. f .. 1n... a...ilz spy , 2. . :..f 1H. y Ln uomnr LMichigan State University w" This is to certify that the dissertation entitled RE-EXAMINING THE INITIAL STEPS OF MEMBRANE AND STORAGE LIPID ASSEMBLY IN PEA LEAVES AND SOYBEAN EMBRYOS: THE DOMINANT FLUX OF NEWLY SYNTHESIZED FATTY ACID INCORPORATION INTO EXTRA-PLASTIDIC GLYCEROLIPIDS IS THROUGH PHOSPHATIDYLCHOLINE ACYL EDITING presented by Philip David Bates has been accepted towards fulfillment of the requirements for the Ph.D. degree in Biochemistry and Molecular Biolgoy @ 5 flew Major Professor’s Signature Sqrf 2’5; goof? Date MSU is an Affirmative Action/Equal Opportunity Employer PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE ABSTRACT RE-EXAMINING THE INITIAL STEPS OF MEMBRANE AND STORAGE LIPID ASSEMBLY IN PEA LEAVES AND SOYBEAN EMBRYOS: THE DOMINANT FLUX OF NEWLY SYNTHESIZED FATTY ACID INCORPORATION INTO EXTRA-PLASTIDIC GLYCEROLIPIDS IS THROUGH PHOSPHATIDYLCHOLINE ACYL EDITING By Philip David Bates In expanding pea leaves and developing soybean embryos over 95% of fatty acids (FA) synthesized in the plastid are exported for assembly of eukaryotic membrane and storage glycerolipids. It is often assumed that the major products of plastid FA synthesis (18:1 and 16:0) are first acylated to glycerol-3-phosphate (G3P) producing 1610/1821 and 18:1/18:1 initial molecular species of phosphatidic acid (PA) which are then converted to phosphatidylcholine (PC), the major eukaryotic phospholipid and site of acyl desaturation. However, by rapid labeling of pea leaf and soybean embryo lipids with [“C]acetate and [14C]glycero| we demonstrate that acyl editing is an integral component of eukaryotic glycerolipid synthesis. First, no precursor-product relationship between diacylglycerol (DAG) and PC [“C]acyl chains was observed at very early time points. Second, analysis of PC molecular species and the positional localization of nascent FA within PC at these early time points showed that the majority of newly synthesized acyl groups (> 90% of the [“C]18:1 and [“C]saturated acyl groups in pea, and >60% in soybeans) were incorporated into PC alongside a previously synthesized unlabeled acyl group (18:2, 18:3 or 16:0). The nascent FA occupied 62% of the sn-2 position in pea leaves and 86% in soybean embryos. Third, [”Clglycerol labeling produced PC molecular species highly enriched with 18:2, 18:3 and saturated acyl chains and not 18:1, the major product of plastid fatty acid synthesis. And fourth, the kinetics of [“C]glycerol and the positional analysis of newly synthesized FA incorporation into triacylglycerol (TAG) of soybean embryos indicated that TAG is synthesized from a DAG molecule generated by removal of the PC phosphocholine head group and not de novo synthesized DAG. In conclusion, we propose that most newly synthesized and plastid exported acyl groups are not immediately utilized for PA synthesis, but instead are incorporated directly into PC through an acyl editing mechanism that operates at both sn-1 and sn-2 positions. The acyl groups removed by acyl—editing are largely used for the net synthesis of PC and other glycerolipids through (G3P) acylation. Very little if any TAG is produced by three consecutive acylations as in the traditional Kennedy pathway. These results provide a more quantitative basis to demonstrate in vivo that FA flux in/out of PC by acyl editing, and the inter-converstion of DAG with PC are dominate fluxes in the assembly of eukaryotic membrane and storage glycerolipids in plants. This dissertation is dedicated to my wife Cassandra. Her loving support, strong encouragement, sacrifice of moving to Michigan, and ability to make me fat and happy has made the perils of graduate school relatively easy to bear. ACKNOWLEDGEMENTS I would like to express my appreciation to my advisor John Ohlrogge. John gave me the opportunity to work in his lab and entrusted me with interesting and challenging research projects. His door was always open when I needed assistance and he was always encouraging, even when the research was not going well. I am especially grateful for his guidance that shifted my research towards the subject of this dissertation. l have learned not only how to do good science, but from his examples also how to be a good leader. Mike Pollard provided me with enormous amounts of overall guidance and help with step by step methods, scientific theory, and data analysis. Without his analytical mind some of the major conclusions and finer details of many experiments would have been passed over. Mike was my main advisor for many aspects of my research. Plus, Mike kept me on my toes by providing a jovial remark to anything I said that was less than very intelligent. I am grateful for all the members of the Ohlrogge lab for their friendship and assistance, especially Doug Allen and Tim Durrett who provided very valuable discussions and research assistance. Finally, I would like to thank my guidance committee for their guidance during my graduate research and their support to head into new directions when it was needed. TABLE OF CONTENTS LIST OF TABLES viii LIST OF FIGURES ix 1 INTRODUCTION 1 1.1 Background overview 3 1.2 Fatty acid synthesis and desaturation 4 1.2.1 Acetyl-CoA carboxylase: Major FAS regulatory point ........................ 6 1.3 Esterification of fatty acids to glycerol-3—phosphate: de novo glycerolipid synthesis. 9 1.3.1 The two pathways for de novo glycerolipid synthesis in plants ........ 10 1.3.2 Transfer of eukaryotic synthesized DAG to the plastid for glycolipid 14 synthesis 1.4 Long chain acyl-CoA pools for building glycerolipids ............................. 16 1.5 Synthesis of eukaryotic membrane lipid classes 18 1.6 TAG synthesis 19 1.6.1 Acyl transfer between lipids is required for a majority of TAG synthesis 21 1.7 Suggestive evidence for acyl editing in plants 23 1.8 The in vivo labeling approach 26 2 INCORPORATION OF NEWLY-SYNTHESIZED FATTY ACIDS INTO CYTOSOLIC GLYCEROLIPIDS IN PEA LEAVES OCCURS VIA ACYL EDITING 30 2,1 Ahefmrt 31 2.2 Introduction 32 2.3 Experimental procedures 36 2.4 Results 41 2.5 Discussion 60 2.6 Footnotes 68 3 THE DOMINANT INITIAL FLUX OF FATTY ACIDS INTO SOYBEAN OIL IS THROUGH PHOSPHATIDYLCHOLINE ACYL EDITING ............................. 70 3.1 Abstract 71 3.2 Introduction 72 3.3 Results 76 3.4 Discussion 101 vi 3.5 Conclusion 119 3.6 Materials and methods 120 3.7 Footnotes 126 4 CONCLUSION 127 4.1 Acyl editing conclusions 127 4.2 Additional significance of acyl editing and PC derived TAG synthesis.133 4.3 Future research on acyl editing and TAG synthesis pathways ............. 135 5 APPENDIX A: CHAPTER 2 SUPPLEMENT 143 6 APPENDIX B: CHAPTER 3 SUPPLEMENT 159 6.1 Supplement Figures for Chapter 3 Results 159 6.2 Supplement: Kinetic modeling of pool filling 170 6.3 Discussion of soybean TAG synthesis flux model Figure 3-10 ............ 178 7 APPENDIX C: PC ACYL EDITING IN ISOLATED PEA CHLOROPLASTS183 8 REFERENCES 187 vii LIST OF TABLES Table 3-1 Endogenous lipid pool sizes and fatty acid composition ..................... 78 Table 6-1 TAG Molecular Species Composition of Soybean Embryos ............. 167 Table 6-2 DAG Molecular Species Composition of Soybean Embryos ............. 168 Table 6-3 PC Molecular Species Composition of Soybean Embryos ................ 169 Table 6-4 TAG synthesis model: nascent FA flux value calculations ................ 179 viii LIST OF FIGURES Figure 1-1 Traditional textbook two pathway model of glycerolipid synthesis ..... 11 Figure 1-2 Possible enzymatic reactions leading to TAG synthesis in oilseeds ..20 Figure 2—1 Time course for [“C]acetate incorporation into the acyl groups of glycerolipid classes by excised pea leaves. 43 Figure 2-2 [14C]Acyl composition of [14c1acetate labeled lipids ........................... 45 Figure 2-3 Time course for incorporation of label from [“C]glycerol into the backbone and acyl groups of glycerolipids by excised pea leaves ...................... 47 Figure 2-4. Molecular species analyses of endogenous PC and PA from pea leaves. 49 Figure 2-5 Analysis of endogenous PC molecular species and those labeled after short time incubations of excised pea leaves with [“C]acetate or [“C]glycero| or of seedlings with [“C]COz. 52 Figure 2-6 Stereochemical distribution of [“C]acyl chains in PC. ....................... 55 Figure 2-7 Comparison of unlabeled fatty acid compositions deduced from [1 C]acetate and [14C]glycerol labeled PC molecular species to endogenous pea leaf PC and lyso-PC compositions. 57 Figure 2-8 Distribution of lipid classes, labeled in their acyl chains, after incubation of pea leaves with [1401002 or [“C]acetate. 59 Figure 2-9 Four possible models describing pathways and fluxes for the incorporation of nascent FA, exported from the chloroplast immediately after de novo FAS, into PC 62 Figure 3-1 Endogenous molecular species compositions of DAG, PC and TAG. 79 Figure 3-2 Time course for [‘4C]acetate incorporation into the acyl groups of the major labeled soybean embryo lipids. 82 Figure 3-3 Fatty acid composition of [“Clacetate labeled lipids .......................... 85 Figure 3—4 Positional analysis of [”C]acetate labeled acyl groups in different lipids. - 87 Figure 3-5 Molecular species compositions of [“C]acetate labeled PC and DAG. 89 Figure 3-6 Six min [“C]acetate labeled TAG molecular species ......................... 92 Figure 3-7 Time course for the incorporation of [14C]glycerol into the backbone and acyl groups of the major labeled soybean embryo lipids. ............................. 96 Figure 3—8 Molecular species composition of [‘4C]glycerol backbone labeled PC, DAG and TAG. 99 Figure 3-9 Representations of mathematical models for TAG synthesis ........... 107 Figure 3—10 Soybean TAG synthesis flux model. 111 Figure 5—1 Supplemental Figures S1A-S1M: Lipidomics of pea leaves and acetate labeled pea leaves. 144 Figure 5-2 Estimates of single and dual acyl-labeled PC molecular species for Model 3 (Figure 2-9) Figure 6-1 [“C]Fatty acid composition of total [“Cjacetate labeled soybean lipids. - 159 Figure 6-2 Alpha oxidation of [‘4C]acetate labeled saturated FA from TAG and PC. 160 Figure 6-3 TLC image examples and flow chart for analysis of radiolabeled glycerolipids. 161 Figure 6-4 [“Clacetate labeled PC molecular species time course. ................. 162 Figure 6-5 [“C]acetate labeled DAG molecular species time course. .............. 163 Figure 6-6 [“C]acetate labeled TAG molecular species time course ................ 164 Figure 6-7 [14C]glycerol labeled molecular species time course of PC and ONT-65 Figure 6-8 Fatty acid composition of [“C]glycerol labeled TAG, DAG and PC molecular species. 166 Figure 6-9 Kinetic modeling of pool filling simulations 174 Figure 7-1 [”Cjacetate labeled PC and PG molecular species from isolated chloroplasts. 184 1 INTRODUCTION Fatty acid (FA) based lipids are essential to life and perform a wide variety of functions. Living cells depend upon a hydrophobic bilayer of amphipathic membrane lipids to separate the cell contents from the aqueous environment. Eukaryotic cells accomplish complex metabolism by compartmentalization of reactions within membrane bound organelles. Membrane lipids can be utilized as substrates for metabolic signaling compounds [1-5]. Multi-cellular plants contain fatty acid derived physical barriers against water, solutes, pathogens and solar UV radiation in the form cutin, suberin, and wax [6-11]. Triacylglycerols (TAG) contain three FA bound to a glycerol backbone and are the major form of reduced carbon used for energy storage in plants and animals. In addition to the biological importance of FA containing lipids within an individual organism, many animals rely on plant TAG as a major source of nutrition. The human population utilizes plant TAG for food, fuel and industrial applications. In 2006, world harvest of plant oil was over 127 million tons [12]. Vegetable oil produced from soybeans represents the largest source of oil for nutritional purposes at approximately 35 million tons produced in 2006 [13]. On average the FA associated with common sources of vegetable oil are composed of 16 and 18 carbon saturated FA (palmitic (16:0) and stearic acid (18:0), respectively) and 18 carbon FA containing 1-3 double bonds (oleic (18:1), linoleic (18:2) and linolenic acid (18:3), respectively). These common FA are useful for cooking oils and high energy food sources but do not contain the most healthy, oxidatively stable or most useful functional groups required in many highly nutritional foods, fuels or industrial applications, respectively. Nutritionally important unusual FA are the very long chain polyunsaturated fatty acids (PUFA) such as docosahexaenoic acid (DHA) produced in the microalgae Crypthecodinium cohnii [14] and usually consumed through fish. An example of industrially important unusual FA is the hydroxyl containing ricinoleic acid produced by the castor plant Ricinus communis [15, 16] which can be used in lubricants and as a feedstock to polymers such as nylon. Most plants cultivated for high yields of vegetable oil (e.g. soybeans and rapeseed) were domesticated based on their value as food/feed and therefore do not contain many of the industrially important unusual FA. Plant species that do accumulate high levels of unusual FA (e.g. castor or tung) typically have agronomical traits that make harvesting large yields difficult and expensive. There is great interest in engineering unusual FA into some of the seasonal high yielding crop species. Additionally, the rising human population and dwindling petroleum reserves increases demand for renewable sources of chemical feed stocks and fuel. Plant oils can be used as direct replacements for diesel fuel and the wide variety of unusual FA could be the base for many industrial applications [12, 13]. Even though there have been many attempts to engineer the FA composition or the quantity of TAG in oilseeds, most cases have been met with only limited success [17-27]. Most of the enzymes responsible for production of the common, and many of the unusual, FA have been identified. However, not all the factors that regulate FA synthesis or TAG accumulation are known and questions remain about the sequence of reactions in the incorporation of FA into membrane and storage lipids. A greater knowledge of the pathway of membrane lipid and TAG synthesis will allow us to design better strategies for engineering oilseeds in the future. This thesis will address the sequence of reactions that are involved in incorporation of FA into extra-plastidic membrane and storage lipids. 1.1 Background overview Assembly of most membrane or storage glycerolipids involves four basic steps: 1) FA synthesis; 2) esterification of FA to a glycerol based backbone; 3) addition of the sn-3 headgroup or FA; 4) desaturation of acyl chains and/or transfer of modified acyl groups to other lipids. Although this process always begins with fatty acid synthesis, the order of the remaining three steps may vary. In plants these activities may be separated within a particular membrane system or duplicated in different compartments, and thus may involve multiple pools of substrates and trafficking between compartments. Additionally, some activities (in particular desaturation) may be more active with a specific glycerolipid substrate and thus require deacylation of the lipid after desaturation and further utilization of the modified FA to make other lipids. An overview of the current state of knowledge on plant membrane and storage glycerolipid synthesis is presented as a background to the research in this dissertation and to set the stage for the evidence that is presented in later chapters for changes in the current view of FA flux through central plant lipid metabolism. 1.2 Fatty acid synthesis and desaturation In plants the major site of de novo fatty acid synthesis (FAS) for production of membrane and storage lipids is the plastid. This is significantly different from animals and fungi where FAS is primarily cytosolic. The non- cytosolic nature of plant FAS was first determined by subcellular-Iocalization of acyl carrier protein (ACP), an essential component of FAS, to the chloroplasts of spinach [28]. Mitochondria can also synthesize fatty acids, however it appears that mitochondrial FAS is mostly for producing octanoate (8 C) for lipoic acid synthesis [29]. Additionally, stable isotope labeling of Brassica napus embryos demonstrated onlya plastidic labeling signature in total lipids [30], indicating that any mitochondrial contribution to bulk lipids would be very minor compared to the plastidic FAS. It is possible that mitochondrial FAS may play a more dominant role in lipid synthesis in specific cell types such as epidermal cells involved in wax and cutin synthesis where modified acyl chains are predominant [31, 32]. Since this work deals with membrane and storage lipids from leaves, total seedings and developing embryos, FAS will refer to the major plastidic activity. The first committed step in FAS is the production of malonyl-CoA by acetyI-CoA carboxylase (ACCase) [33] in the plastid. The three carbon carboxylic acid of malonyl-CoA is transferred to ACP and all further reactions take place with the growing acyl chain on ACP. De novo synthesis of a fully reduced 16 or 18 carbon fatty acid takes place by the sequential condensation of the malonyl- ACP with a growing fatty acyl—ACP. Release of CO; drives each condensation reaction, elongating the acyl chain two carbons at a time. The four steps to make the initial four carbon fully reduced acyl-ACP are [34]: First, condensation of the initial malonyl-ACP wtih acetyl-CoA to produce the four carbon 3-ketobutyryl— ACP by 3-ketoacyI-ACP synthase lll (KAS lll) releasing C02. Second, NADPH- dependent reduction by 3-ketoacyI-ACP reductase produces 3-hydroxybutyryl- ACP. Third, removal of water by 3-hydroxyacyl-ACP dehydratase produces a trans-Az-butenoyl-ACP. Finally, reduction of the double bond by 2,3-trans-enoyl- ACP reductase utilizing NADPH or NADH and producing the saturated four carbon chain, butyryl-ACP. Additional condensations with malonyl-ACP are catalyzed by different enzymes than the original condensation with acetyl-CoA. During elongation from butyryl-ACP (4C) to palmitoyl-ACP (16C) condensation is provided by 3—ketoacyl-ACP synthase l (KAS I). PalmltoyI-ACP can be condensed once more with malonyl-ACP by 3-ketoacyl—ACP synthase ll (KAS II) and reduced to produce the 180 saturated stearoyl-ACP. In this manner saturated acyl chains are elongated two carbons at a time and after each condensation the reduction, dehydration and further reduction of the keto-, hydroxyl-, and enoyl- ACPs are catalyzed by the enzymes listed above. 16:0- and 18:0-ACP are the major products of FAS however, the major newly synthesized acyl chain that is incorporated into glycerolipids is the mono- unsaturated oleic acid (18:1). Oleoyl-ACP is produced by the soluble stroma localized stearoyl-ACP desaturase by inserting a double bond at the A 9 position [35]. Thus de novo FAS and initial desaturation within the plastid of pea leaves, for example, produces 18:1 (~75%), 16:0 (~20%), and 18:0 (~5%) acyl chains for further lipid metabolism. Plant membrane and storage glycerolipids mostly contain acyl chains other than the initial products of plastid FAS. Leaf membrane lipid acyl groups typically contain PUFA of either 16- or 18- carbons in length; most commonly 18:2, 18:3 and 16:3. Production of PUFA is accomplished after removal of the de novo synthesized acyl groups from ACP and incorporation into membrane lipids within the ER or plastid envelop membranes [36, 37]. Utilization of membrane lipid acyl groups for desaturation in plants differs from that of animals, yeast and fungi where the desaturases act on acyl Coenzyme A (CoA) thioesters as substrates [38, 39]. Additionally, longer chain fatty acids (>180) can accumulate in some minor membrane lipids such as phosphatidylserine (PS) or storage TAG. Extra-plastidial elongation of acyl groups beyond 18 carbons may utilize glycerolipids as a primer to donate the acyl substrate [40]. The donated acyl chain is elongated two carbons at a time with the steps analogous to plastid FAS except that the intermediates of elongation are bound to CoA instead of ACP [41]. 