C02-concentrating mechanisms. In Plant Metabolism (2nd edition) (Dennis, D.T., et al., editors), Addison Wesley Longman. Catoni E., Desimone M., Hilpert M., Wipf D., Kunze R., Schneider A., Flugge U.l., Schuhmacher K., and Frommer WB. (2003) Expression pattern of a nuclear encoded mitochondrial arginine-ornithine translocator gene from Arabidopsis. BMC Plant Biology 3: l. Cavero S., Vozza A., del Arco A., Palmieri, L., Villa L., Blanco E., Runswick M.J., Walker J .E., Cerdan S., Palmieri F., and Satrustegui J. (2003) Identification and metabolic role of the mitochondrial aspartate-glutamate transporter in Saccharomyces cerevisiae. Molecular Microbiology 50: 1257-1269. Coschigano K.T., Melo-Oliveira R., Lim J., and Coruzzi G.M. (1998) Arabidopsis gls mutants and distinct F d-GOGAT genes: implications for photorespiration and primary nitrogen assimilation. Plant Cell 10:741-752. Day D.A., and Wiskich, J.T. (1977) Glutamate transport by plant mitochondria. Plant Science Letters 9:33-36. Douce R., and Neuburger, M. (1999) Biochemical dissection of photorespiration. Orrrent Opinion in Plant Biology 2:214-222. Fiermonte G., Dolce V., David L., Santorelli F.M., Dionisi-Vici C., Palmieri F., and Walker J .E. (2003) The mitochondrial ornithine transporter. Bacterial expression, reconstitution, functional characterization, and tissue distribution of two human isoforms. Journal of Biological Chemistry 278:32778-32783. Fiermonte G., Palmieri, L., Todisco S., Agrimi G., Palmieri F., and Walker J.E. (2002) Identification of the mitochondrial glutamate transporter. Bacterial expression, reconstitution, functional characterization, and tissue distribution of two human isoforms. Journal of Biological Chemistry 277. 19289-19294. Indiveri C., Abruzzo G., Stipani 1., and Palmieri F. (1998) Identification and purification of the reconstitutively active glutamine carrier fiom rat kidney mitochondria. Biochem. Journal 333:285-290. Hoyos M.B., Palmieri L., Wertin T., Arrigoni R., Polacco J.C., and Palmieri F. (2003) Identification of a mitochondrial transporter for basic amino acids in Arabidopsis thaliana by ftmctional reconstitution into liposomes and complementation in yeast. Plant Journal 33: 1027-1035. 146 Lancien M., Martin M., Hsieh M.H., Leustek T., Goodman H., and Coruzzi G.M. (2002) Arabidopsis gltI T mutant defines a role of NADH-GOGAT in the non- photorespiratory ammonium assirnilatory pathway. Plant Journal 29:347-358. Ludwig RA. (1993) Arabidopsis chloroplasts dissimilate L-arginine and L-citrulline for use as N source. Plant Physiology 101:429-434. Melo-Oliveira R., Oliveira LC, and Corn- G.M. (1996) Arabidopsis mutant analysis and gene regulation define a nonredundant role for glutamate dehydrogenase in nitrogen assimilation. Proceedings of the National Academy of Sciences of the USA 93:4718- 4723. Ogren W.L. (1984) Photorespiration: Pathways, regulation, and modification. Annual Review of Plant Biology 35:415-442. Palmieri L., Pardo B., Lasorsa F.M., del Arco A., Kobayashi K., Iijima M., Runswick M.J., Walker J .E., Saheki T., Satrustegui J., and Palmieri F. (2001) Citrin and aralarl are Ca(2+)-stimulated aspartate/ glutamate transporters in mitochondria. European Molecular Biology Organization (EMBO) Journal 20:5060-5069. Somerville CR, and Ogren, W.L. (1980) Inhibition of photosynthesis in Arabidopsis mutants lacking leaf glutamate synthase activity. Nature 286:257-259. Taira M., Valtersson U., Burkhardt B., and Ludwig RA. (2004) Arabidopsis thaliana GLNZ-encoded glutamine synthetase is dual targeted to leaf mitochondria and chloroplasts. Plant Cell 16:2048-2058. Vivekananda J., and Oliver, DJ. (1989) Isolation and partial characterization of the glutamate/aspartate transporter fi'om pea leaf mitochondria using a specific monoclonal antibody. Plant Physiology 91:272-277. Weber A.P.M. (2005) Synthesis, export, and partitioning of the end products of photosynthesis. In The Structure and Function of Plastids (Wise, R.R., and Hoober, J.K., eds), Kluwer Academics Publishers. Weber A.P.M., Schneidereit J., and Voll L.M. (2004) Using mutants to probe the in vivo function of plastid envelope membrane metabolite transporters. Journal of Experimental Botany 55: 123 1- 1244. Weber A., and Fliigge U.I. (2002) Interaction of cytosolic and plastidic nitrogen metabolism in plants. Journal of Experimental Botany 53:865-874. Wingler A., Lea P.J., Quick W.P., and Leegood RC. (2000) Photorespiration: metabolic pathways and their role in stress protection. Philos. Trans. R Soc. Lond. Ser. B-Biol. Sci. 355:1517-1529. 147 Yu J., and Woo, K.C. (1988) Glutamine transport and the role of the glutamine translocator in chloroplasts. Plant Physiology 88:1048-1054. 148 CHAPTER 5 A REVERSE GENETIC APPROACH TO IDENTIFY THE PROPOSED METABOLITE TRANSPORT PROTEINS IN THE PHOTORESPIRATORY PATHWAY‘ 5 Dr. Nicole Linka and Andrea Brautigam contributed to Table 5.3 and Dr. David Gagneul provided the metabolite data for Figures 5.4, 5.5, and 5.7. 149 ABSTRACT In C3 plants, the oxygenation reaction of Ribulose-1,5-bisphosphate Carboxylase/ Oxygenase (RubisCO) produces a significant amount of 2-phosphoglycolate, which must be converted to 3-phosphoglycerate in the photorespiratory pathway. The enzymatic steps of this pathway are remarkably coordinated although they are distributed in the three compartments plastids, peroxisomes, and mitochondria. To cope with the high flux of the intermediates between these compartments during the day, a significant number of metabolite transporters are required. To date, forward genetic screens have identified only a single plastidial transporter DiT2.1/DCT1 (dicarboxylate transporter) as a classical photorespiratory mutant. The mutant is non-viable under ambient air conditions and indistinguishable from wild type under C02 enriched air. To identify additional metabolite transporters, publicly available microarray datasets were screened for putative transporter genes that are co-expressed with fifteen known photorespiratory enzymes. We identified 32 candidate genes and present the initial genetic analysis Of five genes. One mutant shows a delayed pale green phenotype after 14 days in ambient air. The corresponding gene was previously identified as a putative mitochondrial carnitine carrier, designated BOU (Bout de Souffle). However, our results indicate that instead of carnitine, BOU presumably imports Vitamin B6 as an essential cofactor required for the function of the photorespiratory enzyme glycine decarboxylase. 150 INTRODUCTION Photorespiration (PR) and the Calvin-Benson-Cycle are closely linked pathways via the bifunctional enzyme ribulose-1,5-bisphosphate carboxylase/oxygenase (RubisCO) (Ogren, 1984; Tolbert, 1997). Under ambient atmospheric conditions, oxygen (02) competes with carbon dioxide (C02) for the reaction with ribulose-1,5-bisphosphate (RuBP) catalyzed by RubisCO. Carboxylation of ribulose-1 ,5-bisphosphate yields two molecules of 3-phosphoglyerate (3-PGA) that are converted to triose phosphates (TP) using chemical energy in the form of ATP and NADPH provided by the photosynthetic light reactions (Calvin, 1962). However, oxygenation yields only one molecule of 3-PGA and one molecule of 2-phosphog1ycolate (2-PG). The latter molecule represents a metabolic dead end in the chloroplast stroma. It must thus be removed from the chloroplast and recycled to 3-PGA in a complex, highly compartmentalized pathway named photorespiration. In photorespiration, two molecules of 2-PG are recycled to one molecule of 3-PGA performed in a series of enzymatic steps, which are distributed in three different subcellular compartments (Figure l; Reumann and Weber, 2006). 2-PG is hydrolyzed to glycolate in the chloroplast (Somerville and Ogren, 1979; Schwarte and Bauwe, 2007), and metabolized to glyoxylate and further to glycine in the peroxisomes (Volokita and Somerville, 1987 ; Igarashi et al., 2003). Glycine oxidation and serine synthesis takes place in the mitochondria (Somerville and Ogren, 1981; Vol] et al., 2006; Engel et aL, 2007) and serine is shuttled back to the peroxisomes to be deaminated and reduced to glycerate (Murray et al., 1989; Liepman and Olsen, 2001). Glycerate is imported and finally phosphorylated to 3-PGA in the chloroplast (Boldt et al., 2005). The resulting 3-PGA is then further metabolized in the Calvin-Benson-Cycle. In summary, 15] two molecules of 2-PG are converted to one molecule each of, 3-PGA, C02, and NH3. The pathway leads to a net loss of C02 and ammonia and therefore, in addition to a carbon cycle, photorespiration is intertwined with a nitrogen cycle (Ogren, 1984; Keys, 2006). Ammonia (NH3) is released during glycine oxidation in the mitochondria (Keys et al., 1978). It has been widely accepted that NH3 enters the chloroplast and is re- assimilated by the concerted action of glutamine synthetase (GS) and glutamate synthase (GOGAT) (Somerville and Ogren, 1980; Coschigano et al., 1998; Wallsgrove et al., 1987). The product glutamate is guided to the peroxisomes and directly utilized to produce glycine fiom glyoxylate — In this way, a new round of nitrogen cycling is initiated. The flux of NH3 re-assimilation exceeds by far the rate of de novo ammonium ion production and has been demonstrated to be absolutely required under photorespiratory conditions to circumvent toxic glyoxylate accumulation (Sharkey, 1988; Keys, 2006). It has been recemly demonstrated that GS is dually targeted to chloroplasts and mitochondria (T aira et al., 2004). This would allow the re-assimilation of the NH3 into glutamine on location. Still, the enzyme GOGAT is exclusively active in the chloroplast (Coschigano et al., 1998) and glutamine instead of ammonia has to be transported (Linka and Weber, 2005). In another scenario, ammonia could be incorporated into carbamoyl phosphate and subsequently into ornithine yielding citrulline in the mitochondria. Citrulline is broken down in the chloroplast to ornithine and ammonia enters the GS/GOGAT cycle (Ludwig, 1993; Linka and Weber, 2005). The identification of enzymes that are required for conversion of 2-PG to 3-PGA has been one of the earliest and, in retrospect, most successful forward genetic screens in 152 the Arabidopsis thaliana research field (Somerville and Ogren, 1979; Somerville, 2001). Classical photorespiratory mutants are indistinguishable from wild-type plants in C02 enriched air but not viable under ambient air conditions (Somerville and Ogren, 1979). PR mutants significantly contributed to establishing A. thaliana as a model organism for plant biochemical geneticists and plant physiologists (Meyerowitz and Pruitt, 1985). Almost three decades later, the enzyme glycerate-3-kinase as the last enzymatic step in the series was identified on the molecular level and published as the “last unknown enzyme in the photorespiratory cycle” (Boldt et al., 2005). As outlined above, photorespiration is a prime example for pathway compartrnentalization in eukaryotic cells, since enzymes for the carbon and nitrogen cycle are partitioned almost in equal parts in chloroplasts, peroxisomes and mitochondria (Kunze et al., 2002). Despite the excellent progress in understanding the photorespiratory pathway, none of the metabolite transporters navigating the photorespiratory intermediates through the cell are mapped on the genome of A rabidopsis thaliana (Reumann and Weber, 2006). Forward genetic screens identified only one afi'rliated transport system: the malate/ glutamate transporter DiT2.l/DCT1 (dicarboxylate transporter; Somerville and Ogren, 1983; Renne et al., 2003). This exports the amino-donor glutamate fi'om the chloroplast for the glyoxylate: glutamate aminotransferase (GGT) reaction in the peroxisomes as described earlier. The released product a-ketoglutarate (or-KG) has to be shuttled back to the chloroplast and converted to glutamate by the GOGAT reactions (Weber and Flugge, 2002) The list of A. thaliana photorespiratory mutants has been continuously expanded. For example the peroxisomal catalase detoxifies H202 released by the glycolate oxidase 153 (Queval et al., 2007) and malate dehydrogenases sustain the NADH pool via an oxaloacetate/malate redox-system for the reduction of hydroxypyruvate to glycerate (Cousins et al., 2008). The mitochondrial glycine decarboxylase is a multienzyme complex consisting of four different subunits (P protein, H protein, T protein, and L protein) and several cofactors [lipoic acid, tetrahydrofolate (T HF), NAD, FAD, and pyridoxal-phosphate (PLP)] (Douce et al., 2001). A null mutant for the P protein subunit is inviable and reduced GDC activities result in classical PR phenotypes (Heineke et al., 2001; Engel et al., 2007). It has been demonstrated that impaired biosynthesis of the cOfactor lipoic acid (Ewald et al., 2007) or THF recycling in mitochondria causes a decrease in GDC activity (Collakova et al., 2008). Import of malonic acid and pyruvate is a prerequisite for lipoic acid synthesis and the three precursors glutamate, pABA (para- aminobenzoic acid) and pterin have to be provided for THF synthesis inside the mitochondria (W ada et al., 1997; Rebeille et al., 2007). Apparently, primary photorespiratory metabolites as well as a great number of additional compounds have to be distributed within the cell to cope with the high photorespiratory rates timing the day. In this paper we attempted to establish a catalog of metabolite transporter candidates by a reverse genetic approach. Routine transcriptional profiling is providing an enormous data set, which is becoming a valuable tool to associate proteins of unknown ftmction to known pathways. We used the comprehensive system-biology database (CSB.DB) to identify the best 5% of transcripts co-regulated with a set of 15 highly co-expressed photorespiratory genes. Candidates, which are co- expressed with several PR genes and contain at least two predicted hydrophobic 154 transmembrane spans, were selected for genetic analysis to identify classical photorespiratory mutants. MATERIAL AND METHODS Chemicals The radiolabeled compounds [l4C(U)]-arginine, [l4C(U)]-g1utamic acid, [14C(U)]- glycine, [l4C(U)]-L-serine, [N-methyl-14C]-L-camitine, [2-14C]-malonic acid and [1- 14C]-L-ornithine were purchased from Moravek Radiochemicals and Biochemicals (Brea, CA, USA). Sigma-Aldrich (St. Louis, MO, USA) supplied the chemicals and the company Bio-Rad (Hercules, CA, USA) provided the Bio-Beads SM-2 hydrophobic adsorbent and the ion exchange resins Dowex AGl-X8 and AG50-X8. Reagents and enzymes for nucleic acid