a (I: . . .r.!... , 3. . x 6 can 9“!“ ,. I . 3: r .1.. n. . r .... numamfibfi x ..~...a.n.« u. a .3... JIM. v I“ ' all It I. ulna». atQ b h kin“ .11. i\ 3.9.3.4 havfifiif‘. J; 5. 19.03174... ~a 1.. a: 23m L TN 3 ao&7 LIBRARY Michigan State University This is to certify that the dissertation entitled BIOCHEMICAL AND FUNCTIONAL CHARACTERIZATION OF ARABIDOPSIS PLASTIDIC PYRUVATE KINASES presented by CARL ANDRE has been accepted towards fulfillment of the requirements for the PhD. degree in PLANT BIOLOGY Major Professor’s Signature) 1/2610? Date MSU is an affirmative-action, equal-opportunity employer n.—.-.—.—-—.— — --.—..—.—.----.---a---n-a--o---n--- — --o-I- -u-n-.- —. o.-.-.--o--.-----.-.—.—.- 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. DAIEDUE DAIEDUE DATEDUE 6/07 p:/C|RCIDaleDue.indd-p.1 BIOCHEMICAL AND FUNCTIONAL CHARACTERIZATION OF ARABIDOPSIS PLASTIDIC PYRUVATE KINASES By Carl Andre A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Plant Biology 2007 ABSTRACT BIOCHEMICAL AND FUNCTIONAL CHARACTERIZATION OF ARABIDOPSIS PLASTIDIC PYRUVATE KINASES By Carl Andre In plants, flux through the Embden-Meyerhoff pathway (glycolysis) is vital for supplying precursors for many biosynthetic pathways, including fatty acid, amino acid, and isoprenoid biosynthesis. Additionally, the production of ATP by glycolysis is expected to be important in non-photosynthetic tissues. Pyruvate kinase (PK) catalyzes the final reaction of glycolysis and in plants many of the upstream glycolytic enzymes are regulated directly or indirectly by its activity. The regulation of individual PK isoforrns is thus tailored such that glycolytic flux is appropriate for the tissue in which they reside. The specific roles of individual PKs are further defined by compartmentation between the cytosol and plastid. The Arabidopsis genome has fourteen putative PK-encoding genes. Three of these encode proteins which are located in the plastid and constitute 470 kDa plastidic PK (PKp) complexes composed of four a- and either four [51- or four Bz-subunits. Interaction of the a-subunit with either B-subunit confers unique regulatory properties to the enzymes which suggest specific roles in plant metabolism. Notably, the (13. enzyme has higher specific activity and is less sensitive to allosteric regulation by metabolites. Developing Arabidopsis embryos synthesize triacyl glycerol (TAG) to fuel seed germination and establishment. A mutant disrupted in the [ii-encoding gene (named pkp 1) has impaired seed-PKp activity and a 60% reduction in TAG accumulation with a reciprocal increase in carbohydrate content. Rescue of this phenotype can be achieved by ectopic overexpression of either B-subunit encoding gene, although [32 is not capable of full restoration. Thus, a specific role for the GB] enzyme is in catabolizing imported sugar to provide the precursors for high rates of fatty acid synthesis in embryos. As expected, pka seeds do not germinate and establish as efficiently as wild type. However, this defect is not completely attributable to a lack of TAG and it seems that PKp activity is also necessary for proper metabolism in germinating seeds. It was found that pka seedlings and mature plants have much in common with mutants deficient in isoprenoid and tocopherol biosynthesis, but there was no obvious defect in fatty acid metabolism. Apparently, the function of PKp in leaves is also to provide precursors for whichever biosynthetic processes are in highest demand. fir tin me 0nl= \I OL ACKNOWLEDGMENTS Firstly, I am gratefully indebted to my thesis advisor, Dr. Christoph Benning. He worked tirelessly to insure that I had a place in his lab to develop my scientific skills, despite a seeming lack of interest from funding agencies in supporting this project. He allowed for creativity and exploration while still establishing clear goals and demanding work to be performed at the best of my abilities. My thesis advisory committee, Dr. J ianping Hu, Dr. John Ohlrogge, and Dr. Andreas Weber also deserve my thanks. They made themselves accessible for discussion at their own inconvenience and contributed thoughtful insight about my research. I am thankful to the Michigan State University US. Department of Energy Plant Research Laboratory and to the Department of Plant Biology for accepting me as a graduate student and for fostering some of the best plant science in the world. I only hope that I can help uphold the reputation for which these departments are known. I would also like to acknowledge the Department of Biochemistry and Molecular Biology for hosting me as one of their own and for allowing me to use their numerous facilities. Included are the labs of Dr. Jack Preiss and Joe Leykam where l invaded and used equipment for extended periods of (sometimes frustrating) time. I have also benefited from the community here at MSU where numerous discussions with graduate students and postdocs, especially the members of the Benning lab, past and present that always made work a pleasant place, helped guide me in the right direction. I can leave MSU knowing that I worked and socialized with some of the best scientists and people there are. Thanks. iv TABLE OF CONTENTS LIST OF TABLES ............................................................................................................. vii LIST OF FIGURES .......................... . ................................................................................ viii CHAPTER 1 - Seed storage compound accumulation and related metabolisms in plants ....................................................................................................... 1 1 .1 . Introduction ......................................................................................................... 2 1.2. Embryo development .......................................................................................... 3 1.3. Signals that control embryo maturation and storage compound accumulation .. 4 1.3.1. Genetic regulators ....................................................................................... 4 1.3.2. Small molecule regulators ........................................................................... 5 1.4. Considerations for seed oil biosynthesis ............................................................. 8 1.4.1. Competition between biochemical pathways .............................................. 8 1.4.2. Supply of energy ....................................................................................... 1 1 1.4.3. Supply of reducing equivalents ................................................................. 14 1.5. The biosynthetic pathway from sucrose to oil .................................................. 16 1.5.1. Synthesis of fatty acids and triacyl glycerol .............................................. 18 1.5.2. Supply of carbon precursors for fatty acid synthesis ................................ 20 1.5.3. Glycolysis in oil seeds .............................................................................. 25 1.6. Pyruvate kinase ................................................................................................. 28 1.6.1. Pyruvate kinase phylogeny ....................................................................... 29 1.6.2. Pyruvate kinase molecular architecture .................................................... 30 1.6.3. Mechanisms of regulation ......................................................................... 31 1.6.4. Analysis of pyruvate kinase mutants ........................................................ 33 1.6.5. Pyruvate kinase moonlighting ................................................................... 34 1.7. Rationale and outlook ....................................................................................... 34 References ..................................................................................................................... 36 CHAPTER 2 - Molecular and kinetic analysis of Arabidopsis plastidic pyruvate kinase ....................................................................................................... 49 Abstract ......................................................................................................................... 50 Introduction ................................................................................................................... 51 Materials and Methods .................................................................................................. 53 Results ............................................................................................................................ 62 Discussion ..................................................................................................................... 82 References ..................................................................................................................... 87 CHAPTER 3 - Analysis of carbon metabolism and storage compound accumulation in seeds of an Arabidopsis mutant deficient in a plastidic pyruvate kinase ....................................................................................................... 9 1 Abstract ..................................................................................................................... 92 Introduction ................................................................................................................... 93 Materials and Methods .................................................................................................. 96 Results .......................................................................................................................... 101 Discussion ................................................................................................................... l 18 References ................................................................................................................... 1 24 CHAPTER 4 - Germination, establishment, and grth of Arabidopsis plants lacking a plastidic pyruvate kinase ......................................................... 128 Abstract ....................................................................................................................... 129 Introduction ................................................................................................................. 1 30 Materials and Methods ................................................................................................ 133 Results .......................................................................................................................... 136 Discussion ................................................................................................................... 149 References ................................................................................................................... 154 CHAPTER 5 - Conclusions and perspectives ................................................................. 157 Appendix A ..................................................................................................................... 166 Appendix B ..................................................................................................................... 181 Appendix C ..................................................................................................................... 192 Appendix D ..................................................................................................................... 204 vi Ial Ial Tat Tat Ial Ia'l LIST OF TABLES Table 2.1. Primer sequences .............................................................................................. 56 Table 2.2. PKp kinetic constants ..... g ................................................................................. 80 Table 2.3. Metabolite effectors of PKp activity ............................................................. 81 Table 3.1. WT and pka seed storage compound accumulation ...................................... 105 Table 3.2. Metabolite levels in WT and pka developing seeds ..................................... 115 Table 4.1. Leaf pigments in 25 day old plants grown on soil or on agar plates .............. 145 vii Fig: Fig; Fig: Fig: Fig; Fig. LIST OF FIGURES Figure 1.1. Competition for resources between seed storage compounds ........................... 9 Figure 1.2. Simplified scheme of carbon metabolism in developing seeds ....................... 17 Figure 1.3. Alternative pathways of pyruvate production ................................................. 23 Figure 1.4. The glycolytic network in green oil seeds ....................................................... 27 Figure 1.5. Pyruvate kinase centrality to plant metabolism ............................................... 28 Figure 2.1. Pyruvate kinase similarity and selected gene expression in Arabidopsis ........ 64 Figure 2.2. Phylogenetic analysis of plant PK protein sequences ..................................... 66 Figure 2.3. Subcellular localizations of pyruvate kinase subunits ..................................... 68 Figure 2.4. PKp protein purification and initial activity assay ........................................... 71 Figure 2.5. In vitro interaction of PKp subunits ................................................................. 72 Figure 2.6. In vivo co-immunoprecipitation of PKp subunits ............................................ 75 Figure 2.7. Kinetics of active PKp complex formation ...................................................... 77 Figure 2.8. PKp subunit inactivation .................................................................................. 79 Figure 3.1. Identification of a SALK T-DNA mutant in PKp-Bl ..................................... 102 Figure 3.2. Reverse transcriptase-PCR analysis of PKp gene expression ........................ 103 Figure 3.3. pka seed phenotypes .................................................................................... 104 Figure 3.4. PK activity in pkp] and wild-type seeds ....................................................... 107 Figure 3.5. Oil and protein phenotype of pkp] seeds ...................................................... 109 Figure 3.6. Rescue of the pka seed oil phenotype .......................................................... 1 11 Figure 3.7. PKp gene expression and enzyme activity in rescued pka lines .................. 1 13 Figure 3.8. Carbohydrate accumulation and phosphofi'uctolinase enzyme activity in pkpl and wild-type seeds ............................................................................................. 116 viii Figure 4.1. Sucrose dependent establishment and root elongation in pka ..................... 137 Figure 4.2. Delayed germination and induction of PKp activity in pkpl ......................... 138 Figure 4.3. Inhibition of pkp] germination by exogenous sucrose .................................. 140 Figure 4.4. Hypocotyl elongation and PKp activity in dark-grown seedlings ................. 141 Figure 4.5. Fatty acid composition in germinating seeds and seedlings .......................... 142 Figure 4.6. Seed germination following accelerated aging treatment ............................. 143 Figure 4.7. Plant growth and morphology of pkpl .......................................................... 144 Figure 4.8. Pyruvate kinase activity in rosette leaves during the day and night .............. 146 Figure 4.9. Lipid profile and composition of pkpl leaves ............................................... 147 Figure 4.10. Carbohydrate content of 25 day old leaves ................................................. 148 Figure A]. Cytosolic localization of G6PD5 and G6PD6 .............................................. 172 Figure A.2. Single and double mutants for G6PD5 and G6PD6 .................................... 174 Figure A.3. G6PDH isoforms and activity in developing siliques .................................. 176 Figure 3.1. Impaired seedling establishment in the wriI mutant .................................... 187 Figure 3.2. Glycolytic enzyme activities in green seedlings ........................................... 188 Figure C.l. Relative gene expression of putative Arabidopsis lipin homologues ........... 198 Figure C.2. Gene structure of the two Arabidopsis LpinI homologues ........................... 199 Figure C.3. Lipid analysis of Arabidopsis lipin single and double mutants .................... 200 Figure D.l. DGTS biosynthetic pathway and involvement of RthaAB ........................ 206 Figure D.2. BtaA-catalyzed DGHS biosynthesis from radiolabeled substrates .............. 212 ix Chapter 1 Seed storage compound accumulation and related metabolisms in plants 1.] TI: en‘ tilt of I65: lht’t Ted} pro mlfi bio SEC mt 1.1. Introduction The reproductive success, and in part evolutionary fitness, of most plants is largely dependent on sexual reproduction involving fertilization of an ovule and subsequent embryo development. The challenges of reproduction vary, and are greatly influenced by the environment in which the plants take root. Regardless of such factors, the packaging of nutrient reserves in seeds has evolved as a major strategy to insure the establishment and survival of the next generation. Gerrninating seeds must metabolize their own storage reserves until the seedlings establish photosynthetic rates capable of sustaining growth. Carbohydrates, oils in the form of triacylglycerol (TAG), and proteins constitute the reserves used for this establishment. In addition to their indispensability for plant survival, these compounds have broad economic importance as agricultural commodities. For this reason, we need to investigate and understand the metabolism which leads to the accumulation of these compounds with the eventual goal of engineering plants to be more productive. Seed oil, in particular, is of utmost interest given the current Zeitgeist. The model plant Arabidopsis serves as an excellent organism in which to study seed oil biosynthesis, as it is very closely related to canola (Brassica napus) which is a major oil seed crop in the Northern hemisphere. In this chapter, I will review the metabolisms involved in storage compound accumulation, with a focus on oil, and I will introduce some of the current hypotheses pertaining to regulation of biosynthesis and turnover. I will establish that the glycolytic enzyme pyruvate kinase is pivotal in these processes and that it represents a step that needs to be researched further to fully understand the metabolism of developing oil seed embryos. em' 3].. dis! 5101 (La 0ft de\ 1.2. Embryo development Plant embryogenesis is initiated by a fertilization event in which one male gametophyte unites with a single female gametophyte to form a zygote. An additional fertilization occurs between another male gametophyte and a specialized diploid cell within the embryo sac resulting in the formation of the triploid endosperm. The function of the endosperm is to nourish the embryo during development and germination and in some cases (e. g. cereals) it is the site of storage product accumulation (Berger 2003, Penfield et al. 2004). Once fertilized, the development of plant embryos generally occurs in three distinct phases: 1) cellularization and establishment of a body plan, 2) deposition of storage reserves and maturation, and 3) acquisition of desiccation tolerance (Goldberg et al. 1994). Immediately following the double fertilization event, the seed rapidly grows mostly due to syncytial development of the endosperm (Berger 1999). During this time (~ 5 days in Arabidopsis) the zygote goes through many rounds of asymmetrical cell division and establishes tissues that will eventually become the hypocotyl and cotyledons (Laux and Jurgens 1997). Upon entering the maturation phase, cell division slows in lieu of expansion as cotyledonary cells begin to fill with storage compounds. This phase of development is depicted in Chapter 3, Figure 3.3A, and is associated with distinct changes in gene expression as discussed below. During this time, the water content of the seed steadily declines (Baud et al. 2002). The acquisition of desiccation tolerance and seed dormancy is characterized by nearly complete loss of water content from the seed and is associated with expression of presumably cryo-protective protein-encoding LEA genes (Wise and Tunnacliffe 2004) and programmed cell death of maternal tissues which will eventually become the seed coat. 31' be ”3 C0 r_ Ups 1.3. Signals that control embryo maturation and storage compound accumulation Embryo maturation, as described above, is the phase during which storage compounds are biosynthesized and deposited in developing seeds. The regulation of this process is at best loosely understood. There are, however, known transcriptional and metabolic cues which exert some control over this process. It is difficult to tease apart these two means of regulation as metabolism is directly influenced by the genes which are active in a given tissue, the expression of which may in turn be controlled by cellular metabolites. 1.3.1. Genetic regulators Arabidopsis has served as the model for the study of the transcriptional control of seed maturation. The changes in metabolic gene expression during the shift from embryo morphogenesis to storage compound accumulation have been documented (Girke et al. 2000, Ruuska et al. 2002, White et al. 2000). Additionally, some genetic factors which control such changes in metabolic gene expression have been elucidated. The most upstream factor identified to date is PKL (PICKLE). The PKL gene encodes a chromatin- remodeling factor that specifically represses embryo associated transcription and metabolism in non-embryo tissues (Ogas et al. 1999, Rider Jr et al. 2003, Rider Jr. et al. 2004). Indeed, pkl mutants ectopically accumulate seed storage compounds in roots (Ogas et al. 1997). Some of the specific targets of PKL-mediated repression encode the transcription factors LECI, LEC2, and F US3. These three LEAF Y COTYLEDON-class transcription factors are regulators of embryo identity and are so named because of the leaf-like appearance of cotyledons of the respective mutants. Overexpression of these genes results in spurious embryo formation in vegetative tissues (Lotan et al. 1998, Stone etal. redtI LEC targt] targ. pOIC CIIL‘I' C0; (10‘ et a1. 2001). These genes also control storage product accumulation as evidenced by reduced amounts of storage protein and oil in mutant seeds (Meinke et a1. 1994). The LEC-class transcription factors are considered master regulators and are hypothesized to target a subset of transcription factors controlling embryo maturation. One potential target of LECl is the WRII gene. Overexpression of WRII in the lac] background rescues the desiccation intolerant phenotype and the promoter of WRII contains a potential DNA binding site of LECl (Cemac and Benning, unpublished). The WRII gene encodes an AP2/EREB domain transcription factor that regulates sugar metabolism (1'. e. glycolysis) in tissues where WRI] is expressed, namely, seeds and roots (Cemac and Benning 2004, Ruuska et al. 2002). The wriI mutant was originally identified in a screen for mutants specifically reduced in seed oil accumulation (F ocks and Benning 1998). Overexpression of WRII in seedlings was subsequently found to cause an embryo identity characterized by the synthesis and accumulation of TAG and seed storage proteins (Cemac and Benning 2004). While this is an incomplete review of all the genes involved in regulating embryo maturation, it is clear that there is a strong relationship between transcriptional regulation of embryo identity and seed storage compound biosynthesis. 1.3.2. Small molecule regulators The list of small molecules which potentially regulate seed maturation, and thus storage compound accumulation, is growing. The phytohorrnone abscisic acid (ABA) is one well documented regulator of these processes. There are two peaks of ABA accumulation during seed maturation of most species (Karssen et al. 1983). The first peak potentially regulates the transition from embryo cell division to cell enlargement in Arabidopsis as ABA is known to induce the expression of cell cycle inhibitory factors (Wang et al. 1998). During this time ABA also induces the expression of seed storage protein genes (Crouch and Sussex 1981). An Arabidopsis mutant defective in the sensing of ABA, abi3-3, fails to accumulate seed storage proteins (Koomneef et al. 1989). However, as mentioned above, it is difficult to tease apart transcriptional and metabolite-based regulation and it has been shown that ABI3 interacts in the signaling mediated by the LEC-class transcription factors described earlier (Parcy et al. 1997). Non-hormonal metabolite signals are gaining recognition as potential regulators of seed maturation. Among these, the most attention has fallen upon sugars. Sugars, especially glucose, have been identified as regulators of photosynthetic gene expression and as components of signal transduction networks in Arabidopsis vegetative tissue (J ang et a1. 1997, Jang and Sheen 1994). Developing seeds import large amounts of sugar from maternal tissue, and it is expected that sugar signaling networks would exist here as well. In legume seeds, the transition from cell division to storage product accumulation is correlated with a shift from a high hexose to sucrose ratio to a high sucrose to hexose ratio (Borisjuk et al. 2003, Weber et al. 1996). It has been proposed that the concentrations of the sugars themselves provide signals that regulate this developmental transition (Weber et a1. 1997). A similar switch in the ratio of hexose to sucrose is observed in Arabidopsis; however, it comes after the onset of high rates of oil biosynthesis and may be a result of changed metabolism rather than triggering those changes (Baud et al. 2002, Hill et al. 2003). The apetalaZ (ap2) mutant of Arabidopsis produces seeds with higher mass and yield (Ohto et a1. 2005). The increased cell size and CI 6T CD fl) (let em m 301: am u“;- U] I... number in ap2 embryos is correlated with an increased hexose to sucrose ratio throughout embryo development. The modulation of sugar metabolism appears to have an effect on embryo development in this mutant, although again, it is unclear whether the altered sugar ratio causes the phenotype or is the result of it. Transgenic tobacco plants with seed-specific overexpression of Invertase were used to examine importance of the switch from hexose to sucrose in oil seed crops. The transgenic lines maintained a very high hexose to sucrose ratio throughout embryogenesis but developed normally, suggesting that the switch from hexose to sucrose accumulation is not a signal (Tomlinson et al. 2004). Another sugar, trehalose-6-phosphate (T6P), is involved in embryo maturation. As opposed to sucrose and hexoses, T6P exists at very low concentrations in Arabidopsis embryos and cannot serve as an important energy source, which buttresses the notion of it as a signaling molecule. The importance of this molecule in regulating embryo development is evidenced by an Arabidopsis mutant deficient in T6P-synthase which is embryo lethal (Eastrnond et al. 2002). Furthermore, an analysis of the aborted embryos revealed a dearth of oil accompanied by an increase in starch accumulation (Gomez et al. 2006). Components involved in the T6P signaling network remain to be identified. The amino acid asparagine may be another metabolite regulator of embryo maturation. An analysis of soybean seeds from lnigh- and low-seed protein lines revealed a tight correlation between the amount of aspargine present in the embryo and the amount of seed storage protein in the dry seed (Hemandez-Sebastia et al. 2005). It is unknown whether this regulation is due to effects on transcription of storage protein genes or due to substrate availability. Ara in I; {on Ber pro: SCCL Sim and liq; PIC of» L“ CI. 1.4. Considerations for seed oil biosynthesis The composition of seed storage compounds varies between plants, but a focus on Arabidopsis and other oil seed crops is pertinent in the context of the research presented in later chapters. Approximately 30% of the dry weight of an Arabidopsis seed is in the form of TAG, while another 30% is composed of seed storage proteins (Focks and Benning 1998). Starch transiently accumulates in developing embryos, but is not a major product in mature Arabidopsis seeds. The timing of these processes has been studied at the metabolic and transcriptional levels using Arabidopsis (Baud et al. 2002, Ruuska et al. 2002). At the onset of embryo cell expansion, which is 5 days after flowering (DAF), starch is synthesized in the plastids and accumulates through 9 DAF at which point it is degraded. Coincident with starch degradation are the synthesis of TAG and seed storage proteins, both of which continue to accumulate linearly until about 17 DAF, when the seed begins to desiccate and become metabolically quiescent. Figure 1.1 depicts a simplified scheme for the biosynthesis of seed storage compounds in oil seeds. Sucrose and some amino acids are provided by the maternal source organs through the endosperrn liquid and are then metabolized by the embryo (Schwender and Ohlrogge 2002). The precursors of each storage compound are derived from connected biochemical pathways, of which fine regulation is necessary to balance the final reserve composition of the seed. Such carbon partitioning will be discussed in more detail in the following section. 1.4.1. Competition between biochemical pathways The biosynthetic pathways in Figure 1.1 are simplified to emphasize the potential for competition for resources between the biosyntheses of starch, oil, and protein. Seed stor: C21110i limit and I fail} fair} deer Ft“? "'1 'I raw-r—a storage protein biosynthesis is the most independent pathway. Stable isotope labeling of canola embryos revealed that 30% of the amino acids used in the synthesis of storage proteins are imported and 70% arise from de novo biosynthesis from sugars (Schwender and Ohlrogge 2002). Labeled amino acids were not catabolized and incorporated into fatty acids. Moreover, there appears to be independent sugar metabolism for protein and fatty acid synthesis, which is supported by numerous Arabidopsis mutants in which decreases in seed oil or protein does not result in compensatory increases in the other major storage compound. The wriI mutant, for instance, is reduced 80% in oil with no effect on the protein content (Focks and Benning 1998). Conversely, abscisic acid biosynthetic and signaling mutants exhibit reductions in seed storage protein content with no affect on oil (Finkelstein and Somerville 1990). Source Tissue 1 r f 1 81101088 .9 hexdse-P starch granule protein body 7‘ 1. " ‘b I L I 5 ,y' . . A! 1} . IE ‘ . 0 . . . -%‘o . 2.: V...” v .0 amino acids —> Protein 00 '. TAG<—<—Pyr ,i . . L @ Embryo J Figure 1.1. Competition for resources between seed storage compounds Sucrose and amino acids are imported from source tissue via the phloem and are incorporated into the seed metabolic network. Images are TEM micrographs of representative storage organelles from Arabidopsis developing embryos. Pyr, pyruvate; TAG, triacylglycerol; TCA, tricarboxylic acid. In oil seeds, the true competition occurs between starch and oil biosynthesis. The transient accumulation of starch in Arabidopsis seeds is hypothesized to serve several possible functions: 1) to maintain the sink strength of the embryonic tissue during fatty acid biosynthesis (Da Silva et al. 1997), 2) to supply carbon precursors for fatty acid synthesis upon its degradation (Norton and Harris 1975), and 3) to supply sugar (and thus osmolarity) needed for the acquisition of desiccation tolerance (Leprince et al. 1990). Two studies have directly addressed the question of the firnction of transitory starch in Arabidopsis seeds. In one a mutant of plastidic phosphoglucomutase (png) which is compromised in starch biosynthesis was analyzed with respect to seed storage compounds (Periappuram et al. 2000). A 40% reduction in seed oil content was found in pgm] . The authors, however, could not conclude whether the reduction in starch compromised the sink strength or the carbon supply for fatty acid synthesis. Another study utilized embryo specific reduction of ADP-glucose pyrophosphorylase to impair transient starch accumulation in canola embryos (Vigeolas et al. 2004). While starch accumulation was severely compromised in the transgenic lines, oil biosynthesis was compromised only during the early stages of development. The authors concluded that starch was not a source of sugar for the production of fatty acid precursors during the main stages of oil biosynthesis. These two studies present no clear conclusions about the role of transitory starch accumulation. There is additional correlative evidence, though, that links starch accumulation to fatty acid synthesis in Arabidopsis embryos. Three mutants defective in oil accumulation, wril, led, and shrunken seed 1 (sseI, encoding a homolog of the peroxisome biogenesis factor Pex16p), all display compensatory increases in starch accumulation (Focks and Benning 1998, Lin et al. 1999, Lin et al. 10 300- deer to h by 1 ml or 3 mai 3‘00 ace 0ft: (Re Ct: Ct 2004, Meinke et al. 1994). In the case of wriI and sseI which both have observed decreases in the rate of fatty acid synthesis, sugars (hexose and sucrose) also accumulated to higher than wild-type levels in the developing seeds. These findings can be explained by two hypotheses: 1) starch biosynthesis functions to lessen the osmotic potential of the embryo by consuming excess sugar, thus maintaining the embryo status as a sink tissue, or 2) excess sugar inhibits the degradation of starch which would normally be needed to maintain a sugar supply to feed fatty acid synthesis. As the starchless mutants described above are desiccation tolerant, it is unlikely that starch serves to facilitate sugar accumulation for this purpose. Yet, no definitive conclusions can be drawn about the role of transitory starch accumulation in oil seeds. 1.4.2. Supply of energy Fatty acid biosynthesis is dependent on a steady supply of adenosine triphosphate (ATP). Starch and amino acid synthesis along with the assembly of storage proteins (4 ATP or GTP per peptide bond) occurring in the embryo no doubt compete for the same ATP' (Regierer et al. 2002). The committed step of fatty acid synthesis catalyzed by acetyl- CoA carboxylase (ACCase) uses ATP in a two step reaction converting acetyl-CoA into malonyl-CoA. It is the first step that is ATP depedent, in which a carboxyl group from bicarbonate is transferred to a biotin prosthetic group on the enzyme (Ohlrogge and Browse 1995). The synthesis of one 18 carbon fatty acid thus requires 9 ATP for the ACCase reaction alone. ACCase is considered the rate limiting reaction of fatty acid synthesis and its activity is light dependent by means of redox-regulation, pH, and Mg2+ concentration'(Hunter and Ohlrogge 1998, Sasaki et a1. 1997). Indeed, fatty acid 11 syn Rm AC not Wilt AT of g mi: rem pht‘ IZIIC Ara fro. Upr but synthesis in canola embryos is stimulated by light (Bao et al. 1998, Goffinan et al. 2005, Ruuska et al. 2004). Such light-dependent activation may be one means of coordinating ACCase activity with the production of ATP by photosynthesis in green tissues. However, not all oil seeds are green, and even then only as much as 30% of ambient light may reach the embryo (Eastrnond et al. 1996, King et a1. 1998). Other means of producing ATP include the lower half of glycolysis and mitochondrial respiration. The contribution of glycolysis to ATP production has not been directly tested to date, however, mitochondrial respiration has been studied in oil seeds. Stable isotope labeling studies revealed that only as much as 22% of the needed ATP can be generated through oxidative phosphorylation in canola embryos (Schwender et al. 2006). Increasing the respiration rate through indirect activation of the mitochondrial pyruvate dehydrogenase complex in Arabidopsis resulted in higher oil yield, suggesting that increased ATP production in mitochondria may stimulate fatty acid synthesis (Marillia et al. 2003). Export of ATP fi'om the mitochondria occurs through the adenine nucleotide carrier and subsequent uptake by plastids would be through a known ATP/ADP transporter (Resier et al. 2004), both in counter exchange with ADP. Oxygen tension may have a role in regulating mitochondrial respiration in seeds. In Arabidopsis siliques, the ambient oxygen concentration is low, and further reduction brought on by growth in sub-ambient [0;] limited seed growth and oil accumulation (Porterfield et al. 1999). Additional studies in canola revealed that oxygen concentration in seeds is also low, and that increasing [0;] in low light conditions results in elevated ATP and UDP concentrations, accompanied by faster lipid synthesis (Vigeolas et al. 2003). Although, when light is not limiting increasing atmospheric [02] has no 12 stimulatory effect, suggesting that in green seeds photosynthesis alone is capable of supplying the needed ATP (Goffman et al. 2005). Glycolysis is known to be induced in hypoxic tissues such as roots and seeds (Plaxton and Podesta 2006); however, its contribution to ATP pools has not been investigated. In any case, ATP must be imported into (if not generated in) the plastid for use in the relevant biosynthetic processes. The Arabidopsis genome encodes two plastidic ADP/ATP transporters, and a double mutant lacking both displays 40% reduced lipid and and about 30% reduced protein accumulation in seeds, presumably due to a lack of ATP in plastids (Reiser et al. 2004). Interestingly, mRNA levels of genes encoding subunits of plastidic pyruvate kinase (examined in this thesis) were higher in this mutant. The authors speculate this was a mechanism to compensate for reduced ATP import by instead producing it in the plastid, although, pyruvate kinase enzyme activity was not measured. The importance of ATP supply is further highlighted by an observed shift from ATP to pyrophosphate (PPi) consuming metabolism during the phase of maximum fatty acid biosynthesis in Arabidopsis seeds. This shifi was observed at the level of transcription and enzyme activity (Baud and Graham 2006, Ruuska et al. 2002). In particular, ATP-dependent phosphofructokinase and invertase (which leads to downstream ATP consumption) activities are replaced by PPi-dependent phosphofructokinase and sucrose synthase (for the cleavage of sucrose). The current hypothesis is that in response to a limited oxygen supply, and thus ATP, this shift is induced in an effort to conserve the adenylate pool for use in fatty acid synthesis. A similar pattern is seen in potato tubers where ATP is required for starch synthesis (Appeldoom et al. 1997). 