1.2.1 AcetyI-CoA carboxylase: Major FAS regulatory point Production of malonyl-CoA by acetyl-CoA carboxylase (ACCase) is the first committed step in de novo FAS [33] and regulation of ACCase activity ultimately determines the flux through all of FAS [42]. All plants except the grass family Gramineae contain two forms of ACCase, a homomeric and heteromeric form [43]. The homomeric form produces the cytosolic malonyl—CoA pool, while the heteromeric form is plastid localized and produces the malonyl-CoA pool for de novo FAS. A second homomeric form that is plastid localized has been found in Brassica napus, however its proportional contribution to FAS is unknown [44]. Since in most plants de novo FAS utilizes the products of the plastid localized heteromeric ACCase, the discussion will be limited to this isoform and will be referred to as just ACCase. There are four subunits that make up ACCase which catalyzes two half reactions. In the first half reaction the biotin carboxylase (BC) catalyzes the ATP and H005 dependent carboxylation of a biotin prosthetic group attached to the biotin carboxyl carrier protein (BCCP). The activated carboxy-biotin is transferred to the active site of the carboxyltransferase (CT) made up of q-CT and B—CT subunits. CT completes the second half reaction by carboxylation of acetyI-CoA producing malonyl-CoA [45]. The B—CT subunit is encoded by the plastid genome [46] and requires RNA editing for functionality [47], while the other three subunits are all encoded by the nuclear genome and each contains a cleavable plastid targeting peptide [48-50]. All four subunits are coordinately expressed and assembled together in the plastid [51, 52]. It is believed that expression of the [3— CT subunit by the plastid coordinates the expression of the other three subunits [53]. The ACCase complex is mostly stroma localized but may be partially associated to the inner envelope membrane in rapidly expanding leaves [54]. ACCase activity can be regulated developmentally or enzymatically. All four subunits of ACCase are developmentally regulated with the highest expression in the organs with the greatest demand for FA synthesis, for example expanding leaves and developing oil seeds [52, 55]. Increased demand for long chain FA induces expression of all subunits, as demonstrated in Brassica napus seeds expressing a lauroyl—acyl carrier protein thioesterase which terminated FAS at 12 carbons starving the cell of long chain FA [56]. The enzymatic activity of ACCase can be activated by the substrate acetyl-CoA [57], and downstream products of FAS such as long chain acyl-CoAs can cause feedback inhibition [58]. Light is a strong regulator of FAS with over five times more FAS in the light than the dark [59]. Some of lights effect on FAS may be due to increases in ' ACCase enzymatic activity without a change in gene expression [46, 57]. Light causes the pH and Mg2+ concentration of the stroma to change toward the enzymatic optima of ACCase [60]. A disulfide bond between the o-CT and B-CT is reduced in the light which increases activity of ACCase. It is believed that the disulfide reduction is mediated through a thioredoxin linked pathway [61-63]. The B—CT subunit of ACCase is also reversibly phosphorylated in the light which has been shown to increase enzymatic activity [64]. Since malonyl-CoA production commits carbon to FAS, this step has been the subject of multiple attempts to increase FAS and thus TAG in oilseeds. Over expression of individual nuclear encoded ACCase subunits has not increased ACCase activity or FAS [65, 66]. Increased expression of the plastid operon that includes B—CT increased leaf longevity and doubled seed yield in tobacco [53]. However, it is unknown if these effects are due to an increase in ACCase alone or if the other gene products in the operon contributed to this phenotype. Targeting the Arabidopsis homomeric ACCase to the plastid of Brassica napus seeds increased in vitro ACCase activity, and caused a 5% increase in seed TAG [67]. The strong control that ACCase has on the flux of carbon through FAS indicates the importance of further understanding its regulation in relation to lipid metabolism. 1.3 Esterification of fatty acids to glycerol-3—phosphate: de novo glycerolipid synthesis. Glycerolipids (membrane and storage) contain a glycerol moiety as a central backbone in which fatty acyl groups and/or polar headgroups are attached. Polar lipids that make up membranes contain two acyl groups esterified to the sn-1 and sn-2 positions of glycerol reserving the sn—3 position for a polar headgroup. TAG contains a third acyl group esterified to the sn-3 position. De novo glycerolipid synthesis starts with two sequential acylations of glycerol-3- phosphate (G3P). The first acylation by G3P acyltransferase (GPAT) produces 1- acyl-2-lyso-glycerol-3—phosphate, commonly known as lyso-phosphatidic acid (LPA). The second acylation by lyso-phosphatidic acid acyltransferase (LPAAT) produces 1,2-diacylglycerol-3-phosphate, referred to as phosphatidic acid (PA). These two acylation steps were first elucidated from microsomes of guinea pig liver, with the energy of the reaction provided by activated acyl-CoA [68]. Later it was shown by Eugene Kennedy in rat liver extracts that the diacylglycerol (DAG) backbone produced from de novo glycerolipid synthesis and dephosphorylation of PA could be used to make the common membrane lipids phosphatidylcholine (PC) and phosphatidylethanolamine (PE) through enzymatic reaction with nucleotide activated headgroups, CDP-choline and CDP-ethanolamine, respectively [69]. Kennedy also demonstrated in chicken liver microsomes that DAG could be acylated with acyl-CoA to produce TAG [70]. The fundamental work by Eugene Kennedy in understanding aspects of glycerolipid synthesis has led to the acylations of G3P en route to TAG synthesis in plants to be commonly referred to as the Kennedy pathway. 1.3.1 The two pathways for de novo glycerolipid synthesis in plants Plants contain two pathways for de novo glycerolipid synthesis (Figure 1- 1) involving the GPAT and LPAAT activities producing the common intermediate PA. 1) The the ER localized “eukaryotic pathway" and, 2) the plastid stroma and envelope localized “prokaryotic pathway". The prokaryotic pathway was named because the plastid LPAAT specifically esterifies a 16 carbon FA to the sn-2 position of the glycerol backbone similar to that of cyanobacteria. In contrast, the eukaryotic pathway acyltransferases localize 16 carbon FA, if present at all, to the sn-1 position similar to animals [71]. The prokaryotic pathway utilizes the activated acyl-ACPs from FAS and stearoyl-ACP desaturase, to make the initial PA molecular species 18:1/16:0 (molecular species refer the combination of FA esterified within a particular lipid). Prokaryotic PA is used to synthesize phosphatidylglycerol (PG) in all plants and a portion of the plastidic glycolipids in some plants [34]. Plastid localized Eukagotic pathway I" PI,PS p5 Dammit,“ G3P Lyso-PA ? PA 5]) 1 2 q . . a s' Endoplasmic reticulum 18:3 plants ~95% of FA 16:3 plants ~62% of FA , PA or DAG transferred Into plastld for eukaryotic glycollpld synthesis Note: traditional textbook model does not include acyl editing Plastid 1 esp Lyso-PA PA FAS 162/(2%:321- E —) {jg—D? Acetyl-CoA m _’ 13:1 13:1 160 \ I 18:3 plants ~5% of FA 16:3 plants ~38% of FA Figure 1-1 Traditional textbook two pathway model of glycerolipid synthesis The two pathway model of leaf glycerolipid synthesis as described in the text. The model shows general localizationof prokaryotic pathway to the membranes of the plastid and eukaryotic pathway to the endoplasmic reticulum membranes. Estimations of the amount of fatty acids that flow through each pathway are based on the fact that 18:3 plants only make PG by the prokaryotic pathway [71], and the 16:3 plant Arabidopsis leaf measurements [72]. Abbreviations: glycerol- 3-phosphate, G3P; lyso-phosphatidic acid, Iyso-PA; phosphatidic acid, PA, diacylglycerol, DAG; phosphatidylcholine, PC; phosphatidylinositol, PI; phosphatidylserine, PS, phosphatidylethanolamine, PE; phosphatidylglycerol, PG; monogalactosyldiacylglycerol, MGDG; digalactosyldiacylglycerol, DGDG; suphoquinovosyldiacylglycerol, SQDG; fatty acid, FA; coenzyme A, CoA; acyl carrier protein, ACP; fatty acid synthesis, FAS. desaturases act on the initial 18:1/16:0 molecular species of PG and glycolipids to produce the highly unsaturated molecular species (eg. 18:3/16:3) that are characteristic of chloroplasts [37, 73]. Glycerolipid synthesis by the eukaryotic pathway is generally assumed to utilize newly synthesized acyl groups exported from the plastid. Acyl chains are hydrolyzed [74] from ACP in the stroma by a thioesterase specific for oleoyl or saturated acyl—ACP [75, 76] and reactivated to acyl-CoAs by long chain acyl-CoA synthetase (LACS) in the outer chloroplast envelope [77, 78]. Experiments with isolated pea chloroplasts suggest a channeled mechanism to LACS rather than simple diffusion across the plastid membranes [79]. It is then assumed that 18:1 and saturated acyl-CoAs are available within the cyctosol for the ER localized eukaryotic pathway acyltransferases. Utilization of nascent exported 18:1 and 16:0 by the eukaryotic pathway predicts the initial PA molecular species to be 18:1/18:1 and 16:0/18:1. PA is rapidly converted to PC which is the major extraplastidial phospholipid in multi—cellular plants [80] and main site of acyl desaturation for production of all other molecular species [36, 37]. The diacylglycerol backbone of PE, phosphatidylserine (PS), phosphatidylinositol (PI) and storage TAG are also produced by the eukaryotic pathway. Additionally, the eukaryotic pathway can produce the “DAG moiety" for plastid glycolipid synthesis. New insights into the trafficking events by which the DAG moiety is transported into the plastid have been achieved recently [80, 81], and will be briefly discussed below. 12 The relative amounts of the eukaryotic or prokaryotic pathway that are used to incorporate the acyl chain products of FAS into glycerolipids differs in different types of plants and tissues within the same plant. The amount of prokaryotic pathway in all plants has been related to the relative amount of PA phosphatase activity in the plastid to produce DAG for glycolipid synthesis [73]. “18:3” plants, so called because the leaf galactolipids contain primarily 18:3 FA at the sn-2 position, contain very little or no PA phosphatase activity. In 18:3 plants only PG is produced from prokaryotic PA. Therefore, greater than 95% of the FAs produced in the plastid are exported to the ER for acylation to G3P by the eukaryotic pathway and all of the DAG moiety for plastidic glycolipids in 18:3 plants must be imported into the plastid from the eukaryotic pathway. “16:3" plants, so called because the leaf galactolipids contain 16:3 FA at the sn-2 position can dephosphorylate PA in the plastid to provide a prokaryotic pathway DAG moiety for glycolipid synthesis. The relative proportion of FAS products utilized by the prokaryotic or eukaryotic pathways varies among different 16:3 plants [82]. In the case of the 16:3 plant Arabidopsis, the flux of acyl chains into leaf glycerolipids is approximately 62% eukaryotic and 38% prokaryotic [72], with about half of the galactolipid DAG moieties contributed by the eukaryotic pathway. The eukaryotic pathway is the major pathway of glycerolipid synthesis in non-leaf and oilseed tissue [83], and it is the dominant pathway among angiosperms, of which 88% are 18:3 plants [82]. The two pathway scheme for de novo glycerolipid synthesis was first elucidated by radiotracer studies of plant tissue and isolated organelles [71, 84] and then confirmed genetically through analysis of pathway mutants [34, 85]. Critical experiments in the elucidation of the two pathways involved pulse—chase labeling of FAS with [14C]acetate or 1"’COz, and [3H]glycerol labeling of the backbone moiety of glycerolipids in leaf tissue of 18:3 and 16:3 plants [86, 87]. Over a period of hours it was evident in the 18:3 plant maize, that newly synthesized FA first accumulated in PC; then during the chase, the labeled FA and glycerol together were lost from PC and accumulated in MGDG. Thus the DAG moiety of eukaryotic synthesized PC was incorporated into a plastid lipid. Rapid synthesis of labeled PC and transfer to MGDG was also evident in spinach, a 16:3 plant, but to a lesser degree due to a portion of the MGDG synthesized by the prokaryotic pathway. Furthermore, isolated chloroplasts from spinach could produce galactolipids from in situ synthesized DAG, however PC was only produced from chloroplast-synthesized acyl-CoAs when microsomes were present [84], indicating the physical separation of the two G3P acylation pathways. The act1 mutant in the 16:3 plant Arabidopsis thaliana provided further evidence of the two independent pathways by blocking prokaryotic glycolipid synthesis by a severely reduced plastidial GPAT activity. This forced newly synthesized FA to enter the eukaryotic pathway for acylation to GSP. In the act1 mutant the eukaryotic pathway provided all of the DAG for glycolipid synthesis, essentially turning a 16:3 plant into a 18:3 plant [88]. 1.3.2 Transfer of eukaryotic synthesized DAG to the plastid for glycolipid synthesis 14 Glycolipids are synthesized from DAG in the plastid envelope. However, in 16:3 and 18:3 plants more than half and all, respectively, of the DAG moiety is produced by the extraplastidial eukaryotic pathway (Figure 1-1). This separation in glycerolipid synthesis activities has led to multiple hypotheses on the mechanism of lipid transfer between the ER and plastid. Pulse-chase labeling in maize demonstrated that the glycerol backbone and acyl chains of PC were a precursor to plastid MGDG synthesis. Careful analysis of radiolabeled glycerol and acyl chains in MGDG molecular species suggested that 18:2/18:2-PC was the donor molecule for the DAG moiety incorporated into MGDG [86]. DAG was hypothesized to be the transfer molecule when a cytosolic phospholipase D and a hypothesized PA phosphatase from cytosolic extracts were added to isolated pea chloroplasts inducing DAG accumulation in the chloroplast [89]. PA (possibly derived from PC) was suggested as the transport molecule when a component of an inner envelope ABC transporter, that could bind PA in vitro, was shown to disrupt plastid import of eukaryotic DAG when mutated [90]. Furthermore, such mutants accumulate PA with a eukaryotic fatty acid composition. An alternative hypothesis to the transfer of a complete DAG moiety originating from PC is the transfer of lyso-PC. In vitro experiments suggest that lyso-PC can spontaneously partition through an aqueous environment from isolated microsomes to isolated chloroplasts [91]. Additionally pulse chase labeling of leek seedlings suggested that only the sn-1 acyl-chain from PC was transferred to MGDG and DGDG [92, 93]. However the extremely long time points of these experiments (24 hr increments) make it difficult to interpret the 15 labeling data in relation to the demonstrated rapid transfer of PC to MDGD in other species [72, 86-88]. The transfer lipid may not have to traverse the hydrophilic cytosol at all, but may travel through an ER plastid connection site. Putative connection sites termed plastid associated membranes (PLAMs) containing ER enzymatic activities have been found associated with isolated chloroplasts [94]. Strong physical connections between the ER and chloroplasts have been demonstrated in vitro utilizing optical tweezers [95]. Connection sites may allow direct transfer of lipid molecules between the ER and plastid membrane systems as has been demonstrated for ER-mitochondria connection sites in rat liver and yeast [96, 97]. The connection site hypothesis, as opposed to the lyso-PC hypothesis, could also give an explanation of how glycolipids are exported from the plastid under times of phosphate stress [80, 98], and may also be involved in the export of newly synthesized FA to the eukaryotic pathway acyltransferases. 1.4 Long chain acyl-CoA pools for building glycerolipids Eukaryotic pathway acyl transferases utilize Coenzyme A (CoA) activated acyl groups to drive the esterification of FA to a glycerol backbone. The composition of the FA esterified to 00A (e.g. 16:0, 18:3, etc) and acyltransferase selectivity can affect the molecular species of lipids produced. Acyl-CoAs do not freely diffuse across membranes and transport may be an active process as in mammalian mitochondria [99]. Production of acyl-CoAs from FFA by LACS in 16 different subcellular compartments may generate multiple pools of acyl groups with different compositions within the cell for different acyltransferase reactions. LACS activity has been shown to take place in the ER [100], chloroplasts [77, 78], mitochondria [101], peroxisomes [102], and lipid bodies [103]. In Arabidopsis LACS is encoded by a family of nine genes [104] of which two are involved in activation of acyl groups for beta—oxidation in peroxisomes [105], at least two located at the plastid envelope are candidates for involvement in export of acyl groups from the plastid to the eukaryotic pathway [106], and one has been shown to be very important for cuticle development [107]. The widespread localization of LACS throughout the cell demonstrates that FFA are generated throughout the cell, almost certainly by lipase activities, and the reutilization of FA by acyltransferases requires the activation to acyl-CoAs. The multiple sites of acyl-CoA synthesis may also be a mechanism for controlling the acyl composition of multiple subpools of acyl-CoA for production of specific lipid molecular species, such as highly saturated or polyunsaturated TAG species. There is evidence that acyl-CoAs involved in lipid metabolism are not freely soluble or present in membrane bound pools but are bound to acyl-CoA binding proteins (ACBP) that deliver the activated acyl groups to acyltransferases. ACBP have been found in all four eukaryotic kingdoms and some bacteria [108]. ACBP are expressed in all plant organs [109]. High levels of ACBP accumulate in oilseeds during oil deposition [110], suggesting their role in TAG synthesis. Arabidopsis contains two ACBPs that are predicted as cytosolic [111], two ACBPs have been localized to the plasma membrane [112, 113], and one is targeted to the outside of the cell [114]. Much like the LACS proteins, the wide distribution of ACBP suggests multiple roles in lipid metabolism. RNAi knockdown of a membrane associated ACBP in Arabidopsis increased the 18:2/18:3 ratio of phospholipids [109]. Since acyl chains are desaturated while bound to PC the knockdown of ACBP probably did not affect the desaturation of 18:2. Reduced ACBP levels may affect the delivery of recycled acyl-CoA to acyltransferases after removal of desaturated FA from PC and reactivation to acyl-CoA. 1.5 Synthesis of eukaryotic membrane lipid classes Leaf glycerolipids are primarily made up of polar membrane lipids. Membrane lipid synthesis starts with PA from de novo glycerolipid synthesis (Figure 1-1) and follows one of two routes for acquiring the polar headgroup. (1) Nucleotide—activated head group dependent sn-3 esterification of DAG produced from PA by a phosphatase. Cytidine 5’-diphoshphate (CDP) activated choline or ethanolamine is used for PC and PE synthesis and uridine 5'-diphosphate (UDP) activated sugars are utilized in monogalactosyldiacylgylcerol (MGDG) and sulfolipid (SL) synthesis [34]. A second galactose can be added to MGDG by reaction of UDP-galactose with the 6-hydoxyl of the MGDG headgroup to produce digalactosyldiacylgycerol (DGDG) [115]. (2) PA is reacted with cytidine 5’-triphosphate to make CDP-DAG. The activated CDP-DAG is then reacted with serine or myo-inositol to produce PS or Pl, respectively. PG is synthesized by the reaction of CDP-DAG with G3P, producing first phosphatidylglycerol-phosphate, 18 and then PG by a phosphatase reaction [34]. Subcellular localization of the enzymes responsible for headgroup addition defines where each lipid is produced. PC, PE, PS, and PI are produced from DAG and PA in the ER [34] while MGDG, DGDG, SL and most of PG are produced within the plastid envelope membranes [73, 81]. 