amplification, purification and cloning were purchased fiom Promega (Madison, WI, USA), Qiagen (Hilden, Germany) and Invitrogen (Carlsbad, CA, USA). Plant Material and Growth Conditions Seeds of the A. thaliana ecotype Columbia-0 and T-DNA insertion lines for At1g54350 (SALK003 891 & SALK069087), At2g42770(SALK118636), At2g48070 (SALK066383) and At3g26710 (SALK055566) were obtained fiom the Arabidopsis Biological Resource Center (ABRC) at Ohio State University (USA). The A. thaliana GABI-Kat line 079D12 was ordered from the Center for Biotechnology (CeBiTec), Bielefeld University (Germany) and the SHMT (shmI) line was previously established in 155 our lab (V 011 et al., 2006). Surface-sterilized seeds were germinated on petri dishes containing half-concentrated MS medium supplemented with 1% (w/v) sucrose and 0.8% agar and plantlets were transferred on soil after two weeks and watered regularly with Hoagland solution. Unless stated otherwise, the plants were kept under a 12 h-light (120 uEinstein per square meter and minute)/ 12 h-dark regime at 22°C in growth chambers with elevated C02 concentrations (0.3%) throughout their complete life cycle. Selection of the transgenic A. thaliana plants was performed with kanamycin (50 ug/mL) or hygromycin (30 ug/mL). Transcript co-response analyses The publicly available comprehensive system-biology database (CSB.DB) (Steinhauser et al., 2004) was used to identify genes transcript, co-expressed with 15 genes known to be involved in the photorespiratory pathway in A. thaliana. AtGenExpress matrices atge0100 and atge0200 comprised of 63 experiments with 12,200 genes of the developmental expression series, and the abiotic stress series (abovegromrd samples) with 13,197 genes, respectively (Schmid et al., 2005), were analyzed. The NASC Arrays database (Nottingham Arabidopsis Stock Centre) with microarray experiments from the matrix nasc02 71 was also included (Craigon et al., 2004). Based on the Spearman’s correlation, the best 5% of co—responding genes from each single gene query were further analyzed with the ARAMEMNON database to identify putative membrane proteins with at least two predicted alpha helix transmembrane spans (Schwacke et al., 2003; Steinhauser et al., 2004). 156 Phylogenetic tree analysis The putative mitochondrial carnitine carrier, designated BOU (Bout de Souffle), belongs to the eukaryotic mitochondrial carrier family (MCF). Homologous sequences of the MCF from yeast and human and Arabidopsis thaliana were obtained from the National Center for Biotechnology Information (NCBI) (mzllwwwncbinlmnihggv). The BLAST program at NCBI (Altschul et al., 1990) identified homologous genes in the genome of the red alga Galdieria sulphuraria (http://genomics.msu.edu/galdieria_/1 using the human and yeast proteins as a query with a cut-off expectation value (E-value) of _<_ 0.0001. The phylogenetic analysis was performed with full-length amino acid sequences of 208 proteins according to Weber et al., 2006. Isolation of homozygous T-DNA insertion lines Genomic DNA was isolated and T-DNA insertions were determined by standard PCR with the following primers: T-DNA specific primer LBal and ML116 were used to amplify fiom the T-DNA left border in SALK lines and the GABI-Kat line, respectively, and the At3g] 8780 (AtACT2)-specific primer pair ML40/ML41 served as a quality control. At1g54350 gene-specific primer pair ML102/ML103, At2g42770 gene-specific primer pair ML104/ML105, At3g26710 gene-specific primer pair ML108/ML109, At2g48070 gene-specific primer pair ML107/ML]20, and At5g46800 gene-specific primer pair ML114/ML115 were designed to analyze the transgenic lines. Primer sequences are shown in Table 5.1. Amplified PCR fi'agments were cloned into the pGEM-T Easy vector (Promega) and the T-DNA insertion sites were verified by sequencing (Research Technology Support Facility, Michigan State University, USA). 157 Homozygotes were further propagated and heterozygous mutants were self-pollinated and analyzed again for homozygous T-DNA insertion. Isolation of total RNA and RT-PCR analysis The method by Chomczynski and Sacchi (1987) was used to extract total RNA from 6- weeks old A. thaliana rosettes. After DNAse treatment, cDNA was synthesized (Superscript II F irst-Strand Synthesis System, Invitrogen). The transcript levels were examined from the putative metabolite transporter BOU (At5g46800) and the control gene AtACT7 (At5g09810) by PCR analysis. The PCR reaction was an initial denaturation step at 94°C for 2 min, followed by 35 cycles of 94°C for 35 s, 58°C for 30 s, 72°C for 2 min, and final extension of 72°C for 3 min and products were visualized on an ethidium brornide-stained, 1% agarose gel. Table 5.1. Primer sequences used in Chapter 5 Primers were used for cloning of heterologous expression (HE) constructs, Reverse- Transcriptase PCR (RT-PCR) expression analyses, complementation studies (COMP) or T-DNA insertion line screens (TDNA). Underlined nucleotides highlight restrictions sites used for cloning. Primer Sequence (5'—3') LBal TGGTTCACGTAGTGGGCCATCG ML 40 GTTGGGATGAACCAGAAGGA ML 41 GAACCACCGATCCAGACACT ML 90 TATAGQAICCAAATGGCGGATGCGTGGAAG ML 91 ATATCTCGAGTTTATCCCAAGCTTGACCT ML 92 TATACflIACQACCATGGCGGATGCGTGGAA ML 93 ATATIQI‘AQATCCCAAGCTTGACCTTGTCA sea asssgggg 158 Table 5.1. (cont’d). Primer Sequence (5'—3') Purpose ML 102 TAGCT CTT CT GATCCACGGAG TDNA ML 103 TCAATGGTGGCATTAACAAGC TDNA ML 104 CCCAACACCCACAAGTAGTTG TDNA ML 105 TTTCC'ITTGAAGCAAGCTGTG TDNA ML 107 AGAGAAAGAAACCAACCTGCC TDNA ML 108 TTCGACCTGAACATCATCTCC TDNA ML 109 TTGGTCTGATTTTATGTCGGG TDNA ML 114 CAAAGGTTTTGATI'TTGGTTTG RTPCR ML 115 CAAACCCC'ITAGCGAACTCTT TDNA ML 116 CCCATTTGGACGTGAATGTAGACAC TDNA ML 120 TTGCCTTAAACCCACGTACC TDNA ML 143 CGCACAATCCCACTATCCTT TDNA ML 164 CAATGCGGTTTTG'ITCACTG RT-PCR ML 165 CCCAAGCTTGACCTTGTCAT RT-PCR ML 166 AAAAGAAGCACCAGCGACTC RT-PCR ML 167 TTCAATGTCCCTGCCATGTA RT-PCR ML 168 TGAACAATCGATGGACCTGA RT-PCR ML 175 TACQAILKECGGATGCGTGGA COMP ML 176 ATM TCCCAAGCTTGACCTTGTC COMP ML 180 CACACICIAQAAGCCGGATCTCAGTG HE ML 181 CACACGAGATATACCCATfiCGGATG HE 159 Complementation of the GABI-Kat line 079D12 For the molecular complementation of the T-DNA insertion line GABI-Kat 079D12 (bad) the hill-length cDNA of BOU (At5g46800) was amplified by PCR with the primer pairs ML175 and ML176 (Table 5.1), cloned into pGEM-T Easy (Promega) and verified by sequencing. The BOU gene was excised with the designed restriction sites Nhe I and Nco 1 and ligated into the plant expression vector pCAMBlA 1302 after removal of the GFP reporter gene with the same restriction enzymes. Homozygous bouI plants were transformed by the Agrobacterium-mediated floral dip method (Clough and Bent, 1998) and several independent lines were selected on MS agar plates containing 20 ug/mL hygromycin. The 35S:BOU construct in the T1 generation was confirmed on the genomic DNA level by PCR with the primers ML143/ML165 and BOU transcripts were analyzed with the primers ML164/ML165 and ML114/ML166. The control transcripts AtACT7 were amplified with the primer ML167 and ML168 (Table 5.1). T2 generations were screened by segregation analysis and homozygous lines were established in the following generation. The standard PCR reaction was an initial denaturation step at 94°C for 2 min, followed by 35 cycles of 94°C for 35 s, 58°C for 30 s, 72°C for 2 min, and final extension of 72°C for 3 min and products were visualized on an ethidium bromide-stained, 1% agarose gel. Antibody production and protein analysis To use BOU as an antigen and for biochemical characterization, the protein was expressed in Escherichia coli strain BL2] Codon Plus (Stratagene, La Jolla, USA). Coding sequence was amplified with the primers ML180/ML181 (Table 5.1) by PCR 160 fi'om cDNA templates. The fiagment was subcloned into pGEM-T Easy (Promega) and via Nde I and Barn HI restriction sites introduced into the pET16b vector (Merck Chemicals Ltd., Darmstadt, Germany). Expression was induced with 1 mM IPT G (isopropyl B-D-l-thiogalactopyranoside) at an OD600 of 0.6 and cells were harvested after two hours. Inclusion bodies containing recombinant BOU protein translationally fused with an N-terminal hexahistidine tag were isolated according to Schroers et al., 1997, and further purified using nickel-affinity chromatography under denaturing conditions (Qiagen). In detail, 100 mL cultures of E. coli cells were harvested, resuspended in 10 mL Buffer A (20 mM TRIS/HC] pH 7.8, 1 mM EDTA. 1 mM PMSF), incubated with lysozyme (1 mg/mL) and burst by a subsequent sonication step. The insoluble inclusion bodies were pelleted by centrifugation (12,000 g, 10 min, 4°C), dissolved by sonication in buffer A and centrifuged again at 1,100 g for 5 minutes. The supernatant (containing the inclusion bodies) was transferred to a new tube, and inclusion bodies were collected by centrifugtion (12,000 g, 5 min, 4°C), and washed three times with a buffer containing 2% (v/v) Triton X-100, 20 mM TRIS/HC] pH 7.8, 0.1 mM EDTA. The inclusion bodies were solubilized in 400 uL SLS buffer [1.67 % (w/v) sodium lauroylsarcosine, 20 mM TRIS/HCl pH 7 .8, 0.1 mM EDTA] and diluted with 800 uL deionized water. The expression and purification steps were colorimetrically confirmed by Coomassie-stained SDS-PAGE and by immunoblot analysis using a monoclonal anti-penta-His antibody (Qiagen, Germany) combined with an alkaline phosphatase (AP)-conjugated secondary antibody (Promega). Cocalico Biologicals Inc. (Reamstown, PA, USA) produced the polyclonal antibodies against the purified BOU in rabbits. Polyclonal antibodies raised against lipoic acid, H- and P-protein subunits of 161 glycine decarboxylase (GDC), and serine-hydroxymethyl-transferase (SHMT) were kindly provided by Prof. Bauwe from the University of Rostock (Germany). For the protein analysis, 5 g of 4-5 weeks old A. thaliana rosettes were harvested and mitochondrial fiactions were prepared as described by Keech et al., 2005. Fifty micrograms of enriched mitochondrial proteins were resolved on a 15% SDS-PAGE, immobilized on PVDF membranes (Bio-Rad) and analyzed with the Amersham ECL (Enhanced Chemiluminescence) Western Blotting detection system with a secondary anti-rabbit antibody conjugated to HRP (horseradish peroxidase, GE Healthcare UK Limited, Buckinghamshire, UK). Protein concentrations were determined with the Pierce BCA protein assay (Pierce Biotechnology, Rockford, IL, USA). Metabolite analysis For time course experiments, mutant and wild type plants were grown for four weeks under high C02 concentrations (0.3% C02) and transferred to ambient air (0.038% CO2). Entire rosettes were harvested prior to and two, four and seven days after the shift. Rosette leaves from control plants, which were kept at 0.3% C02, were sampled in parallel. Experiments were set up in a completely randomized manner. Start of the experiment and sampling were done at midday. Shoots were collected, immediately frozen, and subsequently powdered in liquid nitrogen, and stored at -80°C tmtil firrther processing. Glyoxylate and ammonium were extracted and quantified according to Brautigam et al., 2007. For proteinogenic and non-proteinogenic amino acid analyses, 100 mg fresh tissue were extracted with 700 RL pre-chilled 100% methanol (-10 to - 20°C) supplemented with phenylalanine-d5 20 rtM (internal standard for amino acid 162 profile normalization). Samples were shaken at 70°C for 15 min. After cooling on ice, 700 uL de-ionized water was added to the mixture. Subsequently, 375 uL pre-chilled chloroform was added and the samples were shaken again for 30 min at 4°C. The samples were centrifuged for 10 min at 14,000 g, and the upper, aqueous phase was used for metabolite analysis. Amino acids were characterized and quantified by LC-MS/MS using a Waters Symmetry C18 column (Milford, MA) with a Shimadzu LC-20AD HPLC system (Columbia, MD) coupled to a Waters Quattro micro mass spectrometer (Gu et al., 2007). Amino acid concentrations were calculated by making reference to external calibration curves, and recovery of successive procedures was assessed using the internal standard phenylalanine-d5. Measurement of glycine decarboxylase activity Mitochondria fi'om wild type and ban] mutant plants were isolated as described above (Keech et al., 2005) except that the final mitochondria pellet was resuspended in 500 uL of 50 mM TES/KOH (pH 7 .5), 2 mM EDTA, 5 mM MgC12, 30% (v/v) glycerol and 0.1% (v/v) Triton X-100 to permeabilize the membrane and release the matrix proteins. The 100 RL activity assay contained 50 uL mitochondrial preparation and final concentrations of 50 mM TES/KOH (pH 7.5), 1 mM EDTA, 2.5 mM MgC12, 15% (v/v) glycerol and 0.01% (v/v) Triton X-100, 2 mM B-mercaptoethanol, 0.03 mM pyridoxal- phosphate (PLP), 0.5 mM tetra-hydrofolate (T HF), 1 mM NAD+ and 0.2 mM [l4C(U)]- glycine. The reactions were conducted in an airtight vial. After 15 min the reaction was terminated by the addition of 1 M sulfirric acid (0.5 mL). GDC activity was linear for at least 30 minutes. The released [14C]-C02 was transferred with a stream of nitrogen gas 163 into a new vial filled with 5 mL of a 1 M KOH solution to trap the carbon dioxide as hydrogen bicarbonate (Walker and Oliver, 1986). The released [l4C]-C02 was quantified by liquid scintillation counting (LSC) with a Beckrnan LS6000 counter (Beckman Coulter Inc., Fullerton, CA). Heterologous expression of BOU in Saccharomyces cerevisiae To clone BOU into the yeast expression system pYESZ/NT (Invitrogen) the full-length sequence was amplified from cDNA by PCR as described earlier. The primer pairs ML90/ML91 and ML92/93 (Table 5.1) were used to yield recombinant BOU protein with an N- and C-terminal hexa-histidine epitope. The fiagment obtained with the first primer set was digested with Barn HI and Xho I and cloned into the pYES/NT-A vector. For the C-terminal His-tag fusion, the plasmid pYES/NT-A was digested with Hind III, treated with T4 polymerase (Promega) and subsequently incubated with Xba I. The BOU fragment was processed consecutively with the enzymes Kpn 1, T4 polymerase and Xba I and ligated with the prepared pYES/NT-A vector. Transformed INVScl yeast (Invitrogen) cells were selected and the recombinant protein was expressed and partially purified as described by Linka etal., 2008. Production of the BOU protein was confirmed by immuoblot analysis using an anti-His antibody (Qiagen, Germany) and an AP- conjugated secondary antibody (Promega). Reconstitution into liposomes Two methods were used in this study to functionally reconstitute recombinant BOU protein into liposomes. For the freeze-thaw-sonicate procedure (Palmieri etal., 1995), a 164 2.5% (w/v) Lon-phosphatidylcholine (PC) suspension was prepared in 120 mM TRIS/BC] (pH 8.0) with 30 mM of the substrate or in 150 mM TRIS/HC] (pH8.0). The phospholipids were dissolved by sonication (Branson Sonicator 250, output 2, duty cycle 30%) on ice for three minutes. 