13 1.4.3. Supply of reducing equivalents Reducing equivalents in the form of NADH and NADPH are necessary for the synthesis, elongation, and desaturation of fatty acids. The reductases involved in fatty acid synthesis specifically use NADPH (3-ketoacy1-ACP reductase) and NADH (enoyl-ACP reductase) (Caughey and Kekwick 1982, Shimakata and Stumpf 1982, Slabas et al. 1986). Synthesis of a saturated 18 carbon fatty acid requires 8 NADH and 8 NADPH. Subsequent desaturation of fatty acids indirectly requires the input of NAD(P)H (Shanklin and Cahoon 1998). The very long chain fatty acids found in wax, suberin, spingolipids and seed oil, are elongated in the cytosol via a similar mechanism as fatty acid synthesis in the plastid, and thus also require 2 NAD(P)H for each two carbon addition (Barrett and Harwood 1998). The NADH required for fatty acid biosynthesis could be supplied from the plastidic pyruvate dehydrogenase complex, which generates NADH and fatty acid precursors (acetyl-CoA) in a 1:1 ratio. In green seeds exposed to light photosynthesis could produce the required NADPH, as is the case with ATP. As described above, light stimulates fatty acid synthesis in embryos and this could likely be a result of not only ATP, but also NADPH production (Bao et al. 1998, Goffinan et al. 2005, Ruuska et al. 2004). Indeed, canola embryos contain chloroplasts similar to shade grown leaves and are capable of photosynthesis (Asokanthan et al. 1997, Eastmond et al. 1996, King et al. 1998). Calculations based on reported 02 evolution rates from canola embryos indicate that photosynthesis could provide all of the NADPH required for fatty acid synthesis (Ruuska et al. 2004). Indeed, carbon use efficiency for storage oil synthesis is greatly reduced in dark-cultured canola embryos compared to those grown in even low light 14 COIII plio1 I V. OI conditions (50 umol m_2 s", Goffrnan et al. 2005). This effect could be due to the role of photosynthesis in producing reductant and ATP. The oxidative pentose phosphate pathway (OPPP) generates 2 NADPH for each glucose molecule metabolized, and this could help fulfill the requirements for the reductant necessary for high rates of fatty acid synthesis (Eastmond and Rawsthome 2000). The first enzyme of the OPPP, glucose-6-phosphate dehydrogenase (G6PDH), is feedback inhibited by NADPH and would presumably be inactivated in reducing conditions (Scheibe and Anderson 1981, Wakao and Benning 2005). Additionally, in vivo labeling revealed that a maximum of 25% to 45% of the reductant required for oil biosynthesis is provided by the OPPP (Schwender et al. 2003). Furthermore, an Arabidopsis double mutant which is defective in both cytosolic G6PDH isoforms (resulting in a 50% decrease in total enzyme activity) has no reduction in oil content (see Appendix A). Plastidic NADP-dependent malic enzyme (NADP-ME) also produces NADPH and a role for this enzyme in canola oil metabolism has been proposed (Kang and Rawsthome 1994, Singal et al. 1995). However, metabolic flux analysis of cultured canola embryos revealed that malate is not a major contributor to fatty acid synthesis (Schwender and Ohlrogge 2002). NADP-ME mutants have been isolated from Arabidopsis, and loss of seed specific isoforms has no effect on oil accumulation (Wheeler et al. 2005). Taken together, it seems that the OPPP and NADP-ME play little or no role in green seeds, but in non-photosynthetic seeds there is evidence to the contrary. Labeling studies with sunflower (Helianthus annuus) found that malate is the preferred substrate of fatty acid synthesis when fed to isolated leucoplasts (Pleite et al. 2005). Moreover, the consumption of malate was reduced when co-fed with glucose-6- 15 phosphate (G6P), and vice versa. Labeled G6P was not incorporated into fatty acids, but instead was entered into the OPPP. The balancing of flux between the OPPP and NADP- ME suggests that these pathways both contribute to NADPH pools in isolated leucoplasts. However, more recent steady-state isotopic-labeling experiments using cultured embryos indicated very little flux of malate into oil (Alonso et al. 2007). NADP-ME has also been implicated in the supply of reductant in castor (Ricinus communis) seeds. As with sunflower, malate was the preferred substrate for fatty acid synthesis in isolated leucoplasts (Smith et al. 1992) and a correlation was observed between leucoplast NADP-ME activity and the onset of fatty acid synthesis (Shearer and Dennis 2005). Clearly, green and non-green oil seeds differ in how they supply ATP and reducing equivalents for fatty acid synthesis, but the unifying theme is that in either there are multiple overlapping and interacting pathways for the production of both. 1.5. The biosynthetic pathway from sucrose to oil The biosynthetic pathways leading to the accumulation of TAG in oil seeds have been extensively studied in Arabidopsis and other oil seed crops. Sucrose imported from maternal tissues is the main carbon source for these metabolisms. There is no symplasmic connection between maternal and filial tissues and so sucrose (or hexose) is released to the seed apoplasm and is then imported into cotyledonary cells of the embryo by membrane localized transporters (Rosche et al. 2002). Once inside the embryo hexose is converted eventually to fatty acids and incorporated into TAG which is deposited into oil bodies (see Figure 1.1). Oil bodies originate from the endoplasmic reticulum (ER) and are delineated by a phospholipid monolayer heavily embedded with proteins (Galili et al. 16 SC Fig Em. an. 1998). Figure 1.2 is based on metabolic flux analysis and transcript profiling of green oil seeds (Ruuska et al. 2002, Schwender et al. 2004a, Schwender et al. 2004b) and depicts the major pathway of carbon metabolism from the loading of sucrose to the budding of oil bodies from the ER. Emphasis is given to the supply of carbon precursors for fatty acid synthesis, but all metabolisms will be covered in forthcoming sections. Sucrose PM f " . 1 Cytosol Sucrose—> Hexose-P 66 P . RuBisCO Shunt 3PGA pep *L *L’ ATP Free FA fr ATP Pyr ‘T'PY' PY' ATPxi U NAD(P)H Glycolysis Gluco eo II Mitochondrion L A Plastid LL Jr Figure 1.2. Simplified scheme of carbon metabolism in developing seeds Emphasis is on oil production. Asterisks demark the pyruvate kinase reaction. Single arrows can indicate multiple reactions. 3PGA, 3-phosphoglycerate, ATP, adenosine triphosphate; ER, endoplasmic reticulum; FA, fatty acid; FAS, fatty acid synthase; G6P, glucose 6-phosphate; Gluconeo., gluconeogenesis; PEP, phosphoenolpyruvate; PM, plasma membrane; Pyr, pyruvate; TAG, triacylglycerols; TCA, tricarboxylic acid. 17 L. h! 1.5.1. Synthesis of fatty acids and triacylglycerol The site of almost all de novo fatty acid synthesis in plants is the plastid (Ohlrogge et al. 1979). Fatty acids are synthesized by cycles of elongation in two carbon increments, with malonyl-CoA acting as the carbon donor. The synthesis of malonyl-CoA in plastids is performed by a multimeric bacterial-type ACCase and as mentioned above is subject to regulation by light through various mechanisms. Attempts to boost seed oil yield by increasing the activity of ACCase have had limited success to date and indicate a more complex regulation of fatty acid synthesis (Thelen and Ohlrogge 2002). To be used as a substrate by the fatty acid synthase complex (FAS), malonate must first be transfered to acyl carrier protein (ACP). Malonyl-CoAzACP transacylase performs this transfer and the activity of this enzyme may also be subject to regulation by light through redox mechanisms (Lemaire et al. 2004). All subsequent steps in the synthesis of fatty acids in the plastid involve ACP and a bacterial type II (multimeric) FAS. A carbon-carbon bond and the release of one C02 from malonate are the products of condensation of malonyl- ACP with acyl-ACP (or acetyl-CoA). Three separate condensing enzymes known as 3- ketoacyl-ACP synthases (KAS) can perform this reaction depending on the initial acyl chain length. The first round of condensation to form a four-carbon product, which uses acetyl-CoA as a primer, is carried out by KASIII (J aworski et al. 1989). Elongation from 4 carbons up to 16 is done by KASI and the final addition to make an 18-carbon fatty acid is done by KASII. Each condensation yields a 3-ketoacyl-ACP product which must be reduced (using NADPH), dehydrated, and then reduced again (using NADH) to yield a saturated product ready for the next round of condensation. These steps are catalyzed by 3-ketoacyl-ACP reductase, 3-hydroxyacyl-ACP dehydratase, and enoyl-ACP 18 l'Ct dc EF tht at. dra of. et 2 _V€i Ch; AC CV reductase, which together with KAS make up the multimeric F AS complex. As a final step, a portion of the 18-carbon ACP can be desaturated by a plastid localized A9 desaturase to form oleoyl-ACP. Additional desaturation steps are carried out in both the ER and the plastid by membrane bound desaturases. These modifications occur once fatty acids are esterified to a glycerol backbone and are the targets of metabolic engineering as the production of very long-chain polyunsaturated fatty acids in plants is becoming an attractive alternative to natural sources (Truksa et al. 2006). Fatty acid synthesis is completed by cleavage of a 16- or 18-carbon chain from ACP by an acyl-ACP thioesterase (FAT). Two forms of FAT exist, FATA and F ATB, which preferentially cleave unsaturated and saturated acids, respectively. Overexpression of a lauric acid-specific FATB encoding cDNA in Arabidopsis or canola resulted in a dramatic increase in the lauric acid content of seed oil, thus underscoring the importance of FAT enzyme activity in regulating seed oil composition (Voelker et al. 1992, Voelker et al. 1996). Once cleaved from ACP fatty acids are exported from the plastid in an as of yet undetermined manner and are subsequently esterified to CoA in the cytosol. The cytolsol is the site of fatty acid elongation which yields the 20 to 24-carbon very long- chain fatty acids (VLCFAs) present in many seed oils. A homodimeric, eukaryotic-type ACCase and a membrane bound acyl elongation complex participate in elongation in the cytosol and are regulated differently than their plastidic counterparts (Bao et al. 1998). Although, ACCase is still a bottleneck and reducing the activity of the cytosolic enzyme resulted in reduced VLCFA content in Arabidopsis seed oil (Baud et al. 2003). Many other modifications to fatty acids occur, such as hydroxylation, methylation, etc.., and these unusual fatty acids are typically specific to seed oil (V oelker and Kinney 2001). 19 Theei 2003) edhed tutip phosr 200i) totht the 52 (Ilkl isco padr toI) red; 0\ Cf AI—l "It The enzymes responsible for such modifications are being elucidated (e. g. Bao et al. 2002), and progress is being made in understanding how these unusual fatty acids are edited out of membrane lipids (Voelker and Kinney 2001). The assembly of triacyl glycerol from acyl-CoAs and membrane lipids occurs by two pathways: the Kennedy pathway, and the recently described phosphatidylcholine:diacyl glycerol acyltransferase (PDAT) pathway (Dahlqvist et al. 2000). In both pathways, the initial steps are identical. Fatty acids in the ER are esterified to the sn-1 position of a glycerol-3 phosphate backbone followed by a second transfer to the sn-2 position. The resulting phosphatidic acid is then converted to diacylglycerol (DAG) by the action of a phosphatidic acid phosphatase. In the Kennedy pathway, DAG is converted to TAG by the action of DAG acyltransferase (DAGAT). In the PDAT pathway, an acyl chain from the sn-2 position of phosphatidylcholine (PC) is transferred to DAG to form TAG. An Arabidopsis mutant defective in a seed specific DAGAT has reduced seed oil (Katavic et al. 1995, Routaboul et al. 1999, Zou et al. 1997). Conversely, overexpression of the same gene results in enhanced seed oil content (J ako et al. 2001). While no direct role in seed oil biosynthesis has been attributed to any of the six genes encoding PDAT isoforms in Arabidopsis, one is specifically expressed in developing seeds (Stahl et al. 2004). The PDAT pathway may be a means of discriminating TAG- specific fatty acids out of membrane lipids. In any case, the genetic evidence for Arabidopsis points to the DAGAT pathway as the major source of TAG. 1.5.2. Supply of carbon precursors for fatty acid synthesis Plastidic fatty acid synthesis and subsequent elongation in the cytosol are dependent on a 20 SICE C011 cha C0. Ind nea incl cou SO. U steady supply of carbon precursors. Imported photosynthate in the form of sucrose or hexose is the major source of carbon for a developing oil seed and must be broken down to serve this purpose. The initial steps involve glycolysis and the newly discovered RuBisCO shunt, and these will be discussed later. Here, I will address the direct supply of acetyl—CoA for ACCase in the plastid and cytosol. Because acetyl-CoA is membrane impermeable (Liedvogel 1986), its supply will be considered separately for each compartment. The generation of cytosolic acetyl-CoA has only recently been extensively studied in plants. A cytosolic ATP-citrate lyase (ACL) from Arabidopsis has been identified and characterized (Fatland et al. 2002). This enzyme is the main source of cytosolic acetyl- CoA in animals and it was speculated that a similar situation might be occurring in plants. Indeed, when the activity of this enzyme is ablated by expression of antisense cDNA nearly all anabolic processes involving cytosolic acetyl-CoA pools were perturbed, including VLCFA biosynthesis (Fatland et al. 2005). It appeared that no other pathways could compensate for the loss of ACL activityin leaves suggesting that it is the sole source of cytosolic acetyl-CoA in that tissue. The substrate of this enzyme, citrate, is synthesized in the TCA cycle, and ultimately from glycolysis of imported sugars. It should be noted, however, that the synthesis of VLCFA in seeds was relatively unaffected in the ACL antisense plants. Two enzymes exist in plastids which potentially supply acetyl-CoA to ACCase: the pyruvate dehydrogenase complex (PDC), and acetyl-CoA synthetase (ACS). Expression analysis of ACS and components of the PDC in Arabidopsis suggest that ACS makes little contribution to acetyl-CoA production in plastids of developing seeds 21 acid this PD< PD (lo (151: Hi (Ke et al. 2000). Additionally, pyruvate is preferred over acetate as a substrate for fatty acid synthesis in isolated canola embryos (Kang and Rawsthome 1994). When combined this evidence strongly supports the PDC as being the source of plastidic acetyl-CoA. The PDC reaction also generates NADH at a one to one ratio with acetyl-CoA, which as described earlier could support the demands of fatty acid synthesis. One consequence of PDC activity is the release of C02, which results in high concentrations of C02 in developing seeds (Goffinan et al. 2004). The concentration of C02 is potentially not as damaging though as the fact that one third of fixed carbon in pyruvate is lost at this step. However, a refixation shunt involving RuBisCO has been discovered which partially makes up for the loss of carbon by PDC, at least in green seeds (Schwender et al. 2004a). The amount of acetyl-CoA generated by PDC is entirely dependent on the amount of pyruvate available in the plastid. Figure 1.3 depicts pathways that can contribute to the steady state pool of plastidic pyruvate. The most direct route is the import of pyruvate from the cytosol. A plastid localized pyruvate transporter is hypothesized to exist based on the need for pyruvate translocation during C4 carbon fixation. However, no report has yet been made of the transporter’s identity. This pathway would also require a supply of cytosolic PEP and the activity of cytosolic pyruvate kinase (PKC). The activity of PKc is high in developing embryos of castor and soybean (Glycine max) (Turner et al. 2005). Moreover, isolated plastids from canola embryos are capable of incorporating ld’C-labeled pyruvate into fatty acids (Eastmond and Rawsthome 2000, Kang and Rawsthome 1994). However, combined data from metabolic flux analyses of cultured canola embryos estimates that a maximum of 30% of the pyruvate used in fatty acid synthesis is generated in this manner (Schwender et al. 2004b). 22 1h: 0 I v« PEP ADP C) ATP <> '''''''''''''''''''''''''''''''' f. Pyr NADPH NADP+ NADP+ NADH C) malate NADPH Cytosol PlastidJ .J Figure 1.3. Alternative pathways of pyruvate production Arrows represent reactions catalyzed by single enzymes. Gray ovals are plastid envelope transporters. The pyruvate transporter is marked with “2’. 1) cytosolic pyruvate kinase; 2) plastidic pyruvate kinase; 3) PEP carboxylase; 4) NAD-malate dehydrogenase; 5) NADP- malic enzyme. 0AA, oxaloacetate; PEP, phosphoenolpyruvate; Pyr, pyruvate. Alternatively, plastidic pyruvate kinase (PKp) can generate pyruvate directly in the plastid. In this case, plastidic or cytosolic PEP can serve as the substrate. The import of PEP by plastids of canola embryos has been demonstrated (Kubis et al. 2004) and microarray analysis of developing Arabidopsis embryos suggested that most PEP is generated in the cytosol (Ruuska et al. 2002). A mutant of the seed resident PEP transporter (chlorophyll a binding protein underexpressed, cue!) has been identified, but unfortunately has not been analyzed with a direct focus on the role of this transporter in oil biosynthesis (Knappe et al. 2003, Li et al. 1995, Vol] et a1. 2003). A recent steady- state carbon flux analysis on cultured embryos of canola led to the proposal of a 23 RuBi plast react to P] Alih PG! alto cari \i'h': PET an' RuBisCO shunt involving reactions of the reductive pentose phosphate pathway in the plastid (Schwender et al. 2004a). The proposed pathway bypasses the initial glycolytic reactions in the cytosol and generates PGA in the plastid, which could then be converted to PEP in the plastid, which could compensate for import deficiency in the cue] mutant. Although, the labeling studies of Schwender and others (2004a) cannot define whether PGA is converted to PEP in the plastid or cytosol. Figure 1.3 also contains a three enzyme pathway which bypasses pyruvate kinase altogether. In the first step, cytosolic PEP is converted to oxaloacetate (0AA) by PEP carboxylase (PEPC). Then, malate dehydrogenase (MDH) converts 0AA to malate, which can be imported into the plastid and metabolized by NADP-ME to generate pyruvate. Malate is transported across the plastid envelope by a malate/Pi translocator and supports the highest rate of fatty acid synthesis in isolated plastids from castor embryos (Eastmond et al. 1997, Smith et a1. 1992). However, NADP-ME mutants have been isolated from Arabidopsis, and loss of seed specific isoforms has no effect on oil accumulation (Wheeler et al. 2005). Generally, PEPC is thought to serve an anaplerotic role in developing seeds, replenishing TCA intermediates consumed during storage protein biosynthesis (Chollet et al. 1996). The mystery of the source of pyruvate in developing seeds is further complicated by the need for ATP generation in the cytosol or plastid versus the need for the transfer of reducing potential to the plastid (see section 1.4). The pathways illustrated in Figure 1.3 are capable of doing both. The combined evidence leads to the hypothesis that in green oil seeds, the pyruvate kinase pathway(s) are predominant, but that in non-green or protein-storing seeds, the pathway is flexible and may involve PK bypasses. 24 react glue bjp: Fun 1.5.3. Glycolysis in oil seeds The PEP used for the synthesis of pyruvate is ultimately produced by some or all of the reactions of glycolysis. Stable isotope labeling has been used to demonstrate that 90% of glucose fed to developing canola embryos is converted to pyruvate by the RuBisCO bypass and the second half of glycolysis (Schwender et al. 2002, Schwender et al. 2004a). Furthermore, a link between glycolysis and seed oil metabolism is apparent in the wriI mutant of Arabidopsis, in which a general reduction in glycolytic activity results in an 80% reduction in seed oil (F ocks and Benning 1998). Glycolysis in plants occurs in both the cytosol and the plastid and both compartments are connected through plastid membrane transporters (Plaxton 1996, Weber 2004). The glycolytic intermediates of developing canola embryos appear to be in near equilibrium between the cytosol and plastid (Schwender et al. 2003), and activities of the full glycolytic sequence have been detected in both compartments in embryos (Eastrnond and Rawsthome 2000). It is therefore likely that changes in glycolytic enzyme activities influence the transport of related intermediates across the plastid envelope. This compartmentation raises the question of a preferred route of glucose metabolism in developing oil seed embryos. Expressed sequence tag (EST) analysis of developing Arabidopsis seeds indicated that mRNAs encoding cytosolic enzymes for the entire glycolytic pathway are abundant, but that only mRNAs encoding plastidic enzymes for the second half of the pathway metabolizing trioses are abundant (White et al. 2000). The recently discovered RuBisCO shunt (Figure 1.2, (Schwender et al. 2004a) bypasses the initial reactions of glycolysis in the plastid and could explain the dearth of ESTs for these genes. Microarray data from Arabidopsis extended these findings by detecting a shift in expression from genes 25 encoding cytosolic glycolytic enzymes to those encoding plastidic glycolytic enzymes at the onset of storage compound accumulation (Ruuska et al. 2002). The regulation of glycolysis in oil seeds occurs at least on the transcriptional and metabolic levels. The WRII transcription factor has been shown to induce glycolytic gene expression in germinating seedlings and the same trans-activation is thought to occur in developing seeds (Cemac and Benning 2004). Specifically, the activities of PPi- dependent phosphofructokinase (PFP), enolase, and pyruvate kinase are reduced in the wri] seeds (F ocks and Benning 1998). Furthermore, genes encoding all three enzymes are coordinately expressed with WRII in developing seeds (Schmid et al. 2005). These activities represent the most highly regulated or regulatory steps of plant glycolysis and are shown in Figure 1.4. The most highly regulated step in Figure 1.4 is the conversion of fructose 6-phosphate to fructose 1,6-bisphosphate by the ATP- and PPi-dependent phosphofructokinases (PFK and PFP, respectively). In Arabidopsis PFK activity is high early during embryo morphogenesis while PF P activity is induced during the phase of storage compound accumulation (Baud and Graham 2006). The inhibition of these enzymes (directly or indirectly) by PEP constitutes the bottom-up regulation of glycolysis in plants (Plaxton 1996) and, at least for PFP, this inhibition may be intensified by the hypoxic conditions present in developing seeds (Podesta and Plaxton 2003). Furthermore, these reactions are both reversible and may have roles in controlling the partition of carbon between starch biosynthesis and the production of precursors for fatty acid synthesis. The activities of enolase and PK can have great influence over the concentration of the regulatory metabolite PEP. Little is known about the enolase present in seed tissues, and for that matter, from plants at all. An enolase encoding gene is 26 upregulated during hypoxic conditions in rice and may suggest a mechanism by which this enzyme is induced in seed tissues (Umeda and Uchimiya 1994). On the other hand, much is known about pyruvate kinase and given its potential importance in seed metabolism it will be discussed separately. 1‘ GGP G6P l .l. PFK PFP H 3:113:00 A A F1,6-P2 : E H Plastid C? C? It PEP : 3PGA F6P—>F2,6'-P2 I ll F6P,2K if 2PGA l 2PGA 551qu l ,ENOcH PKp . '2 PEP—>Pyr JJ Figure 1.4. The glycolytic network in green oil seeds Enzymes and transporters are shown as gray ovals. Individual arrows represent single reactions. Dashed lines represent metabolite based regulation. 2PGA, 2- phosphoglycerate; 3PGA, 3-phosphoglycerate; ENO, enolase; F 1,6-P2, fructose 1,6- bisphosphate; F2,6-P2, fructose 2,6-bisphosphate; F6P, fructose 6-phosphate; F6P,2K, fructose 6-phosphate 2 kinase; G6P, glucose 6-phosphate; PEP, phosphoenolpyruvate; PFP, pyrophosphate-dependent phosphofi'uctokinase; Pyr, pyruvate. 1.6. Pyruvate kinase 27 Pyruvate kinase catalyzes the conversion of pyruvate to PEP, coupled to the substrate level phosphorylation of ADP to generate ATP (Figure 1.5). In most eukaryotes studied PK is cytosolic, but in plants the enzyme also occurs in the plastid. Plant PK activities arise from the expression of multiple isozymes with different biochemical properties that depend on the tissue and plant source. Oxaloacetate m PE P mm Shikimic Acid TCA Cycle ADP Aromatic Amino Acids Amino Acids ‘ Phenolic Compounds PK lndole Compounds all? .thawwruw. Pyr-U vate II emu-“W““v‘ valine Alanine Leucine \ Lysine Acetyl CoA lsoprenoids (MEP pathway) Fatty Acids TCA Cycle lsoprenoids (Mevolnate pathway) Figure 1.5. Pyruvate kinase centrality to plant metabolism Solid arrows represent direct metabolic conversions. Metabolic pathways are represented by broken arrows. MEP, methyl-erythritol phosphate; PEP, phosphoenolpyruvate; PK, pyruvate kinase. 28 ArabidOpsis, for instance, has 14 annotated PK genes that likely exhibit a large degree of variation with respect to regulation of gene expression and enzyme activity (Arabidopsis Genome Initiative 2000). The potential influence on the energy status, glycolytic activity, and fatty acid precursor pool in seeds. alone makes PK an interesting enzyme to study. However, as shown in Figure 1.5, PK lies at a crossroad and its activity may have an impact on nearly all of plant metabolism. 1.6.1. Pyruvate kinase phylogeny Pyruvate kinase is an ancient enzyme and occurs in all kingdoms of life. Previous phylogenetic analyses of diverse PK nucleotide (Munoz and Ponce 2003) or amino acid sequences (Hattori et al. 1995, Oria-Hemandez et al. 2006, Schramm et al. 2000) revealed common clustering patterns. Two major families of PK exist; one contains animal, fungal, some bacterial, and plant cytosolic enzymes, while the other includes archaebacterial, cyanobacterial, and plant plastidic enzymes. This segregation suggests distinct ancestral origins of the cytosolic and plastidic enzymes. Most notably, a cyanobacterial enzyme appears to be the origin of plant plastidic PK, which may be a result of the endosymbiotic event which gave rise to plastids. In mammals, two genes encode four separate PK isoenzymes by way of alternative splicing (Noguchi et al. 1986, Satoh et a1. 1988). The phenomenon has not been documented in other systems but raises the possibility of much more PK enzyme diversity than can be assessed by gene sequence alone. 29 1.6.2. Pyruvate kinase molecular architecture In all but one of the organisms studied to date, PK exists as a protein complex. Most non- plant PKs exist as homotetrarners with individual subunits having molecular masses of 56-60 kDa (Muirhead 1990, Munoz and Ponce 2003). The subunit structure of the plant enzymes differs and is somewhat ambiguous. Purified cytosolic PK (PKc) from germinating castor seeds is an (1202 heterotetramer with subunits of 56- and 57 kDa (Plaxton 1988). Leaf PKc from castor has been found as both a homo- and heterotetramer and the enzymes from developing seeds and cotyledons are homoteteramers (Hu and Plaxton 1996, Plaxton 1989). Analyses of PKcs from canola are consistent with this heterogeneity. In one study, PKcs from developing and germinating canola seeds were isolated as heterotetramers with subunits the proportions of which differed between the tissues (Sangwan et al. 1992). However, the enzyme from a suspension cell culture was later shown to be a homotetramer (Smith et a1. 2000). Plastidic PKs (PKp) are also non-unifonn in their subunit organization and in one case independent analyses of the same enzyme have resulted in conflicting conclusions. The PKp from developing castor endosperm is composed of a and B subunits. Studies of recombinant versions of these proteins led to the conclusion that PKp-a and PKp-B are distinct enzymes (Blakeley et a1. 1995, Blakeley and Dennis 1993, Wan et al. 1995). However, others have determined that a native version of the same enzyme is a heterohexarner composed of both subunits (Negm et al. 1995, Plaxton et al. 1990, Plaxton 1991). Analysis of leucoplast PK from canola suspension cells revealed a single heterohexameric enzyme composed of equal amounts of a and [3 subunits (Plaxton et al. 2002). Combined, the data from canola and castor suggest PKp is normally a 30 heterohexamer of (1303 stoichiometry. However, the relative expression and actual amount of the individual subunits varies depending on the tissue (Blakeley et al. 1995, Sangwan et al. 1992) and may indicate variable subunit stoichiometry. 1.6.3 Mechanisms of regulation Pyruvate kinase enzyme activity is dependent on monovalent and divalent cations , with K+ and Mg2+ typically filling the role (Munoz and Ponce 2003). There are some K+- independent enzymes and phylogenic analysis correlates this feature with specific amino acid residues and suggests that K+-dependence evolved from K+-independent enzymes (Oria-Hemandez et al. 2006). The evolution of Kl-dependence is thought to have been driven by the fact that K+ facilitates active conformation acquisition and subsequent binding of ADP (Oria-Hernandez et al. 2005). Most non-plant PK’s are allosterically regulated, often involving activation by adenosine monophosphate (AMP), fi'uctose 1,6-bisphosphate and other hexose metabolism intermediates. The contact region between subunits is responsible for the allostery and mutation of a single amino acid in this region is sufficient to modify the allosteric properties of an enzyme (Ikeda et al. 1997, Valentini et al. 2000). Plant PKs are typically insensitive to allosteric regulation by FBP but are regulated more by central carbon metabolites and pH effects (Plaxton 1996). Plant cytosolic PKs typically have pH optima of approximately pH 7.0, whereas plastidic PKs are most active at pH 8.0 (Hu and Plaxton 1996, Plaxton et al. 2002, Smith et al. 2000). The most common metabolite regulators of plant PK activity are tricarboxylic acid cycle intermediates and amino acids, and they typically act as inhibitors (Plaxton and Podesta 2006). In general, non-plant 31 enzymes are activated by upstream metabolites, whereas plant PKs are inhibited by downstream metabolites. This generalization reflects the role of PK in the respective organisms. In non-plants, PK functions primarily in carbohydrate catabolism and the production of energy. In plants energy can be produced by photosynthesis and the role of PK is mainly to supply various metabolic pathways with carbon precursors. In summary, non-plant enzymes are activated when energy is needed and plant enzymes are inhibited when carbon precursors are abundant. The multimeric structure of PK allows for additional levels of regulation. Catalysis only occurs when individual subunits associate and an active complex is formed. In vitro, this can be stimulated by the addition of polyethylene glycol, resulting in enzyme activation (Podesta and Plaxton 1993). This suggests that in vivo, association and dissociation of subunits may influence PK activity. Each PK subunit has intrinsic characteristics that confer regulatory properties to the complexes in which they are found. Thus, the potential for multiple subunit stoichiometries for heteromeric enzymes introduces the possibility of multiple regulatory states. The novel regulatory properties of PK hybrids formed by combining subunits of normally homomeric mammalian enzymes provides proof of this concept (Dyson and Cardenas 1973, Hubbard and Cardenas 1975). Individual subunits can be further regulated by covalent modification. Marrnnalian liver PK is inhibited by phosphorylation on serine residues and recently a soybean PKc was shown to be targeted for degradation by phosphorylation (Munoz and Ponce 2003, Tang et al. 2003). The availability of a particular subunit for complex formation is also controlled by effects on gene expression, translation, compartmentation, and protein turnover and these (and other) factors likely help govern PK activity in any given tissue. 32 1.6.4 Analysis of pyruvate kinase mutants Pyruvate kinase deficiency occurs or has been induced in many organisms but only one case has been reported in plants. In humans, PK deficiency is the most common glycolytic defect and is the leading cause of hereditary non spherocytic haemolytic anaemia with 180 known mutations which result in the disease (Zanella et al. 2007). Erythrocytes are entirely dependent on glycolysis for ATP production and so a severe reduction in PK activity can result in death at infancy (Zanella et al. 2007). In yeast (Saccharomyces cerevisiae), PK deficiency results in an inability to grow on glucose or other fermentable sugars (Sprague, Jr. 1977). Interestingly though, and counter intuitively, flux through the tricarboxylic acid cycle is stimulated in the absence of PK activity (Pearce et al. 2001 ). A similar glucose non-fermentable phenotype is observed in Escherichia coli PK mutants (Pertierra and Cooper 1977). It is clear from the above cases that PK plays a pivotal role in regulating glycolytic flux and energy production in cells lacking respiratory capacity. Pyruvate kinase deficiency has been induced in tobacco (Nicotiana tabacum) inadvertently by ectopic expression of a potato (Solanum tuberosum) tuber PKc fused to a chloroplast transit peptide (Gottlob-McHugh et al. 1992). Instead of increasing plastid localized PK activity, PKc was silenced. As a result, root biomass was reduced relative to the shoot suggesting an altered sourcezsink relationship (Knowles et al. 1998). Further analysis revealed impaired export of photosynthate from leaves at night along with a higher rate of respiration (Grodzinski et al. 1999). Whether the reasons for higher respiration are the same as those for yeast and bacterial PK mutants is unknown. 33 1.6.5. Pyruvate kinase moonlighting Ancient enzymes such as PK often have roles in addition to their primary enzymatic function. Hexokinasel from Arabidopsis, for instance, has recently been identified as a glucose sensor and can translocate into the nucleus where it mediates expression of sugar responsive genes (Cho et a1. 2006, Moore et a1. 2003). Another glycolytic enzyme, enolase, acts as transcriptional regulator in response to cold stress in Arabidopsis (Lee et al. 2002). Most recently, a mammalian PK was found to translocate into the nucleus in response to interleukin-3 stimulation, where it promotes cell proliferation (Hoshino et al. 2007). Several studies using other systems find PK in other non-traditional roles. Using a rat cDNA library and a KATP channel as bait in a yeast two-hybrid experiment, PK along with glyceraldehyde-3 -phosphate dehydrogenase and triose-phosphate isomerase were found to form a complex which regulates flux of K+ (Dhar-Chowdhury et al. 2005). Another study identified PK as a lysophosphatidic acid (LPA) binding protein (Desmaret et al. 2005). PK activity was inhibited upon binding LPA, implying LPA as a signaling molecule which controls metabolism. No moonlighting functions have yet been determined for a plant PK. 1.7. Rationale and outlook Modeling the seed metabolic network based on transcript abundance and metabolic flux analysis places PKp at an important node that connects catabolic and anabolic processes (Figure 1.2, Figure 1.5). A hypothesis based on this model is that most of the carbon precursors used in fatty acid synthesis are produced in the plastid by PKp, and that PKp also provides energy to the developing embryo in the form of ATP. While native PKps 34 have been isolated and described from castor and canola, little has been done to directly test the function of these enzymes in vivo. Use of the model plant Arabidopsis makes investigating the connection between in vitro properties and in vivo function readily doable. Embryo tissue is not abundant in Arabidopsis but the availability of the entire genome sequence makes possible the rapid cloning of putative PK genes and heterologous production of the respective proteins. Publicly available gene expression data can help this process by hastening the identification of those genes which are actively transcribed in the tissue of interest. By this method I have identified and expressed in E. coli the genes which encode PKps involved in seed metabolism, in lieu of purifying the native enzyme from embryo tissue. Molecular and kinetic characterizations of the recombinant proteins indicated that they perform their annotated function and were used to make predictions about how they function in vivo. A shortcoming of previous studies of seed resident PKps is the lack of any genetic evidence to confirm any such predictions. To this end, I took advantage of the Arabidopsis T-DNA insertion mutant population. Plants that are impaired in PKID activity were used to test the hypothesis that PKp is crucial for the production of precursors for fatty acid synthesis in green oil seeds. Further analysis of the mutants also lead to the discovery of novel functions for PKp. By taking such a two-prong approach, I was able to draw conclusions that unite in vitro and in vivo data for the synthesis of a more complete picture of seed oil metabolism. Finally, I will outline some strategies for the application of my findings towards metabolic engineering for the production of valuable phytochemicals. 35 References Alonso, A.P., Goffman, F.D., Ohlrogge, J.B., and Shachar-Hill, Y., (2007) Carbon conversion efficiency and central metabolic fluxes in developing sunflower (Helianthus annuus L.) embryos. Plant Physiol. In Press. Appeldoorn, N.J.G., deBruijn, S.M., KootGronsveld, E.A.M., Visser, R.G.F., Vreugdenhil, D., and vanderPlas, L.H.W. (1997) Developmental changes of enzymes involved in conversion of sucrose to hexose-phosphate during early tuberisation of potato. Planta 202:220-226. Arabidopsis Genome Initiative (2000) Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature 408:796-815. Asokanthan, P.S., Johnson, R.W., Griffith, M., and Krol, M. (1997) The photosynthetic potential of canola embryos. Physiologia Plantarum 101:353-360. Bao, X., Katz, S., Pollard, M., and Ohlrogge, J. (2002) Carbocyclic fatty acids in plants: biochemical and molecular genetic characterization of cyclopropane fatty acid synthesis of Sterculiafoetida. Proc. Natl. Acad. Sci. USA. 99:7172-7177. Bao, X., Pollard, M., and Ohlrogge, J. (1998) The biosynthesis of erucic acid in developing embryos of brassica rapa. Plant Physiol. 118:183-190. Barrett, RB. and Harwood, J.L. (1998) Characterization of fatty acid elongase enzymes from germinating pea seeds. Phytochemistry 48:1295-1304. Baud, S., Boutin, J ., Miquel, M., Lepiniec ,L., and Rochat, C. (2002) An integrated overview of seed development in Arabidopsis thaliana ecotype WS. Plant Physiol. Biochem. 40:151-160. Baud, S. and Graham, LA. (2006) A spatiotemporal analysis of enzymatic activities associated with carbon metabolism in wild-type and mutant embryos of Arabidopsis using in situ histochemistry. Plant J. 46: 1 55-169. Baud, S., Guyon, V., Kronenberger, J., Wuilleme, S., Miquel, M., Caboche, M., Lepiniec, L., and Rochat, C. (2003) Multifunctional acetyl-CoA carboxylase 1 is essential for very long chain fatty acid elongation and embryo development in Arabidopsis. Plant J. 33:75-86. Berger, F. (1999) Endosperm development. Curr. Opin. Plant Biol. 2:28-32. Berger, F. (2003) Endosperm: the crossroad of seed development. Curr. Opin. Plant Biol. 6:42-50. Blakeley, S.,,Gottlob-McHugh, S., Wan, J., Crews, L., Miki, B., Ko, K., and Dennis, D.T. (1995) Molecular characterization of plastid pyruvate kinase from castor and tobacco. Plant Mol. Biol. 27:79-89. 36 Blakeley, SD. and Dennis, D.T. (1993) Molecular approaches to the manipulation of carbon allocation in plants. Can. J. Bot. 71:765-778. Borisjuk, L., Rolletschek, H., Wobus, U., and Weber, H. (2003) Differentiation of legume cotyledons as related to metabolic gradients and assimilate transport into seeds. J. Exp. Bot. 54:503-512. Caughey, I. and Kekwick, R.G. (1982) The characteristics of some components of the fatty acid synthetase system in the plastids from the mesocarp of avocado (Persea americana) fruit. Eur. J. Biochem. 1233553-561. Cernac, A. and Benning, C. (2004) WRINKLEDI encodes an AP2/EREB domain protein involved in the control of storage compound biosynthesis in Arabidopsis. Plant J. 40:575-585. Cho, Y.H., Yoo, S.D., and Sheen, J. (2006) Regulatory functions of nuclear hexokinasel complex in glucose signaling. Cell 127:579-589. Chollet, R., Vidal, J., and O'Leary, M.H. (1996) Phosphoenolpyruvate carboxylase: A ubiquitous, highly regulated enzyme in plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 47 :273-298. Crouch, M.L. and Sussex, LA. (1981) Development and storage-protein synthesis in Brassica napus L. embryos in vivo and in vitro. Planta 153264-74. Da Silva, P.M.F.R, Eastmond, P.J., Hill, L.M., Smith, A.M., and Rawsthome, S. (1997) Starch metabolism in developing embryos of oilseed rape. Planta 203:480-487. Dahlqvist, A., Stahl, U., Lenman, M., Banas, A., Lee, M., Sandager, L., Ronne, H., and Stymne, S. (2000) Phospholipidzdiacylglycerol acyltransferase: an enzyme that catalyzes the acyl-CoA-independent formation of triacylglycerol in yeast and plants. Proc. Natl. Acad. Sci. USA. 97 :6487-6492. Dean Rider S. Jr, Henderson, J.T., Jerome, R.E., Edenberg, H.J., Romero-Severson, J., and Ogas, J. (2003) Coordinate repression of regulators of embryonic identity by PICKLE during germination in Arabidopsis. Plant J. 35:33-43. Desmaret, S., Qian, L., Vanloo, B., Meerschaert, K., Van Damme, J., Grooten, J., Vandekerckhove, J ., Prestwich, G.D., and Gettemans, J. (2005) Lysophosphatidic acid affinity chromatography reveals pyruvate kinase as a specific LPA-binding protein. Biol. Chem. 386:1137-1147. Dhar-Chowdhury, P., Harrell, M.D., Han, S.Y., Jankowska, D., Parachuru, L., Morrissey, A., Srivastava, S., Liu, W., Malester, B., Yoshida, H., and Coetzee, W.A. (2005) The glycolytic enzymes, g1yceraldehyde-3-phosphate dehydrogenase, triose- phosphate isomerase, and pyruvate kinase are components of the K(ATP) channel macromolecular complex and regulate its function. J. Biol. Chem. 280:38464-3 8470. 