1.6 TAG synthesis The traditional view of TAG synthesis (aka Kennedy Pathway) starts with two consecutive acylations of G3P by the eukaryotic pathway producing PA (Figure 1-2). PA dephosphorylation to form DAG and addition of a third acyl group to the sn-3 position of the glycerol backbone by a diacylglycerol acyltransferase (DGAT), produces TAG and completes the Kennedy Pathway [116]. Since TAG is not a substrate for FA desaturation, the acyl-CoA pool that feeds TAG synthesis acyltransferases must contain all the acyl groups found in TAG. Alternatively, the DAG moiety for TAG synthesis can be produced from PC by the reverse reaction of CDP-choline:diacylglycerol cholinephosphotransferase (CPT) [16, 117, 118]. As with the Kennedy pathway final TAG synthesis can be through an acyl-CoA dependent DGAT. Additionally, an acyl-CoA independent mechanism can directly transfer an acyl group from the sn-2 position of PC to 19 pc~m C GaP.—-—> LPA PA-+ DAG TAG Kennedy pathway 7 DAG MAG Figure 1-2 Possible enzymatic reactions leading to TAG synthesis in oflseeds Reactions that may be involved in TAG synthesis as mentioned in the text. Substrates are as text, enzymes are enclosed in boxes. Acyl-CoA pools and how they are generated are not shown. Enzymes utilizing acyl-CoA as an additional substrate are rectangles all other enzymes are in ovals. The traditional Kennedy pathway is highlighted in gray. TAG, DAG and PC are bold because of their central importance in TAG synthesis. Substrates: G3P, glycerol-3-phosphate; LPA, Iyso-phosphatidic acid; PA, phosphatidic acid; DAG, diacylglycerol; PC, phophatidylcholine; LPC, lyso-phosphatidylcholine; MAG, monoacylglycerol; TAG, triacylglycerol. Enzymes: GPAT, G3P acyltransferase; LPAT, LPA acyltransferase; LPCAT, LPC acyltransferase; PDAT, phospholipid:diacylglycerol acyltransferase; DGAT, diacylglycerol acyltransferase; MAGAT, monoacylglycerol acyltransferase; DDTA, DAG/DAG trans-acylase; PLD, phospholipase D: PLC, phospholipase C; CPT, diacylglycerol:CDP-choline cholinephosphotransferase. 20 DAG producing TAG and Iyso-PC by a phospholipid:diacylglycerol transferase (PDAT) [119-121]. Although their role in oilseeds has been much less investigated a phospholipase mediated pathway to produce DAG for TAG synthesis can be proposed. DAG could be produced from PC by phospholipase C or by phospholipase D plus PA phosphatase, similar to what has been proposed for transfer of a eukaryotic DAG moiety from PC into the plastid for galactolipid synthesis [89, 90]. A DAG/DAG transacylase may also be involved in TAG synthesis [122] and could transfer FA from the sn-2 position of one DAG to the sn-3 of another, producing TAG and a sn-1-acyl—monoacylglycerol (MAG). Re-acylation of MAG by a monoacylglycerol-acyltransferase (MAGAT) would generate a new DAG for TAG synthesis. Despite the multiple pathways that could be involved in TAG synthesis as indicated by their in vitro enzymatic activities, only the DGAT reaction has been conclusively shown to be involved in TAG synthesis in vivo [123, 124]. Further knowledge of the actual enzymes and pathways that lead to TAG accumulation will allow better strategies for engineering of TAG quantity and/or FA composition. 1.6.1 Acyl transfer between lipids is required for a majority of TAG synthesis Synthesis of polyunsaturated TAG molecular species requires significant transfer of acyl chains between lipid species. Phosphatidylcholine is the major site of FA desaturation for membrane lipids and TAG [116], and thus the modified acyl groups need to be removed for incorporation into TAG (at least for 21 incorporation into the sn-3 position). Additionally, production of many very long chain (>18C) polyunsaturated FA requires desaturation on PC prior to elongation and incorporation into TAG [125]. Desaturated FA may be removed by lipases, reactivated to acyl-CoA and thus be utilized by the Kennedy pathway acyl transferases to acylate the sn-1, -2, or -3 position of TAG glycerol backbone. Alternatively, the reverse CPT mechanism of DAG formation is one possible method for incorporation of desaturated acyl chains into the sn-1lsn-2 positions TAG [16, 117, 118]. The relative amount of Kennedy pathway or reverse CPT used to produce TAG in oilseeds is unknown. It has been suggested from labeling of oilseed species that accumulate PUFA in TAG, that reverse CPT may be a mechanism for channeling desaturated FA from PC into TAG [126, 127]. Further utilization of a PDAT that specifically transfers PUFA from sn-2 PC into the sn-3 position of TAG [120] may complete a mechanism to specifically produce tri-PU FA molecular species of TAG. Many plants accumulate unusual FA in seed oil and these unusual FA are mostly excluded from membrane lipids even though most acyl chain modification takes place on the membrane lipid bound acyl chain [128]. An example is ricinolelic acid (12-hydroxyoctadec-Q-enoate) which is produced in castor beans (Ricinus communis) on PC [129, 130] and accumulates up to 90% of FA in seed oil but is mostly excluded from bulk PC. The exact mechanisms that exclude these unusual FA from membrane lipids and segregate them into TAGs are unknown but identification of unusual FA specific lipases [131] suggest that they may be selectively removed from PC and channeled into TAG through the 22 Kennedy pathway [17, 116]. In transgenic plants that produce unusual FA in a non native plant species the unusual FA tend accumulate in PC in addition to TAG suggesting that the mechanisms for selective removal of the unusual FA from membrane lipids and incorporation into TAG are not available in the transgenic plant [128]. The final step in TAG production may be provided by a DGAT that is specific for the unusual FA, as demonstrated by the DGAT2 enzymes of castor and tung that appear to preferentially insert ricinoleic and eleostearic acid, respectively into the sn-3 position of DAG [132, 133]. It is well documented that in oilseeds it is very important for the plant to be able to transfer unusual modified FA from the site of synthesis in PC to the TAG molecule where the modified FA accumulates. The transfer mechanisms and enzymes could be the same or very similar to those for common desaturated FA. Much of the oilseed data suggests that these deacylation-reacylation enzymatic activities are specific for channeling unusual modified FA into TAG. However, a large number of genes in the Arabidopsis genome are annotated as acyltransferases and phospholipases with unknown metabolic function [134], therefore it is possible that similar deacylation-reacylation reactions with common FA may have a more central role in lipid metabolism. 1.7 Suggestive evidence for acyl editing in plants Maintenance and fine tuning of the fatty acid composition of phospholipid membranes appears to be essential in all organisms. Acyl editing also termed 23 "remodeling" or “retailoring” is one mechanism to achieve this. Acyl editing involves the deacylation and reacylation of glycerolipids but does not by itself result in net lipid synthesis. Acyl editing was first recognized as a significant pathway of lipid metabolism in animal tissues [135]. In mammalian brain tissue in _ which phospholipid metabolism consumes 20% of cellular ATP [136], a quarter of that is used for the deacylation-reacylation of phospholipids [137]. Evidence for acyl editing has also been found in prokaryotes and yeast [138, 139]. In plants, acyl editing has not been previously studied as an integral part of central lipid metabolism but there is ample suggestive evidence that it is involved. The acyl editing reaction for phospholipids such as PC can take place through two general mechanisms. The first mechanism involves deacylation by exchange of acyl groups from PC directly onto CoA producing acyl-CoA and lyso-PC (CoAzPC acyl exchange). Deacylation by CoAzPC acyl exchange has been demonstrated in developing safflower cotyledon microsomes, and was attributed to the reverse reaction of Iysoephosphatidylcholine acyltransferase (LPCAT) [140]. LPCAT activity has been extensively reported in isolated plant microsomes [141-144] and chloroplasts [91, 94, 145]. When reported, the specificity of LPCAT activity is mostly at the sn-2 position of PC. The identification of LPCAT activities in plant organelles presents the possibility for CoArPC acyl exchange as an acyl-editing mechanism but does not indicate how prevalent the reverse LPCAT activity is in vivo. The second possible mechanism for acyl editing involves deacylation by hydrolytic cleavage of PC by a phospholipase generating a FFA and a Lyso-PC. To complete the acyl editing 24 cycle the FA must be activated to acyl-CoA by LACS and PC is regenerated by LPCAT activity. It is possible that both acyl chains could be removed by lipases generating a glycerophosphocholine that could be reacylated to PC. However glycerophosphocholine dependent acyl transferases have not been reported in plants. Many plant phospholipases have been identified that could be involved in an acyl editing cycle, of those the majority are PLAz, which remove acyl chains from the sn-2 position [146]. Many of these PLAzS are specific for unusual FA produced in oilseeds [131]. Removal of acyl chains from the sn-1 position has been less prevalent in the plant literature but there are a few examples [147, 148]. Just as for the CoAzPC acyl exchange mechanism, evidence of lipase enzymatic activity that could be involved in acyl editing is not evidence of an actual acyl editing function in vivo. More substantial evidence that acyl editing is involved in central lipid metabolism has come from multiple isotopic labeling studies. Isolated chloroplasts from pea plants could incorporate newly synthesized [14C]acetate labeled FA into PC through a channeled pool of acyl-CoA [79]. Since chloroplasts do not contain the eukaryotic pathway for de novo PC synthesis this result suggests an acyl editing mechanism that can incorporate newly synthesized FA into previously synthesized PC in the outer chloroplast membrane or an attached PLAM region. Additionally, oxygen-18 labeling of the carbonyl oxygen of newly synthesized FA demonstrated that the entire pool of PC acyl chains in spinach leaves is under a substantial flux of hydrolytic deacylation and reacylation [74]. This hydrolytic turnover of PC acyl chains occurs at a rate similar to the overall 25 rate of FAS, and therefore may play a major role in phospholipid metabolism. Additional insightful evidence on acyl editing came from careful comparison of prokaryotic MGDG and eukaryotic PC molecular species after labeling leaf disks of the 16:3 plant Brassica napus with [“Clcarbon dioxide. MGDG contained the ' predicted prokaryotic pathway initial molecular species of dual labeled 18:1/16:0. The eukaryotic pathway predicts the major initial PC molecular species to be dual labeled 18:1/18:1. However carbon dioxide pulse-chase labeling produced PC with a large amount of scrambling of labeled FA with unlabeled FA. The authors concluded that the acyl chains of the PC pool are under constant exchange directly after de novo PC synthesis and during their prolonged 48 hr chase period [149]. These three lines of evidence suggest that the plant PC pool is under a deacylation-reacylation flux, so that the acyl groups esterified to G3P by the eukaryotic pathway and incorporated into PC are not a permanent fixture of PC. However, how acyl editing is tied in with central membrane and storage lipid synthesis is unknown. 1.8 The in vivo labeling approach Modern biological science benefits from various approaches to study the metabolic pathways of life, each with its own strengths and weaknesses. The pathways of plant lipid metabolism were originally elucidated by two approaches: 1) Discovery of enzymatic activities through in vitro assays of plant extracts, 2) step by step pathway development through in vivo labeling. Through these 26 approaches the two pathway hypothesis for de novo glycerolipid synthesis was discovered over 25 years ago and it is still valid today [71]. Since that time the advent of genetic and molecular biology approaches has unleashed very powerful tools for discovery of individual genes/enzymes involved in metabolic pathways through mutational analysis. Additionally, the ability to clone and overexpress genes has greatly increased the ability to study the effects of changing one enzyme on metabolism in vivo and to purify enzymes for in vitro assays. However, the redundancy of genes and metabolic pathways, the lethality of certain gene knockouts, and the ability for plant metabolism to adapt to some changes, can mean that key points of metabolism may not be revealed by genetic or molecular biology techniques alone. Additionally, in vitro assays identify what reactions are present but usually cannot determine their relative contributions to metabolic fluxes in vivo. Recently, metabolic profiling techniques have allowed us to study differences in metabolism due to different growth, environmental or genetic changes. However, analysis of bulk pools of metabolites may not reveal subcellular compartmentation, metabolic cycles or rapid conversions of one metabolite to another. In vivo labeling approaches are still the only way to study the kinetics of metabolites moving through pathways in vivo and the relative contributions of parallel or convergent pathways. Recently the in vivo labeling approaches have been reinvigorated with the use of stable isotope labeling in conjunction with mass spectrometry, nuclear magnetic resonance (NMR) and sophisticated software for analysis. Stable isotope labeling over prolonged periods and 27 analysis of the composition of the label in metabolic end products (eg. membrane and storage lipids, starch and cell walls, and proteins) allows a steady-state determination of the relative fluxes through different known metabolic pathways. Steady-state labeling that does not fit known metabolic pathways can suggest areas of uncertainty in the pathway that need to be revised [150]. The sensitivity of mass spectrometry and NMR often limit stable isotope labeling to prolonged periods and analysis of steady-state metabolism. In vivo labeling with radioisotopes is usually the most sensitive method for detecting small amounts of labeled compounds as they move through metabolic pathways [151]. Rapid dynamic labeling with radioisotopes, as opposed to long term steady-state labeling, involves introducing a labeled compound to the plant or tissue and rapidly stopping metabolism to determine how the labeled compound has been metabolized. Multiple labelings of different time periods allows us to generate kinetic curves for accumulation of label overtime in different metabolic products. Metabolic pathway precursor-product relationships can then be determined by the relative time it takes for each metabolite to accumulate label linearly, which corresponds with the complete labeling of all precursor pools. The research in this dissertation utilizes in vivo dynamic radiolabeling to investigate if acyl editing is involved in central lipid metabolism. The textbook eukaryotic pathway hypothesizes that newly synthesized FA are exported from the plastid, activated to acyl-CoA, and ester'rfed to G3P producing initial PA molecular species of 18:1/18:1 and 16:0/18:1. The PA is then converted to PC for further desaturation of the nascent FA. Additionally, the initial PA may be used by 28 the Kennedy pathway to produce TAG in oilseed tissue. However, evidence that nascent FA are channeled directly into PC in isolated chloroplast [79] through a possible acyl editing mechanism suggest that there may be alternative mechanisms at work for incorporation of newly synthesized FA into membrane lipids. Additionally, some unusual FA produced in the plastid that are not further modified on PC are transiently incorporated into PC prior to accumulation in TAG [152, 153], suggesting there may also be an alternative to direct plastid export of acyl groups into the Kennedy pathway for TAG synthesis. The goals of this dissertation have been to test the glycerolipid synthesis mechanisms in vivo. The next two chapters utilize in vivo kinetic labeling techniques to investigate the initial events by which newly synthesized FA and the glycerol backbone are incorporated into eukaryotic membrane and storage lipids. The results in pea leaves and in developing soybean embryos demonstrate that the initial flux of FA into glycerolipids occurs through PC acyl editing and that this pathway should be considered a major part of eukaryotic glycerolipid synthesis. 29 2 INCORPORATION OF NEWLY-SYNTHESIZED FATTY ACIDS INTO CYTOSOLIC GLYCEROLIPIDS IN PEA LEAVES OCCURS VIA ACYL EDITING Published in: Journal of Biological Chemistry Vol. 282, NO. 43, pp. 31206-31216, October 26, 2007 Philip D. Bates‘, John B. Ohlrogge’, Mike Pollard’ From Department of Biochemistry and Molecular Biology1 and Department of Plant Biologyz, Michigan State University, East Lansing, MI 48824. 30 2.1 Abstract In expanding pea leaves over 95% of fatty acids (FA) synthesized in the plastid are exported for assembly of eukaryotic glycerolipids. It is often assumed that the major products of plastid FA synthesis (18:1 and 16:0) are first incorporated into 16:0/18:1 and 18:1/18:1 molecular species of phosphatidic acid (PA) which are then converted to phosphatidylcholine (PC), the major eukaryotic phospholipid and site of acyl desaturation. However, by labeling lipids of pea leaves with [”C]acetate, [“C]glycerol and [”C]carbon dioxide we demonstrate that acyl editing is an integral component of eukaryotic glycerolipid synthesis. First, no precursor-product relationship between PA and PC [“C]acyl chains was observed at very early time points. Second, analysis of PC molecular species at these early time points showed that >90% of newly synthesized ["C]18:1 and [“C]16:0 acyl groups were incorporated into PC alongside a previously synthesized unlabeled acyl group (18:2, 18:3 or 16:0). And third, [“C]glycerol labeling produced PC molecular species highly enriched with 18:2, 18:3 and 16:0 FA, and not 18:1, the major product of plastid fatty acid synthesis. In conclusion, we propose that most newly synthesized acyl groups are not immediately utilized for PA synthesis, but instead are incorporated directly into PC through an acyl editing mechanism that operates at both sn-1 and sn-2 positions. Additionally, the acyl groups removed by acyl-editing are largely used for the net synthesis of PC through glycerol-3-phosphate acylation. 