50 rtL of the yeast extract or 30 pg of the recombinant, solubilized BOU protein were mixed with the liposome suspension to a final volume of 1 mL. The protein-phospholipid mixtm'e was frozen in liquid nitrogen and thawed at room temperature. To obtain unilamellar and sealed vesicles, the suspension was sonicated with 25 pulses (output 2, duty cycle 30%). Alternatively, in the “cyclic detergent removal procedure” (Kramer and Heberger, 1986), 30-50 ug of the recombinant BOU protein was mixed with 110 uL 10% Triton X-100. A 10% (w/v) PC suspension was prepared in deionized water and 160 RL was added to the . protein/detergent mixture. The concentrated micelles were diluted to a volume of 1 mL with TRIS/HC] (pH 7.8) and the putative substrate to a final concentration of 120 mM and 30 mM, respectively. After an incubation of 15 min on ice, the detergent was gradually removed by passing the mixture twelve times over a column filled with 1 g of a hydrophobic absorbent (Bio-Beads SM-2, Bio-Rad). External substrate was removed by an additional chromatography step using a desalting Sephadex PD-10 column (GE Healthcare), previously equilibrated with 150 mM TRIS/HCl (pH 7.8). Transport assay W: Uptake of 0.1 mM [N-methyl-14C]-L-camitine into proteoliposomes was tested with and without preloading (L-carnitine; 30 mM). After 165 defined time intervals (up to 30 minutes) the vesicles were passed through a PD-lO column to separate the external radiolabeled substrate from the vesicles. Liposomes were collected from the flow-through and the imported [N-methyl-l4C]-L-carnitine was quantified by LSC. Glutamate and malonic acid transmit assay: Liposomes were prepared with 30 mM glutamate, glutamine or malonic acid and tested for the uptake of radiolabeled [l4C(U)]- glutamic acid (0.75 mM) or [2-14C]-malonic acid (0.25 mM), respectively. External radiolabeled substrate was removed by binding to an anion-exchange column (Dowex AGl-X8 Resin acetate form, 100-200 mesh, Bio-Rad), which had been pre-equilibrated with 10 mM TRIS/HCl (pH 7.8) and 140 mM sorbitol. The proteoliposomes-specific radioactivity was quantified by LSC. WW: Putative antiport and uniport transport activities of BOU with the substrates arginine, citrulline and ornithine were investigated. In both cases the experiments were initiated with 0.25 mM [14C]-radiolabeled substrate and stopped by cation-exchange chromatography (AG50W-X8 Resin, 100-200 mesh, BioRad). Prior to the experiment, the hydrogen ion activated resin columns were converted to its lithium form by washing the resin with five volumes of 1 M lithium acetic acid (LiAc). Due to the greater affmity of arginine and ornithine compared to lithium-ions, both substrates could be efficiently removed fi'om the external medium. 9W: As described above, proteoliposomes were produced in the presence of 150 mM Tris/KOH (pH 7.8) or with 120 mM Tris/KOH (pH 7.8) and either 30 mM glycine or serine. A concentration of 0.25 mM [l4C(U)]-L-serine and [14C(U)]-glycine were used to initiate the transport assay. External radiolabeled serine 166 and glycine were removed from the uptake medium by anion-exchange chromatography. Because the affinity of the resin was rather low for these substrates, the hydroxyl ion of the anion-exchange resin AGlW-X8 was replaced with the bufler molecule tricine. Liposomes were eluted from the column with 10 mM Tris/HC] (pH 7.8) and 140 mM sorbitol and quantified by LSC. To study the dependence of glycine and serine import on an electrochemical gradient across the membrane, artificial pH and potassium (K+) gradients were created in the presence of the ionophores nigericin (0.1 pig/mg PC) and valinomycin (1.5 rig/mg PC), respectively, as described by Indiveri et al., 1997. Alternatively, internal or external pH values were adjusted with the buffer compormd MES/KOH (pH 6.0) to analyze the effect of serine and glycine transport by a pH gradient across the liposomal membrane. Giant liposomes preparation and Patch-Clamp recording Recombinant BOU protein was produced in E. coli and inclusion bodies purified as described above. 100 uL of solubilized protein was diluted with 1.4 mL resin buffer [0.1 M sodium phosphate buffer (pH 8), 0.3 mM sodium chloride, 0.0] mM irnidazole, 2% (v/v) Triton X-100]. Nickel-NTA agarose (Qiagen) was added (100 uL) and gently shaken for one hour at 4°C. The agarose beads were pelleted (1,000 g, 0.5 min, 4°C), the supernatant discarded, and the beads were washed five times with resin buffer. The protein was eluted with 1 mL resin buffer containing 250 mM imidazole. The initial mixture for functional reconstitution contained 10 ug protein, 0.25% (v/v) Triton X-100, 0.35% (w/v) PC, 0.1 M KC] and 0.02 M MOPS/KOH (pH 8). Proteoliposomes were prepared by hydrophobic chromatography (Kramer and Heberger, 1986). To yield giant 167 liposomes suitable for patch-clamp measurements, the vesicles were centrifuged at 100,000x g for 1 hour at 4°C and resuspended in 40 RL MOPS/KOH (pH 7.4) containing 5% (v/v) ethylene glycol. Half of the liposomes were partially dehydrated on a microscope slide in a desiccator containing anhydrous CaC12. After 3 hours the proteoliposomes were rehydrated with 20 uL 0.1 M KC1 in a humidified Petri dish overnight to form giant liposomes (Criado and Keller, 1987; Keller et al., 1988; Delcour et al., 1989; Riquehne et al., 1990). Patch-pipettes were prepared fi'om 1.5 mm borosilicate glass with a P-87 F laming/Brown mircopipette puller (Sutter Instrument Company, Novato, CA). A Nikon Eclipse TE2000-S microscope (Nikon Instruments Europe B.V., Amstelveen, The Netherlands) and the Eppendorf PatchMan NP2 micromanipulator (Eppendorf AG, Hamburg, Germany) was used to achieve gigaohm seal formation with the giant liposomes. Patch-clamp recording was performed using the HEKA EPC10 patch clamp amplifier combined with the PatchMaster software (HEKA Instruments Inc. Bellrnore, NY). RESULTS Putative metabolite transporter co-expressed with the photorespiratory pathway The publicly available Comprehensive System-Biology DataBase (CSB.DB) (Steinhauser et al., 2004) provides a statistical tool to scan transcript profiles fi'om the NASC’s Affymetrix facility (matrix nasc02 71; Craigon et al., 2004) and from the AtGenExpress consortium (matrix atge0100 and atge0200, Schmid et al., 2005) for co-expressed genes in the model organism A. thaliana. In particular, we were interested in the enzymes 168 participating in the recycling of 2-phosphog1ycolate to 3-phosphoglycerate. We showed that with the exception of glycerate kinase (GLYK), catalase (CAT2), and the subunits L1 and H2 of the glycine decarboxylase, all known photorespiratory genes form a robust cluster of 15 transcripts (data not shown) and thus represent a highly co-regulated metabolic pathway. We used the CSB.DB platform to retrieve additional candidates that co—respond with the strong gene cluster. Based on the Spearman's rank correlation coefficient (SRCC) the best 5% of each of the 15 individual gene queries from each of the three microarray matrices (atge0100, atge0200, and nasc02 7 I ) were screened for hydrophobic membrane proteins with at least two putative alpha helix transmembrane regions using the plant membrane database ARAMEMN ON (Schwacke et al., 2003). These proteins represent candidate metabolite transporters, which could enable the high flux of metabolites between the subcellular compartments during photorespiration. Initially, 162 different hydrophobic proteins were identified in the top 5% of positively co-regulated transcripts. 32 out of the 162 hydrophobic genes Show at least 30 out of the 45 possible co-responses with the fifteen query genes (Table 5.2) in the three microarray matrices, and are summarized in Table 5.3. 169 Table 5.2. Co-expressed photorespiratory enzymes from Arabidopsis thaliana Fifieen out of twenty known photorespiratory genes are highly co-regulated. AGI Gene Description Localrz' ation Reference At5g36700 2-Phosphoglycolate Phosphatase Chloroplast Schwarte and (PGLP l) Bauwe (2007) At5g35630 Glutamine synthetase Chloroplast Somerville et a1. (GS2/GLN2) (1980) At5g12860 2-Oxoglutarate:malate antiporter Chloroplast Tanigushi et a1. (DiTl/OMT 1) (2002) At5g64290 Glutamatezmalate antiporter Chloroplast Renne et a1. (DiT2.l/DCT1) (2003) At3g14420 Glycolate oxidase Peroxisomes Volokita and (GOX2) Somerville (1987) At1g23310 Gluzglyoxylate aminotransferase Peroxisomes Igarashi et a1. (AOATl/GGAT) (2003) At2gl3360 Ala/Ser:glyoxylate Peroxisomes Liepman et a]. aminotransferase (AGT] /SGT 1) (2001) At] g68010 Hydroxypyruvate reductase Peroxisomes Murray et a1. (HPR) (1989) At2g35370 Glycine decarboxylase Mitochondria Bauwe et a]. H protein (GDHl) (2003) At1g32470 Glycine decarboxylase Mitochondria Bauwe et a]. H protein (GDH3) (2003) At1g48030 Glycine decarboxylase Mitochondria Bauwe et a]. L protein (LPD2) (2003) At4g33010 Glycine decarboxylase Mitochondria Engel et a]. P protein (GDPl) (2007) At2g26080 Glycine decarboxylase Mitochondria Engel et a]. P protein (GDP2) (2007) At1g11860 Glycine decarboxylase Mitochondria Bauwe et a]. T protein (GDTl) (2003) At4g3 7930 Serine hydroxymethyltransferae Mitochondria Vol] et a1. (SHMT 1) (2006) 170 Table 5.3. Putative transporter genes, which are transcriptionally co-regulated with the photorespiratory pathway The comprehensive system-biology database was used to identify the top 5% of genes, which are co-expressed with 15 genes of the photorespiratory pathway (Table 5.2). Putative membrane proteins and transmembrane spans were extracted from this list using the ARAMEMNON database (Schwacke et al., 2003). Genes highlighted in bold were firrther tested in this study. Subcellular localization was experimentally demonstrated (Sun et a1, 2008; Haezlewood et al., 2004) for those proteins with an asterisk (*). Candidate Gene annotation Putative Predicted gene transmembrane subcellular spans localization 1 At1g10830 Sodium symporter-related 6 Chloroplast/ Mitochondria 2 At1g22850 Hypothetical protein 4 _5 Chloroplast 3 At] g32080 putative membrane protein (MEPl) 12 Chloroplast“ (Envelope) 4 At1g44920 Hypothetical protein 4 Chloroplast“ 5 At] g5 0730 Hypothetical protein 2 Mitochondria 6 At] g52870 Putative peroxisomal PMPZZ-type Chloroplast/ protein of unknown function 2 Mitochondria 7 Atlg54350 Putative plastidial ABC-type Chloroplast" transporter (AtPMPI) 5 (Envelope) 8 At] g7 8180 Putative mitochondrial carrier 3 5 Chloroplast! protein - Mitochondria 9 At2g42770 Putative-peroxisomal PMP22-type 2 Chloroplast“ protern of unknown function (Envelope) 10 At2g48070 Protein of unknown fimction 3 _4 Chloroplast function 10 12 At3 g2 15 80 Hypothetical protein 6 Chloroplast l3 At3 g25805 Hypothetical protein 7 Chloroplast/ Mitochondria 14 At3g26710 Hypothetical protein 4 Chloroplast“ (Thylakoid) 15 At3g48200 Hypothetical protein 9 Secretory pathway 16 At3 g50685 Hypothetical protein 3 Chloroplast 17 At3 g5 5250 Hypothetical protein 2 Chloroplast/ Mitochondria 18 At3g56160 Putative bile acid:sodium symporter 7 Chloroplast 171 Table 5.3. (cont’d). Candidate Gene annotation Putative Predicted gene transmembra subcellular ne spans localization 19 At3g59780 Protein of unknown function 3 Chloroplast/ Nucleus 20 At3g61870 Hypothetical protein 5 Chloroplast 21 At4g02920 Hypothetical protem 2 Unknown 22 At4g] 641 0 Hypothetical protein 3 Chloroplast/ Mitochondria 23 At4g24090 Hypothetical protein 3 Chloroplast 24 At4g24460 Protein of unknown function 10 Chloroplast 25 At4g35760 Hypothetical protein 4 Chloroplast 26 At5g13720 Hypothetical protein 3 Chloroplast 27 At5g27290 Hypothetical protein 3 Chloroplast/ Mitochondria 28 At5g38660 APE] (Acclirnation of * Photosynthesis to environment) - 2 Chloroplast unknown function 29 At5g42130 Putative mitochondrial carrier 3 6 Chloroplast‘ protein ' (Envelope) 30 At5g46800 Putative mitochondrial carrier Mitochondria“ protein (BOU) 3-6 31 At5 g5 7345 Protein of unknown fimction Chloroplast/ 2-3 . . Mrtochondna 32 At5g64970 Putative mitochondrial carrier 3 6 Mitochondria“ protein - + chloroplast“ 172 Phenotype of T-DNA insertion lines from candidate genes Classical mutants in the photorespiratory pathway grow asymptomatically under elevated levels of carbon dioxide (C02); however, they turn chlorotic under ambient air conditions (Somerville, 2001). To test the hypothesis that the 32 candidate genes are involved in the photorespiratory pathway, T-DNA insertion lines for several candidate genes were established and their ability to grow under ambient C02 levels were evaluated. To this end, homozygous T-DNA exon insertion lines were established for the putative plastidial ABC-type transporter At1g54350 (pmpI-I & pmpI-Z) and the putative peroxisomal PMP22-type protein of unknown function At2g42770 (SALK118636) (Figure 5.1). Germination under elevated C02 levels for ten days and shifting the seedlings to ambient air for two weeks induced the photorespiratory phenotype in the known serine-hydroxymethyltransferase (SHMT) mutant shmI-Z, which was used as a positive control (Figure 5.2; Vol] et al., 2006). In contrast, mutant lines pmpI-I, pmpI-Z and SALK118636 had no obvious grth phenotype compared to A. thaliana wild type Columbia-0 plants (Figure 5.2). Seeds had also been germinated under ambient air conditions and in the presence and absence of sucrose as an external carbon source. Under all conditions tested, the three mutant lines were indistinguishable from the wild type (data not shown). Heterozygous T-DNA insertion lines SALK055566 (At3g26710, hypothetical protein) and SALK066383 (At2g48070, protein of unknown function) were self-fertilized (Figure 5.1) and repeatedly screened for a T-DNA linked homozygous genotype. In both populations, only heterozygous and wild type-like lines germinated and developed normally, independent of the C02 concentration in the growth chamber (data not shown). 