37 Dyson, R.D. and Cardenas, J.M. (1973) Bovine pyruvate kinases. 3. Hybrids of the liver and skeletal muscle isozymes. J. Biol. Chem. 248:8482-8488. Eastmond, P., Kolacna, L., and Rawsthome, S. (1996) Photosynthesis by developing embryos of oil seed rape (Brassica napus L.). J. Exp. Bot. 47: 1 763-1 769. Eastmond, P.J., Dennis, D.T., and Rawsthome, S. (1997) Evidence that a malate/inorganic phosphate exchange translocator imports carbon across the leucoplast envelope for fatty acid synthesis in developing castor seed endosperm. Plant Physiol. 114:851-856. Eastmond, P.J. and Rawsthome, S. (2000) Coordinate changes in carbon partitioning and plastidial metabolism during the development of oilseed rape embryos. Plant Physiol. 1221767-774. Eastmond, P.J., van Dijken, A.J., Spielman, M., Kerr, A., Tissier, A.F., Dickinson, H.G., Jones, J.D., Smeekens, S.C., and Graham, LA. (2002) Trehalose-6-phosphate synthase 1, which catalyses the first step in trehalose synthesis, is essential for Arabidopsis embryo maturation. Plant J. 29:225-235. Fatland, B.L., Ke, J., Anderson, M.D., Mentzen, W.I., Cui, L.W., Allred, C.C., Johnston, J .L., Nikolau, B.J., and Wurtele, E.S. (2002) Molecular characterization of a heteromeric ATP-citrate lyase that generates cytosolic acetyl-coenzyme A in Arabidopsis. Plant Physiol. 1302740-756. Fatland, B.L., Nikolau, B.J., and Wurtele, E.S. (2005) Reverse genetic characterization of cytosolic acetyl-CoA generation by ATP-citrate lyase in Arabidopsis. Plant Cell 17:182-203. Finkelstein, R. and Somerville, CR. (1990) Three classes of absisic acid (ABA)- insensitive mutations of Arabidopsis define genes that control overlapping subsets of of ABA resposnes. Plant Physiol. 94:1172-1179. Focks, N. and Benning, C. (1998) wrinkled]: A novel, low-seed-oil mutant of Arabidopsis with a deficiency in the seed-specific regulation of carbohydrate metabolism. Plant Physiol. 118:91-101. Galili, G., Sengupta-Gopalan, C., and Ceriotti, A. (1998) The endoplasmic reticulum of plant cells and its role in protein maturation and biogenesis of oil bodies. Plant Mol. Biol. 38: 1 -29. Girke, T., Todd, J., Ruuska, S., White, J ., Benning, C., and Ohlrogge, J. (2000) Microarray analysis of developing Arabidopsis seeds. Plant Physiol. 124: 1 570-1 581. Goffman, F.D., Alonso, A.P., Schwender, J., Shachar-Hill, Y., and Ohlrogge, J.B. (2005) Light enables a very high efficiency of carbon storage in developing embryos of rapeseed. Plant Physiol. 138:2269-2279. 38 Goffman, F.D., Ruckle, M., Ohlrogge, J., and Shachar-Hill, Y. (2004) Carbon dioxide concentrations are very high in developing oilseeds. Plant Physiol. Biochem. 42:703-708. Goldberg, R.B., de Paiva, G., and Yadegari, R. (1994) Plant embryogenesis: zygote to seed. Science 266:605-614. Gomez, L.D., Baud, S., Gilday, A., Li, Y., and Graham, LA. (2006) Delayed embryo development in the Arabidopsis trehalose-6-phosphate synthase 1 mutant is associated with altered cell wall structure, decreased cell division and starch accumulation. Plant J. 46:69-84. Gottlob-McHugh, S.G., Sangwan, R.S., Blakeley, S.D., Vanlerberghe, G.C., Ko, K., Turpin, D.H., Plaxton, W.C., Miki, B.L., and Dennis, D.T. (1992) Normal growth of transgenic tobacco plants in the absence of cytosolic pyruvate kinase. Plant Physiol. 100:820-825. Grodzinski, B., Jiao, J ., Knowles, V.L., and Plaxton, W.C. (1999) Photosynthesis and carbon partitioning in transgenic tobacco plants deficient in leaf cytosolic pyruvate kinase. Plant Physiol. 120:887-896. Hattori, J ., Baum, B.R., Mchugh, S.G., Blakeley, S.D., Dennis, D.T., and Miki, B.L. (1995) Pyruvate kinase isozymes: Ancient diversity retained in modern plant cells. Biochem. Syst. Ecol. 23:773-&. Hemandez-Sebastia, C., Marsolais, F., Saravitz, C., Israel, D., Dewey, R.E., and Huber, S.C. (2005) Free amino acid profiles suggest a possible role for asparagine in the control of storage-product accumulation in developing seeds of low- and hi gh-protein soybean lines. J. Exp. Bot. 56:1951-1963. Hill, L.M., Morley-Smith, E.lR., and Rawsthome, S. (2003) Metabolism of sugars in the endosperm of developing seeds of oilseed rape. Plant Physiol. 1312228-236. Hoshino, A., Hirst, J .A., and Fujii, H. (2007) Regulation of cell proliferation by interleukin-3-induced nuclear translocation of pyruvate kinase. J. Biol. Chem. 282:17706-17711. Hu, Z.Y. and Plaxton, W.C. (1996) Purification and characterization of cytosolic pyruvate kinase from leaves of the castor oil plant. Arch. Biochem. Biophys. 3332298- 307. Hubbard, D.R. and Cardenas, J .M. (1975) Kinetic properties of pyruvate kinase hybrids formed with native type L and inactivated type M subunits. J. Biol. Chem. 250:4931-4936. Hunter, S.C. and Ohlrogge, J.B. (1998) Regulation of spinach chloroplast acetyl-CoA carboxylase. Arch. Biochem. Biophys. 359:170-178. 39 Ikeda, Y., Tanaka, T., and Noguchi, T. (1997) Conversion of non-allosteric pyruvate kinase isozyme into an allosteric enzyme by a single amino acid substitution. J Biol. Chem. 272:20495-20501. Jako, C., Kumar, A., Wei, Y., Zou, J., Barton, D.L., Giblin, E.M., Covello, P.S., and Taylor, D.C. (2001) Seed-specific over-expression of an Arabidopsis cDNA encoding a diacylglycerol acyltransferase enhances seed oil content and seed weight. Plant Physiol. 1262861-874. ‘ Jang, J.C., Leon, P., Zhou, L., and Sheen, J. (1997) Hexokinase as a sugar sensor in higher plants. Plant Cell 9:5-19. Jang, J.C. and Sheen, J. (1994) Sugar sensing in higher plants. Plant Cell 6: 1665-1679. Jaworski, J.C., Clough, R.C., and Barnum, SR. (1989) A Cerulenin insensitive short chain 3-Ketoacyl-Acyl Carrier Protein Synthase in Spinacia oleracea leaves. Plant Physiol. 90:41-44. Kang, F. and Rawsthome, S. (1994) Starch and fatty acid biosynthesis in plastids from developing emrbyos of oil seed rape. Plant J. 6:795-805. Karssen, C.M., Brinkhorst-van der Swan, D., Breekland, A., and Koornneef, M. (1983) Induction of dormancy during seed development by endogenous abscisic acid: studies on abscisic acid deficient genotypes of Arabidopsis thaliana (L.) Heynh. Planta 157:158-165. Katavic, V., Reed, D.W., Taylor, D.C., Giblin, E.M., Barton, D.L., Zou, J., MacKenzie, S.L., Covello, P.S., and Kunst, L. (1995) Alteration of seed fatty acid composition by an ethyl methanesulfonate- induced mutation in Arabidopsis thaliana affecting diacyl glycerol acyltransferase activity. Plant Physiol. 108:399-409. Kc, J., Behal, R.H., Back, S.L., Nikolau, B.J., Wurtele, E.S., and Oliver, DJ. (2000) The role of pyruvate dehydrogenase and acetyl-coenzyme A synthetase in fatty acid synthesis in developing Arabidopsis seeds. Plant Physiol. 123:497-508. King, S.P., Badger, M.R., and Furbank, R.T. (1998) C02 refixation chracteristics of developing canola seeds and silique walls. Aust. J. Plant Physiol. 25:377-3 86. Knappe, S., Lottgert, T., Schneider, A., Vol], L., Flugge, U.I., and Fischer, K. (2003) Characterization of two functional phosphoenolpyruvate/phosphate translocator (PPT) genes in Arabidopsis--AtPPT1 may be involved in the provision of signals for correct mesophyll development. Plant J. 36:411-420. Knowles, V.L., Mchugh, S.G., Hu, Z., Dennis, D.T., Miki, B.L., and Plaxton, W.C. (1998) Altered grth of transgenic tobacco lacking leaf cytosolic pyruvate kinase. Plant Physiol. 116245-51. 40 Koornneef, M., Hanhart, C.J., Hilhorst, H.W., and Karssen, GM. (1989) In vivo inhibition of seed development and reserve protein accumulation in recombinants of abscisic acid biosynthesis and responsiveness mutants in Arabidopsis thaliana. Plant Physiol. 90:463-469. Kubis, S.E., Pike, M.J., Everett, C.J., Hill, L.M., and Rawsthome, S. (2004) The import of phosphoenolpyruvate by plastids from developing embryos of oilseed rape, Brassica napus (L.), and its potential as a substrate for fatty acid synthesis. J. Exp. Bot. 55:1455-1462. Laux, T. and Jurgens, G. (1997) Embryogenesis: A new start in life. Plant Cell 9:989- 1000. Lee, H., Guo, Y., Ohta, M., Xiong, L., Stevenson, B., and Zhu, J.K. (2002) L082, a genetic locus required for cold-responsive gene transcription encodes a bi-functional enolase. EMBO J. 21:2692-2702. Lemaire, S.D., Guillon, B., Le Marechal, P., Keryer, E., Miginiac-Maslow, M., and Decottignies, P. (2004) New thioredoxin targets in the unicellular photosynthetic eukaryote Chlamydomonas reinhardtii. Proc. Natl. Acad. Sci. USA. 101:7475-7480. Leprince, 0., Bronchart, R., and Deltour, R. (1990) Changes in starch and soluble sugars in relation to the acquisition of desiccation tolerance during maturation of Brassica campestris seed. Plant Cell Environ. 13:539-546. Li, H.M., Culligan, K., Dixon, R.A., and Chory, J. (1995) CUEl - A mesophyll cell- specific positive regulator of light-controlled gene-expression in Arabidopsis. Plant Cell 721599-1610. Liedvogel, B. (1986) Acetyl Coenzyrne-A and isopentenylpyrophosphate as lipid precursors in plant cells - biosynthesis and compartmentation. J. Plant Physiol. 1242211- 222. Lin, Y., CIuette-Brown, J.E., and Goodman, HM. (2004) The peroxisome deficient Arabidopsis mutant sseI exhibits impaired fatty acid synthesis. Plant Physiol. 135:814- 827. Lin, Y., Sun, L., Nguyen, L.V., Rachubinski, R.A., and Goodman, HM. (1999) The Pex16p homolog SSE] and storage organelle formation in Arabidopsis seeds. Science 284:328-330. Lotan, T., Ohto, M., Yee, K.M., West, M.A., Lo, R., Kwong, R.W., Yamagishi, K., Fischer, R.L., Goldberg, R.B., and Harada, J.J. (1998) Arabidopsis Leafy Cotyledon 1 is sufficient to induce embryo development in vegetative cells. Cell 93:1195-1205. Marillia, E.F., Micallef, B.J., Micallef, M., Weninger, A., Pedersen, K.K., Zou, J., and Taylor, D.C. (2003) Biochemical and physiological studies of Arabidopsis thaliana 41 transgenic lines with repressed expression of the mitochondrial pyruvate dehydrogenase kinase. J. Exp. Bot. 54:259-270. Meinke, D.W., Franzmann, L.H., Niclde, T.C., and Yeung, EC. (1994) Leafy cotyledon mutants of Arabidopsis. Plant Cell 6:1049-1064. Moore, B., Zhou, L., Rolland, F., Hall, Q., Cheng, W.H., Liu, Y.X., Hwang, 1., Jones, T., and Sheen, J. (2003) Role of the Arabidopsis glucose sensor HXKl in nutrient, light, and hormonal signaling. Science 300:332-336. Muirhead, H. (1990) Isoenzymes of pyruvate kinase. Biochem. Soc. Trans. 18:193-196. Munoz, M.E. and Ponce, E. (2003) Pyruvate kinase: current status of regulatory and functional properties. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 1351197-218. Negm, F.B., Cornel, F.A., and Plaxton, W.C. (1995) Suborganellar localization and molecular characterization of nonproteolytic degraded leukoplast pyruvate kinase from developing castor oil seeds. Plant Physiol. 109:1461-1469. N oguchi, T., Inoue, H., and Tanaka, T. (1986) The M1- and M2-type isozymes of rat pyruvate kinase are produced from the same gene by alternative RNA splicing. J. Biol. Chem. 261:13807-13812. Norton, G. and Harris, J.F. (1975) Compositional changes in developing rape seed (Brassica napus L.). Planta 123:163-174. Ogas, J., Cheng, J.C., Sung, Z.R., and Somerville, C. (1997) Cellular differentiation regulated by gibberellin in the Arabidopsis thaliana pickle mutant. Science 277:91-94. Ogas, J., Kaufmann, 8., Henderson, J., and Somerville, C. (1999) Pickle is a CHD3 chromatin-remodeling factor that regulates the transition from embryonic to vegetative deve10pment in Arabidopsis. Proc. Natl. Acad. Sci. USA. 96:13839-13844. Ohlrogge, J. and Browse, J. (1995) Lipid biosynthesis. Plant Cell 71957-970. Ohlrogge,J.B., Kuhn, lD.N., and Stumpf, P.K. (1979) Subcellular localization of acyl carrier protein in leaf protoplasts of Spinacia oleracea. Proc. Natl. Acad. Sci. USA. 76:1194-1198. Ohto, M.A., Fischer, R.L., Goldberg, R.B., Nakamura, K, and Harada, J.J. (2005) Control of seed mass by Apetala 2. Proc. Natl. Acad. Sci. USA. 102:3123-3128. Oria-Hernandez, J., Cabrera, N ., Perez-Montfort, R., and Ramirez-Silva, L. (2005) Pyruvate kinase revisited: the activating effect of K+. J. Biol. Chem. 280:37924-37929. Oria-Hernandez, J., Riveros-Rosas, H., and Ramirez-Silva, L. (2006) Dichotomic phylogenetic tree of the pyruvate kinase family: K+ -dependent and -independent enzymes. J. Biol. Chem. 281:30717-30724. 42 Parcy, F ., Valon, C., Kohara, A., Misera, S., and Giraudat, J. (1997) The Abscisic Acis-Insensitive3, Fusca3, and Leafy Cotyledon] loci act in concert to control multiple aspects of Arabidopsis seed development. Plant Cell 921265-1277. Pearce, A.K., Crimmins, K., Groussac, E., Hewlins, M.J., Dickinson, J.R., Francois, J., Booth, T.R., and Brown, AJ. (2001) Pyruvate kinase (Pykl) levels influence both the rate and direction of carbon flux in yeast under fermentative conditions. Microbiology 1472391-401. ‘ Penfield, S., Rylott, E.L., Gilday, A.D., Graham, 8., Larson, T.R., and Graham, LA. (2004) Reserve mobilization in the Arabidopsis endosperm fuels hypocotyl elongation in the dark, is independent of abscisic acid, and requires Phosphoenolpyruvate Carboxylasel. Plant Cell 16:2705-2718. Periappuram, C., Steinhauer, L., Barton, D.L., Taylor, D.C., Chatson, B., and Zou, J. (2000) The plastidic phosphoglucomutase from Arabidopsis. A reversible enzyme reaction with an important role in metabolic control. Plant Physiol. 122:1193-1199. Pertierra, A.G. and Cooper, RA. (1977) Pyruvate formation during the catabolism of simple hexose sugars by Escherichia coli: studies with pyruvate kinase-negative mutants. J. Bacteriol. 129:1208-1214. Plaxton, W.C. (1988) Purification of pyruvate kinase fiom germinating castor bean endosperm. Plant Physiol. 86: 1064-1069. Plaxton, W.C. (1989) Molecular and immunological characterization of plastid and cytosolic pyruvate kinase isozymes from castor-oil-plant endosperm and leaf. Eur. J. Biochem. 181:443-451. Plaxton, W.C. (1991) Leucoplast pyruvate kinase from developing castor oil seeds : characterization of the enzyme's degradation by a cysteine endopeptidase. Plant Physiol. 97:1334-1338. Plaxton, W.C. (1996) Organization and regulation of plant glycolysis. Annu. Rev. Plant Physiol. Plant Mol. Biol. 47:185-214. Plaxton, W.C., Dennis, D.T., and Knowles, V.L. (1990) Purification of leucoplast pyruvate kinase from developing castor bean endosperm. Plant Physiol. 94: 1 528-1 534. Plaxton, W.C. and Podesta, RE. (2006) The functional organization and control of plant respiration. Crit. Rev. Plant Sci. 25: 1 59-198. Plaxton, W.C., Smith, C.R., and Knowles, V.L. (2002) Molecular and regulatory properties of leucoplast pyruvate kinase from Brassica napus (rapeseed) suspension cells. Arch. Biochem. Biophys. 400:54-62. 43 Pleite, R., Pike, M.J., Garces, R., Martinez-Force, E., and Rawsthome, S. (2005) The sources of carbon and reducing power for fatty acid synthesis in the heterotrophic plastids of developing sunflower (Helianthus annuus L.) embryos. J. Exp. Bot. 56:1297-1303. Podesta, RE. and Plaxton, W.C. (1993) Activation of cytosolic pyruvate kinase by polyethylene glycol. Plant Physiol. 103:285-288. Podesta, RE. and Plaxton, W.C. (2003) Fluorescence study of ligand binding to potato tuber pyrophosphate-dependent phosphofi'uctokinase: evidence for competitive binding between fi'uctose-l,6-bisphosphate and fructose-2,6-bisphosphate. Arch. Biochem. Biophys. 414:101-107. Porterfield, D.M., Kuang, A., Smith, P.J., Crispi, M.L., and Musgrave, M.E. (1999) Oxygen-depleted zones inside reproductive structures of Brassicaceae: implications for oxygen control of seed development. Can. J. Bot. 77: 1439-1446. Regierer, B., Fernie, A.R., Springer, F., Perez-Malls, A., Leisse, A., Koehl, K., Willmitzer, L., Geigenberger, P., and Kossmann, J. (2002) Starch content and yield increase as a result of altering adenylate pools in transgenic plants. Nat. Biotechnol. 20:1256-1260. Reiser, J ., Linka, N., Lemke, L., Jeblick, W., and Neuhaus, HE. (2004) Molecular physiological analysis of the two plastidic ATP/ADP transporters from Arabidopsis. Plant Physiol. 136:3524-3536. Rider, S.D., Jr., Hemm, M.R., Hostetler, H.A., Li, H.C., Chapple, C., and Ogas, J. (2004) Metabolic profiling of the Arabidopsis pkl mutant reveals selective derepression of embryonic traits. Planta 219:489-499. Rosche, E., Blackmore, D., Tegeder, M., Richardson, T., Schroeder, H., Higgins, T.J., Frommer, W.B., Offler, C.E., and Patrick, J.W. (2002) Seed-specific overexpression of a potato sucrose transporter increases sucrose uptake and growth rates of developing pea cotyledons. Plant J. 30: 165-1 75. Routaboul, J.M., Benning, C., Bechtold, N., Caboche, M., and Lepiniec, IL. (1999) The TAGl locus of Arabidopsis encodes for a diacyl glycerol acyltransferase. Plant Physiol. Biochem. 37 :83 1-840. Ruuska, S.A., Girke, T., Benning, C., and Ohlrogge, J .B. (2002) Contrapuntal networks of gene expression during Arabidopsis seed filling. Plant Cell 14:1191-1206. Ruuska, S.A., Schwender, J., and Ohlrogge, J .B. (2004) The capacity of green oilseeds to utilize photosynthesis to drive biosynthetic processes. Plant Physiol. 136:2700-2709. Sangwan, R.S., Gauthier, D.A., Turpin, D.H., Pomeroy, M.K., and Plaxton, W.C. (1992) Pyruvate kinase isoenzymes from zygotic and microspore-derived embryos of Brassica napus - Developmental profiles and subunit composition. Planta 187:198-202. 44 Sasaki, Y., Kozaki, A., and Hatano, M. (1997) Link between light and fatty acid synthesis: thioredoxin-linked reductive activation of plastidic acetyl-CoA carboxylase. Proc. Natl. Acad. Sci. USA. 94:11096-11101. Satoh, H., Tani, K., Yoshida, M.C., Sasaki, M., Miwa, S., and Fujii, H. (1988) The human liver-type pyruvate kinase (PKL) gene is on chromosome 1 at band q21. Cytogenet. Cell Genet. 47:132-133. Scheibe, R. and Anderson, L.E. (1981) Dark modulation of NADP-dependent malate dehydrogenase and glucose-6-phosphate dehydrogenase in the chloroplast. Biochim. Biophys. Acta 636:58-64. Schmid, M., Davison, T.S., Henz, S.R., Pape, U.J., Demar, M., Vingron, M., Scholkopf, B., Weigel, D., and Lohmann, J.U. (2005) A gene expression map of Arabidopsis thaliana development. Nat. Genet. 37:501-506. Schramm, A., Siebers, B., Tjaden, B., Brinkmann, H., and Hensel, R. (2000) Pyruvate kinase of the hyperthermophilic crenarchaeote T hermoproteus tenax: physiological role and phylogenetic aspects. J. Bacteriol. 18222001-2009. Schwender, J., Goffman, F ., Ohlrogge, J.B., and Shachar-Hill, Y. (2004a) Rubisco without the Calvin cycle improves the carbon efficiency of developing geen seeds. Nature 432:779-782. Schwender, J., Ohlrogge, J., and Shachar-Hill, Y. (2004b) Understanding flux in plant metabolic networks. Curr. Opin. Plant Biol. 7:309-317. Schwender, J. and Ohlrogge, J.B. (2002) Probing in vivo metabolism by stable isotope labeling of storage lipids and proteins in developing Brassica napus embryos. Plant Physiol. 130:347-361. Schwender, J., Ohlrogge, J.B., and Shachar-Hill, Y. (2003) A flux model of glycolysis and the oxidative pentosephosphate pathway in developing Brassica napus embryos. J. Biol. Chem. 278:29442-29453. Schwender, J., Shachar-Hill, Y., and Ohlrogge, J.B. (2006) Mitochondrial metabolism in developing embryos of Brassica napus. J. Biol. Chem. 281:34040-34047. Shanklin, J. and Cahoon, E.B. (1998) Desaturation and related modifications of fatty acids. Annu. Rev. Plant Physiol. Plant Mol. Biol. 49:611-641. Shearer, H.L. and Dennis, D.T. (2005) Characterization and functional expression in yeast of a cDNA encoding NADP-dependent malic enzyme from castor oil seed. Can. J. Bot. 83:237-241. Shimakata, T. and Stumpf, P.K. (1982) Purification and characterizations of beta- Ketoacyl-acyl-carrier-protein reductase, beta-hydroxyacyl-acyl-carrier-protein dehydrase, 45 and enoyl-acyl-carrier-protein reductase from Spinacia oleracea leaves. Arch. Biochem. Biophys. 218277-91. Singal, H.R., Talwar, G., Dua, A., and Singh, R. (1995) Pod photosynthesis and seed dark C02 fixation support oil synthesis in developing Brassica seeds. J .Biosci. 20:49-58. Slabas, A.R., Sidebottom, C., Kessell, R., Hellyer, A., and Tombs, M.P. (1986) Oilseed rape NADH enoyl acyl-carrier protein reductase. Biochem. Soc. Trans. 14:581- 582. Smith, C.R., Knowles, V.L., and Plaxton, W.C. (2000) Purification and characterization of cytosolic pyruvate kinase from Brassica napus (rapeseed) suspension cell cultures: implications for the integration of glycolysis with nitrogen assimilation. Eur. J. Biochem. 267 :4477-4485. Smith, R.G., Gauthier, D.A., Dennis, D.T., and Turpin, D.H. (1992) Malate- and pyruvate-dependent fatty acid synthesis in leucoplasts from developing castor endosperm. Plant Physiol. 98:1233-1238. Sprague, G.F., Jr. (1977) Isolation and characterization of a Saccharomyces cerevisiae mutant deficient in pyruvate kinase activity. J. Bacteriol. 130:232-241. Stahl, U., Carlsson, A.S., Lenman, M., Dahlqvist, A., Huang, B., Banas, W., Banas, A., and Stymne, S. (2004) Cloning and functional characterization of a phospholipid:diacyl glycerol acyltransferase from Arabidopsis. Plant Physiol. 135:1324- 1335. Stone, S.L., Kwong, L.W., Yee, K.M., Pelletier, J ., Lepiniec, L., Fischer, R.L., Goldberg, R.B., and Harada, J.J. (2001) Leafy Cotyledon2 encodes a B3 domain transcription factor that induces embryo development. Proc. Natl. Acad. Sci. USA. 98:11806-11811. Tang, G.Q., Hardin, S.C., Dewey, R., and Huber, S.C. (2003) A novel C-terminal proteolytic processing of cytosolic pyruvate kinase, its phosphorylation and degradation by the proteasome in developing soybean seeds. Plant J. 34:77—93. Thelen, J.J. and Ohlrogge, J .B. (2002) Both antisense and sense expression of biotin carboxyl carrier protein isoform 2 inactivates the plastid acetyl-coenzyme A carboxylase in Arabidopsis thaliana. Plant J. 32:419-431. Tomlinson, K.L., McHugh, S., Labbe, H., Grainger, J .L., James, L.E., Pomeroy, K.M., Mullin, J.W., Miller, S.S., Dennis, D.T., and Miki, B.L. (2004) Evidence that the hexose-to-sucrose ratio does not control the switch to storage product accumulation in oilseeds: analysis of tobacco seed development and effects of overexpressing apoplastic invertase. J. Exp. Bot. 55:2291-2303. Truksa, M., 'Wu, G., Vrinten, P., and Qiu, X. (2006) Metabolic engineering of plants to produce very long-chain polyunsaturated fatty acids. Transgenic Res. 15:131-137. 46 Turner, W.L., Knowles, V.L., and Plaxton, W.C. (2005) Cytosolic pyruvate kinase: subunit composition, activity, and amount in developing castor and soybean seeds, and biochemical characterization of the purified castor seed enzyme. Planta 222: 1051-1062. Umeda, M. and Uchimiya, H. (1994) Differential transcript levels of genes associated with glycolysis and alcohol fermentation in rice plants (Oryza sativa L.) under submergence stress. Plant Physiol. 106:1015-1022. Valentini, G., Chiarelli, L., F ortin, R., Speranza, M.L., Galizzi, A., and Mattevi, A. (2000) The allosteric regulation of pyruvate kinase. J. Biol. Chem. 275: 1 8145-18152. Vigeolas, H., Mohlmann, T., Martini, N., Neuhaus, H.E., and Geigenberger, P. (2004) Embryo-specific reduction of ADP-Glc pyrophosphorylase leads to an inhibition of starch synthesis and a delay in oil accumulation in developing seeds of oilseed rape. Plant Physiol. 13622676-2686. Vigeolas, H., van Dongen, J.T., Waldeck, P., Huhn, D., and Geigenberger, P. (2003) Lipid storage metabolism is limited by the prevailing low oxygen concentrations within developing seeds of oilseed rape. Plant Physiol. 133:2048-2060. Voelker, T. and Kinney, A.J. (2001) Variations in the biosynthesis of seed storage lipids. Annu. Rev. Plant Physiol. Plant Mol. Biol. 52:335-361. Voelker, T.A., Hayes, T.R., Cranmer, A.M., Turner, J .C., and Davies, H.M. (1996) Genetic engineering of quantitative trait - metabolic and genetic parameters influencing the accumulation of laurate in rape seed. Plant J. 92229-241. Voelker, T.A., Worrell, A.C., Anderson, L., Bleibaum, J., Fan, C., Hawkins, D.J., Radke, S.E., and Davies, H.M. (1992) Fatty acid biosynthesis redirected to medium chains in transgenic oilseed plants. Science 257:72-74. Voll, L., Hausler, R.E., Hecker, R., Weber, A., Weissenbock, G., F iene, G., Waffenschmidt, S., and Flugge, U.I. (2003) The phenotype of the Arabidopsis cue] mutant is not simply caused by a general restriction of the shikimate pathway. Plant J. 36:301-317. Wakao, S. and Benning, C. (2005) Genome-wide analysis of glucose-6-phosphate dehydrogenases in Arabidopsis. Plant J. 41:243-256. Wan, J., Blakeley, S.D., Dennis, D.T., and Ko, K. (1995) Import characteristics of a leucoplast pyruvate kinase are influenced by a 19-amino-acid domain within the protein. J. Biol. Chem. 270:16731-16739. Wang, H., Qi, Q., Schorr, P., Cutler, A.J., Crosby, W.L., and Fowke, L.C. (1998) ICKl , a cyclin-dependent protein kinase inhibitor from Arabidopsis thaliana interacts with both Cdc2a and Cch3, and its expression is induced by abscisic acid. Plant J. 15:501-510. 47 Weber, A.P. (2004) Solute transporters as connecting elements between cytosol and plastid stroma. Curr. Opin. Plant Biol. 71247-253. Weber, H., Borisjuk, L., Heim, U., Sauer, N., and Wobus, U. (1997) A role for sugar transporters during seed development: molecular characterization of a hexose and a sucrose carrier in fava bean seeds. Plant Cell 9:895-908. Weber, H., Buchner, P., Borisjuk, L., and Wobus, U. (1996) Sucrose metabolism during cotyledon development of Viciafaba L. is controlled by the concerted action of both sucrose-phosphate synthase and sucrose synthase: expression patterns, metabolic regulation and implications for seed development. Plant J. 9:841-850. Wheeler, M.C.G., Tronconi, M.A., Drincovich, M.F., Andreo, C.S., F lugge, U.I., and Maurino, V.G. (2005) A comprehensive analysis of the NADP-malic enzyme gene family of Arabidopsis. Plant Physiol. 139239-51. White, J.A., Todd, J ., Newman, T., Focks, N., Girke, T., de Ilarduya, D.M., Jaworski, J.G., Ohlrogge, J.B., and Benning, C. (2000) A new set of Arabidopsis expressed sequence tags from developing seeds. The metabolic pathway from carbohydrates to seed oil. Plant Physiol. 12421582-1594. Wise, M.J. and Tunnacliffe, A. (2004) POPP the question: what do LEA proteins do? Trends Plant Sci. 9:13-17. Zanella, A., Fermo, E., Bianchi, P., Chiarelli, LR, and Valentini, G. (2007) Pyruvate kinase deficiency: The genotype-phenotype association. Blood Rev. (article in press) Zou, J., Katavic, V., Giblin, E.M., Barton, D.L., MacKenzie, S.L., Keller, W.A., Hu, X., and Taylor, D.C. (1997) Modification of seed oil content and acyl composition in the brassicaceae by expression of a yeast sn-2 acyltransferase gene. Plant Cell 9:909-923. 48 Chapter 2 Molecular and kinetic analysis of Arabidopsis plastidic pyruvate kinase1 ' This work has (been published in Andre, C., Froehlich, J.E., Moll, M.R., and Benning, C. (2007) A heteromeric plastidic pyruvate kinase complex involved in seed oil biosynthesis in Arabidopsis. Plant Cell 19: 1-17. Figure 2.33 was provided by J .E. Froehlich and M.R. Moll contributed to Figure 2.5A-E 49 Abstract Plastidic pyruvate kinase (PKp) catalyzes a highly regulated, ATP-producing reaction of glycolysis. PKp occupies a highly branched node in the triacyl glycerol biosynthetic network of developing seeds and is expected to be an important point for the regulation of carbon partitioning. A detailed biochemical characterization of this enzyme will provide a framework for future manipulations of seed metabolism. The Arabidopsis genome encodes l4 putative isoforms of pyruvate kinases. Three genes encode subunits a, [31, and [32 of plastidic pyruvate kinase (PKp). Recombinant protein production and subsequent kinetic analysis was used to show that active PKp complexes are composed of a and either [3. or [32 subunits. Enzyme activity is dependent the formation of a complex between a and [3 subunits, although the presence of either [3. or B2 results in unique kinetic and regulatory properties. The plastid enzyme prevalent in developing seeds likely has a subunit composition of 4(1451, is most active at pH 8 and is inhibited by glutamate. 50 Introduction The biochemical reactions leading to the accumulation of seed oil are well characterized (Ohlrogge and Browse 1995). Furthermore, the use of stable isotope labeling and forward genetics has linked the supply of precursors for fatty acid biosynthesis in embryos to glycolysis (F ocks and Benning 1998, Schwender and Ohlrogge 2002). A key regulatory step in plant glycolysis is pyruvate kinase, as its products inhibit the upstream glycolytic reaction catalyzed by phosphofructokinase (Plaxton 1996). Pyruvate kinase (EC 2.7.1.40) occurs as both cytosolic and plastidic isoforms and catalyzes the ADP-dependent conversion of phosphoenolpyruvate (PEP) to pyruvate while producing ATP. With respect to seed oil, plastidic pyruvate kinase (PKp) activity and concentration have been shown to correlate with the most active stage of lipid biosynthesis in developing Brassica napus embryos (Sangwan et al. 1992). Microarray data of developing Arabidopsis seeds show that the transcript level of a putative PKp encoding gene coincides with the most active period of TAG synthesis (Ruuska et al. 2002, Schmid et a1. 2005). In addition, embryo PKp from B. napus is activated by 6-phosphogluconate, an intermediate of the OPPP, suggesting a coordination between the production of precursors and reducing equivalents for fatty acid synthesis (Plaxton et al. 2002). Plant PK activities arise from the expression of multiple isozymes with different biochemical properties that depend on the tissue and plant source. Arabidopsis, for instance, has 14 annotated PK genes that likely exhibit a large degree of variation with respect to regulation of gene expression and enzyme activity (Arabidopsis Genome Initiative 2000). Plastidic pyruvate kinase has been purified and characterized from castor (Ricinus communis) endosperm and B. napus suspension cell cultures Wegm et al. 1995, 51 Plaxton et a1. 1990, Plaxton et al. 2002). Both enzymes consist of a. and [3 subunits and exist as 3a3B heterohexamers. Both are regulated by metabolites of central carbon metabolism and have pH optima of approximately 8.0. In the case of the B. napus, a scheme was formulated in which the (kinetic properties of PKp were used to infer a role for the enzyme in regulating the supply of precursors for fatty acid synthesis. The characterization of orthologous PKp(s) from the model plant Arabidopsis is pertinent as such a similarly derived model could then be tested using available genetic tools. Here, potential PKp s from Arabidopsis were identified and heterologously produced using E. coli. The resulting recombinant proteins were then characterized at the molecular and biochemical levels to gain insight into this enzyme’s role in seed oil biosynthesis. 52 Materials and Methods Bioinformatics Pyruvate kinase genes highly expressed in seeds were initially identified in the seed EST Database (White et al. 2000). The 14 putative Arabidopsis PK-annotated sequences from the TAIR website (www.arabidopsis.org) were used for the current work. Other annotated PK sequences were down loaded from NCBI. Predicted full-length protein sequences were aligned using ClustalW (Li 2003) available from the Biology Workbench (San Diego Supercomputer Center, University of California, San Diego; http2//workbench.sdsc.edu/). A phylogenetic tree was generated in Phylip format and bootstrapped using a random number generator seed of 111 and 1000 bootstrap trials. The phylogram was visualized using the TreeView program (version 1.6.6; Page 1996). No manual adjustments were made to the initial alignment. Global gene expression data was mined from the AtGenExpress developmental database (Schmid et al. 2005). GFP fusion localization The cDNAs for the three PK subunits were amplified using primers specific for GFP fusion construct generation as listed in Table 2.1. Fully sequenced products were then inserted into the T-DNA vector pCAMBIA1302 (CAMBIA, Canberra, Australia), which contains a CaMV 358 promoter and a C-terrninal GF P encoding sequence. The resulting constructs were electroporated into Agrobacterium tumefaciens as described below and were then transiently expressed in Nicotiana benthamiana using a published protocol (Voinnet et al. 2003). After 3-4 days, leaf samples were mounted in water on slides and were directly examined using a Zeiss LSMS confocal microscope. Excitation light was 53 provided by an argon laser at 488 nm. GFP fluorescence was observed with a band-pass filter of 505 to 530 nm and chlorophyll fluorescence with a 650-nm long-pass filter. Enhanced-quality images were acquired with the LSMS imaging system software, and post acquisition image processing was performed with the LSMS image browser and Adobe Photoshop software. This work was performed at the Center for Advanced Microscopy (Michigan State University). Pea chloroplast import assays The cDNAs encoding PKp-a, PKp-BI, and PKp-B2 inserted into pBluescript II (Stratagene) were used in this study. These genes were transcribed/translated, and proteins were subsequently labeled with [3SS]-Met using the TNT-coupled wheat germ extract system according to the manufacturer's recommendations (Promega, Madison, WI, USA). The PKp plasmids were linearized prior to translation with the T3 or T7 RNA Polymerase TNT-coupled wheat germ extract system. The plasmid containing the gene encoding the RuBisCO small subunit used for control purposes has been described (Olsen and Keegstra 1992). Pea plants (Pisum sativum var Little Marvel; Olds Seed Co., Madison, WI, USA) were grown under natural light in the greenhouse at 18 to 20°C. Chloroplasts were isolated from 8- to 12-d-old plants as described previously (Bruce et al., 1994). Binding or import reactions were performed according to published protocols (Tranel et al. 1995). Post-treatrnents of import reactions with either thermolysin or trypsin were performed as described previously (Jackson et al. 1998). All fractions were analyzed by SDS-PAGE (Laemmli 1970) and fluorography (Tranel et al. 1995). 54 cDNA cloning and recombinant protein production The cDNAs corresponding to At3g22960, At5g52920, and Atl g32440 were generated from total silique RNA isolated as previously described (V erwoerd et al. 1989). Reverse transcription was done with the Qiagen (Valencia, CA, USA) Omniscript RT kit and 600 ng of total RNA. Primers listed in Table 2.1 were used for PCR amplification of cDNAs to generate products with and without predicted chloroplast transit peptides (cTPs) and with or without epitope tags. All cDNAs were inserted into pBluescript II (Stratagene, LaJolla, CA, USA) and sequenced at the MSU Research Technology Support Facility. The vector pET-15b (Novagen, San Diego, CA, USA) was used for recombinant protein expression in E. coli strain BL21 (DE3) pLysS (Novagen). The PKp-a encoding fragment was inserted into the BamHI and Klenow-filled NdeI sites of pET-l 5b by digesting with KasI, filling in with Klenow, and then by digesting with BamHI and ligating. The open reading frame for PKp-Bi was inserted into the BamHI and Klenow-filled NdeI sites of pET-l 5b by digesting with BglII, filling in with Klenow, and then by digesting with BamHI and ligating. The PKp-B2 encoding fragment was inserted into the XhoI and Klenow-filled NdeI sites of pET-15b by first digesting with SpeI, filling in with Klenow, and then by digesting with XhoI and ligating. Proteins were expressed at 28°C by inducing at an OD600 of 0.6 with 0.5 mM IPTG and allowing the cultures to grow for 4 more hours. His-tagged proteins were recovered over Ni-NTA resin using standard protocols. Purified proteins were exchanged into a buffer of 50 mM NaxHxPO4 pH 7.9, 150 mM NaCl, 5 mM MgCl2, and 10% glycerol. The 6X-His tags were cleaved using a Thrombin cleavage capture kit available from Novagen. Proteins were quantified with the Bradford method using reagent from Sigrna-Aldrich (St. Louis, Mo, USA). 55 Table 2.1 Primer sequences Gene or SALK Line Primer PKP'Bz SALK_013574 SALK_142845 PKp-a SALK_096141 SALK_024870 PKP'BI P,PE(t) 5 ’-ACTAGTATTAAAATCTCCGAAGATAG-3 ’ P(r) 5 ’ -CTCGAGTCATCCACCTATCTTTATCT-3 ’ PE(r) 5 ’-CTCGAGTCATCCCTTGTCATCGTCATCCTT ATAATCTCCACCTATCTTTATCTT-3 ’ 0,0E(f) 5 ’-GGTACCCCTCAGGTTTCTCTGCTCAT-3 ’ 0(r) 5 ’ -GGTACCACTGTGAGTGATTCAAAAAA-3 ’ 0E(r) 5’- GGTACCTCATCCCTTGTCATCGTCATCCTT ATAATCTCCACCTATCTTTATCTT-3 ’ G(f) 5 ’ -GATATCGCTGCTTATGGTCAAATCTC-3 ’ G(r) 5 ’-GATATCTCCACCTATCTTTATCTTAC-3 ’ RT(f) 5 ’-GGGGATGTACCGCAGCCGATA-3 ’ RT(r) 5 ’-GGATGCCGAGGTTCTGACAGG-3 ’ RP 5’-TTTCACACAACAAATTCGTTCATT-3 ’ LP 5’-CAGCTTCCGCGAGTTTCCAAATCA-3 ’ Same as SALK_013574 P,PE(t) 5’-GGCGCCTCCTCGTCATCATCTCC-3’ P(r)-5’-GGATCCT1‘ACGGGACGTTCATI‘ACCT-3’ PE(r) 5’-GGATCCTTACAAATCCTCCTCACTAATCAA CTTTTGCTCCGGGACG'ITCATTACCTG-3’ 0E(r) 5’-GGTACCAGCCAACTGTCCTGAGATTT-3’ 013(1) 5’-GGTACCTTACAAATCCTCCTCACTAATCAA CTT'I‘TGCTCCGGGACGTTCATTACCTG-B’ G(t) 5’-ACATGTCTCAGTCTAT1"CAATTCTCC-3’ G(r) 5’-ACTAGTCGGGACGTTCAT'I‘ACCTGGA-3’ LP 5’-CCAAATTCAACACTCTCACACTTCG-3’ RP 5’-CCATCCCACCATCAACCAAAA-3’ RT(t) 5’-GCTGCTCGTTCCCGTGGAGG-3’ RT(r) 5’-TTGAAGCGGTACAGACTCAT‘-3’ LP 5’-TCTCGGACATGCTGCAATCAA-3’ RP 5’-TTCGCATCAGTCATCTTCGTCTTC-3’ P,PE(f) 5 ’-AGATCTGCTCGTGTTGAGACTGA-3 ’ P(r) 5 ’ -GGATCCTTAAACCTTGCGGACTTGGA-3 ’ PE(r) 5 ’-GGATCCTTATGCATAATCGGGAACATCATA GGGATAAACCTTGCGGACTTGGAT-3 ’ 0,0E(f) 5 ’ -GGTACCCTTCACTACTCTGTCTCAGC-3 ’ 0(r) 5 ’-GGTACCCAAAAACGAGGTTCTACATA-3 ’ 0E(r) 5 ’-GGTACCTTATGCATAATCGGGAACATCATA GGGATAAACCTTGCGGACTTGGAT-3 ’ G(f) 5 ’ -CCATGGCTCAAGTGGTTGCTACCAGG-3 ’ G(r) 3 ’-ACTAGTAACCTTGCGGACTTGGATGT-3 ’ 56 Table 2.1 Continued SALK_04293 8 LP 5 ’-TGAGATAGCA' I " I “ I CAA I“I"I'GATGCG-3 ’ RP 5 ’ -GGCAAATCATTCACTTAGGATGGA-3 ’ RTI (f) 5 ’-CTGGGATGAATGTTGCTAGG-3 ’ RT2(r) 5 ’ -GTCAACTTTGTTCTCCACTCC-3 ’ SALK_042681 LP 5’-ACAGCCAATCTGGCGATCTCA-3’ RF 5 ’-TAT'TACAGGTCTATTTCTTTCGG-3 ’ T-DNA LB 5’-GTTCACGTAGTGGGCCATCG-3 ’ RT(f) 5'-AACAATCGATGGACCTGACTCG-3' RT(r) 5'-TGCGACAATGGAACTGGAATGG-3' ACTIN-l Primers used for cloning of recombinant proteins without predicted cTP with (PE) or without (P) eptiope tags, for cloning of full length proteins for overexpression in plants with (OE) or without (0) epitope tags, for GFP fusion construct making (G), for genotyping of SALK_KO lines (LP, RP) and for expression analysis (RT) Native-PAGE analysis and gel filtration chromatography Native-polyacrylamide gel electrophoresis (PAGE) was performed using 7.5% acrylamide Ready-Gels from Bio-Rad (Hercules, CA, USA). Freshly prepared protein was used and 5 pmols of each PK subunit were loaded per well. The gels were run at 4°C at 140 V. PK activity staining was done as previously described, except that 50 mM HEPES-KOH pH 8.0 buffer was used (Rivoal et al. 2002). Irnmunoblotting was done using standard protocols and monoclonal anti-c-myc, anti-FLAG, and anti-HA antibodies from Sigma-Aldrich. Antibodies were tested for specificity against epitope tagged and untagged versions of all three PK subunits. Gel filtration was accomplished using a Superdex 200 HR10/30 column with a flow rate of 0.4 mL min'l and a buffer of 50 mM HEPES-KOH pH 8.0, 10% glycerol, 50 mM KCl, 5 mM MgCl2, 1 mM EDTA, and 0.04% NaN3. A standard curve was generated with the HMW gel filtration calibration kit 57 fi'om GE Healthcare (Piscataway, NJ, USA). For the samples, 300 uL of 0.5 mg mL'1 (per subunit) solution was injected and 0.35 mL fractions were collected. Co-immunopreciptation analysis 7 Full length epitope tagged versions of PKp-a, PKp-BI, and PKp-I32 were produced by PCR with primers listed in Table 2.1. The respective DNA fragments were then inserted into the KpnI site of a modified pCAMBIAl 300 vector (CAMBIA), which contained the EcoRI/HindIII expression cassette from pBIN 121 (Clontech, Palo Alto, CA, USA). Arabidopsis was stably transformed with these constructs and protein expression was monitored by immunoblotting. Arabidopsis plants were prepared for transformation as previously described (Cemac and Benning 2004). When ready, plants were transformed using the floral dip method (Clough and Bent 1998). Competent cells of Agrobacterium tumefaciens strain C5 8C1 GV3101 pMP90 (Koncz and Schell 1986) were prepared and transformed as previously described (Shen and F orde 1989). Silique tissue was ground in 3 volumes (w/v) of extraction buffer containing 50 mM HEPES-KOH pH 8.0, 5 mM MgCl2, 10 mM KCl, 1 mM EDTA, 1 mM EGTA, lmM DTT, 0.1% Triton-X100, 10% glycerol, 2 mM benzamidine, 2 mM e-arnino-n-caproic acid, 1 mM PMSF, and 1 mM PEP. Debris was removed by centrifugation at 16,000g for 10 min and the supernatant was used for SDS-PAGE or co-IP. For co-IP, 250 uL of supernatant was pre-cleared by incubation for 1 hour at 4°C with 30 uL of a 50% slurry of Protein-A sepharose. The supernatant was kept and mixed with 5 pg of the appropriate antibody and was then nutated at 4°C. After 1 hour, 30 uL of Protein-A sepharose slurry was added and the mixture rocked for an additional hour. After this time, the Protein-A 58 sepharose with bound antibody and proteins was washed 4 times in extraction buffer and then mixed with SDS-PAGE sample buffer, heated, and the supernatant loaded on gels. Excised bands were submitted to the MSU Research Technology Support Facility for tryptic digest and LC/MS/MS. The generated data was then compared against the Arabidopsis proteome using MASCOT software (Matrix Science, Boston, MA, USA). PK enzyme assays and kinetic analysis All chemicals were from Si grna-Aldrich. Pyruvate kinase activity was detected by coupling the production of pyruvate to the conversion of NADH to NAD+ by lactate dehydrogenase unless otherwise noted. Reactions were kept at 25°C, were started by the addition of enzyme mix, and were linear for at least 5 minutes. Absorbance at 340 nm was monitored using a F LUOstar Optima 96-well plate reader (BMG Labtech, Offenburg, Germany). Standard PKp reaction mixtures contained 50 mM HEPES-KOH pH 8.0, 5% PEG-8000, 50 mM KCI, 15 mM MgCl2, 1 mM DTT, 2 mM PEP, 1 mM ADP, 0.2 mM NADH, and 2 U ml‘1 desalted rabbit muscle lactate dehydrogenase. PEP phosphatase activity was corrected for by omitting ADP from the reaction. Reactions at pH 7.0 were done using 50 mM MOPS pH 7.0 instead of HEPES. For kinetic analysis 2.5 pmol of each subunit were mixed and used per reaction. S05 and Vmax values were determined by fitting the Hill equation to plots of initial velocity versus substrate concentration using origin 7.0 (OriginLab Corporation, Northampton, MA, USA). pH optimum curves were generated using a 25 mM MES, 25 mM Bis-Tris-propane buffer over a range of pH’s. For inhibitor/activator studies metabolite stocks were made equimolar with MgC12 and were pH adjusted to 8.0. 