31 2.2 Introduction In plants acyl carrier protein (ACP)-dependent de novo fatty acid synthesis (FAS) is restricted to organelles [28]. Essentially all acyl chains for membrane and storage lipid synthesis are produced in the plastid [30, 34]. The initial products of FAS, 16- and 18-carbon fatty acids (FAs) esterified to ACP, are incorporated into glycerolipids by one of two routes: (i) acyl-ACP is used directly by acyltransferases of the “prokaryotic” pathway within the plastid, or (ii) the acyl- ACP thioester bond is hydrolyzed during acyl export from the plastid prior to FA reactivation and incorporation into glycerolipids by acyltransferases of the “eukaryotic” pathway outside the plastid. This two pathway mechanism for glycerolipid synthesis was first elucidated by radio-tracer studies [71], and was confirmed genetically by analysis of Arabidopsis mutants [34, 154]. The proportions of nascent FA (i.e. immediate products of FAS) incorporated into the eukaryotic and prokaryotic pathways vary widely among plants and different tissues within the same plant. The eukaryotic pathway predominates in non- photosynthetic tissues of all higher plants. In 18:3 plants, so called because they accumulate predominantly 18:3 and not 16:3 in their leaf galactolipids, 95% of the FA produced in the plastid is exported for assembly into glycerolipids by the eukaryotic pathway. Only phosphatidyl-glycerol (PG) is synthesized directly in the plastid from acyl-ACPs. This is in contrast to leaves of 16:3 plants which can also synthesize glycolipids by the prokaryotic pathway. Consequently, only about 60% of FA produced in the chloroplast by 16:3 plants is exported to the eukaryotic 32 pathway [72]. 18:3 plants, which include pea, make up about 88% of angiosperrn species [82]. It is generally assumed that the major exported products of chloroplast FAS, namely 18:1 and 16:0 FA, are transferred to the outer envelope of the plastid where they are activated to acyl-CoAs by a long chain acyl-CoA synthetase [77, 78]. The eukaryotic pathway for de novo PC synthesis then utilizes this pool of newly synthesized acyl-CoAs for sequential sn-1 and sn-2 acylations of glycerol-3-phosphate to produce 18:1/18:1 and 16:0/18:1 molecular species of phosphatidic acid (PA). PA is rapidly converted to phosphatidylcholine (PC) by the action of PA phosphatase and CDP-choline:1,2-diacyl-sn-glycerol cholinephospho-transferase [34]. Desaturation of 18:1 to 18:2 and then 18:3 on PC produces the abundant polyunsaturated molecular species of PC [36, 37]. However, several lines of evidence suggest that this model may be inadequate and needs to take account of acyl editing. We define acyl editing, often also termed “remodeling” or “re-tailoring," as any process that exchanges acyl groups between polar lipids but which does not by itself result in the net synthesis of the polar lipids. Acyl editing has long been considered an important facet of phospholipid metabolism [135]. Acyl editing relevant to this work can occur through two mechanisms. In plants acyl editing via a CoAzPC acyl exchange mechanism was demonstrated in microsomes isolated from developing seeds and was attributed to a reverse reaction of lyso-phosphatidylcholine acyl- transferase (LPCAT) [140]. LPCAT activity has also been described in isolated chloroplasts [91, 94, 145] as well as microsomes [141-144]. Thus LPCAT allows 33 for a mechanism for acyl editing, although the in vitm results do not indicate how prevalent the reaction might be in vivo. In this context, isolated pea chloroplasts incubated with [“Clacetate immediately label PC with newly synthesized FA through a channeled pool of acyl-CoA [79]. As chloroplasts cannot synthesize PC de novo, this indicated a functional mechanism for PC synthesis with nascent FA through an acyl editing mechanism. A second mechanism for acyl editing involves hydrolysis of the phospholipid, such as PC to lyso-PC or even to glycerolphosphorylcholine (GPC), activation of the released free fatty acid and its re—utilization for phospholipid synthesis from lyso-PC or GPC. Based on 1‘30 labeling there is some indication that acyl-chains esterified to bulk cellular PC are under a constant flux of acyl editing that proceeds wholly or in part through a hydrolytic deacylation- reacylation cycle [74]. The most direct line of evidence of acyl editing in plants comes from a careful analysis of the molecular species of monogalactosyl-diacylglycerol and PC after labeling leaf disks of the 16:3 plant Brassica napus with carbon dioxide [149]. [“ClCarbon dioxide labeling produced initial acyl-labeled species as expected for prokaryotic monogalactosyldiacylglycerol, namely dual labeled 16:0/18:1. By contrast for PC a high degree of scrambling between labeled and unlabeled acyl chains was noted. The authors concluded that there was continuous exchange of acyl groups between all molecular species of PC immediately after labeling and during the prolonged pulse-chase period of 48 hours. In this paper we augment and extend these important observations and conclusions in several ways: 34 (1) We perform rapid kinetic studies to more carefully address PC labeling and that of its precursors, PA and 1,2-diacyl-sn-glycerol (DAG). We address whether the initial incorporation of nascent fatty acids occurs via acyl editing, or if there is a rapid incorporation by de novo PC synthesis via glycerol-3-P (G3P) and DAG, which is followed by rapid acyl editing of PC. (2) We perform both total molecular species and stereochemical analyses on acyl-labeled PC using in vivo experiments with expanding pea leaves and seedlings. Pea is an 18:3 plant, so this complements the analysis done previously with the 16:3 plant B. napus [149]. (3) We track molecular species of PC labeled in the glycerol backbone. When combined with the analysis of acyl labeling, this allows us to propose that sn- 1 acyl editing is as important a component as sn-2 acyl editing. (4) As a control, we perform rapid labeling experiments in planta using carbon dioxide. The results parallel those obtained with excised tissue assays, indicating that there are no wound responses that compromise the metabolic conclusions obtained from excised tissue experiments. From these studies weanalyze possible models by which newly synthesized FA are incorporated into eukaryotic lipids. The data presented in this paper and from other studies do not allow us to unambiguously define one particular model, but do narrow down the possible routes in which nascent FA are incorporated into eukaryotic glycerolipids and suggest future directions to re-examine this important yet poorly understood area of plant lipid metabolism. 35 2.3 Experimental procedures Plant Materials—All experiments were performed with Garden pea (Pisum sativum L. cv Little marvel) grown in soil/perlite/vermiculite (1 :1 :1) mixture at 22- 25 °C under a day/night 8/16-h photoperiod of white light at 185—210 pmol m”2 3“. Leaves, or in some cases, whole seedlings were harvested 8 days after sowing. RadiochemicaIs—[1-‘4C]Acetic acid, sodium salt (specific activity 50 mCi/mmol), [“C(U)]glycerol (specific activity 150 mCi/mmol), and [“Clsodium bicarbonate (specific activity 50 mCi/mmol) were from American Radiolabeled Chemicals, Inc. (St. Louis, MO). Leaf [1 4C]Acetate or f4CJGlyceml Labeling —'Leaf labeling experiments used ~ 0.3 g fresh weight of pea leaf strips per assay, incubated in the light (180- 220 pmol m'2 s") at 22-24 °C with reciprocal shaking in 5 ml media containing 20 mM MES pH 5.5, 0.1x MS salts, and 0.01% Tween-20. Cut leaves were placed in media without radioisotope and pre-incubated in ambient light for 5 min. Labeling was started by the addition of radioactive substrate and strong illumination. The reaction was quenched by transfer of leaves into isopropanol at 80 °C for 10 min. For labeling of acyl groups 250 uCi of [1-“C]acetic acid (1.0 mM) was used per replicate. For [“Clglycerol labeling of PC molecular species each incubation contained 47 pM [”C(U)]glycerol (39 110i). For [“Clglycerol time course the labeling media contained 25 pM [“C(U)]glycerol (18 pCi) and the media was reused for consecutive 1, 3, 6 and 9 min time points. In vivo labeling with excised plant tissue can produce considerable variance between samples due to 36 differences in development and in uptake of substrate. To minimize such variance each data point for total incorporation into lipids was normalized against the trend line for all time points to allow improved kinetic plots (Figures 2-1 and 2- 3). Seedling [14C]Carbon Dioxide Labeling— Assays were conducted with 8- day potted pea seedlings in a closed 9.5 L glass desiccator under 250 umol m'2 s'1white light. [“C]COz was released by injection of 1 ml H2804 through the desiccator lid into a vial containing aqueous [“C]NaHC03. The head space of the vial was quickly flushed with 30 ml air and the desiccator sealed for the desired time. Labeling was stopped by removal of the shoots at the base of the first leaves and quenching in 80 °C isopropanol. The assay used ~ 700 uCi substrate and ~ 40 seedlings. The seedlings from each labeling were split into two replicate samples (~ 20 seedlings each) for analysis. Although each labeling was nominally for 5 min, about 5 min was required to remove all the seedlings and quench, so the assay is reported as of 5-10 minutes duration. General Methods—Lipids were extracted from hot isopropanol quenched tissue with hexane/isopropanol [155] after homogenization with a mortar and pestle. Chlorophyll was determined spectrophotometrically at 652 nm in 20% aqueous acetone [156]. Radioactivity in the total lipid samples, eluted lipids or organic and aqueous phases recovered from transmethylation was quantified by liquid scintillation counting (Beckman Instrument lnc., Fullerton, CA), while radioactivity on TLC plates was visualized and quantified by electronic radiography (Packard Instrument Co., Meriden, CT). AgN03-TLC plates were 37 prepared by impregnating Partisil® K6 silica gel 60 A TLC plates (Whatman, Maidstone, UK) with 10% AgN03 in acetonitrile (wlv), drying in air and activating at 110 °C for 5 min. Fatty acid methyl esters (FAMEs) were quantified by gas chromatography (GC) using a flame ionization detection and a DB-23 capillary column (30 m length x 0.25 mm inner diameter, 0.25 p film thickness; J&W). Lipid Class Analysis—For kinetic analyses polar lipids were separated on a K6 silica TLC plates impregnated with 0.15 M ammonium sulfate and heat : activated (110 °C for 3 hr) prior to lipid loading. Plates were developed in acetone/toluene/water (91 :30:8, v/vlv) [157]. DAG was analyzed by first acetylating an aliquot of the total lipids in acetic anhydride/pyridine, (3:2, v/v) and then separating the 1,2-diacyl-3—acetyl-glycerols by silica TLC developed in hexane/diethyl ether/acetic acid (50:50:1, vlv/v). PA was isolated by a first preparative silica TLC separation using a K6 plate developed in chloroform/methanollwater (6522524, vlv/v), and further purified on a second K6 plate developed in chlorofonnlacetone/methanol/acetic acid/water (10:4:2:2:1, v/vlvlv/v). For preparative TLC all solvents contained 0.01% butylated hydroxyl- toluene antioxidant. Lipids were eluted from TLC silica with chloroform/methanol/water (5:5:1, v/v/v). Chlorofonn and 0.88% aq. KCl were added to give chloroform/methanollwater ratios of 2:1 :1, (v/v/v), resulting in a phase separation. The aqueous phase was back extracted with chloroform and lipids were recovered from the combined chloroform phases. For [“C]glycerol and [“C]carbon dioxide labeled lipids the proportion of label in the acyl groups versus the backbone/head group was determined by transmethylation [158] and 38 scintillation counting of the separated organic and aqueous phases. Aqueous phase radioactivity from PC was determined to be in the glycerol backbone by silica TLC developed in methanol/waterl28°/o NH4OH/3M NaCl (50:26.6:17.9:3.4, vlvlv/v), as sample radioactivity co-migrated with glycerol and glycerol-3— phosphate standards, but not choline. Radiolabeled Acyl Group Composition— FAM Es were prepared from purified individual lipids or total lipids by base-catalyzed transmethylation [158]. Recovered FAMEs were separated based on the number of double bonds by AgNOa-TLC, the plates being developed to % height with hexane/diethyl ether (1/1, vlv), then fully with hexane/diethyl ether (9:1, vlv). PC Molecular Species Detennination—PC was separated from other lipids by silica TLC (K6 plates developed with chloroform/methanol/acetic acid (75:25:8, v/v/v)). DAG was produced from the purified PC by phospholipase C (B. cereus, Sigma) digestion [159]. The DAG was acetylated as described above and the 1,2-diacyl-3—acetyl-glycerols separated into molecular species based on the number of double bonds by argentation-TLC [159], using a triple development (‘/2 then ”/4 development in chloroform/methanol (96:4, vlv), then fully in chloroform/methanol (99:1, vlv)). The proportion of radioactivity in each band was determined by electronic autoradiography, then each band was eluted, the recovered lipids transmethylated, and the [“C]FAME analyzed by AgNOa- TLC, as described above. To determine endogenous acyl groups from isolated PC molecular species, triheptadecanoin was added as an internal standard to each fraction during elution and prior to transmethylation and GC analysis of 39 FAME. When necessary 1,2-diacyl-3-acetyl-glycerol fractions recovered from AgNOa TLC plates were further purified by reverse phase TLC on Partisil® KC18 silica gel 60 A plates (Whatman, Maidstone, UK) developed with methanol/acetonelwater (75:25:2, v/v/v). PC Acyl Group Stereochemistry—PC was isolated as described above and digested with phospholipase A2 (Crotalus atrox, Sigma) [159]. Briefly, PC was dissolved in 1 ml diethyl ether and 0.5 unit PLAz in 0.1 ml of 50 mM Tris-HCI pH 8.7, 5 mM CaClz. The reaction was mixed vigorously for 5 min then the ether was evaporated under N2. To extract lipids 3.8 ml chloroforrnlmethanol (2/1, v/v) and 1 ml 0.15 M acetic acid were added, the mixture vortexed, the chloroform phase collected and the aqueous phase back extracted with 2.5 ml chloroform. Reaction products were separated on silica TLC plates developed with chloroform/methanol/acetic acid/water (50:30:8/4, vlvlvlv). Radioactivity in the free fatty acid and lyso-PC fractions was quantified by electronic autoradiography, then each product eluted and transmethylated by heating at 50 °C in 5% sulfuric acid in methanol for 30 min. Labeled and unlabeled FAME compositions were determined as described above. Lipid Molecular Species by Mass Spectrometry—Pea leaf samples were analyzed by ESl-MS/MS by the Kansas Lipidomics Research Center. Extraction of lipids was conducted by their standard Arabidopsis leaf protocol (Web site: ELI/WWW.k-state.edu/lipid/lipidomics/leaf-extraction.html). The data set from this analysis is available as a supplement in Appendix A, Figure 5-1: S1A-S1M. 40 Net Rate of Fatty Acid Deposition—The net rate of fatty acid synthesis by pea leaves was determined by harvesting 10 leaves from pea seedlings at the start and end of the light cycle for three days. After immediate weighing to determine fresh weight, FAME and chlorophyll contents were measured as described above. Triheptadecanoin was added during the lipid extraction as an internal standard, providing methyl heptadecanoate after transmethylation for FAME analysis by GC. 2.4 Results Kinetics of Glycemlipid Labeling from [“C]Acetate— Relationships between precursor and product pools in metabolic pathways are revealed by kinetic labeling experiments. The linear accumulation of a product, once reached, coincides with steady state labeling of all precursor pools. In this context [“C]acetate is ideal for the study of acyl lipid metabolism as it is rapidly taken up by excised leaf tissue and utilized for FAS, while the lipids are labeled in their acyl groups, with minimal head group labeling [79]. The kinetics of [“Clacetate labeling of glycerolipids from rapidly expanding excised pea leaves over a period of 9 minutes is shown in Figure 2-1. PC was the major radiolabeled glycerolipid with over 65% of the label at all time points. Furthermore, steady state labeling was established very rapidly, with no detectable lag (Figure 2-1A) and thus suggested that PC was a very early product of nascent FA incorporation into eukaryotic glycerolipids. A second independent 9 min time course gave similar 41 results (data not shown). Phosphatidylethanolamine (PE) was labeled in a similar manner (Figure 2-1C), with no obvious lag phase, but PC labeling was 15- fold greater than PE labeling despite endogenous PC levels only being twice that of PE (Appendix A, Figure 5-1:S1A). Both the PA and the DAG pools were still filling after the onset of PC steady state labeling. The label in PA reached a maximum by 3 min whereas DAG labeling increased throughout the time course (Figure 2-1 B). Thus, although PA and DAG are intermediates in the de novo synthesis of PC via acylation of glycerol-3-P, these [‘4C]acyl labeled intermediates do not show a precursor-product relationship relative to PC labeling, and thus represent different glycerolipid pools. Of course, we cannot rule out that very small labeled PA and DAG pools might contribute to PC labeling in a precursor-product relationship, since experimentally, quenching in these kinetic assays limits the temporal resolution of the method. A statistical analysis suggested about 0.5 min variance (95% CI) in rate extrapolations to zero time. In conclusion, we were not able to detect a [”Clacyl labeled glycerolipid precursor pool to the labeling of PC. 42 0/ pmol [14C1C2 w via a: ?e 2 3 4.5 6 7 s 910 Minutes 0/ mg chl 400q pmol [14C1C2 i— [g DJ 0 pmol [14C]C2:0/ mg chl Figure 2-1 Time course for [“Clacetate incorporation into the acyl groups of glycerolipid classes by excised pea leaves. A-C: Results expressed as mean pmoles [“Clacetate incorporation per mg chlorophyll (pmol [“0102:0/ mg chl) for each lipid class :I: S.E. (triplicate determinations). Glycerolipid classes are PC, squares; DAG, triangles; PA, inverted triangles; PG, diamonds; PE, circles. Acyl Compositions from [14C]Acetate Labeling of Lipids—Figure 2-2 shows the composition of labeled fatty acids from the [“C]acetate time course, for total lipids, PC, PA, DAG, PG and PE. Oleate was the major product in total lipids (Figure 2-2A), with saturates, mainly palmitate, decreasing from 32% at 1 min to about 24% for the remainder of the assay period. The reason for this decline is ' unknown. There was no detectable desaturation of 18:1 at the earliest time point but desaturation to 18:2 was detectable from 3 min onwards. PC had less saturates (11% falling to 8%) whereas PE had more saturates (54% falling to 38%). By the end of the time course desaturation produced 7% linoleate in PC (Figure 2-28). In contrast to PC and PE, PA (Figure 2-20) contained high levels of saturates (>60%) throughout the time course and had an acyl composition with the closest match to PG. PA is also an intermediate of the prokaryotic glycerolipid synthesis pathway, and so the labeled PA pool is likely the precursor for prokaryotic PG labeling. In this context, close inspection of Figure 2-1B and 2- 1C shows a kinetic precursor-product relationship between PA (label reaching a maximum at 3 min) and PG (lag phase ending at approximately 3 min), confirming this conclusion. DAG contained an intermediate level of saturates compared to PC and PE. The origin of the labeled DAG is not certain. It is probably largely eukaryotic in origin, and may arise in part from the reverse action of CDP-cholinezDAG cholinephospho-transferase or phospholipase C on labeled PC, or from acyl groups edited from PC re-entering the eukaryotic de novo glycerolipid synthesis pathway. Any of these explanations is supported by 44 A: Total Lipids B PC a 100 g 103 o- n- F '- -— _ g 80- g 80- 0 C) '3, 601 3:, 6o- :: : » i-Iu 40- _ 4 a U U 0 V V ——-i 20" “—1 20“ G 6 1 3 . 6 9 1 3 . 6 9 Minutes Minutes H: > H ”a c: A Q L % [14Clacyl chains 9 a a C l l l 3 6 3 6 Minutes Minutes 9 l1! '1 C) S H sates: eccca gill % [MCjacyI chain 3 . 