173 Seedlings from homozygous T-DNA insertions lines of these genes developed white cotyledons and did not establish any new true leaves (Figure 5.2). An external carbon source in the MS medium or different C02 levels in the growth chamber did not rescue the lethal phenotype (data not shown). The homozygous T-DNA knockout line bouI (Figure 5.1) showed a typical photorespiratory phenotype (Figure 5.3). In ambient air, two week-old boul seedlings displayed a pale green phenotype and were significantly smaller than the wild type Col-0 plants. However, bouI seedlings were indistinguishable from wild type when germinated at 0.3% C02 (Figure 5.3). The C02 dependent phenotype could also be observed after shifting 6 week old plants to ambient air conditions. After 14 days, bouI rosettes displayed a pale green phenotype and enhanced senescence, especially in the older rosette leaves (Figure 5.3). This prompted me to focus on the explanation of the bard phenotype. 174 Figure 5.1 Isolation of A. thaliana T-DNA insertion lines for At1g54350, At2g42770, At2g48070, At3g26710, and At5g46800 Upmr part: Gene structure and relative position of T-DNA insertions. A black box refers to an exon. T-DNA insertion sites are indicated by triangles and arrows indicate the relative positions of the primer for PCR analysis. M Genomic DNA PCR analysis of wild type A. thaliana Col-0 plants (WT) and T-DNA insertion lines. Lane G represents gene specific regions surrounding the T-DNA insertion and lane T shows the PCR reaction with primers specific for T-DNA/ gene flanking region. Lane C refers always to AtACT 2 control gene amplification with primer pair ML40/ML41. (A) PCR analysis of the (At1g54350) gene in pmpI -1 (SALK003891) and pmpI-Z (SALK069087). Gene specific products with primer pair ML102/ML103 are absent in the homozygous pmpI -1 and pmpI-Z mutants and present in WT. Primers LBal/ML103 were used to amplify the T -DNA specific fragment. (B) At2g42770: T-DNA insertion (SALK118636) was verified with primers LBal/MLIOS and absence of the amplified gene fi'agment At2g42770 in the mutant line (ML104/ML105). (C) PCR analysis of the heterozygous T-DNA insertion line SALK055566 (At3g26710). Only plants with amplification of target gene (primer ML108/ML109) and T-DNA flanking region (LBal/ML109) were isolated. (D) PCR analysis of the T-DNA insertion line SALK0663 83 (At2g48070). GenePrimers ML107/ML]20 and T-DNA insertion was verified with primers LBal/ML107. (E) Putative mitochondrial acyl-carnitine carrier protein (BOU) At5g46800. Homozygous T-DNA insertion line boul (GABI Kat line 079D12). ML116fMLl 14 verified T-DNA specific insertion and primer pair ML115/ML114 were used for genomic bou fiagment. 175 A pmp1-1 pmp1-2 IML102 ML102 ‘— ML103 ML103 wr wr pmp1-1 pmp1-2 B Salk118636 c Salk055566 ML1 04 ML1 09 i—i'YI-I— :1.“— 4— <- ML1 05 ML1 08 c r c wr Salk 055566 Figure 5.1. Isolation of A. thaliana T-DNA insertion lines The Figure 5.1 continues on the following page. 176 n Salk066383 Salk 066383 Figure 5.1. (cont’d). 177 E bou1 ML114 + 4- 4- 4- ML ML ML 166 165 115 c r c bout Salk 1 18636 pmp1-2 Salk066383 .> 1. x: Figure 5.2. Growth phenotype of T-DNA insertion lines from the candidate genes At1g54350, At2g42770, At2g48070, and At3g26710 (A) Seedlings from A. thaliana Col-0 (WT), T-DNA insertion lines from Atlg54350 (pmpI -1 , pmpI-Z), At2g42770 (SALK118636) and the photorespiratory mutant shmtI-Z (Voll et a1. 2006) were germinated under elevated C02 concentrations (0.3%) for ten days and the picture was taken two weeks after transfer to ambient air (0.038% C02). (B) Homologous T-DNA insertion lines from gene At3g26710 (SALK055566) and (C) At2g48070 (SALK066383) were sown under elevated C02 concentrations and grown for two weeks. The figures in this dissertation are presented in color. 178 Figure 5.3. Growth phenotype of T-DNA insertion line bouI (A) Seedlings from Arabidopsis thaliana Col-0 (WT) and the T-DNA insertion line bouI were germinated at ambient air (0.038% C02) (upper part) or under elevated C02 concentrations (0.3%) (lower part). (B) Eight-week old WT and ban] plants had been growing under high C02 levels (upper part). In parallel, plants were transferred to ambient air after six weeks (lower part) and photographed two weeks later. The figures in this dissertation are presented in color. 179 ambient air (0.038% C02) elevated C02 (0.3% C02) elevated coz (0.3% coz) s. Shift to ambient air (0.038% 002) Figure 5.3. Growth phenotype of T-DNA insertion line boul 180 Rosette leaves of boul accumulate glycine under ambient and elevated C02 levels For several photorespiratory mutants it had been demonstrated that one or more intermediates of the pathway accumulate after transfer to ambient air (Somerville, 2001). To corroborate the phenotypic analysis of bar], we quantified photorespiratory marker metabolites, such as glyoxylate, glycine, serine, and NH3 in Col-0 (WT) and ban] upon shift from 0.3% C02 to 0.038% C02 (Figure 5.4). The most dramatic change was observed for the amino acid glycine. Within 48 hours the glycine concentration in boa] rapidly increased from 3.43 20.44 nmol per gram fiesh weight to 25.64 11.76 nmol per gram fresh weight (8-fold). Even more dramatic is the comparison between wild type and boul plants: After two days in ambient air, glycine levels in the mutant were l30-times higher than in the wild type (Figure 5.4). Interestingly, even in high C02 conditions, the glycine content in the mutant was 25 — 30 times higher than in the wild type. We observed a constant increase of glyoxylate over the time course experiment under ambient air from 31.61 :4.03 nmol per gram fresh weight to 312.8 :65.8 nmol per gram fresh weight. This was not observed under elevated C02 conditions, where glyoxylate is even less abundant compared to WT. The metabolites ammonium ion and serine (Figure 5.4) and glutamate, glutamine and ornithine (Figure 5.5) did not change dramatically during the first five days of the Shift experiment in WT and haul, respectively. After seven days of ambient air the compounds serine, glutamine and ornithine are slightly higher in the bard plants. Citrulline levels (Figure 5.5) were always higher in bad plants. However, the 10-fold difference in C02 concentration between the growth chambers did not have a large effect on the absolute citrulline levels. 181 Figure 5.4. Concentrations of photorespiratory marker metabolites glyoxylate, ammonium ion, glycine, and serine in bar] leaves Arabidopsis thaliana Col-0 (WT, black bars) and ban] (white bars) plants were grown at elevated C02 concentrations (0.3%) for four weeks. Metabolites were analyzed in plants maintained at elevated C02 levels (left panel) or moved to ambient air (right panel). Samples were taken prior to and two, four and seven days after shifting to ambient air ((1). Results are given as the mean :SD of three biological replicates in nmol per gram fresh weight for glyoxylate and NH4+ or in umol per gram fresh weight in the case of glycine and serine. FW, fresh weight; C02, carbon dioxide. 182 Elevated COZ Ambient air ann 'VUU .wr 2£3oo.Dbou1 22 >5! 53200- go. ’o 100- E ILILILEL C 0d 2d 4d 7d on 2d 4d 7d :E SOC-WT Dbou1 800 IWT Elba” fig soe 600 3L §34oc 400 58 s- 2 200 <2 00‘ 5- e 0 0d 2d 4d 7d 0d 2d 4d 7d E ”um 28 WT Elbou1 E 8‘Db0u1 24 3* 2 l. [’1 a? 6‘ 30 or: 4' 3. O 2- 4- s .11 Jill . I] " o'd 2'd 4'd 7'd " o'tr 21d 4'd 7'd — 1o- EW-wr I-wr E 84:! bou1 8-l:lbou1 I3 £5 5 e '6:— m8 4- 4- T: E 2' 2. H o. o- oo 2d 4d 7d 0d 2d 4d 7d Figure 5.4. Concentrations of photorespiratory marker metabolites glyoxylate, ammonium ion, glycine, and serine in boul leaves 183 Figure 5.5. C02-dependent effect on additional metabolite levels in boul and WT The identical samples as described in Figure 5.4 were analyzed and glutamate, glutamine, ornithine, and citrulline levels were determined. Concentrations are shown as the mean :SD of three biological replicates in nmol per gram fresh weight (F W). C02, carbon dioxide. 184 Elevated COZ WT 1:1 bou1 1c Glutamate [nmol per gram FW] e 9’ or _ 0d 2d 4d 7d IWT Dbou1 .3 G Glutamine [nmol per gram FW] 0d 2d 4d 7d Ornlthlne r Citrulline 0. 0d 2d 4d 7d Ambient air IWI' Elbou1 1C 3.5? E 2‘ l | 0' 0d 2d 4d 7d 1c 84 ever Ell Egg 0d 2d 4d 7d B 0.0.. - WT 0-04‘ I: bout 0,03- 0.02' 0.01- i 0d 2d 4d 7d 0.5 I WT 0.4- 1:1 mm 0.3. 0.2- 0.1‘ 0d 2d 4d 7d Figure 5.5. C02-dependent effect on additional metabolite levels in boul and WT 185 Molecular complementation of the boul mutant Although a truncated cDNA of the BOU gene could still be identified by RT-PCR in boul (Figure 5.6), immunoreactive protein was not detectable in the mitochondrial fiaction (Figure 5.9) of boul plants. A boul mutant expressing BOU under the control of the 358 promoter (boul :BOU) (Figure 5.6) germinated and developed normally under ambient air conditions (data not shown) and no significant glycine accumulation was observed (Figure 5.7), showing that the boul mutation causes the C02-dependent phenotype. Glycine decarboxylase (GDC) activity is reduced in boul Elevated levels of glycine and unaltered concentrations of serine and ammonium ions in the boul mutant suggested impaired photorespiratory glycine oxidation due to either a deficiency in glycine uptake into the mitochondria or a lack of GDC activity. Therefore, mitochondria were isolated from high C02 adapted WT and boul plants, the envelope membranes were permeabilized with detergent, and the rate of [U-14C] -glycine decarboxylation to [14C]-CO2 by GDC was determined (Figure 5.8). In boul, the apparent rate of the GDC reaction was only 15.2 1 5.3% of that Observed for wild type. We further tested the abundance of the GDC and SHMT proteins and the presence of the crucial GDC cofactor lipoic acid (Figure 5.9). The GDC H protein subunit, SHMT and lipoylated proteins were not affected in mitochondrial boul fiactions. However, a subtle negative effect on the stability of the PLP-binding GDC subunit P was observed in the boul mutant (Figure 5.9). 186 Figure 5.6. Molecular complementation of boul The boul mutant plants were transformed with BOU cDNA driven by the constitutive CaMV 358 promoter. (A) PCR analysis of genomic DNA (gDNA) from A. thaliana Col- 0 (WT) and a representative complementation line boul/BOU. Lane 1: Amplification of an endogenous BOU fragment with a BOU promoter-specific primer ML114 and primer MLl66. Lane 2: Primer pair ML1 14/ML165 was used to confirm the boul background in the complementation line. Lane 3: 35S-specific primer (ML143) and primer ML166 demonstrate the presence of the complementation construct in the gDNA. (B) RT-PCR fragments from boul and a representative complementation line. Lane 1: cDNA amplification with primer pair ML164/ML165 demonstrates restored transcript ofBOU. Lane 2: Tnmcated BOU cDNA still present in the boul mutant. RT-PCR was done with primer ML164 and ML166. Lane 3: RT-PCR control with AtACT7 (actin) specific primer set m167M168. 187 A gDNA B RT-PCR ML114 ML143 mm —> z -> -> -I —— 4- 4— 35S promoter ML166 ML165 1 2 3 WT” bou1/ BOU Figure 5.6. Molecular complementation of boy] 188 03 O E 501 IL a, E 4 _, I wr .E 3, 3.. D bou1 g h I bou1lBOU #1 5 g s- l I bou1/BOU#2 _. m bou1lBOU #3 O 4- E .3 2- Elevated C02 Arn-bient air Figure 5.7. Complementation of the bard mutant restores glycine concentrations to wild type levels A 35S:BOU construct was introduced into the bou1 mutant line and several independent homozygous complementation lines (bou1/BOU) were established. A fraction of high C02 adapted wild type (WT, black bars), boul (white bars) and bou1/BOU (grey bars) plants were shifted to ambient air and glycine concentrations were determined after seven days. Glycine concentrations are given as the mean :SD of three technical replicates in nmol per gram fresh weight. C02, carbon dioxide; FW, fresh weight. 189 .9. N (TI 1 00- GIN ?‘l‘ Glycine decarboxylase activity [% of WT activity] N 9' Ill Boiled WT bou1 ‘E Figure 5.8. Mitochondrial glycine oxidation is impaired in bou1 Arabidopsis thaliana Col-0 (WT, black bar) and ban] (white bar) plants were grown under elevated C02 concentrations. Mitochondria were isolated fi'om four week-old leaves and permeabilized with the detergent Triton X-100. An aliquot was incubated with 0.2 mM [U-14C]-glycine for 15 min and the released [14C]-C02 was determined. As a control, boiled mitochondrial proteins were monitored to rule out non-enzymatic breakdown of glycine. Relative rates are given as mean a: SD. Relative glycine decarboxylase activity (GDC) of 100% refers to 110.29 nmol €02 per gram protein. 190 Mitochondria Spec'fic wr bou1 l"—'_| l—‘l antibody a. , t B BOU _> ‘7 Mitochondria GDC ' '* ”EM“ 3 SP°°ifi° WT bou1 Protein P protein _> " " antibody t—t PDH E2a I - Lipoic Acidla Eilg PDH E2b KGDH E2 ‘P “- <- H protein GDC H protein _) . . ' GDC H protein SHMT +9.5 Figure 5.9. Immunoblot analysis of mitochondrial enzymes (A) A mitochondrial preparation from A. thaliana Col-0 (WT) and two preparations of the bou1 mutant line as described in Figure 6 were analyzed for the subunits P and H of the glycine decarboxylase (GDC), BOU and serine methyltransferease (SHMT). (B) A lipoic acid specific antibody analyzed the presence Of the lipoylated subunits E2a and 2b from pyruvate dehydrogenase (PDH), the H subunit fi'om GDC, and the E2 subunit of the a-ketoglutarate dehydrogenase (KGDH). Protein extracts were separated on a 15% SDS- PAGE, immobilized on a PVDF membrane and analyzed with the protein-specific antibodies and ECL. 191 Search for the substrate of the recombinant BOU protein To investigate the transport activities of the putative carnitine carrier BOU, the protein was expressed in E. coli and S cerevisiae (Figure 5.10). The recombinant transport proteins were tagged with a N-terminal 6x His-epitope for detection and purification from the cell extracts. In addition, a C-terminal His-tagged construct was produced in yeast (Figure 5.10). The three constructs were reconstituted by either the fieeze/thaw method (FT) or the cyclic detergent removal procedure (CDRP) resulting in six different proteoliposome preparations. Each was tested with a number of putative substrates in a homo-exchange (identical substrate inside and outside), hetero-exchange (external substrate different fi'om the internal), or 1miport (no counter-substrate inside) experiment (Table 5.4). Because the substrate carnitine was hypothesized earlier by Lawand et al. (2002), our observation that the bou1 mutant accumulates glycine (Figure 5.4), and the fact that serine is capable to inhibit glycine uptake into isolated pea mitochondria (Y 11 et al., 1983), these three substrates were the primary targets of analysis. However, none of these metabolites was taken up into BOU-reconstituted liposomes with significant rates (data not shown). Electrochemical gradients across the liposomal