59 Metabolites were tested at pH 8.0 with 100 uM PEP and 150 uM ADP for GB] and 150 pM PEP and 300 uM ADP for (102. The metabolites tested were: glucose, fructose, 6GP, GlP, F6P, F16bP, DHAP, G3P, 2PG, acetate, 0AA, citrate, iso-citrate, 2-oxoglutarate, succinate, fumerate, malate, gro3P, 2P-glycolate, glycolate, Ala, Arg, Asn, Asp, Cys, Gln, Glu, Gly, His, Leu, Lys, Met, Ser, AMP, ADP, ATP, ADP-Glc, UDP-Glc, UDP-Gal, R5P, 6PG, KPi, NaNO3, and NH4C1, all at 10 mM; MgPPi, NADH, NADPH, NAD+, NADP+, oxalate, Ile, Phe, Pro, Thr, Trp, Tyr, Val, CoA, Mal-CoA, Ac-CoA, and shikimate, all at 0.5 mM; F26bP, Oleoyl-CoA, Oleate, and LPA all at 0.05 mM. Oxalate, glyoxylate, and 0AA were found to inhibit the LDH reaction so PK activity was measured by coupling to the ATP dependent conversion of glucose to glucose-6-phosphate by hexokinase followed by the NAD+-dependent conversion of glucose-6-phosphate to 6- phosphogluconate by glucose-6-phosphate dehydrogenase. The reaction mix was adjusted to contain no NADH or LDH, but instead to have 1 mM NAD+, 5 mM glucose, 2 U mL'1 hexokinase, and 2 U mL'l G6PDH. 150 and Ka values are the concentration of a metabolite required for 50% maximum inhibition or activation, respectively. They were calculated by fitting a modified Hill equation to plots of initial velocity versus effector concentration as previously described (Ballicora et al. 2005). Site-directed mutagenesis of PK subunits was done using the QuikChange-XL Site-Directed Mutagenesis Kit using primers designed to the manufacturer’s specifications (Qiagen). An absolutely conserved lysine residue in the PK active site was mutated to leucine for each subunit (PKp-a, K344L; PKp-Bl, K325L; PKp-B2, K314L). This mutation has previously been shown to abolish PK activity (Sakai 2005). Chemical inactivation was achieved by incubating 5 uM solutions of purified subunits with either 60 water or 100-fold molar excess of 2,4,6-trinitrobenzenesulfonic (TNBS) acid in the dark at room temperature for 1 hour. This treatment has been shown to inactivate PK subunits by covalent modification of lysine residues without abolishing protein interactions (Hollenberg et al. 1971) Excess TNBS was quenched with an equal volume of 100 mM Tris-Cl pH 7.5 for 20 minutes on ice and the proteins were used directly in enzyme assays. Accession numbers Arabidopsis Genome Initiative locus identifiers (www.mabidopsisprg) used in this study are as follows: At3g22960 (encoding PKp-a), At5g52920 (encoding PKp-Bl), Atl g32440 (encoding PKp-le, At2g36580, At3g04050, At3g25960, At3g49160, At3g52990, At3g55650, At3g55810, At4g26390, At5g08570, At5g56350, At5g63680. Genbank accession numbers for non-Arabidopsis protein sequences used are as follows: Nt PKp-A, Q40545; Os PKp-A, NP_001059042; Rc PKp-A, Q43117; NtPKp-G, 040546; Se PCC630] PK, YP_172116; Se WH8102 PK, NP_897391; Pfu PK, NP_578917; Sa PK, YP_256251; Ec PK-l, AAA24392; Hs PK-L, BAA02515; An PK, Q12669; Sc PYKl, NP_009362; Os PKG, BAD81116; Nt PKc, Q42954; St PKC, P22200; Gm PKc, Q42806 61 Results Identification of plastid localized and seed resident pyruvate kinases The Arabidopsis genome encodes 14 putative PK isoforms (Arabidopsis Genome Initiative 2000). All but one (encoded by At3g49160) of the predicted PKs contain a fully conserved PK active site of [LIVAC]-x-[LIVM]-[LIVM]-[SAPCV]-K-[LIV]-E- [NKRST]-x-[DEQHS]-[GSTA]-[LIVM] (listed at European Bioinformatics Institute, . http://www.ebi.ac.uL0 and are presumably active enzymes. Several cross-kingdom PK phylogenies have been published (e. g. Hattori et al. 1995, Munoz and Ponce 2003, Schramm et a1. 2000), but only one putative PKc from Arabidopsis was included in these studies. When the 14 Arabidopsis PK amino acid sequences are aligned with bonafide PKs from other organisms they segregate into cytosolic and plastidic clades (Figure 2.1A). This amino acid sequence similarity-based segregation is supported by the exclusive prediction of chloroplast transit peptides at the N-terrnini of the four predicted PKps using ChloroP and TargetP (Emanuelsson et al. 1999, Emanuelsson et al. 2000). The PKps in Figure 2.1A are further divided between a and [3 subunits. One Arabidopsis PKp subunit (encoded by At3g22960) is most similar to described PKp-as. Two others, which show 63% amino acid identity to each other (encoded by At5g52920 ,PKp-BI; At1g32440, PKp-B2), are most similar to known PKp-Bs (Figure 2.1A). Three of the four predicted Arabidopsis PKp genes were identified as seed expressed by EST analysis of developing seeds (White et al. 2000). Two, PKp-a and PK,,- [31, represent the highest level of induction of any PK gene in seed and are coordinately expressed in all tissues while the PKp-B2 encoding gene has very low transcript accumulation in any tissue (Figure 2. 1 B, Schmid et al. 2005). 62 Figure 2.1. Pyruvate kinase similarity and selected gene expression in Arabidopsis (A) Pyruvate kinase phylogeny. Amino acid sequences of Arabidopsis (gene loci in bold) and other bonafide pyruvate kinases were used. Bootstrap values are indicted at branches; or and [3 represent plastidic PK subunit families. The scale represents 10% difference. An, Aspergillus niger; Ec, Eschericia coli; Gm, Glycine max; Hs, Homo sapiens; Nt, Nicotiana tabacum; Os, 0ryza sativa; Pfu, Pyrococcusfuriosus; Rc, Ricinus communus; Sa, Sulfolobus acidocaldarius; Se, Synechocystis sp.; Sc, Saccharomyces cerevisiae; St, Solanum tuberosum (B) Relative gene expression of putative Arabidopsis PKp-a (At3g22960), PKp-BI (At5g52920), and PKp-B2 (At1g32440) encoding genes (derived from published microarray data (Schmid et al., 2005)). DAF, days after flowering. Values are the mean i so (n=3). 63 A At3g49160 0 .2 79 ‘03 3 CL Se PCC6301 PK 1000 Se WH8102 PK Pfu PK Ec PK-1 325 Hs PK-L 1000 SC PYK1 OS PKG o NtPKc % _ Atsgsssso m 9 At4g2ssso £633 St ch O Gm PKG Athessso 1°00 At5908570 N3904050 _ At3925960 813 10010000 At3g55810 1000 M3955650 J—-—At3952990 01 1WN2936580 B A a? g 7000 E5 6000 g 5000 B 4000 8 3000 g; 2000 9 1000 I E o I I . I. . , . I . . I I I w eaaaaagecm U-U- LLLL a a222§25$§535§§355 FEE-Eat“? “"3 3°88 8 g g 5 .5 (I) Figure 2.1. Pyruvate kinase similarity and selected gene expression in Arabidopsis 64 Figure 2.2. Phylogenetic analysis of plant PK protein sequences. Amino acid sequence-based phylogeny of annotated PKs from plant and algal sources. Some sequences are the same as used in Figure 2.1. The rice (Os), corn (Z. mays), Chlamydomonas (C. reinhardtii), Cyanidioschyzon merolae (C. merolae), and Physcomitrella patens (P. patens) sequences are new and accession numbers are listed after the genus identifier. 65 P. patens 12219237 — P. patens 12159402 01 P. patens 12109480 Os12905110 0511905110 At3952990 A12936580 P. patens 1 :134756 2. mays 5_91860 0511910980 C. reinhardtii C_600010 . ‘P ”1 98597 Se WH8102 PK Se PCC6301 PK P. patens 1:147287 P. patens1z147118 At3g431 60 C. rein. C_550090 C. mero/ae CMP260C __+ C. reinhardtii C_2590004 C. reinhardtii C_670023 C. reinhardtii C_910058 P. patens 12188818 P. patens 1:172921 P. patens12186171 P. patens 12131669 P. patens 12169084 Z. mays 5_3070 Z. mays 5_3071 0504958110 SI PKc Gm PKc A15963680 At5908570 0501916960 Nt PKc At5956350 At4926390 At3904050 At3955810 At3955650 At3925960 Figure 2.2. Phylogenetic analysis of plant PK protein sequences 66 An additional phylogenetic analysis was performed using amino acid sequences of annotated plant and algal PKs to understand the evolutionary history of the PKp-a and PKp-B subfamilies (Figure 2.2). Both the 0t- and B-subunit subfamilies are populated by proteins from algae, moss, and higher plants. Furthermore, both subfamilies originate from a branch containing cyanobacterial PKs. When combined, these observations suggest that the evolution of separate PKp subunits occurred after the secondary endosymbiotic event which gave rise to photosynthetic eukaryotes. Predicted PKp subunits are plastid localized To study the subcellular localization of the putative PKps, C-terminal green fluorescent protein (GFP) fusion constructs were generated and transiently expressed in Nicotiana benthamiana under the control of the CaMV 35$ promoter. Figur2.3A shows that for all three putative PKps a punctate GFP signal was observed at the periphery of chloroplasts. No GFP signal was associated with structures other than chloroplasts. All three proteins in question are predicted to be soluble and stromal. The localized patterns observed in Figure 2.3A could be due to the abundance of the fusion proteins, caused by overexpression and resulting in possible overloading of the import apparatus. In vitro chloroplast import followed by protease protection assays were conducted to obtain independent evidence for plastid localization of the PKp subunits. When in vitro produced 3SS-labeled PKp subunits were incubated with isolated pea chloroplasts, the proteins were imported and processed to their presumably mature forms (Figure 2.3B). All three were resistant to treatment with thermolysin and trypsin and were found in the soluble fraction. 67 - - + Trypsin - + - Thermolysin P S P S P S 4— pPKp-O “I n "K mPKp-o ‘“ PPKp'131 «PPKirflz ~~ ~~~- ----+—mPKp-i32 . . -o—mSS Figure 2.3. Subcellular localizations of pyruvate kinase subunits (A) Transient expression of GFP fusion constructs in N. benthamiana. GF P, green fluorescent protein fluorescence; CAF, chlorophyll autofluorescence; the bar is 5 uM long. (B) In vitro import of PKp subunits into isolated pea chloroplasts. After import, chloroplasts were subjected to either no treatment (-) or to post-treatment (+) with either Thermolysin or Trypsin. Intact chloroplasts were subsequently recovered by centrifugation through 40% Percoll cushion and fractionated into a total membrane (P) and a supernatant (S) fraction. A11 fractions were analyzed by SDS-PAGE and fluorography. MM, molecular masses of precursor and mature proteins based on Rf analysis; TP, represents 10% of translation reaction added; p, precursor or m, mature form; pSS, precursor of the small subunit of RuBisCO included as control. 68 Based on the data, the three putative PKp subunits are localized to chloroplasts and are imported and processed into mature, soluble, stromal proteins as predicted by the analysis of their amino acid sequence. The molecular masses indicated in Figure 2.3B were derived from Rf analysis and are in agreement with the transit peptide cleavage site predictions made by ChloroP. The full length precursor proteins were calculated to be 65.1 kDa, 63.5 kDa, and 62.6 kDa with predicted transit peptides of 47, 63, and 55 amino acids for PKp-a, PKp'BI, and PKp'BZ, respectively. The mobility shifts of all three mature proteins reveal no major discrepancies between the predicted and observed transit peptide cleavage sites. Heteromeric subunit composition of recombinant PKps A recombinant approach was taken to study Arabidopsis seed PKp subunit composition due to the scarcity of embryo tissue for native protein purification. The cDNAs encoding the three putative PKp subunits lacking the predicted transit peptides were isolated by reverse transcription and PCR and were inserted into an E. coli expression vector with an N-terminal 6X-His tag and a thrombin site for removal of the tag after purification. Gel electrophoresis (SDS-PAGE) and subsequent immunoblotting were used to confirm purification of single proteins and complete cleavage of the tag (Figure 2.4A). Initial tests indicated that in liquid assays none of the subunits had PK activity on their own and that only GB] and (102 combinations were active (Figure 2.4B). Epitope-tagged versions of the proteins were also generated that in addition to the N-terrninal His-tag had short C- tenninal epitope tags. PKp-a was fused with c-myc (EQKLISEEDL), PKp'BI With HA (YPYDVPDYA), and PKp-B2 with FLAG (DYKDDDDKG). The antibodies used for 69 detection of the epitopes were shown to be lacking of any cross reactivity. Native-PAGE with epitope-tagged proteins was used to explore this subunit requirement in more detail. Figure 2.5, panels A-E show five identical native-PAGE gels developed in different ways. The gel in Figure 2.5A was stained for PK activity. Only the (113. and (1132 mixtures were active. Moreover, the activities coincided with less mobile bands as shown by the Coomassie brilliant blue (CBB) stained gel (Figure 2.58). In the case of (1131, detection of the individual u-myc and Bl-HA fusion proteins with specific antibodies revealed a higher molecular mass complex coinciding in mobility with the active complex in Figure 2.5A, suggesting that the active complex is composed of both a and BI subunits (Figure 2.5c, D). 70 A uncut thrombin cut 0 I32 I31 0 I32 I31 82 kDa— 64 kDa— -*‘-- .. W = -- _ m coomassie 49 kDa — ' 64 kDa —-. i’ “-mu -; Anti His A OD PK Activity (pmol/sec x 103) w .a N O 0 Br B2 031 0132 131132 051 ClI32 (+)ADP (-)ADP Figure 2.4. PKp protein purification and initial activity assay (A) SDS-PAGE and anti-His immunblot of purified PKp subunits. Individual subunits were subjected to no treatment or treatment with thrombin to remove the His tag. 500 ng protein was loaded per well. Immunoblot shows His tag has been removed after treatment with thrombin. (B) Pyruvate kinase activity of various subunit mixtures. Equal volumes of purified subunits were assayed alone or in combination. Approximately 0.1 ng of each subunit was used in each assay with saturating substrate concentrations. No PEP phosphatase activity was observed in control reactions without ADP. . 5 6° w (i .1 (9 a" $2 52 £3 3 x- o #0 ~z~ A A .2- A A :S f x, .; §§ A 92 Q2 Q2 0 F :2 *. * 2 PKActivity 3 .669 kD °\° 100 +031 . M a V80 +0 Coomassre. . , Y 5 , . - {—232 kDa "5 ‘ ir'“ g.."" <1:40. i ‘ E 20 CA * 01* I i a; 0‘ "t'myc I II - I “3'6 0 1o 20 30 40750 60 7060 D(PKp-0) “ ' I g Fraction Anti HA 0 E(PKp-B1) L‘ I I I G Fractionfl'ZTEIIB' 32 kD 2“ . 64 Kile-jg? ;.:s.::.~.;_«..... . Anti KFLAGW L4. 49 k0 - , ‘ ~~ PKp—Bz) Figure 2.5. In vitro interaction of PKp subunits (A-E) Identically loaded native-PAGE gels showing in vitro interaction of PKp subunits. 10 pmol (~0.6 pg) of each subunit was used per lane (A) PK activity stained gel. (B) Coomassie stained gel. (C) Immunological detection of a-myc with anti c-myc (D) Immunological detection of 131-HA with anti HA (E) Immunological detection of B2- FLAG with anti FLAG. * is used to denote bands corresponding to the active PK complexes. (F) PK activity elution profile after F PLC over Superdex-200 (G) SDS-PAGE gel of most active fractions from (F). 75mg of protein was loaded per lane. 72 The result is less clear for the (1132 complex or and I31 subunits (Figure 2.5C, D). The result is less clear for the (1132 complex because the B2 subunit alone forms a higher molecular mass complex with the same mobility as the active enzyme (Figure 2.5A, B, E). However, the a-myc protein is present in the 0132 higher molecular complex (Figure 2.5C). Thus, it is likely that the higher molecular mass active complex in the 01132 mixture is also composed of both subunits. Gel-filtration chromatography was used to estimate the molecular masses of the active PKp heteromers. Figure 2.5F shows that for both 043. and (102 PK activity eluted as a single peak. The molecular masses of these active complexes were calculated to be 463 i 10 kDa for (1131 and 476 i: 10 kDa for (1132, which is consistent with octomeric complexes of 60 kDa (1 subunits and 57 kDa [3 subunits. Based on SDS-PAGE analysis, the most active F PLC fractions contained equal amounts of or and [3 subunits (Figure 2.5G). Thus, the active PKs appear to be heterooctomers composed of 4 0. and 4 [3 subunits. The in vivo interaction of the PKp subunits was tested using co- irmnunoprecipitation (co-1P). Three constructs containing full-length, epitope-tagged cDNAs encoding the 0., BI, and B2 subunits driven by the CaMV 35S promoter were introduced into Arabidopsis. Only the a-myc fusion protein could be directly immunoprecipitated from plant tissue. One possible explanation is that the epitope tags in the [3 subunits were not accessible in the native complex. The SDS-PAGE gel in Figure 2.6A show the result of co-IP experiments with silique tissue from wild type and three 35S:0t-myc plants. 73 Figure 2.6 In vivo co-immunoprecipitation of PKp subunits (A) Coomassie Brilliant Blue stained gel and anti myc immunoblot of co- immunoprecipitated (Co-1P) proteins. Total proteins were extracted from wild-type and 3SS:a-myc silique tissue and were subjected to Co-IP. About 37.5 pg of crude protein and half of the total eluate were loaded per lane. Bands unique to the 35S20t-myc co-IP lanes (in brackets) were excised and identified. IgG heavy chain is indicated on immunoblot for reference. WT, wild type. (B) Protein sequences, predicted transit peptides, and proteomics coverage of PK,- 0 ,PKp-Bl ,and PKp-B2. Gene Loci, encoded subunit, and percent coverage by proteomics are listed. ChloroP predicted chloroplast transit peptides are in bold. Peptide fragments identified by proteomics are highlighted in gray. 74 A Co-I P Crude BSSzo-myc WT 355:0-myc WT coomaSSIe 64 kDa 49kDa 37 kDa 64kDa _ ,» m _ ___ \Anti myc 49 kDa — ...... .- ..., .2...- o-myc “IgG heavy chain At3922960, PKP-a, 67% coverage A I .IIJ II 'I 'l I— r _' I':"""'"' l I— . l .1 A. KlJL !‘ ‘ ‘II. 1 ‘lv'i- ' '- 'i- . OJ." 5 5L .-. . ._ , iglrairiW ._‘ x . a." Illlullsll UVVI IIBVIIDVIIII ssrpartrsvsydgfaedvrvgdellvdggmvrfevleklg:if ' ' mlp‘tissltdwld‘ idfgiaegv dflavskasaevinh ayl " amvargdlgaq it e mlsgesamgqudkaltvlrtvslriemwreekrhesvplqai safadkif " a'VIv'ytlsghmasIvsrcrpdcpIfafttttsvrrrlnlqwgllpfrlsfsddmasnlnktffillksrgmIksgdlvr avsdmlqsiqvmnvp At5952920, PKp-B1, 21% coverage IIICH' I“ I “a r V C VaVlallvvuIsIJIou r. ' .1 I n x _‘L I 1 I u - A. 1 I l' l. r' "vale 2‘1 9 r 5,. IUUIIIIV _ . ‘ - ' I J] LI I -_n l u . IJAI .n m.- IIIIOIIHUIIOOI J IanIRUI gr :1 ,. 1r row A A: n J I I u .1 r J n I I | a r Iyuufvndhr’“ aa IIIIIIIVIgnoa u r ikkyl - I - .- - HIIDuaUII M .. ' __IT_ lIIIOIIla§UB amvargdlgaelpieevpilqeaunlcrsmgkavtvatnmlesm pm gad av liliagciai v .- a U . r iigqairxili I 1 Iaulllllalliiylaivw i4 I " Liy A icixixiqqi ', .w F., 1.: iaiaiiiixqgilivixixgcuiaiv qsgtqpimsqsthmqwkv At1g32440, PKp'Bz. 10% coverage :01 v r‘l v a a _‘1"""' npna l |_l l .1 .L J I Al ' L' J' I J L JL 1 1 ar .» a L L JI I IL .I II I lA' II .I‘. JJL A w I u ur a l' ‘Ir a 1 a J II I I L Al I J I AI LJ‘ J .I I l .I J‘ \IU I q '— a .1 u 4 . .1 . . .1 .1. . n ,a'ol Vnuunv 'l I‘llnl I] I' V V g r 'I I A A .l .1 I A L I I II LA. I v ,— ,_ T u v v uval _. . 1 u .l L 1 ll_ - n 11.. u ._ 1 l 1. - IIGGOIPVI. .. J a 9‘ ,_ a I ,_ 1 Illlql I I ' I .IJ _IA II II I II L .‘I ' LLI I 'I ' J ‘Ia r a a ‘l a-I '- a —-r ‘I as Figure 2.6 In vivo co-immunoprecipitation of PKp subunits 75 A small amount of a-myc protein was detected by an anti myc immunoblot in crude extracts from 3SSza-myc plants. After immunoprecipitation, the a-myc protein was enriched and became visible on CBB-stained gels. In addition, another protein running slightly faster than a-myc was visible on the CBB-stained gel. The CBB-stained and immuno-reactive bands running at 49 kDa and at a slightly less molecular mass were also present in the wild-type control. These were presumably products of the degradation of the anti-myc IgG during elution from the Protein-A sepharose. All three PKp subunits are very close in size and could co-migrate during SDS-PAGE. Therefore, a gel slice including a section above and below the a-myc protein from the co-IP reaction (indicated by brackets in Figure 2.6A) was excised and the contained proteins were subjected to tryptic digest and mass spectrometry (LC/MS/MS). Analysis of the mass spectrometry data identified peptides of all three PKp subunits with significant individual ion scores (Mowse scores >31, p<0.05): PKp-u (encoded by At3g22960), PKp-Bl (encoded by At5g52920), and PKp-Bz (encoded by At1g32440), with 29, 7, and 3 non-redundant peptides representing 67%, 21%, and 10% coverage of the predicted mature proteins, respectively (Figure 2.68). Kinetic characterization of PK complexes Enzyme activity analysis was performed using reconstituted PKp complexes. The maximum PK activity for the (1B; and 0432 heteromers was reached within 1 minute of mixing the subunits (Figure 2.7A). Reciprocal titrations in saturating substrate conditions showed that plots of PK activity versus subunit equivalents follow hyperbolic curves when one subunit is held constant and the other titrated (Figure 2.78). 76 > DJ 120 «$4 53100- E3 0 £2 9 80‘ g E 31 +Ca-VB1 3360 go- . . . . . .“'.CBl.'V° g E o 1 2 3 4 5 6 7 8 a 8 40] Equivalents < A ‘34 zoi 1; :3 E o E,2 =1 V +CG- C5325 go- . . . . . +PB2'egi :2,+gg1 §o123456789 c + 2 Equivalents E15- '5 g 1 €0.5- .2 *5 < o . . . . . o 1 2 3 4 5 6 pm ol/subunit Figure 2.7 Kinetics of active PKp complex formation (A) Time course of PK activity after mixing subunits. Numbers above bars represent the time (minutes) incubated prior to assay. 2 pmol of each subunit was used per assay under saturating substrate conditions. Activity is expressed relative to the 60 minute sample. (B) Subunit titration curves with PKp subunits. For Cor-VB], the a subunit was held constant and the [3; subunit was variable. For CBl-Va, the B1 subunit was held constant and the a subunit was variable. The same notation applies to the [32 titrations. 2.5 pmol of the constant subunit was used per reaction in saturating substrate conditions. Equivalents refers to the molar ratio of the variable subunit to the fixed one. (C) Pyruvate kinase activity relative to protein concentration. Equal amounts of either a and B. or a and [32 were mixed and assayed under sub-saturating substrate conditions. 77 Furthermore, when equal molar ratios of subunits were used, enzyme activity increased linearly with increasing protein concentration (Figure 2.7C). These results suggest that the association (and thus activity) of the subunits is dependent only on protein concentration and that there is little or no cooperativity of subunit association. Site-directed mutagenesis and chemical inactivation were performed to explore the subunit requirement in more detail (Figure 2.8). Both treatments were directed at a lysine residue in the PK active site required for phosphoryl group transfer. In the case of the site-directed mutant proteins (Figure 2.8A,B), PK activity was only observed when both a wild-type or and a wild-type [3 subunit were present. No activity was seen when only one subunit had a wild-type active site. The inclusion of a mutant protein extract did not inhibit the activity of wild-type complexes, indicating the contaminating bands seen in the SDM lanes of the SDS-PAGE gel of Figure 2.8A did not interfere with PK activity. Thus, the loss of PK activity was due only to mutation of the active site of either subunit of the PKp complex. Treatment with 2,4,6-trinitrobenzenesulfonic acid was employed to chemically inactivate the PK subunits. Inactivation was incomplete as determined by mixing TNBS-treated a and [3 subunits. Nonetheless, maximum PK activity was only observed with mixtures of mock-treated proteins. Activity was reduced 60-80% when either subunit was chemically inactivated. Addition of an inactivated subunit to an active mixture did not affect PK activity, meaning carryover of a component of the TNBS treatment was not responsible for reductions in PK activity. Therefore, the observed reductions in PK activity when using one TNBS-treated subunit in combination with a mock-treated one were due only to modification of lysine residues on the protein. Unmodified active sites in both subunits were required for PK catalytic activity. 78 A SDM WT a I31 I32 0 B1 82 CBB 82 kDa-— 64kDa—- * *W """" *——*-- 49 kDa- 82 kDa— ~ Anti His 64 kDa— a“ 49 kDa— - ‘4“ B 14 C 12 c12 C10 E10 35 3 o 9 5.8 0.6 m6 ‘3’ E E 4 5 4 3 2 2 0 0 N N N W I", N'} \ \ W W \ a a.» a; . a WWW. as We: saw é 5" $0.00 g2 a s s° Figure 2.8. PKp subunit inactivation (A) SDS-PAGE and anti-His immunoblot of affinity purified site-directed mutant (SDM) and wild-type (WT) PKp subunits. Approximately 15 pmol of each subunit was loaded per well. Conditions for protein purification were the same for all six proteins, but contaminating protein bands can be seen in the mutant protein lanes. The immunblot reveals which bands are the actual his-tagged PKp subunits (marked with * on CBB gel). (B) PK activity using wild-type and site-directed mutant (denoted with K—->L) proteins. Assays were done at pH 8.0 with saturating substrates and 2 pmol of each subunit. (C) PK activity using wild-type and chemically inactivated (denoted with i) proteins. Assays were done at pH 8.0 with saturating substrates and 2 pmol of each subunit. 79 Further kinetic experiments were done using equal molar ratios of 0431 or (132 mixtures. The enzymes had a strict requirement for Mg2+ and K+ and were completely inactivated after 3 min incubation at 50-55°C. The pH optima were found to be pH 7.8- 8.0 (Table 2.2) and subsequent reactions were conducted at pH 8.0. The V.,“,m of 01(3) was approximately 2-fold higher than that of uBz and the S0_5(s) for ADP and PEP were 1.5 and 3-fold lower, respectively, for 01131 (Table 2.2). Both complexes displayed sigrnoidal saturation kinetics for ADP and PEP, with 01B. having greater Hill coefficients (Table 2.2). Both enzymes were capable of catalyzing phosphoryl group transfer from PEP to NDPs other then ADP. Activities detected using 10 mM CDP, GDP, or UDP were 85, 57, and 67% and 13, 24, and 19% for 0431 and 0B2, respectively, when compared to an ADP control. Clearly, the (1132 enzyme preferred ADP while (1B; was less discriminatory. Table 2.2 PKp kinetic constants Enzyme Constant 11B. (1132 pH Optimum 8.0 7.8 Vmax (U/mg) 13.5 :t 0.3 7.6 :1: 0.4 PEP 80.5 (11M) 75.1 i 7.0 (2.0) 118.6 fl: 11.5 (1.2) ADP $0.5 (11M) 113.9 i 9.1 (1.9) 303.6 :I: 27.8 (1.2) K...t (s") 9.5 x 104 5.3 x104 Hill coefficients for PEP and ADP are given in parentheses. Values are the mean :1: SD (n=4). PEP, phosphoenolpyruvate; ADP, adenosine diphosphate. 80 Numerous metabolites and signaling compounds (fully listed in Materials and Methods) were tested as effectors of Arabidopsis PKp activity at subsaturating concentrations of PEP and ADP. Those compounds which had a significant effect are listed in Table 2.3. Only one activator, 6-phosphogluconate, was identified and it only activates 01132. The rest of the effectors acted as inhibitors with the most effective being glutamate and oxalate (Table 2.3). Values for the constants 150 and Ka were calculated (see Materials and Methods for definitions) for effectors capable of 50% relative activation or inhibition and they clearly show that (1B2 is the more sensitive enzyme with respect to these compounds. Neither enzyme was sensitive to treatment with dithiothreitol (DTT) or sodium tetrathionate (N aTT) suggesting a lack of redox regulation contrary to what was previously observed for plastidic glucose 6-phosphate dehydrogenase isoforms of Arabidopsis (Wakao and Benning 2005). Table 2.3 Metabolite effectors of PKp activity Concentration Relative Tested (mM) Activity (%) 150 (mM) K“ (mM) Metabolite 0101 (102 a0. 0102 uBl 0102 6PG 0.05 96 192 - - - 0.02 Glutamate 5 59 30 6.2 2.1 - - Oxalate 0.2 71 50 0.41 0.21 - - ISO-Citrate 10 70 75 ND. ND. - - AMP 1 77 96 ND. ND. - - ATP 1 77 87 ND. ND. - - Glyoxylate 5 83 79 ND. ND. - - OAA 2 82 88 ND. ND. - - Values represent the mean of at least four repeats. Activity is percent relative to a no effector control (set at 100). 6-PG, 6-phoshpogluconate; AMP, adenosine monophosphate; ATP, adenosine triphosphate; OAA, oxaloacetate; ND, not determined 81 Discussion Arabidopsis has two heterooctomeric PKps Previous kinetic analyses of plant PKs have been limited to enzymes which can be purified in the native state from dissected tissue samples (e. g. Hu and Plaxton 1996, Plaxton et al. 2002, Smith et al. 2000, Turner et al. 2005). The results of such studies have revealed that PK isoforms vary depending on the tissue and subcellular compartment in question. Arabidopsis tissues, and especially developing seed, represent a challenge to the traditional enzyme purification approach due to the difficulty in amassing the quantities of tissue required. However, this work is facilitated by the available resources and model characteristics of Arabidopsis and the findings are crucial if we want to achieve a complete understanding of the biology of this plant. A bioinformatics approach was taken to identify the candidate seed expressed PKp encoding genes in the Arabidopsis genome sequence. Four potential plastid targeted, seed-resident PKs were identified based on phylogenetics and gene expression data (Figure 2.1). The prediction of plastid localization of these proteins was subsequently confirmed using GFP fusion protein localization and pea chloroplast in viva import assays (Figure 2.3). Indeed, PKp-a and PKp-Bl have been found in the stromal fraction of chloroplasts using a proteomics approach (Friso et al. 2004). PKp-B. was also previously identified in a mitochondrial proteomics study, but plastid contamination could not be ruled out (Giege et al. 2003). Previous research on PKp-u orthologs from B. napus and R. communis detailed the proteins’ transit peptide cleavage sites (Plaxton et a1. 2002, Wan et a1. 1995). The N-terminal sequence of the mature B. napus protein aligns with a region about 40 amino acids c-terminal to the predicted Arabidopsis cleavage site, while the R. communis 82 protein has a transit peptide of 83 amino acids. These data suggest that the transit peptide of PKp-a may be longer than the predicted 55 amino acids. However, analysis of the observed molecular masses of the precursor and mature proteins afier pea chloroplast import support the predictions made byTargetP and ChloroP (Fig 3). Moreover, the Arabidopsis PKp-a protein contains a domain (79% amino acid identity to R. communis PKp-A) which is responsible for altered import characteristics of R. communis PKp-A (Wan et al. 1995). Our results indicate that this sequence does not affect the ability of Arabidopsis PKp-u to be imported into and processed in pea chloroplasts. It was determined that the Arabidopsis PKp enzymes are most likely ~460 kDa heterooctomers of 60 and 57 kDa subunits with 41140 stoichiometry (Figure 2.5). The reconstitution of active PKps from individually purified inactive subunits in this study as well as the in vivo Co-IP results leave little ambiguity as to the heteromeric structure of the Arabidopsis PKps. Some studies of recombinant PKp polypeptides from developing R. communis endosperm concluded that the PKp-u and PKp-B homologs are distinct enzymes (Blakeley and Dennis 1993, Blakeley et al. 1995, Wan et al. 1995), while others have determined that the same proteins are actually subunits of a single heteromeric PKp (N egm et al. 1995, Plaxton 1991, Plaxton et a1. 1990). The data presented here support the conclusion that PKp is a complex composed of two different subunits. Arabidopsis PKP activity is determined by two [3 subunits Kinetic analysis revealed that the reconstituted Arabidopsis enzymes behave much like previously documented PKps (pH optima of approximately 8.0, Mg2+ and K+ requirement, and $0.5 values-for PEP and ADP in the 100-300 11M range). A pH optimum of 8.0 is 83 potentially important for these enzymes. Such a pH is generated in the plastid stroma in response to light. As the plastids of Arabidopsis seeds are green and presumably photosynthetic, it is possible that in vivo PKp is light-activated via alkalinization of the stroma. Such regulation of PKp could contribute to the light-induced stimulation of fatty acid synthesis in green seeds (Goffman et al. 2005). What else was discovered were distinct differences between the two isoforms in Arabidopsis. The 01131 form is a more efficient enzyme with a higher specific activity and lower So. 5 values for both PEP (75 uM vs 120 11M) and ADP (Table 2.2). The concentration of PEP in potato tuber tissue has been estimated to be around 50-100 11M (F arre et al. 2001) and assuming the same is true for metabolically active Arabidopsis tissues, (1131 would be expected to be more active in vivo. In addition, (113. is 3-5 times more efficient at utilizing alternative nucleoside diphosphates. The 0102 enzyme, on the other hand, is more responsive to the strong metabolite effectors glutamate, oxalate, and 6-PG (Table 2.3). In fact, 031 is completely insensitive to the activating effect of 6-PG. This is interesting since what has been described as the major PKp from B. napus is activated by 6-PG (Plaxton et al. 2002). Apparently, Arabidopsis and B. napus have diverged in this regard as the gene for the [32 subunit present in the the 6-PG—regulated PK, (1102 enzyme) from Arabidopsis is hardly expressed in any tissue (Figure 2.18). A distinct feature of the Arabidopsis PKps is their inhibition by glutamate (Table 2.2). Regulation by this effector has been reported for PKcs from other plants but not PKps (Hu and Plaxton 1996, Smith et al. 2000). It should be noted that while the presence of either [31 or [32 in the Arabidopsis PKp complex results in different regulatory properties, these subunits should not be considered purely regulatory. They both contain fully conserved PK active sites in which chemical or 84 genetic modification results in an inactive PK complex (Figure 2.8). These experiments, in combination with the fact that the 01 subunit has no activity when assayed alone (Figure 2.48), support a model in which the both a and [3. or [32 subunits are required for enzyme activity and the specific interactions between the subunits result in differential kinetic properties. The primary structures of the recombinant PKps may not be the same as the native enzymes and thus the kinetic data must be considered with caution. However, as the transit peptide cleavage site predictions agree well with the observed molecular masses of precursor and mature PK subunits (Figure 2.38), the differences are likely to be minimal between the native and recombinant proteins. It should also be noted that specific post- translational modifications could result in changes to the enzyme structure and or function, and these possibilities were not explored. Soybean PKC, for instance, can be partially degraded at the c-terminus in vivo, which results in altered regulatory properties (Tang et al. 2003). Perhaps most significant is the realization that PKp from Arabidopsis exists as differentially regulated heteromers resulting from the interaction of the 01 subunit with one or the other [3 subunit. The two PKp-B subunits of Arabidopsis likely arose through gene duplication as indicated by the similarity of their amino acid sequence and gene structure and subsequently evolved unique regulatory features. The 043. enzyme is likely to be dominant in most tissues based on gene expression (Figure 2.1), but it is possible that under conditions when glycolysis needs to be more regulated the 0102 enzyme is produced at higher rates. The regulatory properties of the 01[3. enzyme, notably the lower 805 for ADP and PEP, higher Vmax, lower sensitivity to metabolic regulation, and ability 85 to utilize other NDPs more efficiently, make it a better enzyme for processing carbon at high rates, independent of the metabolic status of the tissue. The 0t[32 enzyme, however, is less active and more susceptible to metabolite-based regulation and may serve more specialized roles. It will be interesting to study the in vivo relationship of these subunits and whether or not one enzyme complex can contain a mixture of [3 subunits that would further fine tune PK activity for specific metabolic demands. 86 References Arabidopsis Genome Initiative (2000) Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature 408:796-815. Ballicora, M.A., Dubay, J .R., Devillers, C.H., and Preiss, J. (2005) Resurrecting the ancestral enzymatic role of a modulatory subunit. J. Biol. Chem. 280:10189-10195. Blakeley, SD. and Dennis, D.T. (1993) Molecular approaches to the manipulation of carbon allocation in plants. Can. J. Bot. 71:765—778. Blakeley, S., Gottlob—McHugh, S., Wan, J., Crews, L., Miki, B., Ko, K., and Dennis, D.T. (1995) Molecular characterization of plastid pyruvate kinase from castor and tobacco. Plant Mol. Biol. 27:79-89. Bruce, B.D., Perry, 8., Froehlich, J., and Keegstra, K. (1994) In vitro import of protein into chloroplasts. In SB Gelvin, RB Schilperoort, eds, Plant Molecular Biology Manual Kluwer Academic Publishers, Boston, pp 1-15 Cemac, A. and Benning,C. (2004) WRINKLEDI encodes an AP2/EREB domain protein involved in the control of storage compound biosynthesis in Arabidopsis. Plant J. 40:575- 585. Clough, SJ. and Bent, A.F. (1998) Flora] dip: a simplified method for Agrobacterium- mediated transformation of Arabidopsis thaliana. Plant J. 16:735-743. Emanuelsson, 0., Nielsen, H., Brunak, S., and Von Heijne, C. (2000) Predicting subcellular localization of proteins based on their N-terminal amino acid sequence. J. Mol. Biol. 300:1005-1016. Emanuelsson, 0., Nielsen, H., and Von Heijne, G. (1999) ChloroP, a neural network- based method for predicting chloroplast transit peptides and their cleavage sites. Protein Sci. 82978-984. Farre, E.M., Tiessen, A., Roessner, U., Geigenberger, P., Tretheway, R.N., Willmitzer, L. (2001) Analysis of the compartmentation of glycolytic intermediates, nucleotides, sugars, organic acids, amino acids, and sugar alcohols in potato tubers using a nonaqueous fractionation method. Plant Physiol. 127:685-700. Focks, N. and Benning, C. (1998) wrinkled] : A novel, low-seed-oil mutant of Arabidopsis with a deficiency in the seed-specific regulation of carbohydrate metabolism. Plant Physiol. 