6 9 Minutes Figure 2-2 [“ClAcyl composition of [“Clacetate labeled lipids Total lipids and isolated lipid classes from the time course assays run with excised pea leaves, as shown in Fig. 1, were analyzed for their labeled acyl group distribution. Results expressed as mean :t SE. (triplicate determinations). Each bar represents the % of total lipid radioactivity present in saturates (16:0>18:0) (black bars), monoenes (18:1) (white bars), and dienes (18:2) (hatched bars). PG also contained 4% trienes (18:3) at 9 min. A, Total lipid FAMEs; B, PC; C, PA; D, DAG; E, PG; F, PE. 45 the appearance of labeled linoleate in DAG by 9 min (Figure 2-20). Whatever its origin, the difference in labeled acyl group composition of PA and DAG compared to PC supports the conclusion from the kinetic data that they do not represent precursors of initial PC acyl labeling. The Kinetics of Glycerolipid Labeling from [’4CIGcherol—[14C]Glycerol is rapidly taken up by pea leaves and incorporated into the glycerol backbone of glycerolipids. In addition, acyl groups also become labeled because glycerol-3- phosphate equilibrates with glycolytic precursors, leading to plastid acetyl-CoA production [86]. To separately analyze the label from the backbone/head-group and the acyl-chains, isolated lipid classes were transmethylated. Analysis of the aqueous fraction from PC by TLC indicated that the radioactivity was contained in the glycerol backbone and not the choline headgroup, as noted before [86]. In marked contrast to acetate labeling, [“C]glycerol incorporation into lipids exhibited a lag (Figure 2-3A). At the earliest time points, radioactivity in DAG was approximately equal to PC. We assume that the labeled PA includes a large contribution from the plastid component. However, chloroplast lipid assembly in 18:3—plants does not require PA conversion to DAG, so we also assume the labeled DAG is largely associated with extra-plastidial membranes. Two lines of argument support the notion that the biosynthesis of PC from labeled acetate (Figure 2-1) and from labeled glycerol (Figure 2-3) report different metabolic processes. First, for de novo PC synthesis the relative 46 A: [14C]glycerol backbone labeling 32129 b—I 00 c c G I l N h a c c a I I I G 015553655910 Minutes CPM/ mg chloroph B: Proportion of [14Clglycerol incorporated into acyl groups _‘ UPC .DAG Mb) 2'?“ i—Ai—NN Uh: 4I ¢ I U! I % acyl group [14C] 3 . 6 9 Minutes Figure 2-3 Time course for incorporation of label from [“Clglycerol into the backbone and acyl groups of glycerolipids by excised pea leaves. _-_ l 6 I r A: The amount of radioactivity in the glycerol backbone of each lipid. Results expressed as mean :I: S.E. (triplicate determinations): PC, squares; DAG, triangles; PA, inverted triangles. B: The % of total lipid [“Clglycerol labeling that is in the acyl chains. The three samples of PC and DAG (Figure 3A replicates) were each combined for the glycerol to acyl group determination. PC, white bars; DAG, black bars. 47 movement of label from PA to DAG to PC should be same for both glycerol and acyl group labeling strategies. However, at the earliest time points, it is clear that acetate acyl chain labeling produces PC >> DAG (Figure 2-1A) whereas glycerol backbone labeling produces DAG = PC (Figure 2-3A), suggesting separate metabolic processes. Second, when we analyze the different kinetics of acyl chain and glycerol labeling of PC and DAG from [“C]glycerol labeling (Figure 2- 3B), the amount of label in the acyl chains of DAG remained fairly constant (4- 5%) while the acyl label in PC fell from 32% at 1 min to 12% at 9 min. The difference is explained by postulating two different pathways for PC synthesis. About 10% of the label from exogenous glycerol is rapidly utilized for de novo FAS and labeled acyl groups move rapidly to PC, and then to DAG, as described for Figure 2—1. The remainder of the labeled glycerol moves through PA to DAG, which has a half life for pool filling of about 3-4 minutes, and then to PC. Under these conditions the fraction of PC which is acyl labeled is initially high but declines steadily, while this simple model predicts that the acyl labeled fraction of DAG remains low and fairly constant. This is what is observed. Together the glycerol backbone and acyl chain labeling from [“C]glycerol and the [“C]acetate acyl group labeling provide evidence for a separate (acyl-editing) pathway for newly synthesized FA to rapidly enter PC without de novo PC synthesis by the eukaryotic pathway. 48 A: PC molecular species by AgNO; TLC 35 88 so ST SM MM MD MI“ DD 01‘ TT Molecular Species B: PC molecular species by ESI-MS/MS 24 vavvesocsceoooooeoeoeocccc MMMMMMMMMMMMMMMMVVVV Molecular Species C: PA molecular species by ESI-MS/MS 4o GVMN—IQWVMN—I‘OWV’MNWVMN 49 Figure 24. Molecular species analyses of endogenous PC and PA from pealeaves. Data is presented as mole percentage of each molecular species within that lipid class (100%). A: PC molecular species as determined by AgN03 TLC/GC (three determinations). Molecular species are reported as the combination of two acyl groups: S, saturates (16:0>18:0); M, monoenes (18:1); D, dienes (18:2); and T, trienes (18:3). B-C: molecular species as determined by ESI- MS/MS (five separate measurements). Molecular species reported as (# total carbons: # total double bonds). B: PC. CzPA. All error bars are standard error. Glycerolipid Molecular Species in Expanding Pea Leaves—Molecular species of PC from expanding pea leaves were analyzed by GC of FAMEs following AgNOa-TLC separation of the corresponding 1,2-diacyl-3-acetyl- glycerols. The data are shown in Figure 2-4A, and are very similar to a published analysis of PC molecular species from pea leaf microsomes as determined by reverse-phase HPLC [143]. In addition pea leaf PC molecular species from expanding leaves were determined by ESl-MS/MS (Figure 2-4B). The major fatty acids of pea leaf PC are 18:2 (44%), 18:3 (31%), 16:0 (18%), 18:0 (4.2%) and 18:1 (2.6%). The major molecular species that we observe are the expected combinations of these fatty acids, given that saturates are confined to the sn-1 position. Oleic (18:1) and palmitic (16:0) are the major products of chloroplast FAS (Figure 2-2A). We note that if these nascent FA were incorporated via de novo PC synthesis without any contribution from acyl editing then 18:1/18:1-PC would be the most abundant initial species formed. However this species represents less than 1% of the total PC (MM, Figure 2-4A). Likewise, the PA molecular species contain only trace amounts of 16:0/18:1 (034:1) and 18:1/18:1 (C36:2) based on ESl-MS/MS (Figure 2-4C). The majority of the C36:2 PA (and PC) species will be 18:0/18:2. Total PA contributes 0.1-0.4 mole % to total pea leaf membrane lipids. The PA profile is quite similar to that for PC, but it does contain significantly more 16:0/18:3 (34:3). PA may include biosynthetic pools, pools derived from lipid signaling via phospholipase D [4] and DAG kinase, and the putative PA pool for delivery of eukaryotic molecular species to the plastid for 50 galactolipids synthesis [90]. PA also contains a small amount of 34:6, which is presumed to be 18:3/16:3, and may therefore indicate a plastid contribution. Analysis of Molecular Species of PC from (“CJAcetate Labeling—To examine the initial molecular species of PC produced from incorporation of nascent FA, pea leaves were labeled with [“C]acetate for 5 min. This time was a compromise between allowing enough labeling of PC for the analysis, yet keeping the amount of labeled 18:2 to a minimum (Figure 2-23). Each labeled molecular species, analyzed as the 1,2-diacyl-3-acetyl-glycerol derivative, was also analyzed for labeled fatty acid composition. Figures 2-5A and 2-58 show the distribution of PC molecular species by mass and [“C]acyl group, respectively. If acyl chains enter eukaryotic phospholipid synthesis via glycerol-3-phosphate acylation without acyl editing the predicted initial molecular species of PC would be 16:0/18:1 and 18:1/18:1 (SN and M/M, respectively). Each will be dual labeled with newly synthesized FA, and if the saturates are 10% of the total label (Figure 2-28) then the SIM to MlM labeling ratio will be approximately 1:4. This does not occur. Also, the SIM molecular species was only labeled with oleate; the corresponding palmitate was unlabeled. The MM molecular species which would be predicted to contain 80% of the initial label actually represented only 6.5% of all the labeled molecular species. It is unclear if this species contains one or both acyl groups labeled. The data in Figure 2-53 indicate that all other labeled PC molecular species contain one labeled acyl group (18:1 or 16:0/18:0) together with an unlabeled acyl group (18:2, 18:3, or 16:0/18:0). Thus > 90% of 51 A: Leaf PC mass composition 35 SS SD ST SM MM MD MT DD DT TT Molecular Species B: 5 min ['4Clacetate labeled acyl chains in 35 .E a 30+ '5 25¢ g. 20- 2.154 10- 5- % [“c SS SD ; H t I I I l F—I— ST SM MM MD MT DD DT TT Molecular Species C: 5 min [”Clglycerol labeled backbone J “-3 g 30- SS SD ST SM MM MD MT DD DT TT Molecular Spec1es D: 5-10 min l4C02 labeled acyl chains 1E Mao .Jlfln SS SD ,1 T I ST SM MM MD MT DD DT TT Molecular Species 52 Figure 2-5 Analysis of endogenous PC molecular species and those labeled after short time incubations of excised pea leaves with [“Clacetate or [“Clglycerol or of seedlings with [“c1coz. A-D: Molecular species are represented as pairs of FA, with bar heights represent % of species among the total. Bar shading represent % of FA per species (mass or 14C): S, saturates (16:0>18:0), black bars; M, monoenes (18:1), white bars; D, dienes (18:2), hatched bars; and T, trienes (18:3), gray bars. A: Endogenous PC molecular species mass composition from Figure 2-4A. B: 5 min [“C]acetate labeled PC molecular species. Results expressed as mean 1 SE. (triplicate determinations). C: 5 min [“Clglycerol labeled PC molecular species, label was in the glycerol moiety only (label in acyl chains has been subtracted). Results expressed as mean :t: SE. (duplicate determinations). D: 5-10 min [“C]COz acyl chain labeled PC molecular species. Results expressed as mean :I: S.E. (quadruplicate determinations). the molecular species of [‘4C]PC contain one newly synthesized (labeled) FA in the same molecule as a previously synthesized (unlabeled) FA. This finding strongly suggests acyl editing as a mechanism of PC synthesis with nascent FA. Analysis of Molecular Species of PC Synthesized de novo from [14C]Glycerol—Because acyl editing only exchanges acyl groups it does not result in net synthesis of phospholipid. However, in order for a leaf to grow net (i.e. de novo) PC synthesis is required. To determine the PC molecular species and hence ascertain which acyl chains are involved in de novo eukaryotic glycerolipid synthesis, pea leaves were incubated with [“C]glycerol. The PC glycerol backbone will be labeled regardless of whether the acyl chains are nascent FA or from acyl-editing. Figure 2-5C shows the distribution of label among PC molecular species after 5 min [“Clglycerol labeling. The data are corrected to give only the glycerol backbone labeling by subtracting the small amount of acyl labeling (Figure 2-3B) that occurs from [“Clglycerol. The PC molecular species distribution obtained with glycerol backbone labeling (Figure 2- 5C) is quite different from that for acetate acyl labeling (Figure 2-58), but closely resembles the endogenous mass distribution (Figure 2-5A), with the exception of a significant reduction in 18:2-containing molecular species and an increase in 18:3-containing species. The major initial PC molecular species predicted from incorporation of only nascent FAs, 18:1/18:1, was less than 0.2% of labeled species. The total 18:1 content of the glycerol-labeled PC is 5%, some of which may derive from nascent FA. We know that at this time point the ratio of unsaturates to saturates in PC from nascent FA is about 10:1 (Figure 2-23) and 53 that there are negligible polyunsaturates. Thus we infer that at the most there is a 5.5% contribution of newly syntheSized FA to de novo PC biosynthesis, or, conversely, that the bulk (2 95%) of the acyl groups in PC molecular species labeled via glycerol backbone come from acyl editing. Figure 2-50, which describes PC molecular species analysis from [“C]carbon dioxide labeling, will be described in a later section. Stereochemistry of PC acylation—Molecular species of PC can be defined by acyl composition and/or by stereochemistry. To determine the stereospecific incorporation of newly synthesized FA, PC isolated from 5 min incubations of pea leaves with [“C]acetate was digested with phospholipase A2 and the reaction products analyzed (Figure 2-6). 62% of the label was incorporated at the sn-2 position and 38% at the sn-1 position. Saturates were predominately found in the sn—1 position, indicating that the PLA2 digestion was sn-2-specific. Although de novo PC synthesis requires equimolar acylations at the sn-1 and sn-2 position we cannot assume that this will be the case for the 16:0 and 18:1 labeled FA if they are mixing with an endogenous acyl-CoA pool containing 18:2 and 18:3, since we do not know the acyl transferase specificity at each position. Previous workers have reported either a similar sn-1 to sn-2 distribution of acyl label, for PC from 30 minute acetate labeling of spinach leaves [84], or, using a 2 hour 54 an 70 fl '5' 60‘ "'T '5 50- 5:40- 3. 30- :0 20- 10- 0- 38 Elltl.__ail|- sn-l snLZ Stereochemical position Figure 2-6 Stereochemical distribution of [“Clacyl chains in PC. Pea leaves were labeled for 5 min with [“C]acetate. FAMEs generated from the fatty acid and lyso-PC products of phospholipase A2 digestion of isolated PC were separated based on the number of double bonds: saturates (16:0>18:0), black bars; monoenes (18:1), white bars; dienes (18:2), hatched bars. Results expressed as mean t SE. (triplicate determinations). 55 acetate labeling of leek leaves, a 2:1 to 3:2 enrichment in the sn-2 position [92, 93]. The unequal stereochemical labeling complements the conclusion that nascent FAs are incorporated into PC along side previously synthesized FA. Unlabeled FA Composition of Endogenous PC, LPC, and Various Labeled PCs—From Figure 2-53 we have estimated the composition of the unlabeled FAs that are associated with the nascent, [“C]acetate-labeled FAs. The calculated values are shown in Figure 2-7, along with the acyl composition of endogenous PC (calculated from Figure 2-4A) and of endogenous lyso-PC (obtained by ESl-MS/MS, Appendix A Figure 5-1 :81 K). To a first approximation all three profiles are similar. The unlabeled, paired acyl groups from [“C]acetate- labeled PC have a slightly reduced 18:2 and increased saturates compared to endogenous PC. The unlabeled fatty acids must have had an inverted stereochemical distribution to that of the labeled FA measured in Figure 2-6; that is, 62% sn-1 and 38% sn-2. The fact that the FA profile of lyso-PC was similar to that of total PC allows us to propose that the endogenous lyso-PC pool is produced by approximately equal sn-1 and sn-2 deacylations, whether catalyzed by hydrolysis or by PC acyl exchange with CoA. Thus the appropriate lyso—PC pool is available for both sn-1 or sn-2 acylations with newly synthesized FA. The unequal stereochemical acylation with nascent FA found in PC (Figure 26) therefore may arise from sn-2 acyl migration in lyso-PC to the more thermodynamically stable sn-1 position, or possibly a higher reacylation of the 56 UI CD A c I N c I % acyl composition E $ ;\§\\\§\\\\\\\w c I Acyl composition Figure 2-7 Comparison of unlabeled fatty acid compositions deduced from [“Clacetate and [“Clglycerol labeled PC molecular species to endogenous pea leaf PC and lyso-PC compositions. Black bars: [120]FA in the same molecule as [‘4C]FA labeled molecular species, as calculated from figure SB. White bars: endogenous PC FA mass composition, as calculated from Figure 4A. Gray bars: endogenous pea leaf lyso-PC FA mass composition, as calculated from ESl-MS/MS (supplemental Figure S1 K, 5 determinations). Hatched bars: FA composition calculated from [”C]glycerol labeled PC molecular species, as calculated from Fig. 5D. All error bars are standard error. S, saturates (16:0>18:0); M, monoenes (18:1); D, dienes (18:2); T, trienes (18:3). 57 sn-1 position with the unlabeled FA that have been previously removed by acyl editing. It is also possible that the bulk lyso—PC pool has little to do with the incorporation of nascent FA. Figure 2-7 also shows the FA composition of molecular species of PC labeled by [“C]glycerol (calculated from Figure 2-5C). [1 4C]Carbon Dioxide Labeling—The use of excised tissue facilitates rapid kinetic analysis. However, incubation of excised leaf tissue might cause a wound response that could alter metabolism. In particular, lipase activities may be rapidly and transiently induced that could alter phospholipid metabolism. To control for this possibility whole pea seedlings were labeled for 5-10 minutes with [“CJCOZ. Label is rapidly incorporated into both the acyl-chains and glycerol backbone of lipids. Figure 2-8 compares the distribution of acyl group labeling among lipid classes for [”C]COz labeling of whole seedlings and [“C]acetate labeling of cut leaves at approximately equivalent assay times. Both labeling experiments indicated that PC is the major recipient of newly synthesized FA and that the relative acyl group labeling among other lipid classes is similar. Additionally, molecular species of PC were analyzed for acyl-chain labeling from whole seedlings with [14C1002 (Figure 2-50). This distribution should be compared with that from [“Clacetate labeling of cut leaves (Figure 2-53). The slightly longer time of assay (approximately 10 minutes including seedling removal and quenching) for [“Clcarbon dioxide compared to [“Clacetate 58 co 6 I l4C02 I'J [14Cjacetate as ‘P N © I % acyl chain labeling PC PG DAG PE PA Lipid Class Figure 2-8 Distribution of lipid classes, labeled in their acyl chains, after incubation of pea leaves with [“CJCOZ or [“Clacetate. Black bars: ~5-10 min [1401002 labeling of intact pea seedlings, with results expressed as mean :I: S.E. (quadruplicate determinations). White bars: 6 min [“Clacetate labeling of excised pea leaves (Data from Fig. 2-1, results expressed as mean :I: S.E. (triplicate determinations». 59 labeling allowed slightly greater desaturation of 18:1 to 18:2. However, it is clear that the pattern of nascent FA acyl labeling from [”C]COZ matches that of [“C]acetate. We conclude that the relative proportions of individual lipids and initial PC molecular species labeled in incubations of excised leaves was not due to a change in lipid metabolism caused by a wound response but is indicative of the in planta lipid metabolism. 2.5 Discussion Largely through analyses of the kinetics, molecular species and stereochemistry of the immediately labeled products of eukaryotic glycerolipid synthesis we demonstrate that, in rapidly expanding pea leaves, (1) there was no detectable precursor-product relationship for PA and DAG in the incorporation of nascent FA into PC, (2) nascent FA were incorporated into PC more rapidly than the de novo synthesis of PC from glycerol, suggesting two systems for the incorporation of acyl groups into PC, (3) greater than 90% of nascent FA were incorporated into PC molecules in combination with an endogenous FA, (4) greater than 95% of de novo synthesized PC was esterified by recycled acyl groups, (5) acyl editing can take place at both the sn-1 and sn-2 positions, and (6) that pea leaves contain an endogenous lyso-PC pool that could be involved in acyl editing. Experiments with carbon dioxide labeling of whole seedlings demonstrated that the metabolism seen in excised leaves is not due to a wound response. 60 These results lead us to conclude that the major pathway for the incorporation of newly synthesized FA into eukaryotic glycerolipids is largely through acyl editing of PC and that de novo synthesis of eukaryotic glycerolipids primarily utilizes acyl chains recycled from acyl editing. We will discuss these results in terms of the possible mechanisms and flux models (Figure 2-9) that can be proposed to drive acyl editing. 