membrane had no positive effect on the negligible glycine or serine transport activity catalyzed by recombinant BOU (Table 5.5). In a search for other candidates, we next employed a phylogenomics approach. The mitochondrial carrier family (MCF) is widely distributed in eukaryotic organisms. A maximum-likelihood phylogenetic tree was generated with the MCF members from A. thaliana, yeast, human, and the red algae Cyanidioschyzon merolae and Galdieria sulphuraria (Figure 5.11). BOU belongs to a sub-clade of the MCF in which all 192 functionally tested members transport several amino acids (summarized in Table 5.5). We hence tested uptake of these substrates into liposomes reconstituted with BOU. Neither the proteinogenic amino acids glutamate, glutamine, arginine, nor the non- proteinogenic amino acids ornithine, citrulline were accepted as a substrate. Since GDC activity is diminished in the bou1 mutant, it is reasonable to hypothesize, that BOU could be involved in the supply of GDC cofactors or precursors for cofactor biosynthesis, which are needed in the mitochondrial matrix. Table 5.6 lists putative substrates, which likely have to be transported across the mitochondrial membranes to provide the biological active cofactors PLP, NAD, FAD, THF and lipoic acid for glycine oxidation (Douce et al., 2001). Except for malonic acid, none of the compounds are available in a radiolabeled form. Thus, flux analysis in a reconstituted liposome system cannot be applied to these substrates. Hence, a method had to be established that permits the screening on non-radiolabeled BOU substrates. A promising approach is the patch-clamp technique, which however requires reconstitution of BOU into giant liposomes that can be visualized in the light microscope. To this end, E. coli expressed BOU protein was purified to homogeneity by Nickel-NTA chromatography (Figure 5.10) and reconstituted into liposomes. Giant liposomes were then produced from small, unilamellar liposomes by a dehydration/rehydration step. Using these giant liposomes, pipette-membrane seals could be reproducibly obtained with a resistance of 1-4 giga ohm, which is suitable for patch-clamp recording. Figure 5.12 shows that BOU reconstituted into giant liposomes has voltage-dependent ion channel properties with discrete ion current levels between the applied voltages +100 mV and +200 mV. Seals with resistances of 5-10 GS2 were also achieved with protein-flee liposomes, however no voltage-dependent ion channel 193 activities could be detected between membrane potentials of -140 mV and +200 mV (data not shown). This experimental set-up enables us to test the effect of putative substrates on the channel activity of BOU. The current hypothesis is, that the correct substrate has a high affinity for its binding site and will thus block the voltage-dependent ion channel activity of the BOU transporter, which would be detected by a decreased open probability of the channel. The substrates in Table 5.4 and Table 5.6 are under investigation. 194 PDVF SOS-PAGE membrane kDa 1 2 1 2 85— ‘ 49— -. .. 26— B PDVF SOS-PAGE membrane 85—“? v 49" law may w zs—r‘ Figure 5.10. Heterologous expression and purification of BOU protein in E. coli and S. cerevisiae (A) Inclusion bodies of BOU producing E. coli cells (lane 1) and Ni-NTA purified BOU protein (lane 2) were separated by SDS-PAGE and stained with Coomassie Blue. The proteins of an identical gel were blotted on a PDVF membrane and the N-terminal His- tagged BOU was verified with an anti-His antibody. (B) Membrane proteins from yeast cells were separated on a 12% SDS-polyacrylamide gel and expression of N-terminal- His-tagged (lane 1) and C-terminal-His-tagged (lane 2) recombinat BOU protein was demonstrated with an anti-His antibody. Protein extracts from yeast cells, which had been transformed with an empty expression vector are shown in lane 3. 195 Table 5.4. Systematic search for the BOU substrate N-terminal His-tagged BOU protein was expressed in yeast and E. coli. In addition, BOU was also produced with a C-terminal fused His-tag in yeast and tested. The purified protein was reconstituted into liposomes by the freeze/thaw (FT) procedure or by the cyclic detergent removal procedure (CDRP) as described in “Material and Methods”. Radiolabelled . Functional external Liposomgtplreloaded Recognblipant pill-loteln reconstitution substrates ‘1’ essed procedure :[14C1' :mmfine Y & E r FT & CDRP i tine None east . co 1 Glycine {MC} Serine Y & E 1' FT & CDRP Gl cine ea“ "'0' , y None 4 Serine [1 91' Glycme Yeast & E. coli FT& CDRP Senne None [14C] Glutamate ' Glutamine Glutamate None Yeast FT [14C] Glutamine f Glutamate Yeast FT Glutamlne None Arginine [14C]- Omithine None Omithine 14C _. Citrulline OE'nithine Arginine Yea“ FT None [14C]- Malonic acid . M alonic acid None Yeast & E. colt FT & CDRP 196 Table 5.5. Presence of a membrane potential does not stimulate glycine or serine transport by BOU The ionophore valinomycin induced a potassium (K+) diffusion potential gradient to analyze its influence on glycine or serine transport To generate an artifical pH gradient with the K+/H+ exchanger nigericin, the internal (in) medium (outward K+ directed gradient) or the external (out) medium (inward K+ directed gradient) was not buffered (NB), respectively. . . K+ (in) / Ll somes H rn/ lpo . p K" (out) Valinomycin N igericin pre oaded Wlth pH out i “,li . . 8 / 8 l / l Glycme, serine, none 8 / 8 50/ 1 8 / 8 l / 50 . . NB / 7 1 / l Glycrne, serine, none NB / 7 50/ 1 , . 7 /NB 1 / 1 Glycme, senne, none 7 / NB 1 / 50 . _ 8/6 Glycme, serine, none 6/8 197 Table 5.6. Several cofactors must be provided for the mitochondrial glycine decarboxylase (GDC) For cofactor biosynthesis inside the mitochondria, precursors have to be imported If a cofactor biosynthesis occurs outside the mitochondria, the biologically active form must be provided from the cytosol. All steps require protein-mediated transport steps across the inner envelope mitochondrial membrane. Abbreviations: PLP, pyridoxal 5-phosphate; SAM, S-Adenosylmethionine; NaMN, nicotinate mononucleotide; pABA, para- aminobenzoic acid. GDC Co factor Final step cofactor Predicted transport processes subunit biosynthesis across mitochondrial membrane PLP P protein PLP Cytosol Pyridoxine Pyridoxal Malonic acid . . Pyruvate rtftlein 1:12:30 Cytosol “Sulfur” donor (cysteine, SAM) p CoenzymeA (Valine, ketopantoate, CoA) T protein THF Mitochondria Pterin, pABA, glutamate L protein FAD Cytosol FAD Quinolinate L protein NAD Mfioygi‘t’lld/fia NaMN ‘ NAD 198 AthAC1 (At2933820) YMC1/YPRO58W YMC2/YBR104W Gs20540.1 AtBOU (At5946800) | '— HsORNTZ/SLC25A2 HsORNT1/SLC25A15 HsCAC/SLC25A20 HsC14orf68/NP997000 AthACZ (At1g79900) ORT1IARG11/YOR1BOC CRC1/YOR1OOC Figure 5.11. BOU relevant branch from the maximum-likelihood unrooted tree of the mitochondrial carrier family (MF C) MCF sequences from yeast, plants and human were obtained from NCBI 1hng/wwwncbi.nlm.nih.gov[). The red algae sequences were extracted from the Galdieria sulphuraria (http://genomics.msu.edu/galdieria) and Cyanidioschyzon merolae (httns/l’ ‘ hinl s H-tokvo.ac.ip) genome project database. Protein annotations from yeast, human, plants and red algae are written in brown, blue, green and red, respectively. Bootstrap values are pointed out at the corresponding nodes. The figure in this dissertation is presented in color. 199 v [mV] C3 wwwele - WWW it“ __________________________ —————— fiTH-——LJLJe‘w-———-—————————————————- . ' ewwefifw «W ------ +200 C5 MWMMW++ ———————————————————————————————— (:2 "L ------ ”V'mfwmw ww+180 J. r... 1....1 “W ———————— 3—14——TpL————————L—-———— C4 —————————— ___..r ———————————————————————— C4 ' T7‘7“"""""'1+160 ........................ 1"-.'.".'- Weevil ca ........................ J ....... an-.." ”Titiamv‘v‘liwi'fif" N.v.v'.w..|..mwiilir .. fili‘u: L41? quym m?» _..__'._,_ _____ +W ._"'“‘ AMT". _____ Memw-__q_.w+l40 _______________ L..- ___L,__________..___~u_ C2 _________________:i _________________ ext. 00080 L- -L 4‘; 1 H A m A A. n A. . ~1 _ ‘4 - v... .. re . a“: , . r . -‘ Y , 1,. w e —b'-hvw¢1-|o‘qu—apfirb~m 0.49.»...Am-WM-J.‘ unatvon¥ t .---I - *AA ‘4-—~ ‘4- A— -‘ ‘__fl“u~”~d*’+fi-*’WM‘W.TI~N~‘V mural—m +40 A A r— ,_ .— mucv-M.u-_u_.s-. ---— 8...... 0.. — ~.I~.~..».u.--.— Mn..."— .._...-.._. _~..~.A — .- v u... .- ~--~n-‘-_.o~.~ “II—u. mM.—.0 m.-—- ...-¢ 0 --...~_-.--—-—.--c.‘_-...~~. on. .— ,__,... --. . - . .. ». .0 ~.—.—_#. ---.._.._—.— --- ~.- -- --——. . -....-..._ u. .. _,__,.'~____ *u ..__. ---.......--.. _ ....——~—~— »¢-~.-— pen-so .W a' v ‘ v -— a ,. W. A A w-w W'»~w-' — w . " v . v—L “‘7 v _— V v— . fl", . ‘fid~~v‘-~um~*vw-qew-—-rw-wwa m‘vmflfinvwwmmm-aww- - 8 0 5......“ we..." ~~m~rnmmopaa unwifiwww 1 sec Figure 5.12. Reconstituted BOU protein exhibits ion channel properties Purified BOU protein was reconstituted into giant liposomes. Single channel currents were recorded in a symmetrical 100 mM KCl buffer between negative (-80 mV) and positive (+200 mV) pipette potentials (V). Currents levels of open state (0) and closed state (C) channels are shown. 200 DISCUSSION Identification of membrane proteins co-expressed with photorespiration The carbon and nitrogen cycles of the photorespiratory pathway are distributed in three different compartments and acting jointly to recycle two molecules 2-phosphoglycolate to 3-PGA (T olbert, 1997). A prerequisite to facilitate the traffic of the photorespiratory intermediates are numerous shuttles inserted into the plastid, peroxisomal and mitochondrial membranes. As an alternative to identify these transporters by forward genetics, the CSB.DB resource was screened for hydrophobic membrane proteins that are highly co-regulated with known photorespiratory genes (Steinhauser et al., 2004). As commented by Wei et al. (2006) the photorespiration is embedded in the expression pattern of genes from the photosynthetic apparatus and the Calvin-Benson-cycle. Several components of the photosynthetic electron transport chain were indeed found to be repeatedly co—expressed with photorespiration, since the light reaction overlaps temporarily and spatially with the photorespiration and the initial substrate ribulose-1,5- bisphosphate and final product 3-PGA are directly connected to the Calvin-Benson-cycle. The expression profiles of several metabolite transporters were also tested. The Calvin- cycle associated triose-phosphate/phosphate transporter, which exports recently fixed C02 into the cytosol, was retrieved from our CSB.DB search. Other transporter genes such as the mitochondrial basic amino acid transporter (AthACl) or the peroxisomal nucleotide transporter AtPNCl do not cluster with the query genes and point out the applicability to screen for photorespiratory candidate transport proteins. To prioritize the candidates for reverse genetic analysis, already characterized carrier proteins and thylakoid-localized membrane proteins were neglected. Despite the progress in proteomic 201 analysis of all three organelles and GFP studies (Miller et al., 2001; Reumann et al., 2004; Sun et al., 2008), the localization of more than two-third of our candidate genes (23 out of 32) is predicted solely by in silico targeting programs. Seven out of the nine candidate proteins were found by Arabidopsis chloroplast proteomic studies (Sun et al., 2008), one is postulated to be dual targeted to mitochondria and plastid (At5g64970) (Dunkley et al., 2006) and the remaining candidate is inserted into the mitochondrial membrane (At5g46800) (Lawand et al., 2002). Photorespiratory transport systems across the plastid inner envelope membrane Besides the two-translocator shuttle for 2-OG/ glutamate recycling in the plastid (Renne et al., 2003), a single transport protein for glycolate/glycerate antiport or their uniport was demonstrated on intact chloroplasts (Howitz and McCarty, 1991). This flexibility in the transport mode (i.e. uniport vs. antiport) is important, because for every recycled and finally imported glycerate two molecules glycolate have to be exported fi'om the plastid (Figure 1). The nitrogen cycle would require a transport system for the ammonium ion, glutamate/glutamine or ornithine/citrulline performed by three to maximal five proteins across the chloroplast inner envelope membrane (Linka and Weber, 2005). Transport of glutamine and citrulline was demonstrated on intact chloroplast - its relevance in photorespiration, however, is speculative (Y u and Woo, 1988; Ludwig, 1993). Eighteen candidate proteins have a robust plastidial computer-predicted targeting and seven of them were in addition identified by proteomics. T-DNA insertion lines for four plastid-targeted genes did not show a conditional chlorotic phenotype under ambient air conditions (Figure 5.2). Homozygous mutant lines from At2g48070 and At3g26710 202 lack pigments and seedlings did not develop beyond the cotyledon stage. At2g48070 has been identified by a recent proteomics study as a protein targeted to thylakoid membranes (Zybailov et al., 2008). These results direct its function towards the biogenesis of the photosynthetic apparatus. In the case of SALK line 055566 (At3g26710) only heterozygous mutant lines fully developed Exon-insertion lines for Atlg54350 and At2g42770 grew normally in ambient air (Figure 5.2). Atlg54350 (AtPMPl) is annotated as one of 129 ABC transporter proteins in A. thaliana (Sanchez-Fernandez et al., 2001). At2g42770 is related to the peroxisomal pore-forming 22 kDa membrane protein PMP22 from human, however, not further characterized yet (T ugal et al., 1999; Visser et al., 2007). Seven additional proteins are annotated as PMP22 or PMP22-related proteins and targeted to plastid, peroxisome and mitochondria in the genome of A. thaliana (ARAMEMNON database; Schwacke et al., 2003). At this point it cannot be ruled out that the PMP22 members, AtPMPl or gene product At3g26710 are involved in photorespiration and it is tempting to speculate that genetic analyses are masked by functional redundancy or by the severity of the loss of function mutant, respectively. The majority of photorespiratory mutants tln'ive under elevated C02 concentrations. This is not true for mutants deficient in the enzyme glycine decarboxylase (Engel et al., 2007). A complete loss of the enzyme activity is lethal and alternative strategies such as inducible RNAi targeted gene silencing might be necessary to aim for novel photorespiratory mutants. 