118:91-101. Friso, G., Giacomelli, L., Ytterberg, A.J., Peltier, J.B., Rudella, A., Sun, Q., and van Wijk, K.J. (2004) ln-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. 87 Giege, P., Heazlewood, J.L., Roessner-Tunali, U., Millar, A.H., Fernie, A.R., Leaver, C.J., and Sweetlove, L.J. (2003) Enzymes of glycolysis are functionally associated with the mitochondrion in Arabidopsis cells. Plant Cell 15:2140-2151. Goffman, F.D., Alonso, A.P., Schwender, J ., Shachar-Hill, Y., and Ohlrogge, J.B. (2005) Light enables a very high efficiency of carbon storage in developing embryos of rapeseed. Plant Physiol. 138:2267-2279. Hattori, J., Baum, B.R., Mchugh, S.G., Blakeley, S.D., Dennis, D.T., and Miki, B.L. (1995) Pyruvate kinase isozymes: Ancient diversity retained in modern plant cells. Biochem. Syst. Ecol. 23:773-&. Hollenberg, P.F., Flashner, M., and Coon, M.J. (1971) Role of lysyl e-amino groups in adenosine diphosphate binding and catalytic activity of pyruvate kinase. J. Biol. Chem. 246:946-953. Hu, Z.Y. and Plaxton, W.C. (1996) Purification and characterization of cytosolic pyruvate kinase from leaves of the castor oil plant. Arch. Biochem. Biophys. 333298- 307. Jackson, D.T., F roehlich, J.E., and Keegstra, K. (1998) The hydrophilic domain of Ticl 10, an inner envelope membrane component of the chloroplastic protein translocation apparatus, faces the stromal compartment. J. Biol. Chem. 273:16583-16588. Koncz, C. and Schell, J. (1986) The promoter of Tl-Dna gene 5 controls the tissue- specific expression of chimeric genes carried by a novel type of Agrobacterium binary vector. Mol. Gen. Genet. 204:383-396. Laemmli, U.K. (1970) Cleavage of structural proteins during assembly of head of bacteriophage-T4. Nature 227 :680-&. Li, KB. (2003) ClustalW-MPI: ClustalW analysis using distributed and parallel computing. Bioinformatics 19:1585-1586. Munoz, M.E. and Ponce, E. (2003) Pyruvate kinase: current status of regulatory and functional properties. Comp. Biochem. Physiol. B: Biochem. Mol. Biol. 1352197-218. Negm, F .B., Cornel, F.A., and Plaxton, W.C. (1995) Suborganellar localization and molecular characterization of nonproteolytic degraded leukoplastid pyruvate kinase from developing castor-oil seeds. Plant Physiol. 109: 1461-1469. Ohlrogge, J., and Browse, J. (1995) Lipid biosynthesis. Plant Cell. 7 2957-970. Olsen, L.J. and Keegstra, K. (1992) The binding of precursor proteins to chloroplasts requires nucleoside triphosphates in the intermembrane space. J. Biol. Chem. 267 :433- 439. 88 Page, R.D.M. (1996) TreeView: An application to display phylogenetic trees on personal computers. Comput. Appl. Biosci. 12:357-358. Plaxton, W.C. (1991) Leucoplast pyruvate kinase from developing castor oil seeds: Characterization of the enzyme’s degradation by a cysteine endopeptidase. Plant Physiol. 97: 1334-1338. Plaxton, W.C. (1996) Organization and regulation of plant glycolysis. Annu. Rev. Plant Physiol. Plant Mol. Biol. 47:185-214. Plaxton, W.C., Dennis, D.T., and Knowles, V.L. (1990) Purification of leucoplast pyruvate kinase from developing castor bean endosperm. Plant Physiol. 94: 1528-1534. Plaxton, W.C., Smith, C.R., and Knowles, V.L. (2002) Molecular and regulatory properties of leucoplast pyruvate kinase from Brassica napus (rapeseed) suspension cells. Arch. Biochem. Biophys. 400254-62. Rivoal, J., Smith, C.R., Moraes, T.F., Turpin, D.H., and Plaxton, W.C. (2002) A method for activity staining after native polyacrylamide gel electrophoresis using a coupled enzyme assay and fluorescence detection: Application to the analysis of several glycolytic enzymes. Anal. Biochem. 300:94-99. Ruuska, S.A., Girke, T., Benning, C., and Ohlrogge, J.B. (2002) Contrapuntal networks of gene expression during Arabidopsis seed filling. Plant Cell 14:1191-1206. Sakai, H. (2005) Mutagenesis of the active site lysine 221 of the pyruvate kinase from Bacillus stearothermophilus. J. Biochem. (Tokyo) 137:141-145. Sangwan, R.S., Gauthier, D.A., Turpin, D.H., Pomeroy, M.K., and Plaxton, W.C. (1992) Pyruvate kinase isoenzymes from zygotic and microspore-derived embryos of Brassica napus — Developmental profiles and subunit composition. Planta 187:198—202. Schmid, M., Davison, T.S., Henz, S.R., Pape, U.J., Demar, M., Vingron, M., Scholkopf, B., Weigel, D., and Lohmann, J.U. (2005) A gene expression map of Arabidopsis thaliana development. Nature Genet. 37:501-506. Schramm, A., Siebers, B., Tjaden, B., Brinkmann, H., and Hensel, R. (2000) Pyruvate kinase of the hyperthermophilic crenarchaeote Thermoproteus tenax: Physiological role and phylogenetic aspects. J. Bacteriol. 182:2001-2009. Schwender, J. and Ohlrogge, J.B. (2002) Probing in vivo metabolism by stable isotope labeling of storage lipids and proteins in developing Brassica napus embryos. Plant Physiol. 130:347-361. Shen, W.J. and Forde, BC. (1989) Efficient transformation of Agrobacterium spp by high-voltage electroporation. Nucleic Acids Res. 17:83 85. 89 Smith, C.R., Knowles, V.L., and Plaxton, W.C. (2000) Purification and characterization of cytosolic pyruvate kinase from Brassica napus (rapeseed) suspension cell cultures - Implications for the integration of glycolysis with nitrogen assimilation. Eur. J. Biochem. 267:4477-4485. Tang, G.Q., Hardin, S.C., Dewey, R., and Huber, S.C. (2003) A novel C-terminal proteolytic processing of cytosolic pyruvate kinase, its phosphorylation and degradation by the proteasome in developing soybean seeds. Plant J. 34:77-93. Tranel, P.J., Froehlich, J ., Goyal, A., and Keegstra, K. (1995) A component of the chloroplastic protein import apparatus is targeted to the outer envelope membrane via a novel pathway. Embo J. 14:2436-2446. Turner, W.L., Knowles, V.L., and Plaxton, W.C. (2005) Cytosolic pyruvate kinase: subunit composition, activity, and amount in developing castor and soybean seeds, and biochemical characterization of the purified castor seed enzyme. Planta 222:1051-1062. Verwoerd, T.C., Dekker, B.M.M., and Hoekema, A. (1989) A small-scale procedure for the rapid isolation of plant RNAs. Nucleic Acids Res. 17:2362. Wan, J., Blakeley, S.D., Dennis, D.T., K0, K. (1995) Import characteristics of a leucoplast pyruvate kinase are influenced by a 19-amino acid domain within the protein. J. Biol. Chem. 270:16731-16739. Voinnet, O., Rivas, S., Mestre, P., and Baulcombe, D. (2003) An enhanced transient expression system in plants based on suppression of gene silencing by the p19 protein of tomato bushy stunt virus. Plant J. 33:949-956. Wakao, S. and Benning, C. (2005) Genome-wide analysis of glucose-6-phosphate dehydrogenases in Arabidopsis. Plant J. 41:243-256. White, J .A., Todd, J ., Newman, T., Focks, N., Girke, T., de Ilarduya, O.M., Jaworski, J .G., Ohlrogge, J.B., and Benning, C. (2000) A new set of Arabidopsis expressed sequence tags from developing seeds. The metabolic pathway from carbohydrates to seed oil. Plant Physiol. 124:1582-1594. 9O Chapter 3 Analysis of carbon metabolism and storage compound accumulation in seeds of an Arabidopsis mutant deficient in a plastidic pyruvate kinase2 2 This work has been published in Andre, C., Froehlich, J .E., Moll, M.R., and Benning, C. (2007) A heteromeric plastidic pyruvate kinase complex involved in seed oil biosynthesis in Arabidopsis. Plant Cell 1921-17. [performed all of the experiments shown here. 91 Abstract Glycolysis is a ubiquitous pathway thought to be essential for the production of oil in developing seeds of Arabidopsis and oil crops. Compartmentation of primary metabolism in developing embryos poses a significant challenge towards testing this hypothesis and for the engineering of seed biomass production. It also raises the question whether there is a preferred route of carbon from imported photosynthate to seed oil in the embryo. Disruption of the gene encoding the [31 subunit of plastidic pyruvate kinase (PKp) causes a 75% reduction in enzyme activity and a 60% reduction in seed oil content. The seed oil phenotype is fully restored by expression of the [3. subunit-encoding cDNA, and partially by the [32 subunit-encoding cDNA. Additionally, carbohydrates accumulate in the mutant seeds, possibly due to reduced activity of upstream glycolytic enzymes. Therefore, the identified pyruvate kinase catalyzes a crucial step in the conversion of photosynthate into oil suggesting a preferred plastid route from its substrate phosphoenolpyruvate to fatty acids. 92 Introduction An important metabolic function of a developing Arabidopsis (Arabidopsis thaliana) seed is the deposition of storage reserves: oil in the form of triacylglycerols (TAG), but also proteins, oligo- and polysaccharides (Baud et al. 2002). As sucrose is the major photosynthetic product transported in the phloem the embryo is required to catabolize incoming sucrose and convert it into the more efficient storage compounds mentioned above. Glycolysis is central to this process as it converts sugars into precursors for protein and fatty acid synthesis while concomitantly producing ATP by substrate level phosphorylation. In fact, stable isotope labeling has been used to demonstrate that 90% of glucose fed to developing canola (Brassica napus) (a close relative of Arabidopsis and oilseed crop) embryos is converted to pyruvate by the RuBisCO bypass and the lower half of glycolysis (Schwender and Ohlrogge 2002, Schwender et al. 2004a). Furthermore, a clear link between glycolysis and seed metabolism is apparent in the wrinkled] (wriI) mutant of Arabidopsis, in which a general reduction in glycolytic activity results in an 80% reduction in seed oil (Focks and Benning 1998). Glycolysis in plants occurs in both the cytosol and the plastid and both compartments are connected through plastid membrane transporters (Plaxton 1996, Weber 2004). The intermediates of upper glycolysis of developing B. napus embryos appear to be in near equilibrium between the cytosol and plastid (Schwender et al. 2003), and enzymes of the full glycolytic sequence have been detected in both compartments in embryos (Eastmond and Rawsthome 2000). It is therefore likely that changes in glycolytic enzyme activities influence the transport of related intermediates across the plastid envelope. This compartmentation raises the question of a preferred route of 93 glucose metabolism in developing oil seed embryos. Expressed sequence tag (EST) analysis of developing Arabidopsis seeds indicated that mRNAs encoding cytosolic enzymes for the entire glycolytic pathway are abundant, but that only mRNAs encoding plastidic enzymes for the second half of the pathway metabolizing trioses are abundant (White et al. 2000). Microarray experiments extended these findings by detecting a shift in expression from genes encoding cytosolic glycolytic enzymes to those encoding plastidic glycolytic enzymes at the onset of storage compound accumulation. While gene expression does not necessary reflect enzyme activity or metabolite flux, the microarray data led to the hypothesis that glucose is broken down in the cytosol to phosphoenolpyruvate (PEP), which is then imported into the plastid and metabolized by plastidic pyruvate kinase (PKp) (Figure 1.2, Ruuska et al. 2002). Revisions to this scheme were introduced based on more recent experiments (2'. e. RuBisCO bypass, see Schwender et al. 2004a) but in the model the lower half of glycolysis metabolizing PEP remains unaltered. While the import of PEP by plastids of B. napus embryos has been demonstrated (Kubis et al. 2004), the chlorophyll a binding protein underexpressed (cue!) mutant of Arabidopsis lacking a seed-expressed PEP transporter has no reported seed reserve phenotype (Li et al. 1995). A recent steady-state carbon flux analysis on cultured embryos of B. napus led to the proposal of a RuBisCO shunt involving reactions of the reductive pentose phosphate pathway in the plastid (Schwender et al. 2004a). The proposed pathway bypasses the initial glycolytic reactions in the cytosol and generates PEP in the plastid, which could compensate for the PEP import deficiency in the cue] mutant. Alternatively, pyruvate generated by cytosolic pyruvate kinase (PKC) could be imported and uSed directly for fatty acid biosynthesis. Isolated plastids from B. napus 94 embryos are capable of incorporating l4C labeled pyruvate into fatty acids (Eastmond and Rawsthome 2000, Kang and Rawsthome 1994), but no plastidic pyruvate transporter has been reported. A current model of primary metabolism in developing Arabidopsis seeds is shown in Figure 1.2. Direct molecular or genetic corroboration of this scheme is generally lacking and the focusing on PKp in developing Arabidopsis seeds should be highly informative in the testing of this hypothesis. 95 Materials and Methods Plant growth and transformation All Arabidopsis plants were of the Col-2 ecotype, except for the SALK T-DNA lines which were Col-0. All seeds were first sterilized in 20% bleach, 0.05% TritonX-lOO for 20 min and were then rinsed 5 times with water and plated on half-strength MS medium, pH 5.9, 0.9% agar, and 2% sucrose. When appropriate, kanamycin or hygromycin B were included in the medium at 50 pg mL'1 and 25 pg mL", respectively. Seeds were stratified at 4°C for 3 days prior to being germinated in an incubator (AR-75; Percival Scientific, Boone, IA, USA) at a photon flux density of 60—80 pmol m_2 sec-I and a light period of 16 h (22°C), and a dark period of 8 h (18°C). Seedlings were transferred to 3.5 inch square pots and were grown in a soil mix as described previously (Xu et al. 2002) and were grown under a 16-h photoperiod with a day temperature of 22°C and a night temperature of 20°C at a photon flux density of 100—120 umol m—2 sec‘l. The plants were fertilized with half-strength Miracle-Gro (Scotts, Marysville, OH, USA) plant food every fifteen days. Wild-type and mutant Arabidopsis plants were prepared for transformation as previously described (Cemac and Benning 2004). When ready, plants were transformed using the floral dip method (Clough and Bent 1998). Competent cells ongrobacterium tumefaciens strain C58Cl GV3101 pMP90 (Koncz and Schell 1986) were prepared and transformed as previously described (Shen and Forde 1989). T-DNA mutant isolation and characterization T-DNA insertion lines were obtained from the SALK T-DNA insertion population (Alonso et al. 2003). Mutants were selected on growth medium containing kanamycin 96 and T-DNA insertions were confirmed using PCR primers specific for gene sequences and the T-DNA left border. Primers were designed using the i-sect tool (http://signal.salk.edu/tdnaprimers.html). Insertion sites were confirmed by sequencing the PCR product of the left border primer and a gene specific primer. Expression of the target gene was analyzed with RT-PCR using primers listed in Table 2.1. Total RNA was isolated from seedlings using the Qiagen RNeasy kit. 600 ng of total RNA was used for reverse transcription using the Qiagen Omniscript RT kit. PCR was done using 5% of the RT product. Actinl (At2g37620) was used for control purposes. The PCR consisted of 25 cycles of 95°C for 30 sec, 60°C for 45 sec, and 72°C for 1 min, followed by a 10 min 72°C extension. Complementation of the mutant was done using full length cDNAs inserted into the Kpnl site of the CaMV 353 containing pCAMBIAl300 derivative mentioned above. An antisense construct was generated using the same full length cDNA for PKp-B, inserted in the antisense orientation into the Kpnl site of the binary vector pBinAR-Hyg (Dormann and Benning, 1998) containing the 12S seed storage protein promoter from Arabidopsis (Ohlrogge J, Benning C, Gao H, Girke T, and White J, Inventors; Plant seed specific promoters. US patent 7,081,565. 2006 July 25). Analysis of gene expression in the rescued lines was done using RNA gel blots. Total silique RNA was extracted using a previously described protocol (Verwoerd et al. 1989) followed by DNase treatment and cleanup using the Qiagen RNeasy kit. Northern analysis (5 pg total RNA) was performed as previously described (D6rmann and Benning 1998). The blots were analyzed using a phosphor imager (Molecular Dynamics, Amersham, Piscataway, NJ, USA). 97 Analysis of chlorophyll content Chlorophyll was quantified in seeds as previously described (Lichtenthaler 1987). Seeds were imaged using a Leica M2 12.5 dissecting microscope (Leica Microsystems, Wetzlar, Germany) equipped with a Spot Insight color camera (Diagnostic Instruments, Sterling Heights, MI, USA). Transmission electron microscopy For electron microscopy, seeds were dissected out of staged silques and soaked in water for 1 hour. Embryos were then expelled from their seed coats by pressing the soaked seeds between two glass microscope slides and were then embedded in 2% agarose. The embryos were fixed for 2 hours at room temperature with 2.5% glutaraldehyde, 2.5% paraformaldehyde in 0.1 M cacodylate buffer and then post-fixed in 1% (w/v) osmium tetrachloride in 0.1 M cacodylate buffer. The samples were then dehydrated in a graded series of acetone, embedded in Poly BD 812 resin, and sectioned. The thin sections (~70 to 100 nm) were stained with uranyl acetate and lead citrate prior to examination in a J EOL 100CX electron microscope (J EOL, Japan). Developing seed enzyme assays Enzyme activities were measured from developing seed proteins extracted in a buffer of 50 mM Tris-Cl pH 7.5, 5 mM MgC12, 1 mM EDTA, 1 mM EGTA, lmM DTT, 0.1% Triton-X100, 10% glycerol, 2 mM benzamidine, 2 mM e-amino-n-caproic acid, and 1 mM PMSF. Pyruvate kinase activity was detected by coupling the production of pyruvate to the conversion of NADH to NAD+ by lactate dehydrogenase. Reactions were kept at 98 25°C, were started by the addition of enzyme mix, and were linear for at least 5 minutes. Absorbance at 340 nm was monitored using a FLUOstar Optima 96-well plate reader (BMG Labtech, Offenburg, Germany). Standard PKp reaction mixtures contained 50 mM HEPES-KOH pH 8.0, 5% PEG-8000, 50 mM KCl, 15 mM MgC12, 1 mM DTT, 2 mM PEP, 1 mM ADP, 0.2 mM NADH, and 2 U ml'l desalted rabbit muscle lactate dehydrogenase. PEP phosphatase activity was corrected for by omitting ADP from the reaction. Reactions at pH 7.0 were done using 50 mM MOPS pH 7.0 instead of HEPES. ATP-dependent phosphofructokinase (PFK) and pyrophosphate-dependent phosphofructokinase (PFP) activities were measured as previously described (Burrell et al. 1994). Seed metabolite analysis Seed oil quantification by fatty acid methyl ester analysis was done as previously described (Focks and Benning 1998). Seed storage proteins were extracted from 50 mg of dry seeds by first grinding in a mortar and pestle followed by 2 extractions of the tissue with 30 volumes of hexane. The delipidated seed material was pelleted by centrifugation at 13,000 g for 10 min. The pellet was then dried in a speed-vac and extracted twice for 15 min with 0.5 volumes of 50 mM Tris-HCl pH 8, 200 mM NaCl, 5 mM EDTA, 0.1% TWeen-20, 2 mM benzamidine, 2 mM s-amino-n-caproic acid, 1 mM PMSF. Water bath sonication was used to resuspend the pellets. The supernatants from the extractions were combined and a 1:5 dilution was used to quantify protein using the Bio-Rad DC Protein Assay Kit. For SDS-PAGE, the equivalent 3 seeds worth of total protein was loaded per lane. Mature seed free amino acids were extracted from 25 mg of seed tissue 3 times with 99 400 uL of 70% methanol. The supematants were combined and extracted 2 times with an equal volume of chloroform to remove the lipids. The remaining aqueous phase was dried under vacuum and then resuspended in 20 mM HCL. Extracts (20 uL) were then loaded and run on over a strong anion exchange column using a Hitachi Amino Acid Analyzer (Hitachi High Technologies America, Inc., San Jose, CA, USA) at the MSU Macromolecular Structure, Sequencing, and Synthesis Facility. Glucose, fructose, sucrose, and starch were extracted from developing seeds and quantified as previously described (Focks and Benning 1998). PEP and pyruvate were extracted from developing seeds with perchloric acid and quantified using a NADH fluorescence assay as previously described (Hausler et al. 2000). ADP and ATP were measured in the same extracts with an ATP Bioluminescence Assay Kit (Sigma) using a previously described protocol (Ruuska et a1, 2000). 100 Results Disrupting the PKp-Bl encoding gene causes a reduction of PKp activity and seed oil content Multiple independent SALK T-DNA Insertion lines were obtained to study the in vivo function of the Arabidopsis PKps (Alonso et a1. 2003). Six lines were reanalyzed for exact insertion site location (PKp-a; SALK_096141, SALK_024870, PKp-Br; SALK_042938, SALK_042681, PKp-Bz; SALK_013574, SALK_142845), but only one (SALK_042938 in PKp-Bl) carried an insertion in a translated portion of the gene (Figure 3.1A). The line SALK_096141 has an insertion in an intron of the PKp-a encoding gene, but RNA levels were unaffected in these plants. Genotyping by PCR revealed homozygous SALK_042938 mutants completely lacking transcripts from the PKp-B. encoding gene as detected by RT-PCR (Figure 3. l B, C). Transcript amounts for the PKp-a and PKp-Bz encoding genes were unchanged (Figure 3.2). The mutant will be referred to as pka . Notably, of the six analyzed mutants, only 19ka had any visible seed phenotype. The mature seeds of pka were wrinkled when observed under a dissecting microscope and developing mutant embryos contained less chlorophyll (Figure 3.3A, B). Rescue of pka with ectopic expression of the PKp-B. encoding gene restored chlorophyll content to nearly wild-type levels, however, overexpressing the PKp—Bz encoding gene was much less effective. Transmission electron microscopy was done to assess any ultrastructural perturbations in the pkpl mutant. The 5k-magnification (Figure 3.3C, upper panels) micrographs show representative cotyledonary cells from 13 DAF embryos. Wild-type cells are full of oil bodies (ob) by this time and still contain some starch (st) in the plastids. The pka mutant has much smaller oil bodies than wild type, but has much 101 larger starch granules. The 67k-magnification images (Figure 3.3C, lower panels) show in more detail the thylakoid structure of the mutant. While organized similarly, the pkp] thylakoids are less extensive than wild type. A SALK_096141 SALK_024870 SALK_013574 SALK_142845 no . SALK_042938 SALK_042681 TDNA LP RTi RP RT2 Actin1 RT1/RT2 1Kb Figure 3.]. Identification of a SALK T-DNA mutant in PKp-Bl (A) Gene structure of the three PK subunits and locations of T-DNA insertions. Black box on T—DNA is the left border. LP and RP depict locations of primers used for genotyping in (B). RT] and RT2 depict locations of primers used for RT-PCR in (C). ATG, start codon; STOP, stop codon. (B) PCR based genotyping of SALK_042938. LP and RP refer to At5g52920 specific primers shown in (A) and in Table 2.1. LB refers to the T-DNA lefi border primer in Table 4. WT, wild type. (C) RT-PCR to measure PKP-Bl encoding gene expression in SALK_042938. Actinl (At2g3 7620) is the control. RTl and RT2 refer to At5g52920 specific primers shown in (A) and in Table 2.1. WT, wild type 102 Actin WT pkp1 1200bp 800bp 400bp m . 74...... w ~— —_ _ — a——. m ~ ”on Figure 3.2. Reverse transcriptase-PCR analysis of PKp gene expression PKp-a and PKp-Bz gene expression in 11 days afler flowering seeds of the pka and WT background. Actinl (At2g37620) is the control. WT, wild type. 103 ADAF5 7 13 15]dry mlmflflfifl . manual 5 ED —* N N (A) 01 O 01 o r r r 10' 5- ng chlorophyll/seed .5 '7 9 1'1 1'3 1'5 17 Days After Flowering Figure 3.3. pkpl seed phenotypes (A) Seed phenotypes of pka and wild type (WT). Embryos were dissected out of developing seeds at the time (DAF, days after flowering) indicated. Fully dessicated mature seeds are shown as well. The bar represents 0.2 mm. (B) Total chlorophyll content in developing seeds of pka, wild type (WT), and lines rescued with CaMV 3SS-driven expression of either PKp-B. (RBI-23) or PKp-Bz (RB2-3). Forty seeds were measured per sample. Values are the mean i SD (n=6). (C) Electron micrographs of cells from 13 DAF wild-type (WT) and pkp] cotyledons. Starch granules (st) and oil bodies (0b) are marked with arrows in the upper panels. Higher magnification in the lower panels reveals thylakoid membranes (th) inside plastids. 5k and 67k denote magnification used. Asterisks (*) protein bodies. Bars in upper panels represent 2 pm. Bars in lower panels represent 0.5 pm. Lefi panels, wild type; right panels, pkpl mutant. 104 The smaller oil bodies, reduced thylakoid membranes, and wrinkled seeds of pka suggest a reduction in lipid biosynthesis and possibly storage compound accumulation. Therefore, oil and protein were quantified in the mature pka seeds (Table 3.1). The mutant accumulated only 40% as much oil as wild type, yet there was only a 15% reduction in protein. Test crosses and analysis of the F1 progeny indicated that the low oil phenotype of pkpl is a recessive trait largely dependent on the genotype of the embryo. A homozygous pkpl sporophyte did however result in a 15% reduction in seed oil when pollinated with wild-type pollen (Table 3.1) indicating a small maternal effect. Table 3.1 WT and [7ka seed storage compound accumulation WT .0ka 2;)? pkg/469x Wag/”261$ 6.77 i 0.7 2.71 i 0.2 5.68 :t 0.8 6.9 i 0.2 (“Z/13:23 5.23 i 0.6 4.39 :t 0.5 ND. ND. SEE/$23; 19.2 :1: 0.9 14.4 i 1.1 ND. ND. Values are the mean of three repeats 3: standard deviation. Seed mass determined by measuring the weight of 500 seeds three times. ND, not determined Enzyme activities in developing seeds were measured to determine the extent of the PKp defect in pka . The presence of a potentially large number of different cytosolic and plastidic PK isoforms complicates the interpretation of activity assays using crude extracts. However, two factors aide in validating seed PKID activity in crude extracts: 1. all 14 PK encoding genes are not highly expressed in any given tissue at the same time (Schmid et al. 2005), 2. cytosolic PKs typically have pH optima of approximately pH 7.0, whereas plastidic PKs have pH optima of approximately pH 8.0 (Hu and Plaxton 1996, 105 Plaxton et al. 2002, Smith et al. 2000, this work). Figure 3.4A shows the time course of PK specific activity using protein extracts of seed dissected from staged siliques. At pH 8.0 wild-type PK specific activity is greatest at 7 days after flowering (DAF) and steadily declines throughout seed development. This pattern agrees with the expression profiles of PKp-u and PKp-Bl encoding genes as shown in Figure 2.1, but is shifted to later DAF possibly due to differences in growth conditions. The pkp] mutant at pH 8.0, however, has 3-fold reduced PK specific activity at 7 DAF and does not change within experimental limitations throughout the rest of the time course. Pyruvate kinase specific activity was not reduced in the mutant at pH 7.0 but was actually increased at 11 and 15 DAF. Mutant and wild-type protein extracts were also assayed in the presence of 5 mM glutamate or 0.2 mM 6-phosphogluconate (Figure 3.48). Recall that glutamate at 5 mM inhibited recombinant (1B1 by about 40%, while 01132 was inhibited by 70% (Table 2.2). Native PK specific activity in 9-11 DAF wild-type seed extract was inhibited only about 25% by 5 mM glutamate while that from pka was unaffected. As shown in Chapter 2, 6- phosphogluconate is a potent activator of 01132 (Table 2.2), yet had no effect on wild-type or pka seed PK specific activity. 106 > 500 3: 300‘ 350 None 0 g 200‘ .E 300‘ lg'U E 1001 .g 250, -6PG a 0 pH 7' i ' T ' 3’ 200‘ . c 400‘ 5 1501 g E 1001 . , a 300‘ 50 1 7;, :5) 1001 WT9-11DAF pkp19-11DAF o - - - . . 5 7 9 11 13 15 17 Days After Flowering Figure 3.4. PK specific activity in pkp1 and wild-type seeds (A) Total PK specific activity measured at pH 8 and pH 7 in saturating substrate conditions. WT, wild type. One mU is defined a 1 nmol pyruvate formed per minute. Values are the mean i SD (n=4). (B) Native seed PK specific activity at 9-11 DAF in response to metabolite effectors. Activity was measured at pH 8.0 with subsaturating substrate concentrations. WT, wild type. None, no effectors; glu, 5 mM glutamate; 6-PG, 0.2 mM 6-phosphogluconate. One mU is defined a 1 nmol pyruvate formed per minute. Values are the mean :I: SD (n=4). 107 Seed fatty acid and protein profiles in pkp1 Fatty acid methylester (FAME) analysis of developing seeds revealed that pkp1 accumulates oil in the same temporal pattern as wild type, but at a much lower rate (Figure 3.5A). The fatty acid composition of the oil in mature seeds was also analyzed. The pkp1 mutant had a decrease in stearic (18:0), oleic (18:1), and linoleic (18:2) acids along with an increase in linolenic (18:3). The proportion of very long chain (20 and 22 carbons) to long chain (16 and 18 carbons) fatty acids was also increased in pkp1 (Figure 3.5B). Total protein extracts were made from mature wild-type and pkp1 seeds and were run on an SDS-PAGE gel to analyze the storage protein profile. Apparently, there are no differences in the major storage protein profiles between wild type and pkp1 (Figure 3.5C) and the 15% reduction in protein content seen in pkp1 (Table 3.1) is not specific to any one protein. Free amino acid content of mature seeds was also compared between wild type and the mutant. Analysis of amino acid extracts by HPLC revealed increases in the proportions of glycine, arginine, and glutarnine in pkp1 (Figure 3.5D). On the other hand, aspartate, aspargine, glutamate, and valine were in greater proportion in the wild- type seeds. Statistically, however, the differences were very small between wild type and pkp1 for any of the listed amino acids. Restoration of seed oil content in pkp1 expressing B-subunit-encoding cDNAs Dealing with a single T-DNA insertion line, which could potentially harbor secondary mutations that cannot be detected by genotyping, required additional precautions to unambiguously link the target genotype with the observed phenotype, in this case the reduction in oil content. 108 A 7 C WT pkp1 3 +WT —— g6 +pkp1 B15 ,._. M 3 m4‘ E3 “wwfi‘m‘fi’flj—QS-a 52‘ “3--“‘5330—128-5 o 1, .— 5 7 9 11 13 15 '7 Days After Flowering 835 .WT 30 kap1 25 °\° . —020 E15 10 I 5 0 I I I E a ‘P. 0 ‘.'. 9‘. 9. 0" ‘.'. 9 ‘.'. :2 <2 2 e a e a a a. Figure 3.5. Oil and protein phenotype of pkp1 seeds (A) Fatty acid accumulation in developing seeds. DAF, days after flowering; FAME, fatty acid methyl ester; WT, wild type. Values are the mean 5: SD (n=6). (B) Fatty acid profile of desiccated mature seeds. Values obtained from FAME analysis of dry seeds. WT, wild type. Values are the mean :t SD (n=6). (C) CBB stained SDS-PAGE gel of total protein extracts from wild-type (WT) and pkp1 seeds. Major storage protein bands are denoted in the margin. (D) Free amino acid profiles of wild-type (WT) and pkp1 mature seeds. Values are the mean :t SD (n=3). 109 To address this issue we constructed transgenic lines in the pkp1 background expressing cDNAs that encode either PKp-Bl or PKp-Bl subunits and included these transgenic lines in the analysis. Expression of these cDNAs was expected to restore the oil content if this phenotype was due to the disruption in the PKp-B. encoding gene in the pkp1 line thereby confirming the link between genotype and phenotype. A T-DNA construct containing a CaMV 35S driven cDNA encoding PKp-fil was used to rescue the lipid phenotype of pkp1 . Antibiotic resistance was used to select transformants and rescued lines were identified by scoring for visual rescue of the wrinkled seed phenotype. Of 43 independent transformants, 8 appeared to be rescued based on seed morphology. Homozygous T3 seeds from individual rescued lines were subjected to FAME analysis (Figure 3.6A). Overexpression of the PKp-Bl encoding cDNA rescues the lipid phenotype of the mutant, but does not result in an increase in oil amount in excess of wild-type seed oil content in any of the lines. A PKp-Bz encoding cDNA was similarly overexpressed in pkp1 and transformants were selected. Twenty independent transformants were identified and 6 of these had rescued seed morphology. FAME analysis was performed on homozygous T3 seeds from these lines and showed rescue of the low-oil phenotype of pkp1 (Figure 3.6B). However, the restoration of oil content was not as complete as was achieved by overexpressing the PKp-Bl encoding cDNA. Overexpression of the cDNA encoding PKp-Bl was able to restore the fatty acid profile of pkp1 to wild type, while the PKp-Bz encoding cDNA overexpressors had the same fatty acid profile as the pkp1 mutant, despite being almost completely rescued in terms of oil accumulation (Figure 3.6C). 110 > o—xmcohwmxr CU Total FAME (pg/seed) O-INOJAUIO'DNCD Total FAME (pg/seed) MW : RB1-7 : RB1-22 :—_—+ RB1-23 :3 RB1-24 :21~ RB1-26 : RB1-31 : RB1-34 :1 R9143 :9 WT mol % I I ‘5‘. w 1— Figure 3.6. Rescue of the pkp1 seed oil phenotype I ‘3’. ‘7. 9 co 0 N v— N N I I I 9. 9 9. CD CD 0 \— 1— N (A) Oil amounts in mature seeds of pkp1 overexpressing the PKp-B. encoding cDNA. The 8 individual rescued lines are denoted with R131 and a number. WT, wild type. Values are the mean i SD (n=3). (B) Oil amounts in mature seeds of pkp1 overexpressing the PKp-Bz encoding cDNA. The 6 individual rescued lines are denoted with R132 and a number. WT, wild type. Values are the mean d: SD (n=3). (D) Average fatty acid profiles of the 8 RBI and 6 R132 lines compared to wild type (WT) and pkp1. Values are the mean at SD (n=6). 111 The complementation data strongly suggest that the seed oil phenotype of pkp1 is caused by the T-DNA insertion located in the PKp-Bl encoding gene. To fiirther exclude the possibility of second site mutations causing the phenotype, an antisense construct was generated to reduce the transcript amount of the PKp-Bl encoding gene specifically in seeds. The Arabidopsis l2S seed storage protein promoter was used for this purpose. Of twenty wild-type transformants identified, three had seeds with phenotypes reminiscent of the pkp1 mutant. None of the empty vector control lines displayed altered seed morphology (wrinkledness). Analysis of FAMEs was performed on the wrinkled seeds and the three lines contained 3.92 :t 0.6, 2.73 i 0.3, and 2.22 i 0.2 pg of oil per seed. The recapitulation of the pkp1 seed phenotype using antisense repression independently corroborates that the low oil phenotype can be caused by reduction or abolishment of expression of the PKPI gene encoding PKp-Bl. Rescued lines have subunit-specific restoration of PK activity PK gene expression and enzyme activity in the rescued lines were explored. Figure 3.7A shows RNA gel blots of 9-11 DAF silique tissue probed with PKp gene specific probes. The top panel shows no accumulation of PKp-13. transcript in pkp1 (as demonstrated by RT-PCR in Figure 3.1C) and restoration only in the R131 lines. In the middle panel, very little PKp-132 transcript is present in wild type, pkp1, or the R131 lines. However, the R132 lines clearly have increased amounts of the transcript. Based on gene expression, the R131 and R132 lines appear to have 01131 or (1132, respectively, as the dominant PK in silique tissue and thus provided an opportunity to further test the metabolite regulation observed for the recombinant 01131 and 01132 complexes. 112 h~§3 83:518 55:: gEv- «nu: «>0: gear—eaaat—Nmu‘fimu Exannnnuunnnnnnn ammococotcrrroccrncmrrosrr a... 5‘: $1253.. ”4‘ PKp'131 "I PKp-B2 ~_ Actin1 120 120 100 100 .S c 0 ._ §80 .380 360 5,60 E E 340 540 E E 20 20 O h N on v co ‘— 3 co 0' F .— co l0 co co 0) ksgattaat: ssaeeesa gameeeaeee “maxim“ Figure 3.7. PKp gene expression and enzyme activity in rescued pkp1 lines (A) Expression of PKp encoding genes in siliques of wild type (WT), pkp1, and rescued lines. R131 is overexpression of PKp-13. and R132 denotes overexpression of PKp-132. Actin1 probe was used for loading control. (B) Native PK specific activity of 9-1 1 DAF siliques in response to metabolite effectors. Lefi panel includes wild type (WT), pka , and pkp1 lines rescued with overexpression of PKp-13. (R131). Right panel includes wild type (WT), pkp1, and pka lines rescued with overexpression of PKp-132 (R132). Activity was measured at pH 8.0 with subsaturating substrate concentrations. None, no effectors; glu, 5 mM glutamate; 6-PG, 0.2 mM 6- phosphogluconate. One mU is defined a 1 nmol pyruvate formed per minute. Values are the mean 1 SD (n=4). 113 PK specific activity was first measured at pH 8.0 from 9-11 DAF silique material in the absence of effectors. The left and right panels of Figure 3.7B show that PK specific activity was restored in the R13. and R132 lines. The inclusion of 5 mM glutamate in the assay mixture resulted in a moderate inhibition of native PK specific activity in wild type and all of the rescued lines, as was observed for wild type in Figure 3.48. The PK specific activity from the R132 lines, however, responded differently from wild type and the R131 lines to the presence of 0.2 mM 6-phosphogluconate. PK was activated by 6- phosphogluconate in the R132 lines, which provides needed correlation between the observed metabolite regulation of recombinant and native PK specific activity. Altered substrate/product ratios and accumulation of glycolytic precursors in the pka mutant The effects of the pkp1 mutation on the pools of the substrates (PEP and ADP) and products (Pyr and ATP) of PK in developing seeds were analyzed. Metabolite measurements were made at 5-7 DAF and at 11-13 DAF, when PK specific activity and chlorophyll content, respectively, are most affected in the mutant seeds. Table 3.2 shows the results of these analyses. At 5-7 DAF, the amount of Pyr in pkp1 is decreased, resulting in a 40% reduction in the Pyr/PEP ratio. At 11-13 DAF, there is no difference in the ratio of Pyr/PEP between wild type and pkp1. Instead, in pkp1, the absolute amounts of Pyr and PEP are proportionally increased. The PEP and Pyr data show that PK specific activity positively correlates with the Pyr/PEP ratio, due mostly to effects on the steady state levels of Pyr. The ATP/ADP ratio is increased almost 2 fold in pkp1 when both PK specific activity (5-7 DAF) and chlorophyll content (1 l-13 DAF) are the most reduced. 114 Table 3.2 Metabolite levels in WT and pkp1 developing seeds PEP Pyr ADP ATP DAF nmol/ g FW nmol/g FW Pyr/PEP nmng FW nmol/g FW ATP/ADP WT 5-7 17.5 i 2.5 89.1i 7.9 5.1 15.8 :1: 1.6 53.7 d: 3.7 3.4 11-13 13.3 11.6 36.5 i 2.7 2.7 12.8 i1.2 38.8 i 3.7 3.0 k 1 5-7 18.6 i1.5 58.1 :1: 6.1 3.1 13.0 :1: 1.3 66.4 :t 3.4 5.1 p p 11-13 22.2 i 4.1 59.1 i 10.4 2.7 10.9 d: 2.6 57.3 i 7.0 5.3 Values represent the mean i SD of at least three repeats. DAF, days after flowering; PEP, phosphoenolpyruvate; Pyr, pyruvate; ADP, adenosine diphosphate; ATP, adenosine triphosphate; FW, fresh weight It seems that the metabolic perturbations in pkp1 actually result in increased energy status in the developing seeds. This is somewhat surprising as PK activity and photosynthesis are expected to contribute to the ATP pool. Pyruvate kinase is a control point for glycolysis as its activity has a direct impact on ATP and PEP, the latter of which is an inhibitor of phosphofructokinase in anoxic tissues such as seeds (Plaxton 1996). It is reasonable that a reduction in PK activity would result in an inhibition of glycolytic flux and an accumulation of carbohydrate precursors. Thus, hexose, sucrose, and starch were also measured in developing wild-type and pkp1 seed. Figure 3.8A shows that in wild type hexoses accumulated early during development and steadily decreased, while sucrose followed the opposite trend. The pkp1 mutant seeds followed the same trends, but contained twice as much of both compounds during peak accumulation times. Starch in the wild type showed the same transient accumulation pattern as previously documented (Focks and Benning 1998, Baud et al. 2002). However, the pkp1 mutant continued to store starch throughout embryo development (Figure 3.8A). 115 ‘\ IL Ugmkwc f f 1‘ 1U®®m \ml Pl” ‘ Hexoses _._ WT -- pkp1 g 200. PF W K :pkp1 600 Sucrose 2’ 300 100 N o O mU/mg protein Starch 50 . 