18:1/18:1 and 16:0/18:1 Are Not Initial Molecular Species of Eukaryotic Glycerolipid Synthesis—The classical scheme for eukaryotic phospholipid synthesis in leaves [34, 37], as summarized in Figure 2-9, Model 1, predicts the biosynthesis of dual-labeled 16:0/18:1 and 18:1/18:1 PA, DAG and then PC species sequentially. However, our kinetic analysis failed to detect precursor- product relationships between PA, DAG and PC pools for acyl group labeling . (Figures 2-1 and 2-2). Furthermore, we have not been able to observe the expected, dual-labeled 16:0/18:1 and 18:1/18:1 PC species. Based on estimates of FAS fluxes and endogenous PA and PC molecular species pool sizes (see supplement Note 1, Appendix A), labeled 16:0/18:1 and 18:1/18:1 molecular species of PC should have been easily detected by our analysis if Model 1 was correct. Additionally, glycerol labeling of de novo synthesized PC revealed mostly esterification with endogenous saturates and polyunsaturates. and not newly synthesized saturates and oleate. Taken together these three lines of evidence strongly suggest that 16:0/18:1 and 18:1/18:1 molecular species are not the first products of nascent FA incorporation into eukaryotic glycerolipids and that Model 1 is incorrect. 61 G3P -> Lyso-PA -> PA -> DAG V II PC FFA/acyl-CoA 3 Lyso-PC sn-1Isn-2 G3P —> Lyso-PA —> PA —> DAG Mildtl 3 Y Lyso-PC FFA/acyl-CoA sn-l/sn-Z G3P —’ Lyso—PA ‘9 PA "" DAG Modtl 7 Y{ FFA/acyl-CoA {LCyso-P sn-l/sn-Z G3P 9 Lyso-PA 9 PA 9 DAG , FFA/acyl-CoA $C Lyso-PC sn-1Isn-2 Figure 2-9 Four possible models describing pathways and fluxes for the incorporation of nascent FA, exported from the chloroplast immediately after de novo FAS, into PC. For simplicity, the models are set up not to depend on enzyme specificity, and show only PC. The mechanisms for removal of fatty acids from PC and provision of acyl-CoA are not defined, while the width of the arrows gives an indication of relative flux. The grey box represents the chloroplast; G3P, glycerol-3-phosphate; FFA, free fatty acid. Model 1 shows a description lacking any acyl editing. Models 2 and 3 are considered the best representations for the acyl editing that accompanies in the incorporation of nascent FA into PC, but a definitive choice of the most correct model is not yet possible. 62 Most Newly Synthesized FAs Enter PC by sn-1 and sn-2 Acyl Editing, while de novo PC Synthesis Primarily Uses Recycled Acyl Groups—We have shown that the majority of newly synthesized (labeled) FAs in pea leaves are immediately incorporated into PC molecular species where > 90% are paired with endogenous (unlabeled) FAs. A similar observation of >60% incorporation of nascent FA paired with endogenous FA is also reported by Williams et al. [149] for Brassica napus leaves 1hr after a pulse of carbon dioxide labeling. Because we have used rapidly expanding leaves from an 18:3 plant, whereas Williams et al. [149] used mature leaves from a 16:3 plant, it appears that the observation may be a general one. The result implies some form of acyl editing. Either the nascent FA are used directly in an acyl editing process (Figure 2-9, Model 2), or acyl groups are released from endogenous lipids and mix with the nascent FA prior to completing the acyl editing cycle and/or acylation of glycerol-3-P for de novo PC synthesis (Figure 2-9, Models 3 and 4). Models 2-4 will be discussed in more detail later. Whatever the acyl editing mechanism, net (i.e. de novo) phospholipid synthesis from glycerol-3-P is required for the leaf to grow. Glycerol labeling showed that de novo PC synthesis utilized mainly endogenous palmitate (+ stearate), linoleate and linolenate and not newly synthesized FA. Thus the FA used for de novo PC synthesis must be largely (2 95%) recycled from acyl editing, and, almost certainly, largely from PC. The fact that saturated FAs are major acyl groups released by acyl editing suggests that sn-1 hydrolysis or CoAzPC sn-1 acyl exchange contributes in a major way to this process. This is confirmed by the results from the acetate labeling experiments shown in Figure 63 258 and Figure 2-6, namely that > 90% of the molecular species contain only one labeled acyl group and 40% of the labeled acyl groups are at the sn-1 position. For Models 2 and 3, this implies that a sn-1 acyl editing process for incorporation of nascent FA into PC represents 30-40% of the net flux. This conclusion from our in vivo labeling represents a substantial biochemical activity largely overlooked or underrepresented by previous in vitro analyses of -PC acyl metabolism [94, 140] and in vivo labeling studies [149]. Proposed Models for Acyl Editing—Questions then arise as to the pathways, fluxes and mechanisms for the metabolism of PC synthesis utilizing nascent FA from the chloroplast. The puzzle is complex because of the large number of unknowns, and our current data do not completely resolve the models proposed. However, it is instructive to review possible models to highlight the unknowns requiring resolution. Three models are considered which might accommodate a flux of x moles of nascent FA from the chloroplast, producing 0.5x moles of net PC synthesis (Figure 2-9, Models 2-4). For simplicity, the models were initially set up not to depend on enzyme specificity, and show only PC. The mechanisms for removal of fatty acids from PC and provision of acyl- CoA are not defined. They may be via phospholipase action with activation of the released free fatty acid by an acyl-CoA synthetase, or via direct CoAzPC acyl exchange. Model 2 describes a situation whereby nascent FA are channeled to lyso- PC. Their incorporation allows just one labeled FA per re-synthesized PC molecule, as observed experimentally. The endogenous FA released from PC is 64 concomitantly used in the de novo synthesis of PC. The endogenous FA may enter a general acyl-CoA pool, and the model does not rule out other sites of phospholipid acyl editing also supplying this pool. Model 2 implies acyl transfer reactions for nascent FA incorporation and de novo PC synthesis may be spatially distinct metabolic processes. The model cannot accommodate the catabolism of PC to GPC, because the acylation of GPC by the nascent, labeled FA will lead to two labeled acyl groups per PC molecule. As the endogenous, unlabeled acyl distribution in PC molecular species from acetate labeling closely matches the endogenous PC acyl distribution (Figure 2-7), the lyso-PC and the FA pools produced by PC editing must have the similar acyl profiles. This is confirmed by the endogenous LPC analysis. The fact that both lyso-PC and unlabeled FA in FAIFA*-PC have compositions identical to the bulk PC does not absolutely require sn-1 and sn-2 deacylations to be on an equimolar basis. However, the phospholipase/acyl exchange activities (and any lyso-PC isomerization) will control the sn—1 and sn-2 distributions observed in acetate- labeled acyl groups. Thus Model 2 can accommodate the excess sn-2 acylation noted in acetate labeling. Model 2 requires an explanation for the reduction in 18:2 FA content for glycerol-labeled PC, by, for example, proposing a preference for the transfer of 18:2-containing acylglycerol moieties to the chloroplast for galactolipid synthesis [86]. Model 3 describes a situation whereby the nascent FAs are directed to a general acyl-CoA pool, along with FAs from PC acyl editing. Thus the pairing of labeled and unlabeled acyl groups in PC is caused by a high flux of acyl editing, 65 which greatly dilutes out nascent FA in the common acyl-CoA pool. In this model it is possible to propose that PC editing stops at lyso-PC or instead goes through two deacylation steps to produce GPC. However, because GPC-dependent acyltransferases have not been reported in plants, we will focus on the lyso-PC example, as shown in Figure 2-9 Model 3. An analysis of the molecular species of PC labeled from acetate showed that 2 90% had single acyl group labeling. Without any substrate specificity constraints, this requires that > 90% of the flux of acyl groups in the acyl-CoA pool originate from unlabeled fatty acids (Appendix A, Figure 5-2). Consistent with this observation, analysis of the molecular species of PC with glycerol backbone labeling required that 2 95% of the flux for de novo PC synthesis come from acyl editing. With this model the sn- 1 and sn-2 distributions for nascent FA do not have to fit a 1:1 ratio. Also, it is easy to explain the deficit of 18:2 acyl groups in the de novo PC labeling with glycerol. We can simply assert that there is some acyl-CoA selectivity in the de novo synthesis at either sn-1 or sn-2 acylation. A moderate change in the de novo synthesis pathway acyl composition will have little effect on the endogenous acyl group composition measured by acetate labeling. Of course, the simplicity of this explanation does not prove it. In Models 2 and 3 the acyl editing cycle (phospholipid deacylation and reacylation) is distinct from de novo phospholipid synthesis. It is, however, possible to have de novo phsopholipid synthesis be part of the acyl editing cycle (Figure 2-9, Model 4). This requires PC to be completely degraded to glycerol or glycerol-3-P, and choline or choline-P. In this model there is only one pathway for 66 the synthesis of PC, and if the PC turnover rate is high relative to nascent FA production, much of the labeling data is consistent with this model. However, a comparison of labeling kinetics for DAG and PC using either acetate or glycerol substrates strongly suggested that we were dealing with two distinct metabolic processes for PC synthesis. This is inconsistent with Model 4, and on this basis, we rule it out as a dominant pathway of leaf cytosolic glycerolipid synthesis. In this Discussion we have arrived at the conclusion that one of a pair of simple models (Figure 2-9, Models 2 and 3), offers the best description of cytosolic glycerolipid synthesis in pea leaves. The models present quantitatively very different magnitudes for the fluxes involved in acyl editing, and give different views on how such acyl editing integrates with the incorporation into PC of nascent FA exported from the chloroplast. Currently, we cannot categorically rule out either Model 2 or Model 3. Furthermore, it is always possible to build more complexity into models, including substrate specificity and the possibility of some hybrid combination. Future oxygen-18 and acyl-CoA pool labeling studies may yield pertinent information on the mechanism of acyl editing: lipolytic or through CoAzPC acyl exchange. Enzymology with isolated chloroplasts and microsomes may yield useful information on transacylation and de novo acyl transferase mechanisms and selectivity. Such studies must take into account the very substantial sn-1 component identified for acyl editing. Furthermore, endogenous PC is 2-fold higher than PE, but acyl labeling of PC from acetate (and also from glycerol) is 15-fold greater. This is despite the total leaf lyso-PE pool being approximately equivalent to lyso-PC (Appendix A Figure 5-1 :81 B). Thus, there is 67 7.5 or 15 fold preference for incorporation into PC via acyl editing that will have to be taken into account in defining any mechanism. A complete understanding may require the study of metabolism in KO mutants. However, with so many genes annotated as lipases, acyl hydrolases and transacylases [134], most with unknown functions, this may not be a facile approach either to this tricky question. Defining the mechanisms of the incorporation of nascent fatty acids into glycerolipids will help our understanding of lipid turnover and be relevant to research areas with more practical applications such as the integration of seed triacylglycerol synthesis with phospholipid turnover and the effects of biotic and abiotic stresses on phospholipid metabolism. 2.6 Footnotes This work was supported by the US Department of Energy grant DE-FGOZ- 87ER13729. ESl-MS/MS analysis was provided by the Kansas Lipidomics Research Center. Abbreviations: ACP, acyl carrier protein; DAG, diacylglycerol; FA, fatty acid; FAME, fatty acid methyl ester; FAS, fatty acid synthesis; GC, gas chromatography; GPC, glycerophosphorylcholine; LPCAT, lyso- phosphatidylcholine acyltransferase; PA, phosphatidic acid; PC, phosphatidylcholine; PE, phosphatidylethanolamine; PG, phosphatidylglycerol. Fatty acids are described by the convention “carbon number:number of double bonds.” For example, oleic acid is represented as 18:1. The molecular species of 68 phospholipids are described by FAIFA. When known, the sn-1 position is represented first, but often the stereochemistry is uncertain. For example, both 1- oleoyl-2-linoleoyl-sn-phosphatidylcholine and 1-linoleoyl-2-oleoyl-sn- phosphatidylcholine are described as 18:1/18:2-PC. 69 3 THE DOMINANT INITIAL FLUX OF FATTY ACIDS INTO SOYBEAN OIL IS THROUGH PHOSPHATIDYLCHOLINE ACYL EDITING Philip D. Bates‘, Timothy P. Durrett’, John B. Ohlroggez, Mike Pollard2 From Department of Biochemistry and Molecular Biology1 and Department of Plant Biologyz, Michigan State University, East Lansing, MI 48824. The results of this chapter will be submitted for publication in plant physiology. The ESl-MS/MS analysis of endogenous soybean lipid compositions was provided by Tim Durrett. The data is presented in Appendix B Tables 6-2, 6-3, 6- 4. The ESl-MS/MS data was reformatted for discussion of soybean lipid metabolism and comparison to labeled lipid data in Table 3-1 and Figures 3-1, 3- 6, and 3-8. 70 3.1 Abstract The reactions leading to triacylglycerols (TAG) synthesis in oilseeds have been well characterized. However, quantitative analyses of acyl group and glycerol backbone fluxes that comprise extraplastidic phospholipid and TAG synthesis, including acyl editing and phosphatidylcholine-diacylglycerol (PC- DAG) inter-conversion, are lacking. To investigate these fluxes we rapidly labeled soybean (Glycine max) embryos with [“Clacetate and [“Clglycerol. Cultured intact embryos that mimic in planta growth were used to minimize any changes in metabolism due to wounding. The initial kinetics of newly synthesized acyl chain and glycerol backbone incorporation into PC, DAG and TAG were analyzed along with their initial labeled molecular species and regiochemical compositions. Over 60% of the newly synthesized fatty acids enter glycerolipids through PC acyl editing, largely at the sn-2 position. This flux, mostly of oleate. was about 4- fold greater than the flux of nascent [“C]fatty acids into sn-1 and sn-2 positions of DAG through de novo glycerolipid synthesis. The total flux for PC acyl editing was estimated to be 2 to 7 times the flux of nascent fatty acids entering the eukaryotic pathway. Thus recycled acyl groups (16:0, 18:1, 18:2 and 18:3) provide most of the acyl chains for de novo glycerolipid synthesis. Our results also suggest that the DAG utilized for TAG synthesis is mostly derived from PC, whereas de novo synthesized DAG is mostly used for PC synthesis. These results provide a quantitative demonstration that acyl flux inlout of PC by acyl editing is the dominate flux in soybean embryos during TAG synthesis. 71 3.2 Introduction In plants essentially all acyl chains for membrane and storage lipid synthesis are produced in the plastid by acyl carrier protein (ACP) dependent de novo fatty acid synthesis (FAS) [28, 34, 160]. However, in oilseed tissue these acyl groups are utilized almost completely (> 95%) by the extra-plastid pathways of de novo glycerolipid synthesis, collectively termed the “eukaryotic pathway” [34]. The plastid acyl-ACPs are hydrolyzed and the free fatty acid and products (usually 18:1>>16:0>18:0) are transported across the plastid envelope [74]. Once activated to acyl-CoAs [77, 78], these acyl groups are available within the endoplasmic reticulum (ER) for incorporation by the acyltransferases of the eukaryotic pathway. The acyl-CoAs may be used for the sequential sn-1 and sn- 2 acylations of G3P to produce phosphatidic acid (PA) [34]. PA is then converted to sn-1,2-diacylglycerol (DAG) by the action of PA phosphatase. DAG represents an important branch point between neutral and membrane lipid biosynthesis. DAG may be acylated to produce TAG, or converted to phosphatidylcholine (PC) by CDP-choline:1,2-diacyl-sn-glycerol cholinephosphotransferase (CPT), with an analogous reaction to form phosphatidylethanolamine (PE) [34]. However, in both leaves [161] and in oilseeds the flux of nascent fatty acids into PC dominates over the synthesis of PE, PS and other phospholipids of the eukaryotic pathway. Desaturation of 18:1 to 18:2 and then 18:3 on PC produces the abundant polyunsaturated molecular species of PC [36, 37]. 72 Recently we investigated the initial steps in eukaryotic membrane lipid synthesis in expanding pea (Pisum sativum) leaves through in vivo labeling of the acyl groups and backbones of glycerolipids [162]. The results were inconsistent with a pathway in which newly synthesized fatty acids are directly esterified to G3P during de novo glycerolipid synthesis. Instead almost all newly synthesized fatty acids were esterified to a sn-1 or sn-2 lyso-PC through an “acyl editing” process resulting in molecular species of PC with one newly synthesized and one previously synthesized FA. In contrast, de novo glycerolipid synthesis in pea leaves primarily utilized recycled acyl groups released from PC during acyl editing. Acyl editing, often also termed “remodeling” or “re-tailoring,” is defined as any process that exchanges acyl groups between polar lipids but which does not by itself result in the net synthesis of the polar lipids. Acyl editing may proceed by at least two mechanisms. The first mechanism involves CoAzPC acyl exchange producing lyso-PC and acyl-CoA, which was demonstrated in microsomes isolated from developing seeds and was attributed to a reverse reaction of Iyso-phosphatidylcholine acyltransferase (LPCAT) [140]. There are several reports of high LPCAT activity in seeds [130, 140, 142, 144, 163]. A second mechanism for acyl editing involves hydrolysis of PC to lyso-PC, activation of the released free fatty acid and its re-utilization for phospholipid synthesis from lyso-PC by LPCAT. Plants exhibit both acyltransferase and transacylase mechanisms for the acylation of DAG to TAG. Three classes of diacylglycerolzacyl-CoA sn-3 acyltransferase have been identified, namely the ER membrane bound DGAT1 73 [164] and DGAT2 [165, 166] classes, and third a soluble cytosolic DGAT class [167]. In addition, the transacylase PDAT (phospholipid:diacylglycerol acyltransferase) allows the direct transfer of acyl groups from the sn-2 position of PC to DAG, producing TAG and lyso-PC products [119]. A DAGzDAG transacylase activity has also been described [122]. In addition to this diversity of TAG synthesis mechanisms and biosynthetic genes, oil-accumulating tissues may utilize different strategies to enrich for polyunsaturated fatty acids (PUFA) in TAG. PC is the major site of fatty acid desaturation and related reactions, while evidence for a direct role for DAG or TAG as a desaturase substrate is lacking. The conversion of PC to DAG (for TAG synthesis) may enrich for PUFAs in TAG. This conversion may be accomplished by a phospholipase or the reverse action of CTP-choline:diayclglycerol cholinephosphotransferase [1 17, 1 18]. Secondly, PUFA may be enriched in TAG by mechanisms of PC acyl editing. One such general mechanism has already been described above, as PC acyl editing may enrich the PUFA acyl-CoA pool by either a transacylase or lipase/LPCAT mechanism [162]. A second possible mechanism arises from the action of PDAT in TAG synthesis. The LPC produced by PDAT would be re-acylated by the action of LPCAT. From the above outline it is clear that there are a number of alternative metabolic routes for TAG synthesis. These include the recycling of intermediates of membrane lipid biosynthesis. The alternatives may vary with tissue, species and development. Different metabolic labeling experiments have been used to probe the sequence of reactions for TAG synthesis in vivo from a variety of 74 oilseed species. Utilization of a PC-derived DAG moiety has been proposed as the major pathway for TAG synthesis in excised linseed and soybean cotyledons [127]. Glycerol backbone labeling of developing safflower cotyledons suggested that the DAG pool feeding TAG synthesis was in equilibrium with PC, however a similar experiment with avocado mesocarp suggested that PC was not involved in TAG synthesis [168]. This led the authors to propose that PC backbone turnover is used primarily in high PUFA containing TAG synthesis. Phosphocholine head group labeling in linseed cotyledons was also consistent with the concept that DAG is in equilibrium with PC, allowing the production of a highly unsaturated DAG pool prior to incorporation into TAG [117]. The authors suggested that the reversible CPT reaction is most likely responsible for this PC- DAG inter-conversion. Labeling of safflower cotyledon slices with 16:0 and 18:1 free fatty acids allowed an alternative mechanism for PUFA enrichment in TAG to be proposed. An sn-2 specific LPCAT could direct 18:1 towards desaturation on PC and release PUFA to the acyl-CoA pool for the Kennedy pathway [169]. Although there is now significant in vivo evidence for PC acyl editing in the eukaryotic pathway in plant leaves [162], the in vivo evidence for such a cycle in oilseeds is still minimal. We took several steps to extend and improve on previous studies of oilseed metabolism to determine the pathway of FA incorporation into membrane and storage lipids in developing soybeans. To avoid potential artifacts associated with wounding due to excision of tissues we utilized cultured soybean (Glycine max) embryos which closely mimic in planta growth and oil accumulation [170]. 