203 The mitochondrial envelope membrane transport protein BOU is required for photorespiratory function BOU (At5g46800) was the only evident mitochondrial transport candidate that was retrieved from the co—expression data (Table 5.3). The gene belongs to the mitochondrial carrier family (MCF) and was previously published as a putative carnitine/acylcarnitine carrier providing alternatively acetate as a carbon and energy source to the mitochondria (Lawand et al., 2002). Under long-day conditions, the seedlings were not capable to establish photoautotrophic growth. In this study, we isolated a second mutant allele (bou1) and demonstrated that “non-photorespiratory” conditions can completely rescue the seedling’s development (Figure 5.3). The classical pale green phenotype was not apparent until the 14th day, which might explain why Somerville & Ogren (1979) did not isolated an EMS mutagenized plant with a deficient BOU gene. Their genetic screen concentrated on visible chlorotic effects after four days. Importantly, metabolic analysis of boy] demonstrated a massive accumulation of glycine already under 0.3% C02 and even more dramatic increase at the second day at ambient air (Figure 5.4). This suggests that metabolic profiling of photorespiratory marker metabolites is a powerful tool to identify additional photorespiratory mutants that do not display an apparent phenotype. Glycine import and serine export are the most prominent transport processes during photorespiration in the mitochondrial envelope membrane. Two studies on isolated mitochondria showed that both substrates could interfere with each other for uptake (Walker et al., 1982; Yu et al., 1983). Analogous to the plastid glycolate/glycerate shuttle, a single protein is hypothesized to mediate a glycine/serine transport in a stoichiometry of 2:1. Numerous attempts with distinct experimental conditions failed to 204 show transport activity of glycine, serine, or carnitine with recombinant BOU (Table 5.4 and Table 5.5). Elevated citrulline levels (Figure 5.5) and the phylogenetic relation to basic amino acids transporters (Figure 5.11) directed the analysis to test the hypothesized ornithine/citrulline and glutamine/ glutamate shuttle system without any evident activity. Intriguingly, bou1 exhibits a similar glycine accumulation in leaves as knockout mutants affected in the provision of cofactors lipoic acid and THF for the glycine decarboxylase complex (Ewald et al., 2007; Collakova et al., 2008). The GDC activity was diminished (Figure 5.8) in haul and the PLP dependent P protein subunit was partially degraded (Figure 5.9). The Vitamin B6 derivative PLP is synthesized in the cytosol (T ambasco- Studart et al., 2005) and GDC is the most abundant soluble protein in leaf mitochondria - therefore a major sink for PLP (Sarojini and Oliver, 1983; Douce et aL, 2001; Bauwe and Kolukisaoglu, 2003). Taking both arguments into account, we hypothesize that BOU provides PLP to the mitochondria. Since PLP covalently binds to the P protein subunit and is an integral part of its active site, it is tempting to speculate that a distorted P protein lacking PLP is prone to degradation (Figure 5.9). Since additional enzymes such as amino-transferases are dependent on PLP and a GDC knockout is lethal, an impaired, but not complete lack of PLP provision, has to be proposed at this point. Using the patch- clamp technique we demonstrated that the BOU is functional when reconstituted into liposomes and displays voltage-dependent ion channel activity (Figure 5.12). This suggests that the previously tested metabolites, which were reconstituted with the same procedure, are not the substrates of BOU. Future studies will show if PLP, pyridoxal or pyridoxine can inhibit the voltage-dependent ion channel activity of BOU. Identification of the major mitochondrial glycine import protein will be challenging. The A. thaliana 205 genome predicts at least 53 selective or broad-spectrum amino acids transporters distribute in every plant tissue and in most subcellular compartments including plasma membrane, mitochondria, plastid and peroxisomes (Wipf et al., 2002). Complete loss of glycine import would stall the essential GDC activity and in a broader view the mitochondrial protein synthesis (Engel et al., 2007). For protein synthesis a complete set of amino acids is required and a sophisticated and partial overlapping import system for mitochondria can be postulated. Mutant analysis of promising amino acid transporters combined with metabolite profiling might help to resolve the major route for glycine import. Photorespiratory transport systems in the peroxisomal membrane In comparison to mitochondria and plastids, very little is known about peroxisomal membranes (Reumann et al., 2004; Reumann and Weber, 2006). The isolation of pure peroxisomal membrane fraction for proteomic studies is challenging and a target signal of peroxisomal membrane proteins for computer-based analysis is not deciphered yet (Reumann et al., 2004). A porin-like channel is postulated to facilitate the flux of the mono- and dicarboxylates glycolate, oxaloacetate, malate, 2-OG and glycerate (Reumann et al., 1996; Reumann et al., 1998). Transport of amino acids glycine, serine, or glutamate and the bulky cofactors NAD, NADP, or CoenzymeA remains completely unanswered. We could extract from the co-expression analysis a member of the PMP22- related protein family in A. thaliana. However, an obvious visible phenotype could not be detected in the corresponding mutant. The twelve PMP22-related genes might be promising targets for metabolic profiling and multiple gene knockout analyses. 206 OUTLOOK Despite the overrepresentation of plastid-targeted proteins, co-response analysis is a valuable tool to associate genes of tmknown function with photorespiration. BOU was identified as a new player in the photorespiratory network. An initial genetic analysis of the remaining 27 lines is feasible in the model organism A. thaliana. Detailed metabolite analysis of the mutant plants and a selective transcript profiling of the candidate genes in wild type plants adapting to high and low C02 concentration will add important information to identify missing transporters from the photorespiratory network. 207 REFERENCES Altschul S.F., Gish W., Miller W., Myers E.W., and Lipman DJ. (1990) Basic Local Alignment Search Tool. Journal of Molecular Biology 215:403-410. Bauwe H., and Kolukisaoglu U. (2003) Genetic manipulation of glycine decarboxylation. Journal of Experimental Botany 54:1523-1535. Boldt R., Edner C., Kolukisaoglu U., Hagemann M., Weckwerth W., Wienkoop S., Morgenthal K., and Bauwe H. (2005) D-glycerate-S-kinase, the last unknown enzyme in the photorespiratory cycle in Arabidopsis, belongs to a novel kinase family. Plant Cell 17:2413-2420. Brautigam A., Gagneul D., and Weber A.P.M. (2007) High-throughput colorimetric method for the parallel assay of glyoxylic acid and ammonium in a single extract. Analytical Biochemistry 362: 1 5 l-153. Calvin M. (1962) Path of Carbon in Photosynthesis. Science 135:879-889. Chomczynski P., and Sacchi N. (1987) Single-Step Method of Rna Isolation by Acid Guanidinium Thiocyanate Phenol Chloroform Extraction. Analytical Biochemistry 162:156-159. Clough S.J., and Bent A.F. (1998) Flora] dip: a simplified method for Agrobacterium- mediated transformation of Arabidopsis thaliana. Plant Journal 16:735-743. Collakova E., Goyer A., Naponelli V., Krassovskaya 1., Gregory J .F., Hanson A.D., and Shachar-Hill Y. (2008) Arabidopsis lO-formyl tetrahydrofolate deformylases are essential for photorespiration. Plant Cell 20: 1 818-1832. Coschigano K.T., Melo-Oliveira R., Lim J., and Coruzzi G.M. (1998) Arabidopsis gls mutants and distinct Fd-GOGAT genes: Implications for photorespiration and primary nitrogen assimilation. Plant Cell 10: 741-752. Cousins A.D., Pracharoenwattana 1., Zhou W., Smith S.M., and Badger M.R (2008) Peroxisomal malate dehydrogenase is not essential for photorespiration in Arabidopsis but its absence causes an increase in the stoichiometry of photorespiratory C02 release. Plant Physiology 148:786-795. Craigon DJ., James N., Okyere J., Higgins J., Jotham J., and May S. (2004) NASCArrays: a repository for microarray data generated by NASC's transcriptomics service. Nucleic Acids Research 32:D575-577. Criado M., and Keller B.U. (1987) A membrane fusion strategy for single—channel recordings of membranes usually non-accessible to patch-clamp pipette electrodes. Federation of European Biochemical Societies (FEBS) Letters 224: 172-176. 208 Delcour A.H., Martinac B., Adler J., and Kung C. (1989) Modified reconstitution method used in patch-clamp studies of Escherichia coli ion channels. Biophysical Journal 56:631-636. Douce R., Bourguignon J., Neuburger M., and Rebeille F. (2001) The glycine decarboxylase system: a fascinating complex. Trends in Plant Science 6: 167-176. Dunkley T.P.J., Hester S., Shadforth I.P., Runions J., Weimar T., Hanton S.L., Grifi'm J.L., Bessant C., Brandizzi F., Hawes C., Watson R.E., Dupree P., and Lilley KS. (2006) Mapping the Arabidopsis organelle proteome. Proceedings of the National Academy of Sciences of the USA 103:6518-6523. Engel N., van den Daele K., Kolukisaoglu U., Morgenthal K., Weckwerth W., Parnik T., Keerberg O., and Bauwe H. (2007) Deletion of glycine decarboxylase in Arabidopsis is lethal under nonphotorespiratory conditions. Plant Physiology 144:1328- 1335. Ewald R., Kolukisaoglu U., Bauwe U., Mikkat S., and Bauwe H. (2007) Mitochondrial protein lipoylation does not exclusively depend on the thAS pathway of de novo fatty acid synthesis in Arabidopsis. Plant Physiology 145:41-48. Gu L., Jones A.D., and Last R.L. (2007) LC-MS/MS assay for protein amino acids and metabolically related compounds for large-scale screening of metabolic phenotypes. Analytical Chemistry 79:8067-8075. Heazlewood J.L., Tonti-Filippini J.S., Gout A.M., Day D.A., Whelan J., and Miller A.H. (2004) Experimental analysis of the Arabidopsis mitochondrial proteome highlights signaling and regulatory components, provides assessment of targeting prediction programs, and indicates plant specific mitochondrial proteins. Plant Cell 16:241-256. Heineke D., Bykova N., Gardestrtim P., and Bauwe H. (2001) Metabolic response of potato plants to an antisense reduction of the P-protein of glycine decarboxylase. Planta 212:880—887. Hdwitz K.T., and McCarty H.E. (1991) Solubilization, partial purification, and reconstitution of the glycolate/glycerate transporter from chloroplast inner envelope membranes. Plant Physiology 96: 1060-1069. Igarashi D., Miwa T., Seki M., Kobayashi M., Kato T., Tabata S., Shinozaki K., and Ohsumi C. (2003) Identification of photorespiratory glutamate : glyoxylate aminotransferase (GGAT) gene in Arabidopsis. Plant Journal 33:975-987. Indiveri C., Tonazzi A., Stipani 1., and Pahnieri F. ( 1997) The purified and reconstituted ornithine/citrulline carrier from rat liver mitochondria: electrical nature and coupling of the exchange reaction with H+ translocation. Biochemical Journal 327: 349- 355. 209 Keech 0., Dizengremel P., and Gardestrom P. (2005) Preparation of leaf mitochondria from Arabidopsis thaliana. Physiologia Plantarum 124:403-409. Keller B.U., Hedrich R., Vaz W.L.C., and Criado M. (1988) Single channel recordings of reconstituted ion channel proteins: an improved technique. Pfliigers Archiv European Journal of Physiology 411:94-100. Keys M. (2006) The re—assimilation of ammonia produced by photorespiration and the nitrogen economy of 03 higher plants. Photosynthesis Research 87:165-175. Keys A.J., Bird l.F., Cornelius M.J., Lea P.J., Wallsgrove R.M., and Miflin B.J. (1978) Photorespiratory Nitrogen Cycle. Nature 275:741-743. Kramer R., and Heberger C. (1986) Functional reconstitution of carrier proteins by removal of detergent with a hydrophobic ion exchange column. Biochimica et Biophysica Acta (BBA) - Biomembranes 863:289-296. Kunze R., Frommer W.B., and Flugge DJ. (2002) Metabolic engineering of plants: the role of membrane transport. Metabolic Engineering 4:57-66. Lawand S., Dorne A.J., Long D., Coupland G., Mache R., and Carol P. (2002) Arabidopsis a bout de souffle, which is homologous with mammalian carnitine acyl carrier, is required for postembryonic growth in the light. Plant Cell 14:2161-2173. Liepman A.H., and Olsen LJ. (2001) Peroxisomal alanine: glyoxylate aminotransferase (AGT 1) is a photorespiratory enzyme with multiple substrates in Arabidopsis thaliana. Plant Journal 25:487-498. Linka M., and Weber A.P.M. (2005) Shuffling ammonia between mitochondria and plastids during photorespiration. Trends in Plant Science 10:461-465. Ludwig RA. (1993) Arabidopsis chloroplasts dissimilate L-arginine and L-citrulline for use as N-Source. Plant Physiology 101:429-434. Meyerowitz E.M., and Pruitt RE. (1985) Arabidopsis thaliana and Plant Molecular- Genetics. Science 229: 1214-121 8. Millar A.H., Sweetlove L.J., Giege P., and Leaver CJ. (2001) Analysis of the Arabidopsis Mitochondrial Proteome. Plant Physiology 127 : 171 1-17 27. Murray A.J.S., Blackwell R.D., and Lea P.J. (1989) Metabolism of hydroxypyruvate in a mutant of barley lacking NADH-dependent hydroxypyruvate reductase, an important photorespiratory enzyme activity. Plant Physiology 91:395-400. 210 Ogren W.L. (1984) Photorespiration: Pathways, regulation, and modification. Annual Review of Plant Physiology 35:415-442. Palmieri F., Indiveri C., Bisaccia F., and Iacobazzi V. (1995) Mitochondrial metabolite carrier proteins: Purification, reconstitution, and transport studies. Methods in Enzymology 260:349-369. Queval G., Issakidis-Bourguet E., Hoeberichts F.A., Vandorpe M., Gakiere B., Vanacker H., Miginiac-Maslow M., Van Breusegem F., and Noctor G. (2007) Conditional oxidative stress responses in the Arabidopsis photorespiratory mutant cat2 demonstrate that redox state is a key modulator of daylength dependent gene expression, and define photoperiod as a crucial factor in the regulation of H202-induced cell death. Plant Journal 52:640-657. Rebeille F., Alban C., Bourguignon J., Ravanel S., and Douce R. (2007) The role of plant mitochondria in the biosynthesis of coenzymes. Photosynthesis Research 92: 149- 162. Renne P., Dressen U., Hebbeker U., Hille D., Flugge U.l., Westhoff P., and Weber A.P.M. (2003) The Arabidopsis mutant dct is deficient in the plastidic glutamate/malate translocator DiT2. Plant Journal 35:316—33 1 . Reumann S., and Weber A.P.M. (2006) Plant peroxisomes respire in the light: Some gaps of the photorespiratory C-2 cycle have become filled - Others remain. Biochimica Et Biophysica Acta-Molecular Cell Research 1763: 1496-1510. Reumann S., Ma C., Lemke S., and Babujee L. (2004) AraPerox. A database of putative Arabidopsis proteins from plant peroxisomes. Plant Physiology 136:2587-2608. Reumann S., Maier E., Heldt H.W., and Benz R. (1998) Permeability properties of the porin of spinach leaf peroxisomes. European Journal of Biochemistry 251:359-366. Reumann S., Maier E., Benz R., and Heldt H.W. (1996) A specific porin is involved in the malate shuttle of leaf peroxisomes. In. Portland Press, pp754—757. Riquelme G., Lopez E., Garcia-Segura L.M., Farragut J.A., and Gonzalez-Ros J.M. (1990) Giant liposomes: a model system in which to obtain patch-clamp recordings of ionic channels. Biochemistry 29: 1 1215-1 1222. Sanchez-Fernandez R., Davies T.G.E., Coleman J.O.D., and Rea P.A. (2001) The Arabidopsis thaliana ABC protein superfamily, a complete inventory. Journal of Biological Chemistry 276:30231-30244. Sarojini G., and Oliver DJ. ( 1983) Extraction and partial characterization of the glycine decarboxylase multienzyme complex from pea leaf mitochondria. Plant Physiology 72: 1 94-199. 