9 13 9 Days After Flowering 9 0.4 5 7 9 11 13 15 17 Days After Flowering Figure 3.8. Carbohydrate accumulation and phosphofi'uctokinase enzyme activity in pkp1 and wild—type seeds (A) Hexoses (glucose plus fructose), sucrose, and starch levels in developing seeds. WT, wild type. Values are the mean 1: SD (n=4). (B) ATP-dependent phosphofructokinase (PFK) and pyrophosphate-dependent phosphofructokinase (PFP) activity in developing seeds. One mU is defined a 1 nmol pyruvate formed per minute. WT, wild type. 116 ATP-dependent and PPi-dependent phosphofructokinase (PF K and PF P, respectively) activities were measured in developing seeds at 9 and 13 DAF (Figure 3.8B), when pkp1 has increased hexoses only, or increased sucrose and starch, respectively. PF K activity decreases in the wild type between the two timepoints while PFP activity increases. In the pkp1 mutant, the PF K and PFP activities change as in wild type, but are reduced by 25% and 40%, respectively. Taken together, the data in Figure 3.8 suggest that a reduction in PKp activity impairs the catabolism of carbohydrates in developing seeds, possibly due to a reduction in upstream glycolytic activities. 117 Discussion The (1131 form of PKp is dominant in seeds A reverse genetic approach was taken to directly test the metabolic function of PKp in developing Arabidopsis seeds. Mutants in the PKp-a gene, which should result in complete inactivation of both PKps, could not be isolated. Only one mutant, in the PKp-131 encoding gene, was identified which had no transcript accumulation and a seed phenotype (Figure 3.1, Table 3.1). Methods to separate the native Arabidopsis PK isoforms (isoelectric focusing, zymograms, native-PAGE) were unsuccessful and so PKp activity was assayed at pH 8.0 to minimize the background from cytosolic enzymes. The PK specific activity profile at pH 8.0 in wild type very closely matches the expression of the 01 and 13. subunit encoding genes (Figure 2. 1 B), while in pkp1 the activity is greatly reduced and does not change (Figure 3.4A). In addition, wild-type and pkp1 seed PK activities were insensitive to activation by 6PG (Figure 3.4B). In contrast, seed PKp activity in pkp1 lines that have been rescued with overexpression of the PKp-132 encoding cDNA (R132) are activated in the presence of 6PG (Figure 3.7B) All these results agree with the in vitro enzyme characterization presented in Chapter 2 and corroborate a specific inactivation of the 11131 enzyme in the pkp1 mutant. Furthermore, the pkp1 mutant supports the initial hypothesis that the influx of photosynthate into embryo tissue and the high demand for lipid and amino acid precursors requires high PKp activity. The regulatory properties of the 01131 enzyme mentioned above make it a prime candidate for this role. Taken together with the lack of transcript accumulation for the gene encoding PKp-132 (Figure 2. 1 B), these data indicate that the 0113. enzyme is the major PKp isoform present in developing Arabidopsis seeds. The increase in PK specific activity at pH 7.0 in 118 pkp1 suggests a compensatory mechanism in the mutant (Figure 3.4A). It is possible that more pyruvate is being generated in the cytosol by another isoform of PK. A preliminary flux map of carbon metabolism in B. napus embryos shows that 30% of the pyruvate used for fatty acid synthesis is cytosolic in origin (Schwender et al. 2003, Schwender et al. 2004b, Schwender and Ohlrogge 2002). If the same is true for Arabidopsis, the increase in PK specific activity at pH 7.0 only marginally compensates, as pkp1 still has a 60% reduction in seed oil. Based on gene expression data (Figure 2.18, Figure 3.2), a small amount of the PKp-132 subunit could also be present in developing seed. The presence of this subunit could explain the incomplete loss of PK activity and seed oil in the pkp1 mutant. Seed metabolism is dependent on proper PKp function The pkp1 mutant seeds are smaller and less green than wild-type seeds (Figure 3.3). The reduction in chlorophyll in pkp1 developing seeds could be the result of sugar accumulation (Figure 3.8A) as sugars are known to repress chlorophyll accumulation and photosynthetic gene expression (J ang et al. 1997, J ang and Sheen 1994). It is also possible that a pleiotropic effect of the pkp1 mutation is reduced chlorophyll biosynthesis. Dark treatment has been shown to decrease fatty acid synthesis by 23% in B. napus embryos (Ruuska et al. 2004). However, all light-activated processes are affected by this treatment and the reduced biosynthetic capability was linked to a lack of light induced activation of certain enzymes. In the case of pkp], light still penetrates the seed and is capable of inducing enzyme activities and other processes. Thus, the primary metabolic 119 defect brought on by a reduction in PKp activity is likely the major factor contributing to the low oil phenotype of pkp1 . In both wild-type and pkp1 seeds, fatty acids accumulate in a linear time course from about 5 DAF until at least 15 DAF (Figure 3.5A). The rate of accumulation, though, is reduced by about 60% in the mutant, which correlates with the reduction in seed oil and the reduction in total PK specific activity at pH 8.0. The fatty acid profile of mature pkp1 seeds (Figure 3.5B) is very similar to that of the wriI mutant and plants with altered biotin carboxyl carrier protein gene expression (Focks and Benning 1998, Thelen and Ohlrogge 2002). These mutations impair fatty acid synthesis either by reducing the supply of precursors or by inhibiting the ACCase reaction, respectively. It is likely that in the pkp1 mutant a reduction in precursors for fatty acid synthesis is the cause of the altered fatty acid profile. Indeed, the steady state level of pyruvate is reduced in pkp1 seeds compared to wild type at 5-7 DAF, which correlates with the onset of fatty acid biosynthesis (Table 3.2). It should be noted that these measurements do not distinguish the subcellular compartmentation of the metabolites in question. Interestingly, the amounts of PEP and pyruvate, but not the ratio of the two, are increased in pkp1 at 11-13 DAF (Table 3.2). The idea of a reduction in the supply of precursors for fatty acid biosynthesis is further supported by the accumulation of carbohydrates in the mutant seed, a phenotype similar to that of wriI (Figure 3.8A). It is also possible that the reduction in the rate of fatty acid synthesis is brought on by a lack of ATP. Seeds of B. napus (and likely Arabidopsis) are a low oxygen environment (Vigeolas et al. 2003) and even with photosynthesis PK could have an important role in the production of ATP. A mutant defective in a plastidic ATP/ADP transporter with reduced ATP import capacity into 120 plastids has reduced oil content in its seeds (Reiser et al. 2004). Moreover, this mutant was shown to compensate for reduced ATP import by increasing the expression of the genes encoding PKp-B. and PKp-132. A similar increase in transcript level for any of the PKp subunits is not seen in pkp1 seeds (Figure 3.2). Additionally, steady state levels of ATP in the mutant seeds were actually increased relative to wild type (Table 3.2), possibly as a result of increased cytosolic PK (PKC) activity in the mutant (Figure 3.4A). Elevated PKc activity in pkp1 might be an indicator of enhanced cytosolic glycolytic flux into mitochondrial respiration, which could also elevate the ATP/ADP ratio. Based on the pkp1 mutant phenotype it is unlikely that ATP production is a major function of PKp in developing Arabidopsis seeds. The pkp1 mutant is rescued by ectopic overexpression of the PKp-Bl and PKp-132 encoding cDNAs except that oil accumulation is recovered less fully in the R132 lines (Figure 3.6). A similar pattern of rescue was observed for seed chlorophyll accumulation (Figure 3.3B) and fatty acid profile (Figure 3.6C), in which the R132 lines were not rescued as completely as the R131 lines. The maximum activity observed in seed extracts was similar in all of the rescued lines (Figure 3.7B). Thus, it is likely that distinct regulatory features determined for the two PKp complexes (detailed in Chapter 2) account for the observed physiological differences between the R131 and R132 rescued plants. No transgenics were observed that had an increase in seed oil content or in PK activity, despite higher than wild-type expression levels of the PKp-BI or PKp-132 encoding genes (Fig 3.7A). This might indicate that in these lines the amount of PKp-u is limiting as it is required for activity. Concomitant with the reduction in seed oil in pkp1 is the accumulation of increased amounts of hexoses, sucrose, and starch (Figure 3.8A). It seems that there is a 121 redirection of carbon partitioning in pkp1 in which less hexose and sucrose are catabolized via glycolysis, but are instead incorporated into starch. However, the starch accumulated in pkp1 seeds at 15 DAF only accounts for approximately 20% of the carbon not incorporated into fatty acids (Table 3.2, Figure 3.8A). One potential mechanism for the observed accumulation of carbohydrates is that elevated PEP (Table 3.2), which is a potent inhibitor of plant phosphofructokinases (Plaxton and Podesta 2006), could slow the entry of hexose—phosphates into glycolysis, resulting in a misregulation of metabolism. The reduction of PFK and PFP activities in pkp1 support this hypothesis (Figure 3.8B). It should be noted that the assays for PFK and PF P are set up in such a way as to maximally activate the enzymes, and thus measure activity as it relates to total protein amounts. The observed reductions in activities therefore do not reflect any potential additional regulation of the enzymes caused by altered metabolite levels. The mechanisms that down-regulate these activities in pkp1 are unknown, but one could speculate that sugar signaling is somehow involved. It is interesting that in pkp1 excess carbon is not redirected into storage protein synthesis. Instead, protein levels are slightly decreased (Table 3.1), which could be a result of decreased glycolytic flux and reduced substrate availability. Two plastid-localized enzymes of branched chain amino acid biosynthesis, acetolactate synthase (ALS) and dihydrodipicolinate synthase (DHPS), use pyruvate as a substrate and have Km values of 1 614 mM for ALS and about 1.7 mM for DHPS (Dumer and Boger 1990, Dereppe et al. 1992). The plastidic pyruvate dehydrogenase (ptPDC) complex has a Km for pyruvate of about 300 11M (Camp et al. 1988). Thus, amino acid synthesis could be out competed by ptPDC in situations of limiting pyruvate availability. However, based on these properties alone, one would expect a decrease in 122 protein content greater than that of oil in pkp1. One possible explanation for the observed phenotype is that branched chain amino acids are only a minor component of total seed protein (Figure 3.5D) and restriction of their synthesis has little effect on protein content. Increased cytosolic PK specific activity in pkp1 (PK at pH 7.0, Figure 3.4A) could also help maintain almost wild-type protein levels by feeding into the TCA cycle and increasing the supply of carbon skeletons available for the synthesis of other amino acids. The results of this study are in support of the metabolic model depicted in Figure 1.2 in which PEP metabolized by PKp in the plastid is the main source of pyruvate for fatty acid and amino acid syntheses. The compartmentation of metabolism apparently serves to isolate metabolic pathways such that specific products can be generated from distinct pools of substrates. In the case of seed oil metabolism, pyruvate generated in the plastid is used mainly for fatty acid synthesis and cannot be fully replaced by cytosolic pools. It seems that Arabidopsis seeds are programmed to make oil from plastidic pyruvate and if that pathway is perturbed, as is the case in the pkp1 mutant, some of the carbon (20% in pkp1) is stored in a different form, e. g. starch. As such the pkp1 mutant provides an example for a plant with altered carbon partitioning in developing seeds. 123 References Alonso, J.M., Stepanova, A.N., Leisse, T.J., Kim, C.J., Chen, H.M., Shinn, P., Stevenson, D.K., Zimmerman, J., Barajas, P., Cheuk, R., Gadrinab, C., Heller, C., Jeske, A., Koesema, E., Meyers, C.C., Parker, H., Prednis, L., Ansari, Y., Choy, N., Deen, H., Geralt, M., Hazari, N., Hom, E., Karnes, M., Mulholland, C., Ndubaku, R., Schmidt, 1., Guzman, P., Aguilar-Henonin, L., Schmid, M., Weigel, D., Carter, D.E., Marchand, T., Risseeuw, E., Brogden, D., Zeko, A., Crosby, W.L., Berry, CC, and Ecker, JR. (2003) Genome-wide Insertional mutagenesis of Arabidopsis thaliana. Science 301 :653-657. Baud, S., Boutin, J ., Miquel, M., Lepiniec, L., and Rochat, C. (2002) An integrated overview of seed development in Arabidopsis thaliana ecotype WS. Plant Physiol. Biochem. 40: 1 51-160. Burrell, M.M., Mooney, P.J., Blundy, M., Carter, D., Wilson, F., Green, J., Blundy, K.S., and Aprees, T. (1994) Genetic Manipulation of 6-Phosphofructokinase in Potato- Tubers. Planta 194295-101. Camp, P.J., Miernyk, J.A., Randall, DD. (1988) Some kinetic and regulatory properties of the pea chloroplast pyruvate dehydrogenase complex. Biochim. Biophys. Acta. 993:269-275. Cernac, A. and Benning,C. (2004) WRINKLEDI encodes an AP2/EREB domain protein involved in the control of storage compound biosynthesis in Arabidopsis. Plant J. 40:575- 585. Clough, SJ. and Bent, A.F. (1998) Floral dip: a simplified method for Agrobacterium- mediated transformation of Arabidopsis thaliana. Plant J. 16:735-743. Dereppe, C., Bold, G., Ghisalba, 0., Ebert, E., and Schar, H. (1992) Purification and characterization of dihydrodipicolinate synthase from pea. Plant Physiol. 98:813-822. Ddrmann, P. and Benning, C. (1998) The role of UDP-glucose epimerase in carbohydrate metabolism of Arabidopsis. Plant J. 13:641—652. Durner, J., and Boger, P. (1990) Oligomeric forms of plant acetolactate synthase depend on flavine adenine dinucleotide. Plant Physiol. 93: 1027-103 1. Eastmond, P.J. and Rawsthorne, S. (2000) Coordinate changes in carbon partitioning and plastidial metabolism during the development of oilseed rape embryos. Plant Physiol. 122:767-774. Focks, N. and Benning, C. (1998) wrinkled]: A novel, low-seed-oil mutant of Arabidopsis with a deficiency in the seed-specific regulation of carbohydrate metabolism. Plant Physiol. 118:91-101. 124 Hausler, R.E., Fischer, K.L., and Flugge, U.I. (2000) Determination of low-abundant metabolites in plant extracts by NAD(P)H fluorescence with a microtiter plate reader. Anal. Biochem. 28121-8. Hu, Z.Y. and Plaxton, W.C. (1996) Purification and characterization of cytosolic pyruvate kinase from leaves of the castor oil plant. Arch. Biochem. Biophys. 3332298- 307. Jang, J.C., Leon, P., Zhou, L., and Sheen, J. (1997) Hexokinase as a sugar sensor in higher plants. Plant Cell 9:5-19. Jang, J.C. and Sheen, J. (1994) Sugar sensing in higher-plants. Plant Cell 6:1665-1679. Kang, F. and Rawsthome, S. (1994) Starch and fatty acid biosynthesis in plastids from developing embryos of oil seed rape. Plant J. 6:795-805. Koncz, C. and Schell, J. (1986) The promoter of Tl-Dna gene 5 controls the tissue- specific expression of chimeric genes carried by a novel type of Agrobacterium binary vector. Mol. Gen. Genet. 204:383-396. Kubis, S.E., Pike, M.J., Everett, C.J., Hill, L.M., and Rawsthome, S. (2004) The import of phosphoenolpyruvate by plastids from developing embryos of oilseed rape, Brassica napus (L.), and its potential as a substrate for fatty acid synthesis. J. Exp. Bot. 55:1455-1462. Li, H.M., Culligan, K., Dixon, R.A., and Chory, J. (1995) Cuel - A mesophyll cell- specific positive regulator of light-controlled gene-expression in Arabidopsis. Plant Cell 7:1599-1610. Lichtenthaler, H.K. (1987). Chlorophylls and carotenoids: Pigments of photosynthetic membranes. Methods Enzymol. 148, 350—3 82. Plaxton, W.C. (1996) Organization and regulation of plant glycolysis. Annu. Rev. Plant Physiol. Plant Mol. Biol. 47:185-214. Plaxton, W.C. and Podesta, EB. (2006) The functional organization and control of plant respiration. Crit. Rev. Plant Sci. 25: 159-198. Plaxton, W.C., Smith, C.R., and Knowles, V.L. (2002) Molecular and regulatory properties of leucoplast pyruvate kinase from Brassica napus (rapeseed) suspension cells. Arch. Biochem. Biophys. 400:54-62. Reiser, J ., Linka, N., Lemke, L., Jeblick, W., and Neuhaus, H.E. (2004) Molecular physiological analysis of the two plastidic ATP/ADP transporters from Arabidopsis. Plant Physiol. 136:3524-3536. 125 Ruuska, S.A., Andrews, T.J., Badger, M.R., Price, G.D., and von Caemmerer, S. (2000) The role of chloroplast electron transport and metabolites in modulating Rubisco activity in tobacco. Insights from transgenic plants with reduced amounts of cytochrome b/f complex or glyceraldehyde 3-phosphate dehydrogenase. Plant Physiol. 122: 491-504. Ruuska, S.A., Girke, T., Benning, C., and Ohlrogge, J.B. (2002) Contrapuntal networks of gene expression during Arabidopsis seed filling. Plant Cell 14:1191-1206. Ruuska, S.A., Schwender, J ., and Ohlrogge, J.B. (2004) The capacity of green oilseeds to utilize photosynthesis to drive biosynthetic processes. Plant Physiol. 136:2700-2709. Schmid, M., Davison, T.S., Henz, S.R., Pape, U.J., Demar, M., Vingron, M., Scholkopf, B., Weigel, D., and Lohmann, J .U. (2005) A gene expression map of Arabidopsis thaliana development. Nature Genet. 37 :501-506. Schwender, J ., Goffman, F., Ohlrogge, J .B., and Shachar—Hill, Y. (2004a) Rubisco without the Calvin cycle improves the carbon efficiency of developing green seeds. Nature 432:779-782. Schwender, J., Ohlrogge, J., and Shachar-Hill, Y. (2004b) Understanding flux in plant metabolic networks. Curr. Opin. Plant Biol. 72309-317. Schwender, J. and Ohlrogge, J.B. (2002) Probing in vivo metabolism by stable isotope labeling of storage lipids and proteins in developing Brassica napus embryos. Plant Physiol. 130:347-361. Schwender, J ., Ohlrogge, J.B., and Shachar—Hill, Y. (2003) A flux model of glycolysis and the oxidative pentosephosphate pathway in developing Brassica napus embryos. J. Biol Chem. 278:29442-29453. Shen, W.J. and Forde, B.G. (1989) Efficient transformation of Agrobacterium spp by hi gh-voltage electroporation. Nucleic Acids Res. 17:8385. Smith, C.R., Knowles, V.L., and Plaxton, W.C. (2000) Purification and characterization of cytosolic pyruvate kinase from Brassica napus (rapeseed) suspension cell cultures - Implications for the integration of glycolysis with nitrogen assimilation. Eur. J. Biochem. 267:4477-4485. Thelen, J.J. and Ohlrogge, J .B. (2002) Both antisense and sense expression of biotin carboxyl carrier protein isoform 2 inactivates the plastid acetyl-coenzyme A carboxylase in Arabidopsis thaliana. Plant J. 32:419-431. Verwoerd, T.C., Dekker, B.M.M., and Hoekema, A. (1989) A small-scale procedure for the rapid isolation of plant RNAs. Nucleic Acids Res. 17:2362. Vigeolas, H., van Dongen, J .T., Waldeck, P., Huhn, D., and Geigenberger, P. (2003) Lipid storage metabolism is limited by the prevailing low oxygen concentrations oilseed rape. Plant Physiol. 133:2048-2060. 126 Weber, A.P. (2004) Solute transporters as connecting elements between cytosol and plastid stroma. Curr. Opin. Plant Biol. 7:247-253. White, J.A., Todd, J., Newman, T., Focks, N., Girke, T., de Ilarduya, O.M., Jaworski, J .G., Ohlrogge, J.B., and Benning, C. (2000) A new set of Arabidopsis expressed sequence tags from developing seeds. The metabolic pathway from carbohydrates to seed oil. Plant Physiol. 124:1582-1594. Xu, C.C., Hartel, H., Wada, H., Hagio, M., Yu, B., Eakin, C., and Benning, C. (2002) The pgpl mutant locus of Arabidopsis encodes a phosphatidylglycerolphosphate synthase with impaired activity. Plant Physiol. 129:594-604. 127 Chapter 4 Germination, establishment, and growth of Arabidopsis plants lacking a plastidic pyruvate kinase 128 Abstract Catabolism of storage reserves is essential for seed germination and establishment. An Arabidopsis mutant (pkp1) deficient in plastidic pyruvate kinase (PKp) which is unable to amass storage oil to the same extent as wild type is abnormal in these processes. Germination is delayed in the mutant and seedling establishment is dependent on an exogenous sugar supply. It appears, however, as though these phenotypes are not entirely caused specifically by a lack of seed oil and may be related to reduced PKp activity. Gerrninating seeds of pkp1 are unable to metabolize storage oil and cannot utilize applied sucrose for hypocotyl elongation in the dark. Additionally, seed longevity is greatly reduced in pkp1 indicating a potential lack of seed tocopherols. Mature pkp1 plants are slightly chlorotic and contain less glucose and fructose. Thus, it appears as though PKp is necessary for proper metabolic function during all aspects of plant growth. 129 Introduction Pyruvate kinase (PK) is a ubiquitous enzyme located at a major branch point in carbon metabolism (see Figure 1.5). Native PKs have been purified and characterized fi'om a variety of organisms and these studies have revealed a wide range of kinetic and regulatory properties for these enzymes (Munoz and Ponce 2003). This diversity of molecular and biochemical characteristics seem to define specific functions for individual PK isoforms within a single organism. For instance, mammals have four PKs which are differentially expressed depending on the metabolic demands of the tissue (Yarnada and Noguchi 1999). Plant metabolism, however, is more complex than that of mammals and this is reflected by the fact that the Arabidopsis genome encodes 14 putative PKs which reside in both the cytosol (PKC) and the plastid (PKp; Arabidopsis Genome Initiative 2000). Its location in the metabolic network and sheer redundancy indicate that PK is a critical enzyme for plants, but few studies have focused on this enzyme with regard to its function in vivo. Only one instance of a plant deficient in PK has been reported. An attempt to engineer transgenic tobacco (Nicotiana tabacum) with increased PKp inadvertently resulted in co-suppression of an endogenous PKc-encoding gene and consequent loss of enzyme activity in leaves (Gottlob-McHugh et al. 1992). At the time the only apparent phenotype was a shift from a high pyruvate/phosphoenolpyruvate (Pyr/PEP) ratio to a low one. More detailed analysis led to the discovery of a root growth defect, despite the PKc deficiency being localized exclusively to leaves (Knowles et al. 1998). This phenotype was exacerbated in low light conditions. Later, it was observed that PKc functions in regulating photoassimilate export at night by controlling carbon flow into 130 respiration (Grodzinski et al. 1999). This result is somewhat confusing as reduced PKc activity actually resulted in increased respiratory CO2 release. However, a yeast (Saccharomyces cerevisiae) mutant with a similar reduction in PK activity also has elevated respiratory flux (Pearce et al. 2001). An induction of a PK bypass (as shown in Figure 1.3) could explain this phenomenon. The occurrence of plant PKs in the plastid and cytosol almost certainly reflects the unique roles for the respective enzymes (Plaxton 1996). For example, Pyr transported into mitochondria for entry into respiration is most likely derived from PKc, especially when considering the lack of a reported plastidic Pyr transporter (Weber 2004). Plastid- localized metabolisms which use PEP and Pyr are no doubt influenced by PKp activity. Two enzymes in the Shikimate pathway, 3-deoxy-D-arabino-heptulosonate 7-phosphate synthase, and 5-enolpyruvylshikimate 3-phosphate synthase, use PEP as a substrate (Herrmann and Weaver 1999). The final product of this pathway, chorismate, is metabolized into aromatic amino acids which themselves are the starting points for the synthesis of a variety of secondary metabolites (e. g. anthocyanins). The plastidic pyruvate dehydrogenase complex, which produces acetyl-CoA for fatty acid synthesis acts on pyruvate. The first enzyme of the methylerythritol-4-phosphate (MEP) pathway, l-deoxy-D-xylulose-5-phosphate synthase, uses pyruvate and glyceraldehyde-3- phosphate to synthesize the first intermediate of plastidic isoprenoid synthesis (Lichtenthaler 1999). Plastid-derived isoprenoids include the carotenoids and phytol used in the biosynthesis of chlorophyll and tocopherol. In addition, pyruvate is a substrate in the biosynthesic pathways of valine, lysine, and isoleucine, and can be directly converted to alanine by a transaminase (AraCyc metabolic map, www.arabidopsis.org/tools/aracyc/). 131 Clearly, the ratio of PEP to Pyr must be balanced, in large part by PKp, such that these biochemical pathways function properly. In Chapter 3, I detailed the fatty acid biosynthetic defect (a plastid localized metabolism) of an Arabidopsis mutant (pkp1) deficient in seed PKp activity. Additionally, pkp1 seeds had much less chlorophyll than wild type and this could indicate substrate- restricted flux through the MEP pathway (which produces the phytol side chain of chlorophyll). Here, I performed experiments to determine the extent to which the lack of seed oil and perturbation of plastidic metabolism affects germination, establishment, and grth of pkp1 offspring. 132 Materials and Methods Plant growth conditions Wild-type plants were of the Col-2 ecotype while pkp1 and the respective rescued lines were in the Col-0 background. All seeds were sterilized with 20% bleach, 0.05% TritonX-IOO for 15 min and were rinsed 5 times in sterile water. Medium used for germination and growth on agar plates was full strength MS, pH 5.8, 0.9% agar and included 0, 2, or 4% sucrose when appropriate. Seeds were stratified at 4°C for 3 (1 prior to being put into an incubator with a photon flux density of 60—80 pmol m"2 sec—1 and a light period of 16 h (22°C), and a dark period of 8 h (18°C). After 10 d, seedlings were either transferred to soil or to fresh agar plates. Soil grown plants were put into a 16-h photoperiod with a day temperature of 22°C and a night temperature of 20°C at a photon flux density of 100—120 pmol rn-2 sec—l. Plant growth measurements were on plants transferred to soil at 10 d after sowing. The aerial portion of six individuals was used for each time point. Root and hypocotyl elongation assays For these assays, seeds were sown in a straight line and the agar plates were arranged vertically. Root lengths were measured to the nearest mm every 24 h. The same set of agar plates was used throughout the experiment. Hypocotyl elongation assays were performed using 7 d after sowing seedlings as previously described (Penfield et al. 2004). Once exposed to the light hypocotyls were immediately measured and those plates were not used again. 133 Seed germination assays All seeds used were produced from mother plants grown in identical conditions and were of the same age. Germination assays were routinely performed using the plant growth procedures mentioned above. When appropriate, 0, 2, or 4% sucrose was added to the medium prior to being autoclaved. Germination was scored as radicle emergence from the seed coat and was determined every 24 h using the same set of horizontally grown agar plates. Accelerated aging treatment was done as previously described (Sattler et al. 2004) except that 42°C was used instead of 40°C. Pyruvate kinase enzyme activity measurements Seedling and leaf crude protein extracts were prepared by grinding tissue in approximately 10 volumes (uL/mg) of buffer containing 50 mM Tris-Cl pH 7.5, 5 mM MgCl2, 1 mM EDTA, 1 mM EGTA, lmM DTT, 0.1% Triton-X100, 10% glycerol, and a protease inhibitor mix (Complete mini, Roche). Pyruvate kinase activity was coupled to the conversion of NADH to NAD+ by lactate dehydrogenase. Reactions were kept at 25°C, were started by the addition of enzyme mix, and were linear for at least 5 minutes. Absorbance at 340 nm was measured using a FLUOstar Optima 96-well plate reader (BMG Labtech, Offenburg, Germany). The PKp reaction mixtures contained 50 mM HEPES-KOH pH 8.0, 5% PEG-8000, 50 mM KCI, 15 mM MgCl2, 1 mM DTT, 2 mM PEP, 1 mM ADP, 0.2 mM NADH, and 2 U ml'l desalted rabbit muscle lactate dehydrogenase. PEP phosphatase activity was corrected for by omitting ADP from the reaction. Reactions at pH 7.0 were done using 50 mM MOPS pH 7.0 instead of HEPES. Protein was quantified using Bradford reagent (Sigma). 134 Lipid and carbohydrate analysis Total leaf lipids were extracted from pre-weighed tissue by vigorously shaking for 5 min in 500 11L of methanol/chloroform/formate (22120.1, v/v). Then 250 11L of 1 M KCl, 0.2 M H3PO4 was added and the tubes were vortexed. The phases were separated by centrifugation at 16,000 g for 5 min. The organic phase was loaded quantitatively onto a treated (soaked in 0.15 M (N H4)2SO4 and dried, then heated to 120°C for 2.5 hrs) silica- 60 TLC plate (Baker). The solvent system used was acetone/toluene/water (91 :3027, v/v) and staining was done with iodine and a-naphthol. Lipid composition was determined by fatty acid methyl ester (FAME) analysis as described previously (Focks and Benning 1998). Ten seedlings or two whole leaves were used for each sample. Glucose, fructose, sucrose, and starch were extracted and quantified as previously described (F ocks and Benning 1998). Five 25 (1 old plants were homogenized together and about 50 mg (fresh weight) of leaf tissue was used for each extraction. Soluble sugars were resuspended in 200 1.1L of water and 15 11L was used for each measurement. Insoluble carbohydrate pellets were resuspended in 300 11L of 0.2N KOH and the remaining volumes were adjusted proportionally. Starch assays were done with 15 11L of the final preparation. Leaf pigment quantification Chlorophyll was extracted from leaves and seedlings using 100 volumes (uL/mg) of 80% acetone and was quantified as previously described (Lichtenthaler 1987). Anthocyanins were extracted and quantified using a published protocol (Martin et al. 2002). 135 Results Seed germination and seedling establishment are aberrant in pkp1 Chapter 3 describes the seed-specific phenotypes of wild type and the pkp1 mutant and includes some data for pkp1 rescued by ectopic overexpression of PKp-Bl (RBI-23) and pkp1 rescued by ectopic overexpression of PKp- 132 (R132-3). During routine grth of these plants it was observed that pkp1 seedlings do not establish (defined as development of true leaves and root elongation) when sown directly on soil. In a single repetition, 100% of wild-type seedlings and 0% of pkp1 seedlings established after 12 days when sown on soil. This defect was rescued 100% in RB1-23 but only 80% in R132-3. A common phenotype of low oil mutants such as pkp1 is an inability to establish in the absence of an exogenous sugar source (Cemac et al. 2006, Lu and Hills 2002). Indeed, pkp1 will not establish unless provided with 2% sucrose in the medium (Figure 4.1A) and even then growth is slow in the mutant. Root elongation assays were performed to quantify this defect. As shown in Figure 4.1 B, pkp1 roots do not grow at the same rate as wild type or either of the rescued lines. The results also indicate that the initiation of root growth may be delayed in pkp1 as at 3 days after sowing (DAS) its roots have not elongated past 1 mm. One possible explanation is that germination itself is inhibited in these seedlings. Germination assays were conducted in the light in the presence of 2% sucrose to explore the possibility that pkp1 seeds do not germinate as fast as wild type. Images of seedlings taken at 3 and 6 DAS qualitatively demonstrate an abnormality in pkp1 seedlings (Figure 4.2A). By 3 DAS, wild type seeds have germinated (defined by radicle emergence‘from the seed coat) and the roots are already elongating. 136 > CD 8 DAS WT 10 DAS pkp1 70 - Suc + 8010 - Suc + Suc Root Length (mm) A o 2 3 4 5 6 7 8 9 10 Days After Sowing Figure 4.1. Sucrose dependent establishment and root elongation in pkp] (A) Germination and seedling establishment in the presence (+) or absence (-) of 2% sucrose (Suc). Ten days after sowing (DAS) pkp1 seedlings are at the same developmental stage as 8 DAS wild-type seedlings. (B) Root elongation in the presence of 2% sucrose. Without sucrose pkp1 roots do not elongate (shown in A). Values are the mean :t SD of 25 measurements. In p190], however, only a portion of the seeds display radicle emergence at 3 DAS and not until 6 DAS have all of the seeds germinated. Figure 4.2B shows the germination rates of wild type, pkp1, and the rescued lines determined from independent experiments using seeds of the same age. After 1 day almost 100% of wild type and R131-23 and about 75% of R132-3 seeds have germinated while in pkp1 less than 20% of the radicles have emerged. Not until 5 DAS does pkp1 reach its maximum germination percent, which is about 90%. Pyruvate kinase (PK) activity was measured at pH 8.0 to ascertain if reductions in enzyme activity are correlated with the delayed germination in pkp1. 137 A B 120 3 DAS C 100 --------------- .9 4.. E 80 E F a) : 6 DAS 9 60 ~ C OJ 2 40~ d) +WT 20‘ +Pkp1 “I“R131-23 Days After Sowing 140 120- g 100- o a 80‘ a, . E 60 . , .. D E 40 . +WT -; ........ - +pkp1 201 "'"RB1-23 0 . . . , 0 2 4 6 8 10 Days After Sowing Figure 41.2. Delayed germination and induction of PKp activity in pkp1 (A) Seed germination is delayed in pkp1 in the light on MS medium with 2% sucrose. DAS, days afier sowing. Un-germinated seeds are marked with white arrows. (B) Seed germination rates on medium with 2% sucrose. Values are the mean i SD (n=6). (C) Pyruvate kinase activity measured at pH 8.0 using protein extracts from germinating seeds and seedlings grown in the light. Values are the mean 3: SD (n=3). One mU is 1 nmol pyruvate formed per min. 138 lmbibed and stratified seeds served as the 0 DAS time point and at this time all of the lines have about the same PK specific activity (Figure 4.2C). Twenty-four h later, there is a doubling of PK specific activity in wild type and a similar, although less intense, induction in the rescued lines. By 5 DAS, activity has returned to the starting point where it remains for the rest of the time course. The pattern of PK induction is skewed in pkp1 and does not peak until 3 DAS, after which it steadily declines. There is indeed a correlation between delayed induction of PK specific activity and germination in pkp1. Developing seeds of pkp1 accumulate carbohydrates in the form of sucrose and starch late during embryogenesis (see Figure 3.8). High sugar concentration in the medium has been shown to delay seed germination (Dekkers et al. 2004, Zhou et al. 1998). Therefore, it is reasonable to postulate that the sugar accumulated in pkp1 seeds is partly responsible for the observed germination defect and that the mutant will be more sensitive to exogenous sugar in the medium. To test this hypothesis, germination assays were performed on agar plates containing 0, 2, or 4% sucrose. A very subtle delay in germination is observed for wild type with increasing sugar concentration (Figure 4.3A). In all treatments though, germination of wild-type seeds reached a maximum by 2 DAS. As predicted, pkp1 is more sensitive to sugar in the medium. In the absence of sucrose, the maximum number of pkp1 seeds has germinated by 3 DAS (Figure 4.3B). With 2%, this time is extended to 5 DAS. On 4% sucrose only 50% of the pkp1 seeds have germinated at 5 DAS and not until 10 DAS is the maximum achieved. 139 > w 120 120 WT pkp1 c 100‘ A I I c 100‘ .9 .9 E 89 - .2 89 1 E E 8 60 - 8 60 4 § 40 - § 40 - 8? +01% SUC 5 +0% SUC 20 . +2% Suc ‘L 20 1 +2% Suc +4% Suc +4% Soc 0 . . . . . O - . . . . 0 1 2 3 4 5 6 0 1 2 3 4 5 6 Days After Sowing Days After Sowing Figure 4.3. Inhibition of pkp1 germination by exogenous sucrose (A) Wild type germination rates in the presence of 0, 2, or 4% sucrose. (B) Germination of pkp1 in the presence of increasing concentrations of sucrose. Storage lipid metabolism is defective in pkp1 Hypocotyl elongation assays in the dark are a standard means of examining storage oil metabolism in germinating seeds. When grown in the dark in the presence or absence of sucrose wild type hypocotyls elongate (Figure 4.4A, 4.4B). pkp1 hypocotyls do not elongate in medium without an exogenous carbon source, which is typical of oil deficient mutants such as wrinkled] (wriI ; Cemac et al. 2006). However, even when provided with 2% sucrose pkp1 seedlings do not elongate their hypocotyls; a phenotype which is rescued in lines overexpressing either B-subunit encoding gene (Figure 4.4A, 4.4B). Dark-grown seedlings must generate ATP mainly by glycolysis and respiration, whether fueled by endogenous storage reserves or by uptake of an exogenous carbon source. 140 A B 25 E 20 .5, 2% Suc % 15 C 2 2 § 104 0. >4 .5 5. 00/0 SUC 0 1 + - + - + - + - Suc WT pkp1 RB1-23 R132-3 C D 200 250 2% Suc +WT 0% Suc 175- 1501 E .5 93 1251 2 9 e G- 1001 o. ‘e” E’ 5 75 +wr 5 E 50 +pkp1 E ...... RB _23 25 ,M‘ R132-3 0 . . 9 . - . . . . . 0 2 4 6 8 10 0 2 4 6 8 10 Days After Sowing Days After Sowing Figure 4.4. Hypocotyl elongation and PKp activity in dark-grown seedlings (A) Dark-grown seedlings at 7 days after sowing on MS medium with 0 or 2% sucrose. (B) Quantification of hypocotyl lengths from seedlings in (A). Values are the mean :t SD of 25 measurments. (+), 2% sucrose; (-), 0% sucrose. (C-D) Pyruvate kinase activity measured at pH 8.0 from dark-grown hypocotyls in the presence (C) or absence (D) of 2% sucrose. Values are the mean :t SD (n=3). One mU is 1 nmol pyruvate formed per min. 141 In either case, PK is important for the substrate-level phosphorylative generation of ATP and for the production of respiratory precursors. Thus, a reduction in PK activity could help explain the hypocotyl elongation phenotype of pkp1 . Indeed, PK specific activity is relatively low in pkp1 etiolated seedlings when grown with or without exogenous sucrose (Figure 4.4C, 4.4D). Seedling establishment is largely fueled by seed storage oil breakdown. Figure 4.5A illustrates this process in seedlings grown on agar plates containing 2% sucrose. Very long chain fatty acids (VLCFAs), which in Arabidopsis are specific to seed triacyl glycerol (TAG), were used as markers for storage oil content. In wild type and the rescued lines storage oil begins to be metabolized at 2 DAS and is essentially depleted by 6 DAS. Interestingly, storage lipids are not used by pkp1 seedlings (Figure 4.5A). A 2 B 50- ’g‘ 40- 8 9". g 92 30- " o\° E ‘6 20 +wr 0 E § +pkp1 10 -----R131-23 0 .M‘RB2'3 0 1 2 3 4 6 6 7 0 1 2 3 4 5 6 7 Days After Sowing Days After Sowing Figure 4.5. Fatty acid composition in germinating seeds and seedlings (A) Seed oil-specific very long chain fatty acid (VLCFA; 20:0, 20:1, 21 :0, 21:1) content in seedlings grown on 2% sucrose. Values are the mean :t SD (n=4). (B) mol% of linolenic acid (18:3) in seedlings grown on 2% sucrose. Values are the mean :t SD (n=4). 142 An additional marker for seedling establishment is an increase in the proportion of linolenic acid (18:3) in membrane lipids, which is a major component of thylakoid membranes. By 5 DAS, the proportion of 18:3 to other fatty acids has increased by about 30% in wild type and the rescued lines (Figure 4.5B). In pkp1 the increase is limited to about 5%. It is clear that lipid metabolism is abnormal in pkp1 seedlings. The lipid metabolism phenotype of pkp1 seedlings is reminiscent of Arabidopsis mutants deficient in seed tocopherol, vitamin e] and vitamin e2 (vteI, vte2; Sattler et al. 2004). Seed vitamin E deficiency leads to irreversible lipid peroxidation and thus, reduced seed longevity. Treatment with high temperature and relative humidity simulates natural aging and was used here to determine the extent to which pkp1 resembles the vte mutants. Figure 4.6 shows the results of this experiment. +VVT +pkp1 .5 100‘ '3' ..... I ..... I ..... I ..... I ..... I ..... ... .......... .RB1-23 E A R624, .9 80 . E 1 l m J l 060 .1, ................. E ................. 240- <0 0. 20--’ 0 I "L " 5 5 = = H 01234567891011 Days After Sowing Figure 4.6. Seed germination following accelerated aging treatment Seeds were the same as were used in previous germination assays and were produced from plants grown in identical conditions. Values are the mean :1: SD (n=8). 143 Wild type germination was delayed relative to untreated seeds (compare to Figure 4.2B), but still reached nearly 100%. Germination of pkp1 was almost completely arrested. Surprisingly, R131-23 seeds were unaffected by the treatment while R132-3 showed a phenotype intermediate between wild type and pkp1. Altered growth and leaf metabolism in pkp1 The pkp1 seedlings that successfully establish continue to experience difficulties in growth after being transferred to soil. Biomass production is limited and around the time of flowering initiation (25 DAS) pkp1 aerial parts weigh approximately half compared to wild type (Figure 4.7A). A B 100 1&7 +WT T“; 80 . +pkp1 Er . r: 60‘ r .C O') '6‘: 40 - 3 g 20 - LL 0 . . . . 