75 To provide an analysis of the initial events in eukaryotic glycerolipid synthesis we performed kinetic labeling studies with time points shorter than previous studies. Labeling was conducted with both [“Clacetate and [14C]glycerol and the labeled compositions, regiochemistry, and molecular species of labeled PC, DAG and TAG, along with endogenous molecular species compositions were determined. Together these approaches have allowed us to develop a more quantitative flux model of glycerolipid synthesis in vivo that demonstrates that acyl fluxes in and out of PC by acyl editing is a dominate flux in soybean during TAG synthesis. A better understanding of the pathways of TAG biosynthesis in developing soybeans may aid future efforts to engineer soybean oil with improved or novel properties. 3.3 Results Embryo culture and endogenous lipid compositions Accurate interpretation of metabolic labeling studies requires that during labeling the tissue functions as it would in vivo. In this context the use of cultured embryos (as opposed to embryo or seed slices in simple buffered media) minimizes lipase activation and other perturbations of lipid metabolism due to wounding. Thus we have utilized a soybean embryo culture system that allows dissected zygotic embryos to develop as in planta for over two weeks [170]. Radioactive precursors are incubated at low concentrations (acetate and glycerol at 0.5-1 mM) relative to the culture media, which contains sucrose (140 mM), 76 glucose (70 mM) and amino acids (48 mM) at levels that mimic the apoplastic concentration surrounding the embryo in planta. This minimizes changes in metabolism due to alternative carbon source utilization during the labeling. Relationships between precursor and product pools in metabolic pathways are revealed by kinetic labeling experiments. The linear accumulation of a product, once reached, coincides with steady state labeling of all precursor pools. In this context the duration of labeling needs to be short enough so that precursor pools have not filled before time point sampling begins. Soybean embryo DAG and PC had very similar FA compositions (Table 3-1) and molecular species profiles (Figure 3-1A). FA are desaturated on PC, but it is DAG that is the precursor for TAG. The combined PClDAG pool is expected to provide precursor DAG for TAG synthesis. TAG FA composition and molecular species distribution are shown in Table 3-1 and Figure 3-1 B, respectively. Developing soybean embryos of ~15-20 mg dry weight were pre-cultured for ~3 days before addition of radioactive substrates, to allow metabolism to equilibrate past any transient response due to the initiation of embryo culture. Cultured embryos accumulate ~6 mg total biomass per day [170] therefore each embryo doubles in biomass during the pre-culture. If the ratio of DAG:PC:TAG (Table 3-1) stays the same during pre-culture the initial pool sizes during labeling would be 208254229260 nmoles lipid embryo" respectively (~ 29230 nmoles total FA embryo"). Developing soybeans produce ~2.7 nmoles FA min'1 embryo‘1 [170]. As about 93% of newly synthesized FA are used to produce TAG, of which 1/3 are utilized 77 Table 3-1 Endogenous lipid pool sizes and fatty acid composition nmol lipid DAG :l: PC :I: TAG :r embryo'1 20.8 0.63 54.2 2.8 926 17.9 °/o mol FA 16:0 14.42 0.43 17.14 0.42 14.93 0.25 18:0 8.21 0.38 9.40 0.31 4.55 0.04 18:1 30.60 0.34 28.56 0.24 26.35 0.2 18:2 36.89 0.33 35.04 0.49 41.18 0.2 18:3 9.88 0.27 9.87 0.34 12.33 0.12 20:0 0.66 0.03 DAG, PC and TAG amounts as measured by ESl-MS/MS of 13-14 mg dry weight developing embryos, before pre-culture. The fatty acid composition as calculated from the molecular species presented in Figure 3-1. Fatty acid values are represented as percent mol among each lipid class. 78 A: DAG and PC molecular species 25 IDAG 20 - oPc g 15- B E 10 - 5 i L 0 .. SS SM SD ST MM MD MT DD DT TT B: TAG molecular species 14 12- 1o- 3 3. 6 E 6' 'l 4‘ ll ll 2- l 0' l (0:5l: F 2 CIF-CDI-I: 2 o P-cni-tZID F t w «>02 0 0 Cl 0 m $ m m a a a 8 w w E g E 2 2 2 8 o 0 Figure 3-1 Endogenous molecular species compositions of DAG, PC and TAG. A-B, Endogenous glycerolipid molecular species from developing soybean embryos of 13-14 mg dry weight per embryo as determined by ESl-MSIMS. Molecular species are represented as two or three FA, with bar heights represent % of species among the total (random stereochemistry). For comparison to molecular species separated by argentation TLC the molecular species are grouped based on the individual FA number of double bonds: Total saturates (16:0, 18:0, 20:0, etc), S; monoenes (18:1), M; dienes (18:2), D; trienes (18:3), T. Actual molecular species measurements of the non grouped lipid molecular species for TAG, DAG, and PC are in Appendix B Tables 6-1, 6-2, 6-3, respectively. A, DAG (black bars) and PC (white bars) were analyzed as their sn- 1,2-diacyl-3-acetyl-glycerol derivatives. B, TAG. 79 for sn-3 acylation, then ~1.7 nmoles FA min“ embryo" flux through the DAG moiety precursor pool into TAG. Therefore turnover time of the combined DAG/PC pool (1500 nmoles FA) would be 882 min (14.7 hrs) if all the DAG and PC takes part in acyl metabolism. Thus our assays of 2-30 min will be effective measures of initial rates and products but will not directly measure the long term precursor-product relationships of FA flux through the DAG moiety into TAG. However, these longer term relationships may be inferred through full analysis of the molecular species of products observed and consideration of the end point compositions. [“ClAcetate labeled fatty acid products [“Clacetate is an ideal substrate for the study of acyl lipid metabolism because it rapidly enters plant tissue and is highly specific for incorporation into newly synthesized FA . Transmethylation of total lipid extracts resulted in >96% of the radioactivity recovered in the hexane soluble fraction. TLC analysis of total lipids at 10 min of labeling indicated <9% of the radioactivity migrated with free sterols or waxes. Thus, >87% of [“Clacetate incorporated into soybean lipids is in the acyl moiety of glycerolipids (examples of TLC separation of labeled lipid classes are in Appendix B Figure 6-3). The rate of [“Clacetate incorporation into soybean lipid acyl fraction was 3.2 pmol [“C]acetate embryo“ day‘1 (data taken from experiment shown in Figure 3-2), and is approximately 1/10 of total embryo fatty acid synthesis rate expressed on a C2 unit incorporation basis (35 pmol acetyl-CoA embryo'1 day'1 [170]. Therefore, assuming an equal distribution and utilization of acetate within the embryo tissue, statistically most of the fatty acids 80 synthesized during the labeling period contained a single [“C]C2 unit. [“C]Oleate (18:1) and [‘4clsaturated FA (16:0, 18:0) (Appendix B Figure 6-1) are the immediate products of fatty acid synthesis. At 30 min of labeling less than 3% of the labeled FA were desaturated to [“Cllinoleate (18:2); therefore short time point labeling represents the initial acyl transferase reactions of newly synthesized FA, prior to further metabolism. Total labeled saturates increased from 23% of total FA at 2 min to 42% by 30 min of labeling. Most of the increase was due to an increase in stearic acid (18:0) labeling. Chemical or-oxidation was employed to determine the position of the label along the [“C]18:0 acyl chain (Appendix B Figure 6-2). Label was evenly distributed along the chain. This result suggested that the 18:0 was a direct product of fatty acid synthesis and not labeled by a high specific activity cytosolic elongation as reported for other very long chain FA [171, 172]. The reason for 18:0>16:0 production when endogenous composition is 16:0>18:0 is unknown. The 18:0>16:0 labeling at early time points in soybean embryos is not expected to affect the total saturates acyl group fluxes as 16:0 and 18:0 behave in a similar fashion in acyl transfer reactions, and have the similar positional distributions in the major lipid classes. Kinetics of [“Clacetate labeling indicate independent rates of lipid class acylation To determine the initial rates and precursor-product relationships for the incorporation of newly synthesized FA into soybean glycerolipids we followed 81 35 5 30 : PC 5‘ 25 . E g 20 - .- TAG tc . . § 15 I“ - 9' 1o . a 5 - DAG g 0 3 r T I I l I C O 5 10 15 20 25 30 35 Minutes Figure 3-2 Time course for [“Clacetate incorporation into the acyl groups of the major labeled soybean embryo lipids. Results are based on the total radioactivity in each lipid class isolated by TLC and measured by liquid scintillation counting. Results expressed as nmol of ["C]acetate incorporated per embryo for each lipid. Glycerolipid classes are PC, diamonds; TAG, triangles; DAG, squares. Time points in minutes are, 2, 4, 6, 10, 20, 30 and 120. The graph shows the first 35 min to focus on initial kinetics. 82 [“Clacetate labeling into the major labeled soybean embryo lipids (Figure 3-2). At all time points PC, DAG and TAG represent ~85% of labeled glycerolipids. PC labeling was linear and gave the highest initial rate (1.15 nmol [“Clacetate embryo“ min“). DAG labeling was also linear over the 30 minute period and gave a lower initial rate (0.36 nmol [“Clacetate embryo“ min“). The initial rate of TAG labeling was the lowest of the three major species (0.26 nmol [“Clacetate embryo“ min“), however TAG labeling accelerated and accumulated more total radioactivity than DAG over time. Other membrane lipids contained very little of the newly synthesized FA. Phosphatidylethanolamine (PE) contained less than 8% of the radioactivity observed in PC even though PE mass abundance is approximately half that of PC. Phosphatidic acid (PA) is an intermediate of de novo glycerolipid synthesis, but PA labeling was barely detectable (less than 2% of that for PC). The high rate of fatty acid incorporation into PC is similar to acyl editing described in pea leaves where PC is the first incorporation product of chloroplast exported nascent FA [162]. The rate of incorporation of nascent FA into PC is independent of the incorporation of nascent FA into DAG or TAG indicating independent acylation reactions. As it is reasonable to assume that the specific radioactivity of the fatty acids for each process is identical, the relative initial acylation rates are about 10:3:2 respectively. The accelerating rate for TAG synthesis over the 30 minute period may represent the combination of three possible effects: (i) increasing contribution from very long-chain saturates, (ii) an accelerated initial rate, possibly due to higher levels of stearate, and (iii) the movement of labeled 83 diacylglycerol moieties from DAG and PC pools to TAG becomes measurable, although longer term assays will be required to detect concomitant reductions in DAG and PC labeling. In the next five sections we describe the acetate labeled PC, DAG and TAG products in more detail and summarize the product structures and kinetic results. [“ClAcetate labeled FA compositions differ among glycerolipid classes Fatty acid methyl esters (FAME) of PC, DAG and TAG were separated by argentation TLC to determine [“ClFA composition based on the number of double bonds (Figure 3-3). At 2 min of labeling [14011821 made up >95% of the labeled FA in PC and total labeled saturates slowly increased to approximately 25% at 30 min. By contrast total labeled saturates were relatively constant in DAG and TAG over the first 30 min of labeling. The average amount of saturate radioactivity between 4 and 30 min of labeling is 29% in DAG and 75% in TAG. The lower amount of labeled saturates at 2 min most likely represented acyl group pool filling. By 120 min radioactivity is detected in polyunsaturated fatty acids (PUFA) in all lipids, and the relative proportions of saturates have dropped as the newly synthesized FA are slowly equilibrated through metabolism to the endogenous FA compositions. The very different initial [“ClFA compositions among PC, DAG and TAG reflects different metabolic processes for incorporation of nascent FA into glycerolipids, which is consistent with their independent acyl labeling kinetics. 84 .3.’ % [1 4C] FAME % [14C]FAME F? % [14C]FAME A res-0500017 000000 80 40 20 TAG 80 6O 40 20 6 10 Minutes 20 30 120 6 10 Minutes 20 30 120 6 10 Minutes 20 30 120 Figure 3-3 Fatty acid composition of [“Clacetate labeled lipids. A-C. Lipid classes from the time course assays were transmethylated and the FAME separated based on the number of double bonds by AgNOa-TLC, and quantified by electronic autoradiography. Each bar shows the distribution of radioactivity among different acyl groups at each time point. Bar shading: total saturates, S, black; monoenes (18:1), M, white; dienes (18:2), D, gray. A, PC; B, DAG; C, TAG. Differential positional acylation of [“Clacetate labeled FA among lipid classes The regiochemistry of labeled FA distributions in PC, DAG and TAG was determined by enzymatic digestion (Figure 3-4). At 2 min 86% of the newly synthesized FA in PC was esterified at the sn-2 position (Figure 3-4A). Over 10 min of labeling, the label at sn-1 increased from 14% to 26%, which was mostly due to an increase in saturates. The unequal distribution of nascent FA incorporation into PC is reminiscent of our previous results in pea leaves at 5 minutes, in which acyl editing produced PC labeled at sn-1 and sn-2 positions of 38 % and 62%, respectively [162]. However, acyl editing in soybean embryos appears to be more sn-2 specific than in pea leaves. The increase in PC sn-1 saturate labeling may reflect a slow conversion of de novo synthesized DAG into PC. An alternative explanation is that the high [“C]sn-2 composition of early labeled PC may reflect a large amount of sn-1 re-acylation with unlabeled recycled saturated FA. As nascent [“Clstearate replaces unlabeled palmitate in the acyl editing FA pool an increasing amount of labeled saturates are incorporated into the sn-1 position of PC over time. In this scenario sn-1 acyl editing may be just as active as sn-2 acyl editing. The position of [14C]FA acylation in DAG differs significantly from PC, in that at 2 min of labeling there is more sn-1 labeled FA than sn-2 (Figure3-43). The relative sn-1 to sn-2 positional labeling equilibrated in DAG by 20 min. Two possible causes for this equilibration are: (i) equilibration of newly synthesized FA into a reacycled acyl group pool feeding de novo glycerolipid synthesis; or 86 E? % [14C]FAME % [14C]FAME % 14C PC ID UM IS 100 80- H " — I 60- 40- 20- n a ll O‘F'Di T ‘T ‘1‘ ‘7 ‘1‘ '7 ‘1‘ ‘T ‘1' C C C C C C C C In in In in m (D (D m 2 min 4 min 6 min 10 min DAG I 100 ID UM S 801 60- 40‘ ' " 20~ 0. ‘TS'T‘P‘TS‘TS‘TW‘TS‘ CCCCCCCCCCCC mmmmmmmmwmmm Min: 2 4 6 10 20 30 ' TAG IFFA DDAG IMAG 100 80- 60- 40- 205 0. 2 4 6 10 20 30 Minutes 87 Figure 3-4 Positional analysis of [“Clacetate labeled acyl groups in different lipids. Major labeled lipids from the [“Clacetate labeling time course were subject to lipase positional analysis. Bar heights represent percent of lipid label in that position. A-B. Bar shading represents the labeled FA as a percent of the whole lipid: total saturates, S, black; monoenes, M, (18:1), white; dienes, D, (18:2), gray. A, PC PLAz digestion. FAMEs generated from lipase products lyso-PC (sn-1) and FFA (sn-2) were further analyzed by AgNOa-TLC. B, Acetylated DAG pancreatic 1,3 TAG lipase digestion. FAMEs generated from lipase products, FFA (Sh-1) and 1-O-2- acyl-3-acetyl glycerol (sn-2) were further analyzed by AgNOa-TLC. C. TAG pancreatic 1,3 TAG lipase digestion. Bar shading: FFA (sn-1 or sn-3), black; DAG (sn-1,2 DAG or sn- 2,3 DAG), white; MAG (sn-2), gray. (ii) highly sn-2 labeled PC conversion to DAG to balance out the higher sn-1 de novo synthesized DAG. Since PC labeling at 20 min is over 4 times that of DAG (Figure 3-2), the proportion of labeled PC moving back to DAG must be minor othenlvise the DAG stereochemistry would reflect PC stereochemistry. The results of the positional analysis of nascent [14C]FA incorporation into TAG differ from both that of PC and DAG. Lipase digestion of TAG indicates that almost the entire label at early time points was sn-1 or sn-3 (lipase product: free fatty acid, Figure 3-4C). Labeled FA can be incorporated into only the sn-3 position of TAG by acylation of an unlabeled DAG; however any sn-1 labeling must come from de novo synthesis of DAG. Since the DAG regiochemistry is approximately equal at sn-1 and sn-2 positions, the amount of label at sn-1 of TAG is predicted to be at most equal to the amount of label at sn-2. However, the end product of the lipase digestion, MAG, which contains the sn-2 label, is less than 2% of the total end products (Figure 3-4). If any of the highly sn-2 labeled PC has been converted to DAG and then TAG, this would further reduce the amount of label at sn-1. Therefore we conclude that initial incorporation of nascent FA into TAG is almost completely localized to the sn-3 position. 88 A: 2 min labeled PC 40 35 q - ID 30. EIM 25- ‘ 'S 20- 15- 1o- 50 f] 0'1 1 .» F— 2 U) <08 5% % 14C MD .— (D B: 6 min labeled DAG 40 35 1 ID 30 . BM 25 ~ ' IS 20 a 15 - 10 d 3; l .ll. % 14C SS SD ST SM MM MD MT DD DT TT Figure 3-5 Molecular species compositions of [“Clacetate labeled PC and DAG. A-B, molecular species are represented as pairs of FA, with bar heights represent % of species among the total labeled lipid. Bar shading represent % of [1 C]FA among species. In species with only one color it means that only the colored FA is labeled the other FA within the molecular species is unlabeled. Bar shading: S, total saturates, black bars; M, monoenes (18:1), white bars; D, dienes (18:2), gray bars. Molecular species of labeled lipids were determined for the earliest time points that had enough radioactivity for the analysis. Examples of TLC systems used in molecular species analysis are in Appendix B Figure 6-3. A, 2 min labeled PC. B, 6 min labeled DAG. Additional time point labeled molecular species compositions for PC and DAG are in Appendix B Figures 6-4 and 6-5, respectively. 