211 Schmid M., Davison T.S., Henz S.R., Pape U.J., Demar M., Vingron M., Scholkopf B., Weigel D., and Lehmann J .U. (2005) A gene expression map of Arabidopsis thaliana development. Nature Genetics 37:501-506. Schroers A., Kramer R., and Wohlrab H. (1997) The reversible antiport/uniport conversion of the phosphate carrier from yeast mitochondria depends on the presence of a single cysteine. Jomnal of Biological Chemistry 272:10558-10564. Schwacke R., Schneider A., van der Graafi‘ E., Fischer K., Catoni E., Desimone M., Frommer W.B., Flugge U.l., and Kunze R. (2003) ARAMEMNON, a novel database for Arabidopsis integral membrane proteins. Plant Physiology 131:16-26. Schwarte S., and Bauwe H. (2007) Identification of the photorespiratory 2- phosphoglycolate phosphatase, PGLPl , in Arabidopsis. Plant Physiology 144:1580-1586. Sharkey TD. (1988) Estimating the Rate of Photorespiration in Leaves. Physiologia Plantarum 73:147-152. Somerville CR. (2001) An early Arabidopsis demonstration. Resolving a few issues concerning photorespiration. Plant Physiology 125:20-24. Somerville S.C., and Ogren W.L. (1983) AnArabidopsis thaliana mutant defective in chloroplast dicarboxylate transport. Proceedings of the National Academy of Sciences of the United States of America-Biological Sciences 80: 1290-1294. Somerville CR, and Ogren W.L. (1981) Photorespiration deficient mutants of Arabidopsis thaliana lacking mitochondrial serine transhydroxymethylase activity. Plant Physiology 67:666-671. Somerville CR, and Ogren W.L. (1980) Inhibition of Photosynthesis in Arabidopsis Mutants Lacking Leaf Glutamate Synthase Activity. Nature 286:257-259. Somerville C.R., and Ogren W.L. (1979) Phosphoglycolate Phosphatase-Deficient Mutant of Arabidopsis. Nature 280:833-836. Steinhauser D., Usadel B., Luedemann A., Thimm O., and Kopka J. (2004) CSB.DB: a comprehensive systems-biology database. Bioinformatics 20:3647-3651. Sun Q., Zybailov B., Majeran W., Friso G., Olinares P.D.B., and van Wijk K.J. (2008) PPDB, the Plant Proteomics Database at Cornell. Nucleic Acids Research Published online on October 2nd 2008. Taira M., Valtersson U., Burkhardt B., and Ludwig RA. (2004) Arabidopsis thaliana GLN2-encoded glutamine synthetase is dual targeted to leaf mitochondria and chloroplasts. Plant Cell 16:2048-2058. 212 Tambasco-Studart M., Titiz O., Raschle T., Forster G., Amrhein N., and Fitzpatrick T.B. (2005) Vitamin B6 biosynthesis in higher plants. Proceedings of the National Academy of Sciences 102: 13687-13692. Taniguchi, M., Taniguchi, Y., Kawasaki, M.,Takeda S., Kato, T., Sato S., Tabata S., Miyake H., Sugiyama T. (2002) Identifying and characterizing plastidic 2- oxoglutarate/malate and dicarboxylate transporters in Arabidopsis thaliana. Plant Cell Physiology 43:706-717. Tolbert NE. (1997) The C2 oxidative photosynthetic carbon cycle. Annual Review of Plant Physiology and Plant Molecular Biology 48:1-25. Tugal H.B., Pool M., and Baker A. (1999) Arabidopsis 22-kiloda1ton peroxisomal membrane protein. Nucleotide sequence analysis and biochemical characterization. Plant Physiology 120:309-320. Visser W.F., van Roermund C.W.T., Ijlst L., Waterham ER, and Wanders RJ.A. (2007) Metabolite transport across the peroxisomal membrane. Biochemical Journal 401:365-375. Voll L.M., Jamai A., Renne P., Voll H., McClung C.R., and Weber A.P.M. (2006) The photorespiratory Arabidopsis shmI mutant is deficient in Sl-lMl. Plant Physiology 140:59-66. Volokita M., and Somerville CR. (1987) The primary structure of spinach glycolate oxidase deduced fi'om the DNA sequence of a cDNA clone. Journal of Biological Chemistry 262:15825-15828. Wada H., Shintani D., and Ohlrogge J. (1997) Why do mitochondria synthesize fatty acids? Evidence for involvement in lipoic acid production. Proceedings of the National Academy of Sciences of the United States of America 94:1591-1596. Walker J .L., and Oliver DJ. (1986) Glycine decarboxylase multienzyme complex. Purification and partial characterization from pea leaf mitochondria. Journal of Biological Chemistry 261:2214-2221. Walker G.H., Sarojini G., and Oliver DJ. (1982) Identification of a Glycine Transporter from Pea Leaf Mitochondria. Biochemical and Biophysical Research Communications 107:856-861. Wallsgrove R.M., Turner J.C., Hall N.P., Kendall A.C., and Bright S.W.J. (1987) Barley mutants lacking chloroplast glutamine synthetase - Biochemical and Genetic- Analysis. Plant Physiology 83: 155-158. 213 Weber A.P., Linka M., and Bhattacharya D. (2006) Single, ancient origin of a plastid metabolite translocator family in Plantae from an endomembrane—derived ancestor. Eukaryotic Cell 5:609-612. Weber A., and Flugge U.L (2002) Interaction of cytosolic and plastidic nitrogen metabolism in plants. Journal of Experimental Botany 53:865-874. Wei H.R., Persson S., Mehta T., Srinivasasainagendra V., Chen L., Page G.P., Somerville C., and Loraine A. (2006) Transcriptional coordination of the metabolic network in Arabidopsis. Plant Physiology 142:762-774. Wipf D., Ludewig U., Tegeder M., Rentsch D., Koch W., and Frommer W.B. (2002) Conservation of amino acid transporters in ftmgi, plants and animals. Trends in Biochemical Sciences 27:139-147. Yu C., Claybrook D.L., and Huang A.H.C. (1983) Transport of glycine, serine, and proline into spinach leaf mitochondria Archives of Biochemistry and Biophysics 227: 180- 187. Yu J.W., and Woo K.C. (1988) Glutamine transport and the role of the glutamine translocator in chloroplasts. Plant Physiology 88: 1048-1054. Zybailov B., Rutschow H., Friso G., Rudella A., Emanuelsson 0., Sun Q., and van Wijk K.J. (2008) Sorting signals, N-terminal modifications and abundance of the chloroplast proteome. Public Library of Science (PLOS) ONE 3:el994. 214 CHAPTER 6 CONCLUSIONS AND PERSPECTIVES 215 CONCLUSIONS AND PERSPECTIVES The long-term goal of studying metabolite transporters is to link the uptake, distribution and secretion or excretion of a metabolite with its particular transport protein. This knowledge is essential for understanding the cellular metabolic networks by providing insights into the control of fluxes of key metabolites between metabolic pathways -from intracellular compartments up to the organismal level. This knowledge is also required for rational engineering of valuable natural products and plant biomass. Proteins, fatty acids, starch, and cellulose have been our central agricultural food and energy resource in the past and will be in firture. The importance of solute carriers for the synthesis of the latter metabolites is underestimated mainly due to the limited molecular, biochemical, and structural information on many of these transport processes, as compared to soluble proteins (Xie, 2008). The findings presented in this dissertation validate that transport systems are rather specific enzymes, which are precisely placed in a eukaryotic cell to connect or even regulate the compartmentalized metabolic network, and not simply pore- like structures in biological membranes that are permeable for most of the small molecules. In chapters one and three I characterized an ancient sugar phosphate transporter system in the plastid envelope membrane of the red alga G. sulphuraria. These transporters originated during the endosymbiosis between a photosynthetic cyanobacterium and its eukaryotic host. The insertion of a single pre-existing carrier into the inner membrane of the endosymbiont provided a selective advantage by reallocating photosynthates into the cytosol of the host, thereby directing the symbiosis to a permanent enslavement of the cyanobiont The bacterial envelope membrane had been 216 massively altered with transport proteins predominantly derived fiom the host. Importantly, the plastid phosphate translocator family (pPT) evolved fi'om an ancient nucleotide-sugar transporter to a versatile group of transporters mediating the exchange of Pi, triose-phosphate (TP), phosphoenolpyruvate (PEP), and glucose-6-phosphate (Glc6P) between cytosol and plastid of higher plants (F lfigge, 1999). It testifies the dynamic adaptation and the potential of carrier proteins to influence metabolic networks. The identification of the triose-phosphate/phosphate transporter fiom G. sulphuraria (GsTPT) with a high specificty and high affinity for its substrate triose- phosphate opens up a new tool to investigate carbon partioning into soluble and insoluble sugars in higher plants. The GsTPT has been introduced into the genome of A. thaliana wild type and (pt knockout plants (Sclmeider et al., 2002) and stable transformants are currently being established. In green plants, carbon partitioning is primarily organized in the plastid. Recently fixed carbon has three major routes in the Calvin-Benson—cycle. It is exported to the cytosol in form of triose-phosphates, further metabolized to produce hexose-phosphates for starch synthesis, or recycled to the C02 receptor ribulose-5- phosphate. The TPT is positioned at the metabolic branch of TP. Triose-phosphate export supplies carbon for sucrose synthesis during the day and on the other hand starch reserves are built up for carbon provision during the night. This assures continuous growth and development of the plant (Smith and Stitt, 2007). Over-expressing a TPT protein in tobacco plants did not significantly alter carbon partitioning, but showed that TPT activity limits the mardmal rate of photosynthesis (Hausler et al., 2000a; Hausler et al., 2000b). The GsTPT will allow us to investigate the capacity of TP export to 217 influence sucrose and starch synthesis. Accumulation of sucrose and hexose-phosphate in the cytosol sequesters cytosolic Pi, which is needed for the counter-exchange with the plastid-localized triose-phosphate. As a result, TP transport rate slows down and 3-PGA levels increase relatively to Pi in the plastid. The altered 3-PGA/Pi ratio activates allosterically the enzyme ADP-glucose pyrophosphorylase (AGP), the committed step in starch synthesis (Ballicora et al., 2004). A three and two-fold higher affinity for Pi and triose-phosphate, respectively, could increase and extend TP export during the day, thereby promoting prolonged sucrose synthesis. TPT is not the solely regulator to control the carbon flux. Thioredoxin and trehalose-6-phosphate or the regulatory molecule fructose-2,6-bisphosphate influence the activity of AGP and sucrose synthesis, respectively (Stitt et al., 1984; Tiessen et al., 2002; Lunn et al., 2006). Monitoring the rate of sucrose and starch accumulation throughout the day and a comprehensive analysis of enzyme activities and levels of soluble sugar and sugar-phosphates will demonstrate if modulating TPT kinetic constants can indeed influence the robust carbon assimilation network. Activities of AGP (Srnidansky et al., 2002) or the sucrose degrading enzymes invertase and sucrose synthase in storage organs such as seeds, fruits, and tubers (Herbers and Sonnewald, 1998) contribute to their sink strength and have 3 effect on organ number, size and dry weight. The triose-phosphate transport protein fi-om G. sulphuraria may be a promising candidate for manipulating the source strength of a crop plant. If this is true, the additive effects of an increased source and sink strength are worth to be evaluated in future. Sink strength is a quantitative trait and an ultimate goal will be to identify the “least gene 218 alleles combination” to improve the agricultm‘al yield. Another difference between the TPT proteins of higher plants and the red algae are their affinity to the metabolite 3-PGA. G. sulphuraria membranes lost or never evolved the capability for 3-PGA transport. This clearly points towards an important physiological role of 3-PGA transport in higher plant chloroplasts. By comparing the A. thaliana wild type and tpt knockout plants expressing the red algal protein, we will be able to assess the role of 3-PGA transport across the chloroplast envelope inA. thaliana leaves. Based on the results I provided in chapter three, the GsTPT cannot restore the 3- PGA transport activity in the knockout plants. We can use this approach to investigate the proposed TP/3-PGA reduction equivalent shuttle and in addition the postulated provision of 3-PGA for the cytosol to fuel the glycolytic pathway for PEP and pyruvate metabolism. We will focus our investigation on the NADP-dependent malate dehydrogenase in the chloroplast and the glycolytic enzymes in the cytosol. Reduction of oxaloacetate (0AA) to malate regenerates the terminal electron acceptor NADP+ of the photosynthetic system and malate is exported to the cytosol and subsequently oxidized to 0AA (Scheibe, 2004). Thus, the redox status of the plastid is tightly connected to the NAD(P)H pool of cytosol and the mitochondria (Raghavcndra and Padmasree, 2003). The so-called “malate/0AA valve” and the TP/3-PGA shuttle are the main routes to export reduction equivalent during the day (Heineke et al., 1991). The effect of the impaired TP/3-PGA shuttle on the photosynthetic performance can be studied in detail. In chapter three I could not assign glucose-6-phosphate as an in vivo substrate for 219 the homologous GPT protein fi'om G. sulphuraria. This finding strongly suggests that GsTPT is a versatile exporter and importer of triose-phosphate depending on the autotrophic or heterotrophic growing conditions of the cell, respectively. I tested more than 20 commercially available sugar-phosphates without success. It will be challenging to identify the biological substrate of the “enigmatic” transport protein. Metabolic profiling is currently being pursued in the laboratory and not yet tested metabolites, such as phosphorylated disaccharides, trehalose, floridoside or sulfated soluble sugars might be recognized in the analysis as putative candidate. The characterization of the pPT family and the comprehensive phylogenetic analysis of the metabolite transport proteins demonstrated the plasticity of plastid organelle evolution. The majority of the plastid metabolite transport systems have been established in the last common ancestor of the Archaeplastida. Nevertheless, distantly related photosynthetic organisms evolved alternative solutions to organize and optimize their compartmentalized metabolic network within the last one billion years. Investigation of these homologous proteins offers a great potential to discover a variety of regulatory mechanisms to fine tune the cellular metabolism and increase agricultural productivity and efficiency. In addition, chapter two provides a new perspective on plastid evolution. Connecting the cytosol of host and cyanobacterium by existing host solute transporters was a key feature for the early success of the endosymbiosis or even reflects a turning point to a permanent partnership of the endosymbiont by the protist. The export of photosynthates was a central element in the integration of plastid-host metabolism. 