5 10 15 20 25 30 Days After Sowing WT pkp1 35 Days After Sowing Figure 4.7. Plant growth and morphology of pkp1 (A) Plant growth curve of wild type and pkp1. Measurements were initiated at 10 days after sowing, when seedlings were transferred from MS medium with 2% sucrose to soil. Each value is the mean i SD (n=6). (B) Whole plant morphology of wild type and pkp1. 144 Taking into account the delay in germination, pkp1 follows the same developmental time course as wild type. At 35 DAS, both genotypes are well into flowering and pkp1 is nearly equal to wild type in size (Figure 4.7B). Another noticeable morphology of pkp1 is slight chlorosis. When grown on soil, total chlorophyll content is reduced by 30%, accompanied by a moderate increase in anthocyanins (Table 4.1). This phenotype is exaggerated in plants propagated on agar plates in lower light conditions (photon flux densities of 100—120 and 60-80 umol m”2 560—1 for soil and agar growth, respectively). Table 4.1 Leaf pigments in 25 day old plants grown on soil or on agar plates Soil-grown plants Agar plate-grown plants Chlorophyll Anthocyanins Chlorophyll Anthocyanins (ug/ mg FW) (A530-A657/ g FW) (148/ mg FW) (A530'A657/g FW) WT 1.06 i 0.05 0.69 i 0.01 1.18 i 0.01 0.22 i 0.03 pkp1 0.76 i 0.09 0.80 i: 0.01 0.42 :t 0.02 0.54 :1: 0.02 Values represent the mean :t SD of at least three repeats. FW, fresh weight Pyruvate kinase enzyme assays were performed to see if the morphological differences of pkp1 are correlated with a reduction in activity. When measured at pH 7.0, which is more specific for cytosolic enzymes, enzyme activity is the same between wild type and pkp1 (Figure 4.8A). Plastidic PK’s typically have a pH optimum of 8.0, and so any differences directly related to the pkp1 mutation are expected to be seen at this pH. The data in Figure 4.8B establishes that PKp activity in pkp1 leaves is ablated and only reaches about 60% of wild-type levels. 145 > u: 60 70- pH 7.0 .WT pH 8.0 .WT C 50 upkp1 : 60 upkp1 '5 “ 50 3, 30 S, E E 30 2 20 2 20 10 10 0 , 0 day night day night day night day night 15 DAS 25 DAS 15 DAS 25 DAS Figure 4.8. Pyruvate kinase activity in rosette leaves during the day and night (A) Pyruvate kinase activity measured at pH 7.0 For each time point, 5 whole plants were homogenized and used to prepare protein extracts. Day = 8 hours afier lights on. Night = 8 hours after lights off. Values are the mean i SD (n=4). DAS, days after sowing. (B) Pyruvate kinase activity measured at pH 8.0 using the same protein extracts described in (A). Lipid metabolism is aberrant in pkp1 seeds and seedlings and so it is logical to presume that there is also an effect in leaves of mature plants. However, limited examination of leaf lipids revealed very little qualitative, if any, difference between wild type and pkp1. Figure 4.9A is a thin layer-chromatogram of total leaf lipids extracted from 25 DAS plants. It is clear that there is no defect in the membrane lipid profile of pkp1. Analysis of the fatty acid profile revealed a very subtle, but significant decrease in 1823 content which is accompanied by an increase in the proportion of linoleic acid (18:2; Figure 4.98). This result is not totally unexpected as the same trend is seen in seedlings, although to a much greater extent (Figure 4.5B). 146 A B 70 IWT 60‘ upkp1 MGDG 50- o 40- DGDG i 0 E30. 20- 10- origin 0- _ 169 16fl 162 189 163 18fl 182 183 WT pkp1 Figure 4.9. Lipid profile and composition of pkp1 leaves (A) Thin Iayer-chromatogram of total lipids extracted quantitatively from 25 days after sowing wild type and pkp1 leaves. Plate was stained with iodine vapor and a-naphthol to maximize the number of lipids detected. DGDG, digalactosyldiacyl glycerol; MGDG, monogalactosyldiacyl glycerol. (B) Fatty acid composition of total leaf lipids from 25 days after sowing plants. Values are the mean 3: SD (n=3). Reduced biomass production and chlorophyll content of pkp] could reflect or cause altered carbohydrate metabolism. Therefore, soluble sugars and starch were extracted from leaves during the day and night and were quantified. Hexose content (glucose and fructose only) was reduced by more than half in pkp1 during both the day and night. On the other hand, sucrose and starch accumulation were unaffected. 147 > CD 0 ..s O .b N UT IWT IWT ' IWT upkp1 3.5‘ upm‘l upkp1 A 8 i A 2 ‘ E 3 ‘ E §6~ 825 316 e 3’ 2 . 2’ g 4‘ 8 1.5, § 1 ' 2 5 .9 a: 3 1 1 m I 2 . (I) 0.5‘ 0.51 01 r . 0 1 ~ , 0 . Day Night Day Night Day Night Figure 4.10. Carbohydrate content of 25 day old leaves (A) Hexose (glucose and fructose) content of rosette leaves. 5 whole plants were homogenized and used for extraction. Day night cycle used was 16 hrs light and 8 hrs dark. Day = 8 hours after lights on. Night = 8 hours after lights off. Values are the mean :t SD (n=4). FW, fresh weight. (B) Sucrose content of the samples described in (A). Values are the mean :1: SD (n=4). (C) Starch accumulation in the same samples as detailed in (A). Values are the mean :t SD (n=4). 148 Discussion PKp activity is required to fuel seed germination Seed germination in Arabidopsis is largely driven by the metabolism of storage reserves other than lipids, while seed oil is more important for fueling subsequent seedling establishment (Cemac et al. 2006). Breakdown of carbohydrates to produce energy proceeds through glycolysis, with PK generating half of the ATP and providing precursors for respiration. In germinating wild-type seeds, a rapid increase of PKp activity is observed coincident with radicle emergence (0-1 DAS, Figure 4.2C). In pkp1 there is a delay in this induction, which could explain or result from the impaired germination rate (Figure 4.2A, 4.2B). The eventual increase in PKp activity in pkp1 can be explained by two possibilities: 1) enhanced expression of the PKp-132-encoding gene in response to unfavorable energy status, or 2) induction of a cytosolic enzyme with higher than normal pH optimum. The first hypothesis is supported by elevated expression of the PKp-132- encoding gene in dark-grown seedlings of a plastidic ATP/ADP transporter mutant, supposedly to compensate for reduced ATP import into plastids (Reiser et al. 2004). Induction of a PKc is also not unreasonable. In seeds of pkp1 PKC activity is increased relative to wild type (Figure 3.4). Addition of an exogenous carbon source does not rescue the germination defect of pkp1. In fact, increasing the amount of sucrose actually inhibits germination (Figure 4.3). It was not tested whether this is a result of osmotic potential in the medium. The oil deficient wri] and triacylglycerol] (tagI) mutants have a similar phenotype but in those cases it was likened to heightened sensitivity to osmolarity (Cemac et al. 2006, Lu and Hills 2002). However, pkp1 differs from these mutants in that there is 50% reduction in 149 PKp activity at the time when nearly 100% of wild-type seeds have germinated (Figure 4.2C). Pyruvate kinase activity in germinating wri] seeds is no different from wild type (Cemac et al. 2006). Therefore it is possible that in pkp1 an accumulation of carbohydrates in the seedling brought on by a reduction in glycolytic flux is responsible for increased sensitivity to sugar in the medium. The incomplete rescue of root elongation by sugar application bolsters the idea that pkp1 is less capable of metabolizing sucrose (Figure 4.1). Storage compound utilization is dependent on PKp either in seeds or seedlings Seedling establishment and hypocotyl elongation in the dark are driven by the catabolism of seed storage oil. Arabidopsis mutants defective in 13-oxidation (Footitt et a1. 2002, Germain et al. 2001) fail to establish unless provided with an exogenous carbon source such as sucrose. Similar phenotypes are observed in glyoxylate cycle and gluconeogenesis mutants which are unable to convert storage reserves into carbohydrates (Comah et al. 2004, Eastmond et al. 2000, Penfield et al. 2004, Rylott et al. 2003). The pkp1 mutant has 60% less seed oil than wild type and does not elongate its hypocotyls in the absence of sucrose (Figure 4.4A, 4.4B). In contrast to the B-oxidation, glyoxylate cycle, and gluconeo genesis mutants, exogenous sucrose does not rescue this phenotype. Without sucrose PKp activity is only marginally reduced in pkp] and hypocotyl elongation is likely inhibited due to a lack of storage reserves (Figure 4.4D). However, on 2% sucrose PKp activity is significantly reduced and this may contribute to the apparent inability to utilize the supplied sugar (Figure 4.4C). It could also be that pkp1 cannot 150 synthesize the membrane lipids necessary for expansive growth, which supported by the fact that the mol% of 1823 does not increase in pkp1 germinating seeds (Figure 4.5B). The pkp1 mutant does not efficiently metabolize its seed oil reserves (Figure 4.5A) and this could also cause faulty seedling establishment and hypocotyl elongation. Storage lipid metabolism is not completely restored in the RBI-23 and R132-3 rescued lines and this is correlated with incomplete rescue of PKp activity (Figure 4.2C). Since PKp is not directly involved in any metabolisms associated with storage lipid breakdown another explanation for this phenotype was sought. Accelerated aging treatment followed by germination assays revealed that pkp1 viability is severely compromised over time (Figure 4.6). This result together with the storage lipid breakdown defect is consistent with a reduction in seed tocopherol (Sattler et al. 2004). Tocopherol is synthesized in the plastid and uses phytyldiphosphate (phytyl-PP) as a precursor (Collakova and DellaPenna 2001). Phytyl-PP is synthesized exclusively in the plastid as a downsteam product of the methylerythritol-4-phosphate (MEP) pathway and as a salvage product released during chlorophyll degradation (Ischebeck et al. 2006). Developing seeds of pkp1 expectedly have lower flux through PKp and therefore less Pyr for entry into the MEP pathway. Additionally, seed chlorophyll content is drastically reduced in developing seeds possibly to do a lack of phytol (see Table 3.2). Moreover, the cloroplastos alterado (claI) mutant of Arabidopsis, deficient in 1-deoxy-D-xylulose-5-phosphate synthase which catalyzes the first step of the MEP pathway (using Pyr as one substrate) has retarded germination and reduced tocopherol content at least in leaves (Estevez et al. 2001). The combined data are in agreement with reduced flux through the MEP pathway leading to lower tocopherol content in pkp1 seeds. Most interestingly, RBI-23 appears to be completely 151 resistant to the accelerated aging treatment employed, suggesting an increase in seed tocopherols (Figure 4.6). However, seed PKp activity and chlorophyll content are not higher than wild type in this line (see Chapter 3). An increase in tocopherol could potentially be explained by altered timing of PKp activity in these seeds. Typically, PKp activity peaks just prior the maximum rate of oil biosynthesis. Tocopherol is synthesized later, once all of the oil has been deposited and seeds begin to degrade and recycle chlorophyll (Valentin et al. 2006). Expression of the PKp-Bl-encoding gene in RBI-23 is driven by the constitutive CaMV 358 promoter and it is possible that spurious PKp activity late during development leads to elevated tocopherol biosynthesis. PKp activity influences chlorophyll biosynthesis and sugar accumulation in leaves The effects of the pkp1 mutation are evident in many aspects of whole plant morphology and physiology. The establishment and growth phenotypes of pkp1 could not be rescued by the application of aromatic amino acids, alanine, branched chain amino acids, or with combinations of those (data not shown). Therefore, impaired amino acid biosynthesis brought on by altered PEP and pyruvate metabolism seems not to be responsible for the pkp1 phenotypes. Leaves of pkp1 contain 30 to 60% less chlorophyll than wild type, depending of if they were grown on soil or on agar plates (Table 4.1). As with seeds, reduced PKp activity (Figure 4.8B) in leaves and resultant impairment of isoprenoid biosynthesis could explain this. Again, the cla] mutant supports this hypothesis in that it too has reduced chlorophyll content in leaves (Mandel et al. 1996). Increased anthocyanins in pkp1 (Table 4.1) may also be related to PKp activity as PEP is a substrate for the Shikimic acid pathway which gives rise to precursors for anthocyanin biosynthesis. 152 Unlike seeds, hexose content is reduced in pkp1 leaves (Figure 4.10A) and likely has no role in repression of chlorophyll biosynthesis. Instead, reduced photosynthesis due to a lack of chlorophyll might be the reason for reduced glucose and fructose. Photosynthetic limitations could also explain the dwindled biomass of pkp1 (Figure 4.7A). Sucrose and starch appear to be unaffected in pkp1 suggesting that any problems with carbohydrate metabolism are localized in or around glycolysis (Figure 4.10B, 4.10C). Altered sugar levels would be expected to result in additional pleiotropic effects as 30 to 50% of Arabidopsis genes are at least partially transcriptionally regulated by sugar concentration (Blasing et al. 2005). It is interesting that lipid content and composition is roughly the same in pkp1 as in wild type. It remains to be determined if pkp1 has a lower rate of fatty acid synthesis, as is the casein developing seeds. In conclusion, the pkp1 mutant has little in common with previously described tobacco plants lacking PKG, thus detailing the unique roles of plastid- and cytosolic-localized PKs. Much work remains to be done to fully understand the role of PKp in plant growth and development. 153 References Arabidopsis Genome Initiative (2000) Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature 408:796-815. Blasing, O.E., Gibon, Y., Gunther, M., Hohne, M., Morcuende, R., Osuna, D., Thimm, O., Usadel, B., Scheible, W.R., and Stitt, M. (2005) Sugars and circadian regulation make major contributions'to the global regulation of diurnal gene expression in Arabidopsis. Plant Cell 17:3257-3281. Cernac, A., Andre, C., Hoffmann-Benning, S., and Benning, C. (2006) WRIl is required for seed germination and seedling establishment. Plant Physiol. 141:745-757. Collakova, E. and DellaPenna, D. (2001) Isolation and functional analysis of homogentisate phytyltransferase from Synechocystis sp. PCC 6803 and Arabidopsis. Plant Physiol. 12721113-1124. Cornah, J .E., Germain, V., Ward, J.L., Beale, M.H., and Smith, S.M. (2004) Lipid utilization, gluconeo genesis, and seedling growth in Arabidopsis mutants lacking the glyoxylate cycle enzyme malate synthase. J. Biol. Chem. 279:42916-42923. Dekkers, B.J., Schuurmans, J.A., and Smeekens, S.C. (2004) Glucose delays seed germination in Arabidopsis thaliana. Planta 218:579-588. Eastmond, P.J., Germain, V., Lange, P.R., Bryce, J.H., Smith, S.M., and Graham, LA. (2000) Postgerrninative growth and lipid catabolism in oilseeds lacking the glyoxylate cycle. Proc. Natl. Acad. Sci. USA. 97:5669-5674. Estevez, J .M., Cantero, A., Reindl, A., Reichler, S., and Leon, P. (2001) l-Deoxy-D- xylulose-S-phosphate synthase, a limiting enzyme for plastidic isoprenoid biosynthesis in plants. J. Biol. Chem. 276:22901-22909. Focks, N. and Benning, C. (1998) wrinkled] 2 A novel, low-seed-oil mutant of Arabidopsis with a deficiency in the seed-specific regulation of carbohydrate metabolism. Plant Physiol. 118:91-101. Footitt, S., Slocombe, S.P., Larner, V., Kurup, S., Wu, Y., Larson, T., Graham, 1., Baker, A., and Holdsworth, M. (2002) Control of germination and lipid mobilization by COMATOSE, the Arabidopsis homologue of human ALDP. EMBO J. 21:2912-2922. Germain, V., Rylott, E.L., Larson, T.R., Sherson, S.M., Bechtold, N., Carde, J.P., Bryce, J.H., Graham, I.A., and Smith, S.M. (2001) Requirement for 3-ketoacyl-COA thiolase-2 in peroxisome development, fatty acid beta-oxidation and breakdown of triacylglycerol in lipid bodies of Arabidopsis seedlings. Plant J. 2821-12. 154 Gottlob-McHugh, S.G., Sangwan, R.S., Blakeley, S.D., Vanlerberghe, G.C., Ko, K., Turpin, D.H., Plaxton, W.C., Miki, B.L., and Dennis, D.T. (1992) Normal growth of transgenic tobacco plants in the absence of cytosolic pyruvate kinase. Plant Physiol. 100:820-825. Grodzinski, B., Jiao, J., Knowles, V.L., and Plaxton, W.C. (1999) Photosynthesis and carbon partitioning in transgenic tobacco plants deficient in leaf cytosolic pyruvate kinase. Plant Physiol. 120:887-896. Herrmann, KM. and Weaver L.M. (1999) The Shikimate pathway. Annu. Rev. Plant Physiol. Plant Mol. Biol. 50:473-503. Ischebeck, T., Zbierzak, A.M., Kanwischer, M., and Dormann, P. (2006) A salvage pathway for phytol metabolism in Arabidopsis. J. Biol. Chem. 281:2470-2477. Knowles, V.L., Mchugh, S.C., Hu, Z., Dennis, D.T., Miki, B.L., and Plaxton, W.C. (1998) Altered growth of transgenic tobacco lacking leaf cytosolic pyruvate kinase. Plant Physiol. 116245-51. Lichtenthaler, H.K. (1987) Chlorophylls and carotenoids: Pigments of photosynthetic membranes. Methods Enzymol. 148, 350—382. Lichtenthaler, H.K. (1999) The l-deoxy-D-xylulose-5-phosphate pathway of isoprenoid biosynthesis in plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 50247-65. Lu, C. and Hills, M.J. (2002) Arabidopsis mutants deficient in diacylglycerol acyltransferase display increased sensitivity to abscisic acid, sugars, and osmotic stress during germination and seedling development. Plant Physiol. 129:1352-1358. Mandel, M.A., F eldmann, K.A., Herrera-Estrella, L., Rocha-Sosa, M., and Leon, P. (1996) CLAl, a novel gene required for chloroplast development, is highly conserved in evolution. Plant J. 92649-658. Martin, T., Oswald, O., and Graham, LA. (2002) Arabidopsis seedling growth, storage lipid mobilization, and photosynthetic gene expression are regulated by carbonznitrogen availability. Plant Physiol. 1282472-481. Munoz, M.E. and Ponce, E. (2003) Pyruvate kinase: current status of regulatory and functional properties. Comp Biochem. Physiol. B Biochem. Mol. Biol. 135:197-218. Pearce, A.K., Crimmins, K., Groussac, E., Hewlins, M.J., Dickinson, J.R., Francois, J., Booth, LR, and Brown, A.J. (2001) Pyruvate kinase (Pykl) levels influence both the rate and direction of carbon flux in yeast under fermentative conditions. Microbiology 1472391-401. 155 Penfield, S., Rylott, E.L., Gilday, A.D., Graham, S., Larson, T.R., and Graham, LA. (2004) Reserve mobilization in the Arabidopsis endosperm fuels hypocotyl elongation in the dark, is independent of abscisic acid, and requires PHOSPHOENOLPYRUVATE CARBOXYKINASEI. Plant Cell 16:2705-2718. Plaxton, W.C. (1996) Organization and regulation of plant glycolysis. Annu. Rev. Plant Physiol. Plant Mol. Biol. 472185-214. Reiser, J., Linka, N., Lemke, L., Jeblick, W., and Neuhaus, H.E. (2004) Molecular physiological analysis of the two plastidic ATP/ADP transporters from Arabidopsis. Plant Physiol. 136:3524-3536. Rylott, E.L., Gilday, A.D., and Graham, LA. (2003) The gluconeogenic enzyme phosphoenolpyruvate carboxykinase in Arabidopsis is essential for seedling establishment. Plant Physiol. 131:1834-1842. Sattler, S.E., Gilliland, L.U., Magallanes-Lundback, M., Pollard, M., and DellaPenna, D. (2004) Vitamin E is essential for seed longevity and for preventing lipid peroxidation during germination. Plant Cell 16: 1419-1432. Valentin, H.E., Lincoln, K., Moshiri, F., Jensen, P.K., Qi, Q., Venkatesh, T.V., Karunanandaa, B., Baszis, S.R., Norris, S.R., Savidge, B., Gruys, K.J., and Last, R.L. (2006) The Arabidopsis vitamin E pathway geneS-l mutant reveals a critical role for phytol kinase in seed tocopherol biosynthesis. Plant Cell 18:212-224. Weber, A.P. (2004) Solute transporters as connecting elements between cytosol and plastid stroma. Curr. Opin. Plant Biol. 7 2247-253. Yamada, K. and Noguchi, T. (1999) Nutrient and hormonal regulation of pyruvate kinase gene expression. Biochem. J. 337 ( Pt 1):1-11. Zhou, L., Jang, J.C., Jones, T.L., and Sheen, J. (1998) Glucose and ethylene signal transduction crosstalk revealed by an Arabidopsis glucose-insensitive mutant. Proc. Natl. Acad. Sci. USA. 95:10294-10299. 156 Chapter 5 Conclusions and perspectives 157 Major Conclusions This multifaceted study of Arabidopsis PKp has confirmed its central location in the seed metabolic network (see Figure 1.2). Of two PKp complexes, the one that is most active and least sensitive to feedback inhibition has been naturally selected for in seeds. Based on analysis of the pkp1 mutant, one can conclude that at least 60% of the fatty acids in Arabidopsis seed oil are ultimately derived from a single PKp (specifically 0113;). Carbohydrates accumulated in place of 20% of the oil in pkp1 seeds. Therefore, PKp represents a regulatory step that controls carbon partitioning in developing embryos. Connectivity between the cytosol and plastid provided by metabolite transporters was insufficient for increased PKC activity to fully compensate for the loss of PKp. This result implies that pyruvate is not a major metabolite that is transported across the plastid envelope in seed tissue. Remaining questions and directions for future research will be discussed in the forthcoming sections. Some additional questions The now bonafide PKps described here account for only a fraction of the annotated PK- encoding genes in Arabidopsis. Phylogenetic analysis suggested the presence of one other plastid targeted PK (At3g49160; Figure 2. 1 A), but pilot experiments with recombinant protein were unable to detect any catalytic activity. The question remains whether the enzymes used in this study are the only PKps in Arabidopsis. If so, the presence of 10 genes encoding PKcs implies either great redundancy or the ability to very precisely regulate cytosolic glycolysis. In any case, the lack of redundancy for PKp is surprising compared to‘ the cytosolic enzymes. 158 The pkp1 mutant still accumulates up to 40% wild-type seed oil and the source of carbon precursors for this residual TAG is unknown. Basal expression of PKp-flg is one explanation. The contribution of the 01132 complex to seed oil biosynthesis could be tested by crossing pkp1 with a null mutantpof PKp-flz. It could be that metabolic redundancy is provided by PKc, as evidenced by induction of its activity in pkp1 seeds (Figure 3.4A). It is also possible that the PK bypasses discussed in Chapter 1 (Figure 1.3) compensate for the reduction in PKp activity. Preliminary experiments detected no changes in the activities of PEP carboxylase, NAD-malate dehydrogenase, or NADP-malic enzyme in pkp1, but more accurate measurements are needed to confirm this. There is also the question of to what extent PKp-132 can function in place of PKp-131. The kinetic parameters of the 01132 complex (Table 2.2) and the inability of PKp-flz overexpression to fully restore oil content in rescued pkp1 lines (Figure 3.6B) suggests that PKp-132 is not able to functionally replace PKp-Bl. However, other aspects of plant growth and development seem to be restored in the 35S2PKp- ,82-rescued pkp1 lines. It seems that only in situations of very high rates of fatty acid synthesis is it advantageous to have 01131 as the dominant PKp. In this work, enzyme assays at pH 8.0 were used to enrich detection of PKp activity. This method was useful in sensing total activity, but was not able to pinpoint the loss of a specific PKp isoform in pkp1. Procedures used to separate and detect individual PKs were unsuccessful and so loss of a specific isoform had to be assumed. Perfection of a zymogram technique as was used for glucose-6-phosphate dehydrogenases would be useful here (see Appendix A). Analysis of the reconstituted enzymes also raised some questions that deserve attention. For instance, it was not tested whether a single complex 159 can contain both 131- and 132-subunits. A mixed-composition enzyme might have regulatory properties distinct from the 01131 and 01132 enzymes, as is the case with hybrids of mammalian PKs (Hubbard and Cardenas 1975). The association of the individual PK subunits is likely a dynamic process and the factors that control the strength of protein- protein interaction were not analyzed in much detail. It was observed that K+ and PEP were required for co-immunoprecipitation of native PKp complexes, suggesting that these are important for subunit association. However, the effects of pH and metabolite effectors, for example, were not examined. It could be that regulation of enzyme activity by certain metabolites is mediated by abrogation of subunit interaction. Several phenotypes of the plcp] mutant require additional examination. The effects of reduced fatty acid synthesis are no doubt extended beyond TAG accumulation. Other lipids such as those making up the hydrophobic layer on the seed coat might also be affected. Preliminary staining of mature seeds suggested increased permeability in pkp1, thus warranting investigation of lipids other than TAG. The reduction in chlorophyll content in pkp1 is also intriguing. Reduced chlorophyll in seeds is correlated with less extensive thylakoid membranes (Figure 3.3). It would be interesting to know if the situation in leaf chloroplasts is the same. Furthermore, the cause of reduced chlorophyll is still unknown. Is it a chlorophyll biosynthetic defect or is it the result of a lack of thylakoid membranes brought on by impaired fatty acid synthesis? Either is possible. Chlorophyll biosynthesis is likely not inhibited by sugar concentration in pkp1 leaves, since hexoses are actually decreased relative to wild type (Figure 4.10). Several lines of evidence also hint at a lack of seed tocopherols in pkp1, but measurements have not yet been'made. Both chlorophyll and tocopherol biosynthetic defects could arise from 160 reduced phytol synthesis, and so it would also be informative to measure flux through the methylerythritol-4-phosphate (MEP) pathway in plastids. Future Directions In addition to answering the immediate questions listed above, future research on PKp should be focused on systems biology and metabolic engineering. The biochemical properties (S0, 5, Vmax, 150, K.) obtained for the PKp complexes could be incorporated into a kinetic model of seed metabolism. The details of many other enzymes in the network would be needed and this dissertation provides a framework for such characterizations. Once made, a kinetic model could be used to predict the effects of specific metabolic perturbations or enhancements. For example, the effects of having 0113. versus 01132 as the dominant seed PKp could be predicted and then tested using pkp1 and the respective rescued lines. A similar approach has been taken to direct the engineering of glycine betaine metabolism in tobacco (McNeil et al. 2001). Simple metabolic profiling of pkp1 seeds and seedlings would also be informative. It is expected that the defect in central carbon metabolism results in pleiotropic effects that would be readily elucidated with this method. Developing Arabidopsis seeds have temporally distinct metabolic fingerprints (Fait et al. 2006) and it would be interesting to see if PKp controls the balance of metabolites beyond what was presented in previous chapters. Metabolic flux analysis using stable isotopomers would also be informative for pkp1 . This method could help determine any rerouting of metabolic fluxes that are not evident based on enzyme activity alone. For example, alternative routes of pyruvate production may be more active in pkp1, despite there being no changes in extractable enzyme activities. In general, a systems 161 biology approach to investigating metabolism in pkp1 would be the most efficient for answering the obvious questions while at the same time discovering unexpected perturbations. Increasing oil yield in crop plants is the main driving force for this research. As PKp is a regulatory enzyme in this process, the next logical step is to increase PKp activity in developing seeds. This was not possible by overexpression of either 13-subunit- encoding cDNA, probably because the amount of 01 was limiting. Therefore, simultaneous overexpression of both 01- and 13-genes should be pursued. This can be achieved using stacked expression constructs with seed-specific promoters. Increasing PKp activity in non-seed tissues would also be a worthwhile endeavor. Ectopic oil production is emerging as a means to greatly increase the yield of a single plant, as seeds have evolved to be at or near their maximum potential. Use of tissue specific or inducible promoters is one means of achieving this. Another way is to take advantage of endogenous regulatory networks. The WRINKLEDI (WRll) transcription factor induces expression of at least PK p-a and PKp-fl ,, possibly by binding directly to their promoters. Thus, overexpression of WRII in the desired tissues could be used to increase PKp activity, with the advantage that WRII also regulates downstream components of fatty acid and TAG biosynthesis (Ruuska et al. 2002). A similar approach was taken to stimulate nitrogen assimilation in maize (Zea mays). Ectopic production of the ZmDofl transcription factor (normally involved in light response) was used to activate anaplerotic carbon metabolism and resulted in a 30% improvement in nitrogen assimilation (Yanagisawa et al. 2004). Increased PKp activity could also be used for the production of useful secondary metabolites. However, the range of metabolic influence of PKp is not 162 fully known and there is (weak) precedent only for the engineering of plastid localized isoprenoid metabolism. Another approach to modulating PKp activity is to engineer the enzyme itself. Targeting of specific residues for increased activity or altered regulation would be greatly facilitated by a crystal structure for the PKp in question. The crystal structure of a mammalian PK has been determined (Muirhead et al. 1986) and it was used to direct mutagenesis for the conversion of a non-allosteric PK into an allosteric enzyme (Ikeda et al. 1997). Comparison of (preferably) crystal structures or the primary sequences of 01131 and 01132 to each other and to additional plant PKs could indicate which amino acids are good targets for modification. A non-targeted approach such as directed evolution using error-prone PCR and DNA shuffling could also yield similarly modified enzymes. Error prone PCR introduces random mutations and resultant protein libraries must then be screened for the desired qualities. In DNA shuffling, PCR is used to randomly combine domains from different genes to generate a library of chimeras which must then be screened. These procedures have been very successful for improving the biodegradation pathways of microorganisms (Parales and Ditty 2005). Pyruvate kinases have distinct domains for substrate and effector binding (Munoz and Ponce 2003) and combination of these from diverse enzymes could prove fi'uitful. A non-mutagenic approach to altering PKp activity would be to mix whole subunits from various organisms. In this work it was determined that the presence of either 13-subunit influences the biochemical properties of the respective PKp complexes. It is therefore possible that hybrid enzymes with novel kinetics and regulation could be constructed by taking advantage of natural variation. This method has been successful for the assembly of a hybrid ADP-glucose 163 pyrophosphorylase out of Arabidopsis and potato subunits (Ventriglia et al. 2007). With so many options for the engineering of PKp enzymes, one should carefully weigh the benefits and pitfalls of each. Targeted mutagenesis involves the least amount of work, but would require inferences to be drawn from crystal structures which have not yet been determined for any plant PK. Directed evolution has a high potential for success, but the outcomes are less predictable and are entirely dependent on the screening method used to analyze the mutant proteins (which should number in the thousands). Combining subunits from various organisms takes advantage of existing enzyme diversity, but one must obtain the purified subunits to perform any experiments. In conclusion, I feel that future research on Arabidopsis PKp should be from multiple directions and should involve collaboration with experts in various fields. Eventually, what is learned needs to be applied to a real crop plant so that humanity can benefit from their investment in science. 164 References Fait, A., Angelovici, R., Less, H., Chad, 1., Urbanczyk-Wochniak, E., Fernie, A.R., and Galili, G. (2006) Arabidopsis seed development and germination is associated with temporally distinct metabolic switches. Plant Physiol. 142:839-854. Hubbard, D.R. and Cardenas, J.M. (1975) Kinetic properties of pyruvate kinase hybrids formed with native type L and inactivated type M subunits. J. Biol. Chem. 250:4931-4936. Ikeda, Y., Tanaka, T., and Noguchi, T. (1997) Conversion of non-allosteric pyruvate kinase isozyme into an allosteric enzyme by a single amino acid substitution. J. Biol. Chem. 272:20495-20501. McNeil, S.D., Nuccio, M.L., Ziemak, M.J., and Hanson, AD. (2001) Enhanced synthesis of choline and glycine betaine in transgenic tobacco plants that overexpress phosphoethanolamine N-methyltransferase. Proc. Natl. Acad. Sci. USA. 98210001- 10005. Muirhead, H., Clayden, D.A., Barford, D., Lorimer, C.G., FothergilI-Gilmore, L.A., Schiltz, E., and Schnritt, W. (1986) The structure of cat muscle pyruvate kinase. EMBO J. 52475-481. Munoz, M.E. and Ponce, E. (2003) Pyruvate kinase: current status of regulatory and functional properties. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 135:197-218. Parales, RE. and Ditty, J .L. (2005) Laboratory evolution of catabolic enzymes and pathways. Curr. Opin. Biotechnol. 16:315-325. Ruuska, S.A., Girke, T., Benning, C., and Ohlrogge, J.B. (2002) Contrapuntal networks of gene expression during Arabidopsis seed filling. Plant Cell 14:1191-1206. Ventriglia, T., Ballicora, M.A., Crevillen, P., Preiss, J., and Romero, J .M. (2007) Regulatory properties of potato-Arabidopsis hybrid ADP-glucose pyrophosphorylase. Plant Cell Physiol. 48:875-880. Yanagisawa, S., Akiyama, A., Kisaka, H., Uchinriya, H., and Miwa, T. (2004) Metabolic engineering with Dofl transcription factor in plants: Improved nitrogen assimilation and growth under low-nitrogen conditions. Proc. Natl. Acad. Sci. USA. 10127833-7838. 165 Appendix A Analysis of glucose-6-phosphate dehydrogenase activity in Arabidopsis T-DNA insertion mutants3 3 This work was done in collaboration with Dr. Setsuko Wakao and is being prepared for publication in: Wakao, S., Andre, C., and Benning C. (2007) Functional analyses of cytosolic G6PDHs and their contribution to seed oil accumulation in Arabidopsis. Plant Physiol. I contributed Figures A. ID, A.2C, and all of A.3. 166 Introduction Glucose-6-phosphate dehydrogenase (G6PDH) is one of the two NADPH generating enzymes of the oxidative pentose phosphate pathway (OPPP). All eukaryotic G6PDHs studied are feed-back inhibited by NADPH, and together with the fact that this enzyme catalyzes a committed step makes G6PDH the regulatory enzyme of the OPPP. In addition to the feed-back inhibition by NADPH, the plastidic isoforms of G6PDH in plants and algae are subject to regulation by the thioredoxin/ferredoxin system (Graeve et al. 1994, Lendzian 1980, Scheibe and Anderson 1981, Wenderoth et al. 1997, Wendt et al. 2000). Hence, they are presumed to act as cellular redox sensors and are inactivated to prevent unnecessary oxidation of carbon when photosynthesis is sufficient for NADPH generation. In plants G6PDH has been frequently described in connection to its involvement in nitrogen assimilation. The induction of its activity or transcript has been described in various systems including pea roots (Bowsher et al. 1992), barley roots (Wright et al. 1997), maize roots (Redinbaugh and Campbell 1998), tobacco roots and leaves (Debnam et al. 2004) and Arabidopsis (Wang et al. 2003). G6PDH activity is necessary to supply the reducing equivalents required for nitrogen assimilation in root cells that lack photosynthesis (Bowsher et al. 1992, Esposito et al. 2001, Esposito et al. 2003, J in et al. 1998, Wright et al. 1997). Aside from nitrogen assimilation, G6PDH has been hypothesized to be an important source of NADPH in other non-photosynthetic tissues (Emes and Neuhaus 1997) and in those that synthesize large quantities of fatty acids (for incorporation into triacylglycerol), such as pollen (N iewiadomski et al. 2005) and oil seeds (Eastrnond and Rawsthome 1998). Green oil seeds such as those of canola 167 (Brassica napus) contain plastids similar to those of shade-adapted leaves (Asokanthan et al. 1997) and are capable of photosynthetic NADPH production, however, only 20-30% of ambient light penetrates the silique walls and reaches the embryo (Eastmond et al. 1996, King et al. 1998). Reported O2 evolution rates from canola embryos were used to calculate that all of the NADPH required for fatty acid synthesis could be provided by photosynthesis (Ruuska et al. 2004). However, in controlled experiments only a 25% decrease in fatty acid synthesis is observed in the dark. Thus, 75% of the required NADPH could be generated by other reactions such as those of the OPPP. In vi v0 stable isotope labeling revealed that a maximum 25% to 45% of the reductant required for oil biosynthesis could come from the OPPP (Schwender et al. 2003). A recent report of G6PDHs from Arabidopsis reported the isolation of T-DNA insertion mutants for both cytosolic isoforms (Wakao and Benning 2005). Here, I examined the effects of these mutations on total G6PDH activity in relation to seed oil biosynthesis. 168 Materials and Methods Plant growth conditions and transformation All seeds were surface sterilized by incubating in 20% bleach, 0.05% Triton-X. The tubes containing the seeds were inverted for 15 min and washed three times with water. The seeds were suspended in 0.1% agar and plated onto MS medium (pH 5.8) (Murashige and Skoog 1962) with 1% sucrose, 0.9% agar and were transferred to soil after 9 days. Wild- type and mutant Arabidopsis plants were prepared for transformation as previously described (Cemac and Benning 2004). Transient expression of G6PDH::GF P for subcellular localization analysis The coding region of G6PD5 and G6PD6 were amplified with the following primers; for G6PD5, (+) 5’-GGACTAGTATGGGTI‘CTGGTCAATGGCA, (-) 5’ GGACTAGTCAATGTAGGAGGGATCCAAA, and for G6PD6, (+) 5’- GGACTAGTATGGGATCTGGTCAATGGCA, (-) 5’- GGACTAGTTAGTGTAGGAGGGATCCAG. The cDNAs were cloned into the Spel site of pCAMBIAl302 (Genbank accession no. AF 234298). Onion epidermal peels were bombarded following the methods previously described (Varagona et al. 1992) using 1100 psi (pounds per square inch) rupture discs at ~4 cm distance using a biolistic gene delivery system (Dupont). For each construct, three peels were bombarded and incubated overnight at 22 C in the dark. The peels were observed with a Leica DMR A2 microscope in the fluorescence mode with the L5 filter cube (Leica Microsystems, Wetzlar, Germany). 169 Construction of complementation vectors for G6PD5 and G6PD6 The T—DNA insertion lines were transformed with BAC clones containing the regions of G6PD5 and G6PD6 that were isolated from a genomic library in a cosmid vector (pBIC20) (Meyer et al. 1994). The 3’-UTR of the respective genes was used as a probe. The T-DNA insertion lines for both GéPDS and G6PD6 have lost their kanamycin resistance, and thus the transformants were selected by kanamycin resistance introduced by the cosmid vector. G6PDH activity assay Protein extraction from various tissues and electrophoresis on cellulose acetate plates (zymogram) were performed as described previously (Wakao and Benning 2005). Roughly 8 volumes (v/w) of extaction buffer were used per sample for homogenization. Liquid assay of G6PDH was performed as described previously (Wakao and Benning 2005). For identification of the middle band, intact plastids were isolated from 5 week old wild-type Arabidopsis plants grown on soil. Five to 10 g of tissue was used per isolation. Intact plastids were isolated from homogenized tissue using a discontinuous Percoll gradient as previously described (Xu et al. 2002). 170 Results and Discussion To test that G6PD5 and G6PD6 indeed encode cytosolic isoforms the subcellular localization of the proteins was examined by transient expression of the respective cDNAs fused to a green fluorescence protein (GFP) gene. For both constructs containing G6PD5 and G6PD6, the green fluorescence was observed dispersed in the cytosol and surrounding what is presumably the nucleus (Figure A. l A, Al .8). The same patterns were observed in multiple experiments. A similar pattern was observed in cells expressing GFP alone, which localizes to the cytosol and to the nucleus (Figure A. 1C). This result together with the lack of a potential targeting sequence in the proteins suggests that G6PD5 and G6PD6 are both likely to be cytosolic proteins. There are three major active G6PDH isoforms in vivo, G6PD5, G6PD6 and an unidentified isoform that is ubiquitous (Wakao and Benning 2005). To help rule out the possibility that the third isoform is localized in the cytosol, we examined G6PDH activity in isolated chloroplasts from Arabidopsis leaves. As shown in Figure A.1D, isolated chloroplasts contain a single band on a zymogram with similar mobility as the unidentified band detected in protein extract from buds. This result suggests that the unidentified ubiquitous isoform is localized in the plastid and that G6PD5 and G6PD6 are the only cytosolic isoforms with major activity in Arabidopsis. To specifically examine the in vivo roles of the cytosolic G6PDHs T-DNA insertion lines for the two genes were obtained from the SALK institute. Their insertion sites were identified using PCR as previously described (Figure A28 and Wakao and Benning 2005). 171 Figure A.l. Cytosolic localization of G6PD5 and G6PD6 (A-C) Onion cells were bombarded with either with pCAMBIA1302 inserted with (A) G6PD5 or (B) G6PD6 coding sequence or (C) the vector alone. (D) Zymogram with isolated chloroplasts shows enrichment in the G6PDH band that is neither G6PD5 nor G6PD6. The arrow indicates origin and direction of electrophoresis. Lm, standard G6PDH from Leuconostoc mesenteroides. The single mutants did not have any obvious morphological phenotypes (Figure A.2A). To address whether this was because of the redundant functions of the two G6PDH isoforms, crosses between the two lines were performed to generate plants homozygous for both T-DNA insertions. GéPDH activity was examined in plants of different 172 genotypes using zymograms of bud protein extracts (Figure A.2C). In each single mutant, g6pd5 and g6pd6, a band is lost from the zymogram. In the double mutant, both G6PD5 and G6PD6 were lost. Interestingly, activity from the plastidic isoform is reduced in all of the mutants. Residual activity of G6PD5 was observed in the double mutant, consistent with previous observations indicating g6pd5 is not a null mutant (W akao and Benning 2005). In the g6pd6 mutant, the uppermost band on the zymogram (G6PD5) is intensified relative to wild type. It is possible that an alternative G6PDH is induced in this mutant. However, in the double mutant this induction is lost indicating that the upper band in the g6pd6 zymogram is indeed G6PD5. Surprisingly the double mutant was also indistinguishable from the WT plant, despite the loss of most of the cytosolic G6PDH activity as observed in zymograms. Therefore we conclude that a nearly complete loss of cytosolic G6PDH activity does not result in severe morphological phenotypes of the plant under normal conditions. Attempts to complement the single and double mutants with a cDNA or G6PDH fused to GFP at the N-terminal have been unsuccessful (data not shown). Only when genomic fragments containing G6PD5 or G6PD6 were introduced did we observe the recovery of the lost bands on zymograms (Figure A.2C). This result together with the zymogram pattern of the double mutant proves the previously defined zymogram bands (Wakao and Benning 2005) were indeed coded by G6PD5 and G6PD6. Seed triacyl glycerol (TAG) was quantified to determine the effects, if any, of loss of one or both cytosolic G6PDHs. The single mutants had no changes in seed oil content. The double mutant had a small, albeit significant, increase in TAG (data not shown) which was associated with an increase in seed mass and not specifically oil accumulation. 173 C 96 Lo Lm g6 rescue “WT 6pd6 g6pd5 Double b 500 bp AtGGPDS SALK_045083 -p p+ p- Lb AtG6PD6 SALK_016157 Y" "pt p- B Figure A.2. Single and double mutants for G6PD5 and G6PD6 (A) Morphological phenotypes of the single and double mutants. (B) Gene structure of G6PD5, G6PD6 and T-DNA insertion sites. Primers used for genotyping PCR were designed as shown. (C) G6PDH zymograms of single and double mutants and lines rescued by Cosmid complementation. The arrow indicates origin and direction of electrophoresis. Lm, standard G6PDH fi'om Leuconostoc mesenteroides. 174 To better understand the increase in seed mass for the double mutant the G6PDH isoforms present in developing seeds were examined (Figure A.3A). In wild type, the dominant isoforms are G6PD6 and the plastidic one. As expected, the double mutant has no activity from any isoforms. The g6pd5 mutant is similar to wild type, except for a reduction in the plastidic isoform. The g6pd6 mutant, on the other hand, has lost the bands present in wild type, but shows a reciprocal increase in G6PD5 activity. This reciprocal induction could explain why only the double mutant and not g6pd6 has a seed mass phenotype. Liquid enzyme assays were performed with the same extracts to quantify any changes in total G6PDH activity (Figure A.3B). It is clear that G6PDH activity is only compromised in the double mutant and that the reciprocal induction of G6PD5 makes up for the loss of G6PD6 in the g6pd6 mutant. The unique attributes of plant OPPP, such as dual localization, complicate the interpretation of G6PDHs role in whole cells. The discovery of a plastidic pentose- phosphate transporter (Eicks et al. 2002) demonstrated a physical connection between plastidic and cytosolic pentose phosphate pathway but how the cellular supply of NADPH is coordinated remains unknown. Additionally, connectivity between biochemical pathways makes interpretation difficult. Overlap of intermediates with other metabolic pathways such as glycolysis, the TCA cycle, and amino acid and nucleotide biosynthesis makes it difficult to discern whether the effect is primarily due to reduced OPPP flux, NADPH supply, or something else. We speculate that in the double mutant but not in the single mutants (because of the G6PDH activity compensation) there is altered carbon metabolism and one of the effects is larger seeds. 175 A WT ngd5/96pd6 g6pd5 g6pd6 DAFLm’l 9 «N «'5 \‘3 Lm’l 9 «N «’5 «‘3 Lm’l e x“ «’5 «‘9 Lm’l c2, \’\ <5 <0 Lo 3:251:31 ' . 9i; i an an. A}: r ‘GGPD5 1 " ' " < plastidic 5 0 2 .“:u— - 2° 30 '53 §25 <7 £20 10 $9 3315 8 if: 810 UJE _ o g 40 w >140 24 5 230 920 16 E $10 8 we we) 0:6) «Meg. etc“ Q) 0 0 50 0: 50 Sucrose (mM) Figure 8.2. Glycolytic enzyme activities in green seedlings wri]-1, wild type, and WRII overexpressing seedlings in the presence or absence of sucrose. Values are the mean i SD, n=5. The seedlings shown in Figure Bl .A are representative of the tissue that was used for enzyme activity measurements. The data obtained from liquid enzyme assays is depicted in Figure B2. In wild type, PK activity in roots grown on Sue was reduced. On the contrary, with Suc present the gluco- and fructokinase activities representing HXK were slightly increased. There were no gross differences between the three genotypes. 188 Based on these results we concluded that glycolysis is active in ll-d-old seedlings that have progressed through the establishment phase during which sugar supplementation could alleviate the deficiency in the wri]-1 mutant. Therefore, WRIl does not to seem to be directly involved in regulation of glycolysis once seedlings start growing during the process of establishment. However, what is presumably limiting during this phase of development in the mutant is the sparse amount of seed storage oil, which is reduced at least 80% in wri]—1 seeds (Focks and Benning 1998). Offering Suc that can provide energy and building blocks at this stage completely restores seedling establishment (Figure B.1A). It seems that the role of WRIl on seedling establishment is indirect, by way of its involvement in the up-regulation of carbohydrate metabolism in the developing seed thereby ensuring the presence of storage oil needed for seedling establishment. 189 References Brocard-Gifford, 1., Lynch, T.J., Garcia, M.E., Malhotra, B., and Finkelstein, RR. (2004) The Arabidopsis thaliana ABSCISIC ACID—INSENSITIVE8 encodes a novel protein mediating abscisic acid and sugar responses essential for growth. Plant Cell 16:406421. Cernac, A., Andre, C., Hoffmann-Benning, S., and Benning, C. (2006)WR11 is required for seed germination and seedling establishment. Plant Physiol. 141:745-757. Cernac, A. and Benning, C. (2004) WRINKLEDI encodes an AP2/EREB domain protein involved in the control of storage compound biosynthesis in Arabidopsis. Plant J. 40:575-5 85. Cheng, J.C., Lertpiriyapong, K., Wang, S., and Sung, Z.R. (2000) The role of the Arabidopsis ELDI gene in cell development and photomorphogenesis in darkness. Plant Physiol. 123:509-520. Comah, J.E., Germain, V., Ward, J.L., Beale, M.H., and Smith, S.M. (2004) Lipid utilization, gluconeogenesis, and seedling growth in Arabidopsis mutants lacking the glyoxylate cycle enzyme malate synthase. J. Biol. Chem. 279:42916-42923. Eastmond, P.J., Germain, V., Lange, P.R., Bryce, J.H., Smith, S.M., and Graham, LA. (2000) Postgerminative growth and lipid catabolism in oilseeds lacking the glyoxylate cycle. Proc. Natl. Acad. Sci. U.S.A. 97:5669-5674. Focks, N. and Benning, C. (1998) wrinkled]: A novel, low-seed-oil mutant of Arabidopsis with a deficiency in the seed-specific regulation of carbohydrate metabolism. Plant Physiol. 118291-101. Footitt, S., Slocombe, S.P., Larner, V., Kurup, S., Wu, Y., Larson, T., Graham, 1., Baker, A., and Holdsworth, M. (2002) Control of germination and lipid mobilization by COMATOSE, the Arabidopsis homologue of human ALDP. EMBO J. 21 :2912-2922. Germain, V., Rylott, E.L., Larson, T.R., Sherson, S.M., Bechtold, N., Carde, J.P., Bryce, J.H., Graham, I.A., and Smith, S.M. (2001) Requirement for 3-ketoacyl-C0A thiolase-2 in peroxisome development, fatty acid beta-oxidation and breakdown of triacylglycerol in lipid bodies of Arabidopsis seedlings. Plant J. 2821-12. Hayashi, M., Toriyama, K., Kondo, M., and Nishimura, M. (1998) 2,4- Dichlorophenoxybutyric acid-resistant mutants of Arabidopsis have defects in glyoxysomal fatty acid beta-oxidation. Plant Cell 10:183-195. Hayashi, M., Yagi, M., Nito, K., Kamada, T., and Nishimura, M. (2005) Differential contribution of two peroxisomal protein receptors to the maintenance of peroxisomal functions in Arabidopsis. J. Biol. Chem. 280: 14829-14835. 190 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 camitine acyl carrier, is required for postembryonic grth in the light. Plant Cell 14:2161-2173. Masaki, T., Mitsui, N., Tsukagoshi, H., Nishii, T., Morikami, A., and N akamura, K. (2005) ACTIVATOR of Spo minzzLUCl/WRINKLEDI of Arabidopsis thaliana transactivates sugar-inducible promoters. Plant Cell Physiol. 46:547-556. Penfield, S., Graham, S., and Graham, LA. (2005) Storage reserve mobilization in germinating oilseeds: Arabidopsis as a model system. Biochem. Soc. Trans. 33:380-383. Penfield, S., Rylott, E.L., Gilday, A.D., Graham, S., Larson, T.R., and Graham, LA. (2004) Reserve mobilization in the Arabidopsis endosperm fuels hypocotyl elongation in the dark, is independent of abscisic acid, and requires PHOSPHOENOLPYRUVATE CARBOXYKINASEI. Plant Cell 16:2705-2718. Reiser, J., Linka, N., Lemke, L., Jeblick, W., and Neuhaus, H.E. (2004) Molecular physiological analysis of the two plastidic ATP/ADP transporters from Arabidopsis. Plant Physiol. 136:3524-3536. Ruuska, S.A., Girke, T., Benning, C., and Ohlrogge, J.B. (2002) Contrapuntal networks of gene expression during Arabidopsis seed filling. Plant Cell 14:1191-1206. Rylott, E.L., Gilday, A.D., and Graham, LA. (2003) The gluconeogenic enzyme phosphoenolpyruvate carboxykinase in Arabidopsis is essential for seedling establishment. Plant Physiol. 131:1834-1842. Wiese, A., Groner, F., Sonnewald, U., Deppner, H., Lerchl, J., Hebbeker, U., Flugge, U., and Weber, A. (1999) Spinach hexokinase I is located in the outer envelope membrane of plastids. F EBS Lett. 461113-18. Zolman, B.K., Monroe-Augustus, M., Thompson, B., Hawes, J.W., Krukenberg, K.A., Matsuda, S.P., and Bartel, B. (2001) chyI , an Arabidopsis mutant with impaired beta-oxidation, is defective in a peroxisomal beta-hydroxyisobutyryl-CoA hydrolase. J. Biol. Chem. 276:31037-31046. 191 Appendix C Leaf and seed lipid analysis of Arabidopsis mutants disrupted in Lipin genes 192 Introduction The biochemical reactions involved in lipid homeostasis have been described, but the regulators of these processes are more elusive. Lipodystrophy is a human metabolic disorder which is characterized by the loss of body fat and concomitant deposition of fat in the liver (Reitman 2005). Mice carrying mutations in the fatty liver dystrophy (fld) gene have symptoms of lipodystrophy and serve as a model for the human disease. Positional cloning was used to identify the gene responsible for mouse fld (Peterfy et al. 2001). The gene, called Lipin-1 (Lpinl) has homologues in all eukaryotic kingdoms. In mice LpinI controls triacyl glycerol (TAG) accumulation in adipose tissue. LpinI knock outs have no TAG and do not develop mature functioning adipocytes, while overexpressors of Lpinl contain more TAG and are prone to obesity (Peterfy et al. 2005, Phan et al. 2004). A yeast lipin mutant is also compromised in TAG accumulation (Han et al. 2006). Combined, these phenotypes are consistent with a role for lipin in TAG biosynthesis. The yeast and, more recently, the mammalian lipin genes have been shown to encode members of the Mg2+-dependent phosphatidate (PA) phosphatase (PAPl) family (Donkor et al. 2007, Han et al. 2006). The generation of diacylglycerol from PA is a step in the synthesis of TAG in both the Kennedy and the phosphocholinezdiacylglycerol acyltransferase pathways (outlined in Chapter 1). While lipin is required for TAG accumulation in yeast and mice, its contribution to TAG biosynthesis in plants has not been investigated. In addition, PA is recognized as a secondary messenger in plant (and other eukaryotic) signal transduction networks (Wang 2004), and an emerging role for PA is as a membrane lipid biosynthetic intermediate which is involved in the trafficking 193 of lipids from the endoplasmic reticulum to the plastid (Awai et al. 2006, Xu et al. 2005). The manipulation of PA levels through modulation of lipin gene expression provides an opportunity to explore the dynamics of such lipid signaling and trafficking. Moreover, the same mutants would provide a chance to investigate whether or not lipin plays a similar role in seeds as in adipose tissue, with respect to TAG synthesis. The relative ease of obtaining T-DNA insertion mutants for nearly any gene in the Arabidopsis genome provided enough impetus to pursue these questions. 194 Materials and Methods Gene identification Arabidopsis lipin homologues (At3g09560 and At5g42870;LPN1 and LPLI, respectively) were identified using. the deduced amino acid sequence of the Mus muscqu Lpz'n] gene (accession NM_172950). Homologues were identified using the basic local alignment search tool (BLAST) available at the TAIR website (wwwarabidopsisorg). Global gene expression data was mined from the AtGenExpress developmental database (Schmid et al. 2005). T-DNA mutant isolation Arabidopsis lipin mutants were selected from the SALK T-DNA insertion population (Alonso et a1. 2003). Selection of mutants was on half-strength MS medium containing 25 pg mL'l Kanamycin. T-DNA insertion sites were determined by sequencing PCR products made using primers designed with the help of the i-sect tool available on the Salk Institute website (http://signal.salk.edu/tdnaprimers.html). Primers for LPN] were 5’-TgTTATTgTTCTCTAATTTTg-3’ and 5’-gTTTngTCAgCTCTgACTgC-3’ and primers used for LPLI were 5’-gAATTCgCgCATAgTTngTC-3’ and 5’- AACAAgCCCCgTATCTCCTgT-3’. The left border primer used was LBal as suggested by the SALK Institute website (http://signal.salk.edu/). PCR conditions used standard buffer conditions and 40 cycles of 95°C for 30 sec, 52°C for 40 sec, and 72°C for 90 sec, followed by a 10 min 72°C extension. All subsequent genotyping was done using the same PCR protocol. 195 Lipid analysis Leaf samples (50-150 mg) were collected from individual plants and were weighed and frozen at -80°C. Lipids were extracted from leaves by vigorously shaking for 5 min in 500 uL of methanol/chloroform/formate (2:1 :0.1, v/v). Then 250 pL of 1 M KCl, 0.2 M H 3PO4 was added and the tubes were vortexed. The phases were separated by centrifugation at 16,000 g for 5 min. The organic phase was removed and loaded quantitatively onto a pre-treated (soaked in 0.15 M (N H4)2SO4 and dried, then heated to 120°C for 2.5 hrs) silica-60 TLC plates (Baker). Lipids were developed in a solvent system of acetone/toluene/water (91:30:7, v/v) and were stained with iodine and a- naphthol. Fatty acid methyl ester analysis of leaf and seed lipids was done as previously described (F ocks and Benning 1998). One small leaf or 10 seeds were used for each replicate. 196 Results and Discussion Lipin homologue identification and analysis of gene expression Arabidopsis lipin homologues were identified based on amino acid sequence similarity to Mus musculus Lpinl. Two genes were identified which encode Lpinl homologues, At3g09560 and At5g42870. The At3g09560 gene encodes a protein most similar to Lpinl (23% identical, 38% similar) and was named lipin] (LPNI). The other gene, At5g42870, encodes a protein less similar to Lpinl (22% identical, 36% similar) and thus was named lipinI—like (LPLI). Lipin proteins contain a DEXT motif characteristic of Mg2+-dependent phosphatases. In the Lpinl homologues from mammals, chicken, fish, Caenorhabditis elegans, Drosophila, Ciona, and S. cerevisiae, the motif is exactly DIDGT (Donkor et al. 2007). The Arabidopsis proteins vary only slightly and both contain a signature sequence of DVDGT. It is therefore likely that the Arabidopsis proteins are also Mg2+-dependent phosphatases. Publicly available Arabidopsis gene expression data reveals the tissue specific expression patterns for both LPN] and LPL] (Figure C.1). In general, LPN] is expressed at the same or higher level as LPLI in any given tissue. Both genes are induced during embryo development and have maximal expression just prior to the onset of TAG biosynthesis and accumulation. The gene expression data suggests that both LPN] and LPLI have roles in embryo development. 197 600 I At3909560 - LPN1 500‘ lAt5g42870-LPL1 $400- _1 S '5300‘ g (D 1‘3 93 Q. 2.. £500 1oo~ _ lllll 5% 0. 3 giiigfi‘éz. , mm”“’“‘“‘“‘“-a>8 LLLLLLLLLLLLLLLL og°——-°’_‘”3300000000 3 VmFfiEQ 0° v into 5) ego C G: o .s (D Figure C.l. Relative gene expression of putative Arabidopsis lipin homologues Figure is derived from published microarray data (Schmid et al., 2005). DAF, days after flowering. T-DNA mutant isolation and lipid analysis The Arabidopsis T-DNA mutant population was utilized to study the in vitro function of the Arabidopsis lipins (Alonso et al. 2003). Three lines were obtained which had inserts in the desired genes (Figure C.2). Genotyping of putative mutants was done with PCR as described in materials and methods and exact insertion sites were determined. As it turns out, SALK_146637 and SALK_042850, both located in LPN], had identical insertion sites. For all three lines, the T-DNA was inserted into a translated portion of an exon which almost certainly causes reduced transcript abundance in homozygous lines. However, it remains to be tested whether or not transcription is actually affected by these 198 insertions. There were no obvious morphological or developmental differences between the single mutants and wild type. A double mutant was generated by crossing lpn] and [pl] and it too was wild-type in appearance. SALK_146637 SALK_042850 At3909560 - LPN 1 STOP SALK_047457 At5 42870 - LPL1 ATG 9 STOP 1 Kb Figure C.2. Gene structure of the two Arabidopsis Lpinl homologues Locations of the T-DNA insertions are indicated. Black box on T-DNA is the left border. ATG, start codon; STOP, stop codon. Lipid analysis was carried out for the individual lipin mutants as well as for the double mutant, despite not knowing the degree of mRNA reduction in any of the lines. First, total leaf lipids were extracted quantitatively and analyzed by thin layer chromatography (TLC). Figure C.3A shows the results of this experiment. Two staining methods were used on the same plate to maximize the number of lipids detected. Membrane galacto- and phospholipids dominate the stained TLC plate. It is clear that there are no gross differences in the lipid composition between wild type and either the single or double mutants. 199 front ' 70 IWT 60‘ u/pn1 "'; nip/1 'f’ :5 _-. ,; 50. n/pn1/Ip/1 :1 MGDG . °\o 40“ g» ‘o E 30- fi'; DGDG 2o- 16:0 16:1 16:2 18:0 16:3 18:1 18:2 18:3 C D 7.0 30 60 .WT A . u/pn1 g 25 urn/1 a: 5.0 Ipn1/lpl1 £ 20 0') 34.0 s: UJ °315 E 3.0 g < '-__L 10 9 2.0 O "‘ 1.0 5 O . 0 ~ 7 . WT Ipn1 Ip/1 Ipn1/lp/1 16:018:018:1 18:2 20:018:3 20.1 22.0 22:1 Figure C.3. Lipid analysis of Arabidopsis lipin single and double mutants (A) Thin layer chromatograrn of total leaf lipids stained with iodine and a sugar-sensitive stain. DGDG, digalactosyldiacylglycerol; MGDG, monogalactosyldiacylglycerol. (B) Fatty acid composition of total leaf lipids. Single leaves were subjected to fatty acid methyl ester analysis using gas chromatography. WT, wild type. (C) Seed oil content of wild type (WT) and lipin single and double mutants as determined by fatty acid methyl ester (FAME) analysis. (D) Fatty acid profile from seeds used for oil content determinations in (C). 200 The fatty acid profiles of total lipids from leaves were determined as well (Figure C.3B). Again, the mutants and wild type were remarkably similar. Seed oil content and composition were examined in mature seeds of the lipin mutants. As shown in Figure C .3C, there were no differences observed between any of the lines. And again, the fatty acid profile of the lipids in questions was determined, and again, the wild type and mutant were the same (Figure C.3D). The lack of a lipid phenotype for these mutant lines is not totally surprising. Gene expression levels in the mutants were not analyzed and it is possible that the T-DNA insertions have no affect on mRNA or protein abundance. A simple RT-PCR or RNA blot experiment would begin to address this issue. Also, there was no confirmation of the encoded proteins’ putative catalytic fimction. Nonetheless, even if the genes in question do encode active Mg2+-dependent phosphatidate phosphatases, and even if gene expression is completely knocked out, the mutants might not have an easily detectable phenotype. Phosphatidate occurs at very low concentrations in plant cells and is not readily measured by the methods used here. Any changes in the steady state amount of PA would have to be detected by other means. Furthermore, the fatty acid profile data in Figures C.3B and C.3D is of total lipids and does not take into account any changes at the level of individual lipid species (i.e. PG, MGDG, DGDG). Beyond these obvious methodological shortcomings, the flexibility of plant metabolism could also account for the apparent lack of aberration in the lipin mutants. For instance, certainly LPN1 and LPLl are not the only PA phosphatases in Arabidopsis. Future studies of lipin in Arabidopsis should be focused on mutants with documented reductions in gene expression in combination with other lipid metabolism mutants, such as TGDl -3. 201 References Alonso, J.M., Stepanova, A.N., Leisse, T.J., Kim, C.J., Chen, H.M., Shinn, P., Stevenson, D.K., Zimmerman, J ., Barajas, P., Cheuk, R., Gadrinab, C., Heller, C., Jeske, A., Koesema, E., Meyers, C.C., Parker, H., Prednis, L., Ansari, Y., Choy, N., Deen, H., Geralt, M., Hazari, N ., Hom, E., Karnes, M., Mulholland, C., Ndubaku, R., Schmidt, 1., Guzman, P., Aguilar-Henonin, L., Schmid, M., Weigel, D., Carter, D.E., Marchand, T., Risseeuw, E., Bragden, D., Zeko, A., Crosby, W.L., Berry, CC, and Ecker, J.R. (2003) Genome-wide insertional mutagenesis of Arabidopsis thaliana. Science 301:653-657. Awai, K., Xu, C., Tamot, B., and Benning, C. (2006) A phosphatidic acid-binding protein of the chloroplast inner envelope membrane involved in lipid trafficking. Proc. Natl. Acad. Sci. U.S.A. 103:10817-10822. Donkor, J ., Sariahmetoglu, M., Dewald, J ., Brindley, D.N., and Rene, K. (2007) Three mammalian lipins act as phosphatidate phosphatases with distinct tissue expression patterns. J. Biol. Chem. 282:3450-3457. Focks, N. and Benning, C. (1998) wrinkled]: A novel, low—seed-oil mutant of Arabidopsis with a deficiency in the seed-specific regulation of carbohydrate metabolism. Plant Physiol. 118:91-101. Han, C.S., Wu, W.H., and Carman, GM. (2006) The Saccharomyces cerevisiae Lipin homolog is a M g2+-dependent phosphatidate phosphatase enzyme. 1. Biol. Chem. 281:9210-9218. Peterfy, M., Phan, J ., and Reue, K. (2005) Alternatively spliced lipin isoforms exhibit distinct expression pattern, subcellular localization, and role in adipogenesis. J. Biol. Chem. 280:32883-32889. Peterfy, M., Phan, J ., Xu, P., and Rene, K. (2001) Lipodystrophy in the fld mouse results from mutation of a new gene encoding a nuclear protein, lipin. Nat. Genet. 27: 121-124. lPhan, J., Peterfy, M., and Reue, K. (2004) Lipin expression preceding peroxisome proliferator-activated receptor-gamma is critical for adipogenesis in vivo and in vitro. J. Biol. Chem. 279:29558-29564. Reitman, M.L. (2005) The fat and thin of lipin. Cell Metab. 1:5-6. Schmid, M., Davison, T.S., Henz, S.R., Pape, U.J., Demar, M., Vingron, M., Scholkopf, B., Weigel, D., and Lohmann, J.U. (2005) A gene expression map of Arabidopsis thaliana development. Nature Gen. 37:501-506. Wang, X. (2004) Lipid signaling. Curr. Opin. Plant Biol. 7 2329-336. 202 Xu, C., Fan, J ., F roehlich, J .E., Awai, K., and Benning, C. (2005) Mutation of the TGDl chloroplast envelope protein affects phosphatidate metabolism in Arabidopsis. Plant Cell 17 23094-3] 10. 203 Appendix D In vitro substrate specificity of the Rhodobacter sphaeroides betaine lipid biosynthetic enzyme BtaA5 5 This work, was done in collaboration with Dr. Wayne Riekhof and has been published in: Riekhof, W.R.., Andre, C., and Benning, C. (2005) Two enzymes, BtaA and BtaB, are sufficient for betaine lipid biosynthesis in bacteria. Arch. Biochem. Biophys. 441:96-105. I contributed Figure D.2. 204 ‘Introduction The sessile lifestyle of many plants, bacteria, and fungi has led to the development of complex means of coping with specific mineral deficiencies. For instance, the replacement of abundant cellular phospholipids with alternative, non-phosphorous lipids is one means of dealing with environmental phosphate limitation. This phenomenon has been the most extensively documented for the a-proteobacteria Rhodobacter sphaeroides and Sinorhizobium meliloti, which synthesize the betaine lipid diacylglyceryl-N,N,N- trimethylhomoserine (DGTS) and omithine containing lipids to replace depleted membrane phospholipids (Benning et al. 1995, Geiger et al. 1999). DGTS was first discovered in a unicellular alga (Brown and Elovson 1974), but has since been found in lower plants and fungi (Kunzler and Eichenberger 1997, Rozentsvet et al. 2000). The structure and zwitterionic nature (Figure D.1A) of DGTS led to the hypothesis that it replaces phosphatydylcholine in the organisms in which it occurs. The use of radioisotope labeling in Chlamydomonas reinhardtii and in R. sphaeroides led to the identification of methionine as the biosynthetic precursor for the homoserine moiety and the adjoining methyl groups (Hofmann and Eichenberger 1996, Sato 1988). Subsequently, a genetic study using R. sphaeroides resulted in the identification of a two gene operon responsible for DGTS biosynthesis during phosphate starvation (Klug and Benning 2001). The two genes, designated btaA and btaB, respectively encode a proposed AdoMet/diacylglycerol 3-amino-3-carboxypropyl transferase producing the intermediate DGHS, and a putative methyltransferase adding three methyl units to the amino group of DGHS to form DGTS. 205 NH; Q7, Q; a? be ~32 .5 (go MetK 0v (on 0&7 0Q DAG Q' Q- ‘o to AdoMet . ml 2 O;;'/\/CSIH3 “H R1 2 < DGHS NH3 OH OH BtaA 5'MTA R2 DGHS R1 i w... E;§:..’.-..'L;§?'+PtdEtn NH; inspim (3x. DGTS R1 Figure D.1. DGTS biosynthetic pathway and involvement of RthaAB (A) BtaA is proposed to catalyze the transfer of the 3-amino-3-carboxypropyl group of AdoMet to the 3-hydroxyl of DAG to form DGHS. DGHS is then N-methylated by BtaB to form DGTS. AdoHcy, S—adenosylhomocysteine; AdoMet, S-adenosylmethionine; DAG, diacylglycerol; DGHS, diacylglycerylhomoserine; DGTS, diacylglyceryl-(N,N,N)- trimethylhomoserine; MetK, AdoMet synthetase; 5'MTA, 5'-methylthioadenosine. (B) E. coli expressing btaA accumulate DGHS. The btaA expression strain (third lane) shows a band co-migrating with authentic DGHS from a phosphate stressed R. sphaeroides btaB KO strain (first lane). Phosphate-replete R. sphaeroides btaB K0 and empty pQE-31 served as negative controls (second and fourth lanes, respectively). 206 This proposed DGTS biosynthetic scheme involving these two enzymes is depicted in Figure D. 1 A. While these genes have been shown to be necessary for DGTS biosynthesis in vivo, a biochemical analysis of their protein products has not been done. This work focuses on the BtaA enzyme, due to its unusual reaction, and describes an in vitro experiment which helped confirm its identity as an AdoMet/diacyl glycerol 3-amino-3- carboxypropyl transferase. 207 Materials and Methods In vitro activity assays of RthaA Escherichia coli TOPIO F' harboring thaA (constructed as described in Riekhof et al. 2005) was grown in 250 ml LB-ampicillin at 37 °C to an OD of 0.7, and induced with 0.25 mM IPTG followed by an additional 4 h of growth at 28 °C. Cells were harvested by centrifugation and the cell pellet was suspended in 10 ml of cold buffer (50 mM Hepes, 1 mM DTT, 1 mM EDTA, pH 7.3). The resuspended cells were sonicated 3—4 times, 30 3 each with a microprobe tip, and the lysate was centrifuged at 2000 g for 10 min to remove unbroken cells and cellular debris. Aliquots (1 ml) of the cell-free extract were then frozen in liquid N2 and stored at -80 °C prior to use. Activity under these storage conditions did not decrease appreciably for at least 1 month. Assays were conducted in 100 pl final volume by combining 48.75 ul of cell-fiee extract with 48.75 pl of 100 mM Hepes, Tris—Cl, or MES, 1 mM DTT, 1 mM EDTA, at varying initial pH to give a final pH in the range of 5.5—8.6 when mixed with the cell-free extract (initial, pH 7.3). Reactions were initiated by addition of 25,000 dpm 1- [I4C]AdoMet (American Radiolabeled Chemicals, 2.5 u], final concentration of AdoMet of 2.1 uM, specific activity 7.14 MBq/nmol), or 75,000 dpm [14C]DAG (dioleoyl-rac- glycerol, [oleoyl l-MC]; American Radiolabeled Chemicals, final concentration 6.3 uM, specific activity 7.14 MBq/nmol). For the reaction initiated with labeled DAG, it was essential to dry [14C]DAG, which was delivered suspended in toluene/ethanol (1 :1, v/v) at room temperature under a stream of nitrogen. A sonicating water bath was used to disperse the DAG at the desired concentration in a buffer of 50 mM Hepes, pH 7.8, 1 mM EDTA, lmM EGTA, 0.1% Triton X-100. Reactions were incubated at 28 C, and 208 terminated by addition of 400 pl of chloroform/methanol (1 : 1, v/v) and 100 pl 0.9% (w/v) NaCl to separate aqueous and organic phases. The organic phase was transferred to a new tube, dried under a stream of nitrogen, and dissolved in 50 pl chloroform/methanol (1 :1, v/v). This lipid extract was then spotted onto silica-6O TLC plates (Baker) and developed with the solvent chloroform/acetone/methanol/acetic acid/water (10:42222:1, v/v), followed by quantification of the signal for DGHS on a Molecular Dynamics phosphorimager screen (Amersham, Piscataway, NJ, USA) with the ImageQuant software package. Alternatively, TLC plates were stained with ninhydrin reagent to visualize primary amine-containing lipids. 209 Results and Discussion The btaA and btaB genes have previously been shown to be necessary for DGTS accumulation in phosphate stressed R. sphaeroides (Klug and Benning 2001). Expression of btaA in E. coli resulted in the accumulation of a primary amine-containing lipid band which co-mi grated with authentic DGHS produced from phosphate stressed R. sphaeroides btaB KO (Figure D. 1 B). We expected that E. coli expressing RthaA could use endogenous diacyl glycerol (DAG) and S-adenosylmethionine (AdoMet) to synthesize DGHS. Heterologous expression of RthaA confirmed the catalytic role of the encoded protein, but did not confirm DAG or AdoMet as the substrates used for DGHS biosynthesis. In vitro assays of RthaA enzyme activity were conducted using 1-[I4C]AdoMet or ['4C]DAG to follow product formation. While we were able to utilize the engineered Hisé-tag to purify small amounts of apparently soluble RthaA protein using various detergents in the lysis buffer, we were unable to demonstrate activity in a reconstituted liposome system. To circumvent this problem, we developed a system to minimize the steps between expression and enzyme assay to demonstrate the proposed reaction catalyzed by RthaA. As we wanted to test the incorporation of label from 1- [14C]AdoMet as well as ['4C]DAG, two sets of controlled reactions were set up keeping all conditions the same except for the compound carrying the label. Both sets of reactions were supplied with the two substrates, AdoMet at 2.1 pM and DAG at 6.3 pM. In the two sets of reactions, the specific activity for one or the other substrate was kept the same. In both sets a 3-fold higher amount of DAG was used because unlike AdoMet, DAG is offered as a micelle suspension and a number of other reactions consume this precursor 210 as well (see below). This most suitable 1—3 ratio of substrates was empirically determined. Given that the specific activity of both substrates and their total concentrations in the two sets of reactions were identical, we expected that the rates for both sets of reactions were identical if both compounds are direct substrates of the enzyme. The result is shown in Figure D.2A. Label from AdoMetwas efficiently incorporated into DGHS by transfer of the 3-amino-3-carboxylpropyl moiety to a lipid acceptor giving rise to a single labeled compound in the lipid fraction of the reaction extract. This compound was not present in the vector control consistent with it being DGHS. When labeled DAG was used, label was observed in a number of polar lipids as DAG is a general precursor for polar lipid biosynthesis. These labeled compounds were also present in the vector control. One compound co-chromatographing with DGHS present in the AdoMet reaction (Figure D.2A, box) was also present, but unlike other lipids, it was absent from the vector control thereby suggesting that it was DGHS. Thus, labeled DAG appeared to serve as a precursor for DGHS biosynthesis under the employed conditions. To determine the relative rates of incorporation, the phosphor imager outputs were quantified (Figure D28). The rates were linear for both labeled substrates (r2 = 0.9964 for the AdoMet reaction and r2 = 0.9743 for the DAG reaction) over the incubation time suggesting that the enzyme was present at non-saturating amounts in the reaction mixture. The rates were very similar (29.6 relative units/min for the AdoMet reaction and 27.7 relative units/min for the DAG reaction). Given that the specific activity for the labeled compounds and the total substrate concentrations in both reactions were identical, one can cautiously conclude that both compounds were incorporated into DGHS with similar efficiency as expected for the direct substrates of the reaction. 211 [14C]-AdoMet [14C]-DAG A thaA \fl: thaA Vic B 15 30 6012012015 30 60120120 6000 32.9! DAG y=29.62x+1784 . , .. .. " r - = 1—DGHS R2=o.9964 § 33000 . a: = 27.749x + 1333 .. t s Q 3 {-i 2000‘ yR2: 0.9743 I 1000* 0 . DAG -Ori 0 2'0 40 60 80 100120140 Time (min) Figure D.2. BtaA-catalyzed DGHS biosynthesis from radiolabeled substrates (A) E. coli cells expressing RthaA were assayed for the ability to incorporate label from 1-[I4C]AdoMet or [”C]DAG into DGHS. Cells either contained the thaA plasmid or an empty pQE31 vector control (Vec). Time courses are shown with incubation times (min) indicated. Reaction products were separated by TLC and detected using a phosphorimager. The product bands containing DGHS (missing in the vector controls) are shown inside the box. DAG, diacylglycerol; DGHS, diacylglycerylhomoserine; Ori, origin. (B) Plot of DGHS synthesized from 1-[14C]AdoMet or ['4C]DAG over time. DGHS amount (expressed as pixel density) was determined from the phosphorimage in (A). The rates of DGHS synthesis from radiolabeled DAG or SAM are linear and have roughly the same value. 212 The RthaA protein in a crude membrane preparation was shown to transfer the 3-amino-3-carboxypropyl moiety from AdoMet t0 DAG by demonstrating that under identical conditions incorporation of label into DGHS from either labeled AdoMet or DAG proceeded at the same rate (Figure D.2). This result is in agreement with the role of RthaA that had been tentatively assigned based on mutagenesis and heterologous expression, and rules out other contingencies for the activity of RthaA. Previous work in a cell-free system on the cognate BtaA-type transferase activity in C. reinhardtii using radiolabeled AdoMet to follow the reaction had given conflicting results as to the identity of the hydrophobic substrate, indicating that DAG might not be the direct substrate because the addition of unlabeled DAG strongly inhibited the reaction (Moore et a1. 2001). However, the point was raised that excess DAG might simply disrupt the membrane environment in which the enzyme(s) are working, and the decrease in activity might not be a result of enzyme inhibition, per se. The results presented here for recombinant RthaA seemingly provide a solution to this question. 213 References Benning, C., Huang, Z.H., and Gage, D.A. (1995) Accumulation of a novel glycolipid and a betaine lipid in cells of Rhodobacter sphaeroides grown under phosphate limitation. Arch. Biochem. Biophys. 317:103-1 11. Brown, A.E. and Elovson, J. (1974) Isolation and characterization of a novel lipid, 1(3),2-diacylglyceryl-(3)-O-4'-(N,N,N-trimethyl)homoserine, from Ochromonas danica. Biochemistry 13:3476-3482. Geiger, 0., Rohrs, V., Weissenmayer, B., F inan, T.M., and Thomas-Oates, J.B. (1999) The regulator gene phoB mediates phosphate stress-controlled synthesis of the membrane lipid diacylglyceryl-N,N,N-trimethylhomoserine in Rhizobium (Sinorhizobium) meliloti. Mol. Microbiol. 32:63-73. Hofmann, M. and Eichenberger, W. (1996) Biosynthesis of diacylglyceryl-N,N,N- trimethylhomoserine in Rhodobacter sphaeroides and evidence for lipid-linked N methylation. J. Bacteriol. 17826140-6144. Klug, R.M. and Benning, C. (2001) Two enzymes of diacylglyceryl-O-4'-(N,N,N,- trimethyl)homoserine biosynthesis are encoded by btaA and btaB in the purple bacterium Rhodobacter sphaeroides. Proc. Natl. Acad. Sci. U.S.A. 98:5910-5915. Kunzler, K. and Eichenberger, W. (1997) Betaine lipids and zwitterionic phospholipids in plants and fungi. Phytochemistry 46:883-892. Moore, T.S., Du, Z., and Chen, Z. (2001) Membrane lipid biosynthesis in Chlamydomonas reinhardtii. in vitro biosynthesis of diacylglyceryltrimethylhomoserine. Plant Physiol. 125:423-429 Riekhof, W.R., Andre, C., and Benning, C. (2005) Two enzymes, BtaA and BtaB, are sufficient for betaine lipid biosynthesis in bacteria. Arch. Biochem. Biophys. 441 :96-105. Rozentsvet, O.A., Dembitsky, V.M., and Saksonov, S.V. (2000) Occurrence of diacylglyceryltrimethylhomoserines and major phospholipids in some plants. Phytochemistry 54:401-407. Sato, N. (1988) Dual role of methionine in the biosynthesis of diacylglyceryltrimethylhomoserine in Chlamydomonas reinhardtii. Plant Physiol. 86:93 1-934. 214 IIIIIIIIIIIIIIIIIIIIIIIIIIIIIII 1|411111111111111111111111111111111111