220 Importantly, the pPT family was not an exception. Transporters derived primarily fiom eukaryotic genes and provide a selective advantage for the protist. When this analysis was initiated, several studies investigated the result of the endosymbiotic gene transfer (EGT) on the nuclear genome of photosynthetic eukaryotes (Timmis et al., 2004). Here we demonstrated that the eukaryotic host made a major contribution for the proteome of the organelle. Host genes diversified and were targeted to new locations in the cell with alternative functions as described for cyanobacterial genes by EGT. In essence, we concluded that the initial connection of independent metabolite entities (i.e., cyanobacterium and eukaryotic host cell) by metabolite transport proteins, was a crucial early event in endosymbiosis, providing the selective advantage that permitted the step- wise completion of the process by EGT and establishment of an efficient protein translocation machinery. Identification of metabolite transporters participating in the photorespiratory pathway As stated in the first paragraph, the challenge will be to characterize the metabolite carrier families predicted fiom computer-based analyses of several genomes, membrane- specific proteomics approaches and the transport capacity of different membrane fractions (Barbier-Brygoo et al., 2001). Chapter two, four and five represent typical examples of the discrepancy between the scientific progress made for soluble and insoluble proteins participating in a specific pathway. One limitation of our phylogenomic analysis is the not yet characterized physiological roles of the majority of the transport proteins. By knowing the substrate of each individual carrier we could 221 functionally categorize the host, cyanobacterial, and chlamydial contribution for the transport of inorganic ions and metabolites across the envelope of the endosymbiont and draw the “ancient plastid-cytosol interaction map” of the proto-alga. The enzymatic steps to recycle 2-phosphoglycolate in the photorespiratory pathway are described in detail, however, the molecular nature of the transport systems in the plastidial, peroxisomal and mitochondrial membranes are mainly unknown (Reumann and Weber, 2006). A first attempt has been already published to successfully apply the basic knowledge of glycolate metabolism and design an alternative pathway to recycle 2— phosphoglycolate and increase biomass (Kebeish et al., 2007). The glycolate catabolic pathway from E. coli was targeted to the plastid inA. thaliana to short-circuit the endogenous photorespiratory pathway. Is the photorespiratory pathway and the identification of the transport proteins already “a thing of the past” (Sage and Coleman, 2001)? A detailed investigation of the transporters can contribute to the understanding of the photorespiratory network. For example, if the plastid glycolate exporter would have been identified, a targeted reduction presumably increases the flux through the biotechnologically introduced prokaryotic glycolate pathway. We could challenge the biotechnological approach by testing the capacity of glyoxylate detoxification in the plastid. Is the engineered photorespiratory bypass able to cope with an increased glycolate oxidation and result in an enhanced biomass production? In chapter four I addressed the nitrogen cycle during photorespiration. The mitochondrial glycine decarboxylase reaction releases C02 and ammonia. It is the “most inefficient step” in terms of the recycling of glycolate during the photorespiration. Nevertheless, GDC is essential and knockout mutants are lethal even under non- 222 photorespiratory conditions (Engel et al., 2007). An alternative approach to improve photorespiration could be the re-assimilation of CO2 and ammonia on site. Future experiments have to elucidate, if glutamine synthetase (GS) and carbamoyl-phosphate synthase are required for the assimilation of both gases during the day (T aira et al., 2004; Keys, 2006). To this end, it is still controversial if a glutamine synthetase (GS) is active in the mitochondria. 35S-driven GF P analysis of GS and purified mitochondria exhibit GS activity (Taira et al., 2004). On the other hand, to cope with the significant ammonia release, GS should be an abundant protein in the mitochondrial matrix. However, numerous organelle specific proteomics studies could not identify a mitochondrial GS (Heazlewood et al., 2004; Eubel et al., 2007). Over-expressing GS alone or in combination with the carbamoyl-phosphate synthase in the mitochondria could be an alternative approach to decrease the loss of the ammonia and C02 gas to the environment. The classical forward genetic approach identified only a single metabolite transporter linked to photorespiration (Somerville, 2001). The aim in chapter five was to initiate an alternative reverse genetic approach to identify membrane protein candidates mediating the flux through the photorespiratory pathway. We isolated 32 putative metabolite transporters and I initiated the genetic analysis of five of them. Two knockout mutants were indistinguishable from wild type in ambient air. Two insertion lines were lethal and did not develop beyond the cotyledon stage. The phenotype was independent of the C02 concentrations in the growth chamber. Lastly, the bou1 mutant showed a delayed classical photorespiratory phenotype. Similar to the GsGPT protein, despite the knowledge gained fi'om genetic, bioinformatic and biochemical analyses of the bou1 223 mutant, it has been difficult to assign a substrate for the mitochondrial transporter. The next step will be to demonstrate the vitamin B6 transport by the BOU protein across the mitochondrial membrane. A future goal will be to establish the remaining knockout plants and systematically screen for accumulation of photorespiratory marker metabolites to prioritize the candidate gene list. Essential genes could be targeted with an inducible RNAi approach. To characterize transport proteins, the genes are heterologous expressed, purified, and ftmctionally reconstituted in artificial vesicles (Palmieri et al., 1995). During my PhD thesis, I established additional systems to fimctionally characterize transport proteins in the laboratory of Prof. Andreas Weber. We are able to express the proteins in yeast, E. coli and in cell-free systems based on extracts of wheat germ and E. coli. I established an alternative method, the so called “cyclic detergent removal procedure” (Kraemer and Heberger, 1986), to reconstitute membrane proteins in the laboratory. Transport assays are also limited due to availability of radiolabeled substrates. With the patch-clamp technique it will be possible to test non-radiolabeled substrates. This will allow uptake assays with candidate metabolite, such as vitamin B6 derivatives, which are not commercially available radiolabeled. A combination of all methods will help to overcome the difficulties in transport research to express, preserve the activity of the purified protein and identify the in vivo substrate. 224 REFERENCES Ballicora M.A., Iglesias A.A., and Preiss J. (2004) ADP-glucose pyrophosphorylase: a regulatory enzyme for plant starch synthesis. Photosynthesis Research 79:1-24. Barbier-Brygoo H., Gaymard F., Rolland N., and Joyard J. (2001) Strategies to identify transport systems in plants. Trends in Plant Science 6:577-585. Engel N., van den Daele K., Kolukisaoglu U., Morgenthal K., Weckwerth W., Parnik T., Keerberg 0., and Bauwe H. (2007) Deletion of glycine decarboxylase in arabidopsis is lethal under nonphotorespiratory conditions. Plant Physiology 144:1328- l 3 3 5. Eubel H., Lee C.P., Kuo J., Meyer E.H., Taylor N .L., and Millar A.H. (2007) Free- flow electrophoresis for purification of plant mitochondria by surface charge. Plant Journal 52:583-594. Flugge U.L (1999) Phosphate translocator in plastids. Annual Review of Plant Physiology and Plant Molecular Biology 50:27-45. Hausler R.E., Schlieben N.H., Nicolay P., Fischer K., Fischer K.L., and Flugge U.L (2000a) Control of carbon partitioning and photosynthesis by the triose- phosphate/phosphate translocator in transgenic tobacco plants (Nicotiana tabacum L.). 1. Comparative physiological analysis of tobacco plants with antisense repression and overexpression of the triose-phosphate/phosphate translocator. Planta 210:371-3 82. Hausler R.E., Schlieben N.H., and Flugge U.l.(2000b) Control of carbon partitioning and photosynthesis by the triose phosphate/phosphate translocator in transgenic tobacco plants (Nicotiana tabacum). 11. Assessment of control coefficients of the triose phosphate/phosphate translocator. Planta 210:383-390. Heazlewood J.L., Tonti-Filippini J.S., Gout A.M., Day D.A., Whelan J., and Millar A.H.(2004) Experimental analysis of the Arabidopsis mitochondrial proteome highlights signaling and regulatory components, provides assessment of targeting prediction programs, and indicates plant-specific mitochondrial proteins. Plant Cell 16:241. Heineke D., Riens B., Grosse H., Hoferichter P., Peter U., Flugge U.l., and Heldt H.W. (1991) Redox transfer across the inner chloroplast envelope membrane. Plant Physiology 95:1131-1137. Herbers K., and Sonnewald U. (1998) Molecular determinants of sink strength. Current Opinion in Plant Biology 1:207-216. Kebeish R., Niessen M., Thiruveedhi K., Bari R., Hirsch HJ., Rosenkranz R., Stabler N., Schonfeld B., Kreuzaler F., and Peterhansel C. (2007) Chloroplastic 225 photorespiratory bypass increases photosynthesis and biomass production in Arabidopsis thaliana. Nature Biotechnology 25:593-599. Keys A.J. (2006) The re-assimilation of ammonia produced by photorespiration and the nitrogen economy of C—3 higher plants. Photosynthesis Research 87:165-175. Kraemer R., and Heberger C. (1986) Functional reconstitution of carrier proteins by removal of detergent with a hydrophobic ion exchange column. Biochimica et Biophysica Acta (BBA) - Biomembranes 863:289-296. Lunn J.E., Feil R., Hendriks J.B.M” Gibon Y., Morcuende R., Osuna D., Scheible W.R., Carillo P., Hajirezaei M.R., and Stitt M. (2006) Sugar—induced increases in trehalose-6-phosphate are correlated with redox activation of ADPglucose pyrophosphorylase and higher rates of starch synthesis in Arabidopsis thaliana. Biochemical Journal 397: 139-148. Palmieri F., Indiveri C., Bisaccia F., and Iacobazzi V. (1995) Mitochondrial metabolite carrier proteins: Purification, reconstitution, and transport studies. Methods in Enzymology 260:349-369. Raghavendra A.S., and Padmasree K. (2003) Beneficial interactions of mitochondrial metabolism with photosynthetic carbon assimilation. Trends in Plant Science 8:546—553. Reumann S., and Weber A.P.M. (2006) Plant peroxisomes respire in the light: Some gaps of the photorespiratory C-2 cycle have become filled - Others remain. Biochimica Et Biophysica Acta-Molecular Cell Research 1763:1496-1510. Sage RE, and Coleman J .R. (2001) Effects of low atmospheric C02 on plants: more than a thing of the past. Trends in Plant Science 6:18-24. Scheibe R. (2004) Malate valves to balance cellular energy supply. Physiologia Plantarum 120:21-26. Schneider A., Hausler R.E., Kolukisaoglu U., Kunze R., van der Graafl' E., Schwacke R., Catoni E., Desimone M., and Fliigge U.L (2002) An Arabidopsis thaliana knock-out mutant of the chloroplast triose-phosphate/phosphate translocator is severely compromised only when starch synthesis, but not starch mobilization is abolished Plant Journal 32:685-699. Smidansky E.D., Clancy M., Meyer F.D., Lanning S.P., Blake N.K., Talbert L.E., and Giroux MJ. (2002) Enhanced ADP-glucose pyrophosphorylase activity in wheat endosperm increases seed yield. Proceedings of the National Academy of Sciences of the United States of America 99:1724-1729. Smith A.M., and Stitt M. (2007) Coordination of carbon supply and plant growth. Plant, Cell & Enviromnent 30:1 126-1149. 226 Somerville GR. (2001) An early Arabidopsis demonstration. Resolving a few issues concerning photorespiration. Plant Physiology 125:20-24. Stitt M., Kurzel B., and Heldt H.W. (1984) Control of photosynthetic sucrose synthesis by fructose 2,6—bisphosphate. 2. Partitioning between sucrose and starch. Plant Physiology 75:554-560. Taira M., Valtersson U., Burkhardt B., and Ludwig RA. (2004) Arabidopsis thaliana GLN2-encoded glutamine synthetase is dual targeted to leaf mitochondria and chloroplasts. Plant Cell 16:2048-2058. Tiessen A., Hendriks J.H.M., Stitt M., Branscheid A., Gibon Y., Farre E.M., and Geigenberger P. (2002) Starch synthesis in potato tubers is regulated by post- translational redox modification of ADP-glucose pyrophosphorylase: A novel regulatory mechanism linking starch synthesis to the sucrose supply. Plant Cell 14:2191-2213. Timmis J.N., Ayliffe M.A., Huang C.Y., and Martin W. (2004) Endosymbiotic gene transfer: Organelle genomes forge eukaryotic chromosomes. Nature Reviews Genetics 5:123-U116. Xie H. (2008) Activity assay of membrane transport proteins. Acta Biochimica Et Biophysica Sinica 40:269-277. 227 IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII illLILLILLILLHLLIINILMILllLUllllLllHllIII 303062 480 9