89 [“Clacetate labeled PC and DAG molecular species contain one newly synthesized and one previously synthesized FA The unequal [“C]FA positional labeling of DAG and PC suggest at least some nascent, labeled FA are incorporated into DAG and PC next to an unlabeled FA. To further investigate this facet of product structure the molecular species of [“C]acetate labeled PC and DAG were analyzed (Figure 3-5). Examples of the TLC systems used in molecular species analysis are given in Appendix B Figure 6-3, while the shorthand nomenclature used to describe DAG, PC and TAG molecular species is contained in the legend to Figure 3-1. At 2 min the labeled molecular species SM made up 37% of total radioactivity in PC, with almost all the label as [“C]M (Figure 3-5A). Thus 35% of the total PC must be [‘Zcisr‘tcw, while only up to 2% of total PC could be the dual labeled species [“C]S[“C]M. The majority of other species (40%) consisted of unlabeled PUFA alongside a nascent labeled S or M. Assigning the labeled MM molecular species of PC (22%) precisely as single or dual labeled is more problematic. Given that labeled oleate was highly incorporated at the sn-2 position in total PC the possibility that MM was largely the dual labeled species is highly unlikely. Furthermore, at least 76% of nascent FA were incorporated into PC alongside unlabeled S, D or T in proportions relative to their endogenous composition and the position of acylation. Endogenous PC contained 30% M (Table 1), which was almost as much as S and T together. Therefore it is highly likely that nascent 90 [”018 and [“C]M were mostly incorporated next to endogenous unlabeled M in the labeled SM and MM molecular species, respectively. [14C]Fatty acid positional analysis of DAG indicated that up to 90% of molecular species could be produced solely from matching sn-1 and sn-2 acylations with nascent [“018 and [“ClM (Figure 3-4B). However, molecular species of [“C]acetate labeled DAG at 6 min (3- 58) indicated that 56% of initial labeled molecular species contained a nascent labeled FA next to an unlabeled PUFA. In the labeled DAG molecular species SM, the unequal S and M labeling indicated that at least an additional 15% of total DAG is [‘ZC]S[“C]M. Therefore molecular species analysis of DAG indicates that at least 71% of DAG contains a newly synthesized FA incorporated in the same molecule as previously synthesized FA. As with the PC, this analysis cannot reveal how much of the remaining SM and MM are dual labeled but it is probably much less than 29%. We conclude that initial [“C]acetate labeled PC and DAG are comprised predominantly of molecular species that contain one labeled newly synthesized FA and one unlabeled previously synthesized FA. [“ClAcetate labeled TAG molecular species indicates bulk DAG was used for sn-3 acylation with nascent [“C]FA Six minutes is the earliest time point where enough radioactivity has accumulated in TAG to conduct a full molecular species analysis. At this time the TAG molecular species distribution labeled with nascent FA does not match the 91 A: To 18 16 - 14 - 12 r 1 “lo 14C ONAODQO I l tal 6 min [14C]acetate TAG ITotal14C UESI-MS/MS mEDPEDFOPEEDPOPtDFt (D a) o o o magmgaggwwggggzzgoo B: [14o]s and mom labeled TAG C: [12C]DAG from [14C]acetate TAG 25 I 120 DAG D DAG ESI I PC ESI SS SM SD ST MM MD MT DD DT TT 92 Figure 3-6 Six min [“clacetate labeled TAG molecular species. A, Total 6 min [14C]acetate labeled molecular species (Black bars), compared to endogenous molecular species (White bars).B, 6 min [“Clacetate labeled TAG molecular species with [“C]FA labeling. Bar shading represents the labeled FA within each molecular species. Saturates, 8, black; monoenes (18:1), M, white; dienes (18:2), D, gray; trienes, T. C, the calculated unlabeled DAG molecular species associated with labeled TAG in panel B, assuming one sn-3 labeled FA per molecular species, black bars. Endogenous molecular species from Figure 3-1, DAG (white bars), PC (gray bars). endogenous molecular species distribution (Figure 3-6A) but, as expected, is enriched in species containing 8 and M and is deficient in species containing two or three polyunsaturated fatty acids. The positional analysis described above indicated that nascent FA incorporation into TAG was primarily in the sn-3 position. Therefore the [14C]acetate-labeled TAG molecular species in Figure 3- 6A each contain one labeled S or M and two unlabeled FA. The relative proportions of [“013 and [“ClM labeled TAG are shown in Figure 353. The calculated unlabeled ['ZClDAG molecular species composition associated with the sn-3 [14C]FA labeled TAG was very similar to the endogenous bulk DAG and PC molecular species compositions (Figure 3-6C). Therefore it appears that DAG generated from the bulk PC pool is utilized for TAG synthesis with newly synthesized FA. Furthermore, inspection of Figure 3-68 reveals that there is little DAG selectivity when considering either [“cls or [“clm individually for esterification at the sn-3 position of DAG. However, this does not rule out selectivity of DAG molecular species for esterification of 18:2 or 18:3 to the sn-3 position. Conclusions from [“Clacetate labeling The [“Clacetate labeling kinetics, and the different labeled FA composition, positional acylation and molecular species of DAG, PC and TAG reveal that newly synthesized DAG is not a direct immediate precursor for nascent FA incorporation into PC or TAG. Three different acyltransferase pathways (de novo sn-1 and sn-2 acylations to produce DAG, PC acyl editing, 93 and sn-3 acyl transfer to produce TAG) act independently of each other for the incorporation of nascent FA into glycerolipids. This clarifies ambiguities in previous seed labeling studies about the overlap of these pathways, at least for developing soybean embryos. The rapid labeling of PC largely at the sn-2 position suggests an acyl editing mechanism in which FA is largely removed from the sn—2 position of PC to generate sn-1-acyl-2-Iyso-PC, which is then re- esterified with nascent FA. The GPAT and LPAT of de novo glycerolipid synthesis utilize nascent acyl groups mixed with unlabeled acyl groups to produce the equal regiochemical labeling and mixed molecular species labeling of [“C]aoetate labeled DAG. Presumably some of the unlabeled acyl groups released from PC during acyl editing are recycled for de novo glycerolipid synthesis. The labeled acyl composition, positional, and molecular species analyses of TAG at early time points demonstrate clearly that nascent FA are esterified to the sn-3 position of DAG by an acyltransferase system that is selective towards saturates but not to DAG molecular species. Furthermore, the DAG is likely derived from the bulk PC pool. Direct incorporation of nascent FA by PC acyl editing, de novo DAG synthesis and TAG synthesis from preformed DAG together account for about 85% of all newly synthesized FA incorporated into cytosolic glycerolipids, while the ratio of initial rates of incorporation (approximately 10:32, respectively), demonstrates that the major flux is a rapid acyl editing mechanism with PC, similar to our previous results in pea leaves [162]. 94 [“Clglycerol labeling of lipid backbones and acyl chains indicate a common backbone pathway and separate acylation pathways among lipid classes Lipids produced by de novo glycerolipid synthesis can be tracked with [“Clglycerol, which is rapidly taken up by developing soybeans and incorporated into the backbone of glycerolipids through G3P acylation. The polar head-group of PC is typically not labeled from glycerol in short duration incubations (lasting a few hours or less) [86, 162]. However, in addition to backbone labeling, acyl groups are also rapidly labeled because G3P can provide precursors for plastidic acetyl-CoA synthesis. The glycerol backbone and acyl moiety labeling can be measured separately after lipid transmethylation. For cultured soybean embryos the labeled nascent FA composition produced by [“C]glycerol was very similar to that from [“Clacetate (data not shown) indicating that de novo fatty acid synthesis products were not dependent on the substrate. The kinetics of [“Clglycerol backbone labeling of PC, DAG and TAG (Figure 3-7A) was distinctly different from acyl labeling of glycerolipids by [14C]acetate (Figure 3-2). At 2 min of [“C]glycerol labeling the backbone incorporation into DAG was already linear and was approximately four times the initial rate for PC and more than eight times that of TAG. However, PC [“Clbackbone labeling rapidly accelerated such that by 30 min PC contained more label than DAG and by 120 min it was approaching the ~2:1 molar ratio of endogenous PC:DAG. Thus the DAG and PC labeling showed the expected 95 A: [“Clglycerol backbone labeling .- 700 g, 600 - E 500 ~ g 400 « DAG 3 300 - E 200 - PC 1 3 100 i TAG E. 03 r T r r E o 5 1o 15 20 25 30 35 Minutes B: Relative acyllbackbone labeling 45 40 . IPC 35 - D DAG 30 ' ITAG 25 ' 20 - 15 s 10 ' 5 . 0 . % [14C]acyl of total lipid 2 6 1O 20 30 120 Minutes Figure 3-7 Time course for the incorporation of [“Clglycerol into the backbone and acyl groups of the major labeled soybean embryo lipids. [“Clglycerol labeled lipids were isolated by TLC and transmethylated to separate the glycerol moiety from the acyl groups, each fraction quantified by liquid scintillation counting. A, time course of only glycerol backbone moiety labeling. Results expressed as pmol [“Clglycerol incorporated per embryo over the time course (2, 6, 10, 20, 30 120 min). The graph shows the first 35 min to focus on initial kinetics. B, % of total lipid [“clglycerol labeling that is in the acyl chains over the time course. Bar shading: PC, black bars; DAG, white bars; TAG, gray bars. 96 precursor-product relationship kinetics. Incorporation of [“C]glycerol into the backbone of TAG lagged significantly behind that of DAG and PC. TAG labeling was still accelerating after 120 min but had not reached the level of labeling observed for PC or DAG even though endogenous levels of TAG accumulate to ~20 times that of PC. The extreme lag of TAG (as compared to PC) behind DAG labeling suggests that the bulk of TAG synthesis (> 98%) is not related to rapidly labeled de novo synthesis of DAG. TAG labeling is more consistent with a precursor-product relationship with PC. The backbone labeling kinetics in developing soybeans are consistent with a major TAG synthesis pathway of G3P to DAG, to PC, to DAG, and finally to TAG and complement the molecular species analysis of [“Clacetate labeled TAG, that had an unlabeled DAG moiety very similar to bulk DAG and PC. The Discussion elaborates on this hypothesis and includes evidence from mathematical modeling of the labeling data through different possible pathways. Approximately 5-10% of [“C]glycerol is rapidly incorporated into acyl chains. Steady-state [“C]glycerol labeling of acyl and backbone precursor pools coincides with a constant acyl to backbone labeling ratio. The dual acyl/backbone labeling from [“C]glycerol confirms the conclusions made from separate [14C]acetate acyl and [“C]glycerol backbone labeling experiments. Figure 3-7B shows the percent of acyl labeling from [“C]glycerol acyl/backbone labeled lipids. At 2 min PC contains 22% of the label in the acyl chains, a value which drops to a steady-state level of ~5% by 20 min. This is consistent with rapid incorporation of nascent FA into PC through acyl editing while the backbone label lags through 97 the de novo glycerolipid synthesis via the DAG pool. The constant ~5% acyl labeling of DAG over the time course represents nascent FA and backbone entering glycerolipids together by de novo G3P acylation. Initially TAG contains the largest proportion of radiolabel in acyl chains and it takes over 120 min to equilibrate to the level of PC and DAG. This result supports the hypothesis that nascent FA can be esterified at the sn-3 position to a DAG pool that is not immediately produced from de novo synthesized DAG and that TAG steady-state labeling is consistent with backbone lagging through the large PClDAG bulk pool. Analysis of acyl chains utilized by de novo glycerolipid synthesis through molecular species separation of [“Clglycerol backbone labeled lipid classes Separation of the backbone-labeled lipid molecular species at early time points allows for analysis of the acyl groups used for acylation of G3P. The initial molecular species of backbone-labeled DAG and PC are very similar (Figure 3-8 A-B, respectively), indicating de novo synthesized DAG is converted to PC without molecular species selectivity. The slightly higher SD and lower MM levels in labeled DAG compared to labeled PC might indicate a differential utilization of DAG molecular species by a low flux pathway, such as for PE or Pl synthesis. The newly synthesized backbone labeled molecular species composition (Figure 3-8A, B) of DAG and PC contain relatively less polyunsaturated molecular species than the endogenous PC and DAG molecular species, 98 A: [14C]glycerol labeled DAG 35 .6 min 140 _ 3° ‘ DESI-MS/MS gn- fw- O o 15 - I". °\o 10 "' 5 . 0 q I 88 SM so ST MM MD MT DD DT TT B: [14C]glycerol labeled PC 35 30* I6 min 14C D ESI-MS/MS SS SM SD ST MM MD MT DD DT TI' C: [14C]glycerol labeled TAG 16 14. I30min14C -12 . DESI-MS/MS o 310- 58' 06' 3... =22: ol WZOl-ZDl-Dl-tZOl-Ol-tOI-I: a) a) o o o w$8w3558mw§§§§228oo 99 Figure 3-8 Molecular species composition of [“Clglycerol backbone labeled PC, DAG and TAG. A-C, molecular species are represented as a combination of two or three FA. Total saturates, S; monoenes (18:1), M; dienes (18:2), D; trienes (18:3), T. [14C]glycerol labeled molecular species are backbone labeling only (acyl chain labeling has been subtracted), black bars. Molecular species were determined at the earliest time point with enough radioactivity for analysis (Figure 3-7A). The endogenous mass molecular species compositions are from ESI-MS/MS measurements, white bars. A, 6 min labeled DAG, 10 & 30 min time points are in Appendix B Figure 6-7B. B, 6 min labeled PC, 10 & 30 min time points in Appendix Figure 6-7A. C, 30 min labeled TAG was the minimum labeling time with enough radioactivity for analysis. respectively. Because acyl group desaturation is believed to occur on PC, but not DAG, de novo DAG has to be converted to PC for further desaturation. And as mentioned earlier the endogenous bulk PC and bulk DAG molecular species compositions are very similar (Figure 3-1A). From this deduction and the observation that the bulk DAG pool molecular species profile (Figure 3-1 A) is similar to that used for TAG synthesis (Figure 3-6B) and not the de novo synthesized DAG composition (Figure 3-8A), we conclude that the immediately de novo synthesized DAG can contribute only a relatively small fraction to the total endogenous DAG pool. Instead, the bulk of the endogenous DAG must be generated by another route. That route involves the removal of the phosphocholine head-group from PC to produce DAG, by one of several possible mechanisms such as a reversal of the cholinephosphotransferase reaction or by phospholipase C action. [“ClGlycerol labeling of the TAG backbone was insufficient for molecular species analysis until 30 min of labeling. However, at 30 min the [14C]glycerol backbone labeled TAG molecular species (Figure 3-80) showed a correspondence to that of initial [“Clglycerol labeled PC and DAG in that the labeled molecular species composition was relatively less polyunsaturated as compared to the endogenous composition of the corresponding lipid. The calculated FA composition esterified to backbone labeled PC, DAG and TAG was also very similar (Appendix B Figure 6-8). However, the FA composition of TAG labeled with [14C]glycerol in the backbone was enriched in S and M, and depleted 100 in D and T, when compared to the FA composition of endogenous TAG. Further interpretation of these results is left to the Discussion section. 3.4 Discussion The goal of this study is to provide a quantitative analysis of acyl fluxes in the developing embryo of an oilseed, from the point where free fatty acids are exported from the plastid to the formation of TAG. It is a timely exercise, given the advances in identification of acyltransferase genes in recent years. The literature describing in vivo labeling of developing seeds suggests various metabolic pathways. However, collectively, it suffers from a number of problems. (1) The tissue is excised and immediately incubated, often in the absence of an osmoticum or nutrients. Thus its physiological state relative to the in planta situation is uncertain. (2) It has been particularly difficult to assess flux directions and pools for the DAG-PC conversions. (3) Interpretation based on initial rates of reaction is under-utilized, and there is often a lack of comprehensive product analysis. (4) Quantitative conclusions have rarely been possible. In order to undertake initial rate studies we have used C2/C3 precursors. These are rapidly taken up and utilized, allowing the system to quickly reach a (quasi) steady state labeling condition and hence the opportunity to obtain initial rate data. By contrast labeled hexose substrates face very large dilutions and take too long to reach a steady state. At the acetate concentrations employed a significant 101 absolute rate of incorporation (about 10%) relative to the endogenous rate of C2 utilization for fatty acid synthesis was measured. Newly synthesized FA are incorporated into cytosolic glycerolipids through three independent acyltransferase pathways Through analysis of the kinetics of glycerolipid acyl labeling from acetate, and analysis of the FA composition, position of acylation, and molecular species of these products, we demonstrate that three different acyltransferase systems are responsible for incorporation of newly synthesized FA into cytosolic glycerolipids. 1) The major flux of nascent FA is through acyl editing of PC. The highest rate of FA incorporation was into PC. There was no detectable kinetic lag to the onset of steady-state labeling of PC, indicating that PC is the immediate product of an acyltransferase utilizing nascent FA. Additionally the labeled FA composition and stereochemistry of PC indicated it was not related to the initial labeled DAG, a possible precursor. Molecular species and FA positional analyses of PC demonstrated that at least 80% of newly synthesized FA are esterified in the same molecule as pre-existing FA. Finally, [14C]glycerol acyl and backbone labeling showed that initial labeled acyl groups and backbone are incorporated into PC independently. Together these results reveal that about 57% of newly synthesized FA are directly esterified to the sn-1 or sn-2 position of PC through acyl-editing mechanisms. Because GPCAT activity has not been described in developing oilseeds, whereas microsomal sn-2 LPCAT and acyl exchange activities are high, we ascribe the acyl editing to acylation of lyso-PC 102 [130, 140, 142, 144, 163]. 2) Newly synthesized FA mixed with recycled FA to acylate G3P for de novo glycerolipid synthesis. Approximately 17% of newly synthesized FAs in glycerolipids were incorporated into DAG, approximately equally distributed between the sn-1 and sn-2 positions. At least 71% of DAG molecular species contained newly synthesized and pre-existing FA in the same molecule. [14C]glycerol acyl/backbone labeling demonstrated that nascent acyl groups and backbone are incorporated into DAG simultaneously. These results demonstrate that the GPAT and LPAT of the eukaryotic de novo glycerolipid synthesis pathway utilize a mixed pool of acyl-CoA containing newly synthesized and recycled FA. The recycled FA most likely originate from PC acyl editing. 3) Newly synthesized FA were utilized for sn-3 acylation of pre-existing DAG to generate TAG. Approximately 11% of newly synthesized FA was directly incorporated into the sn-3 position of TAG, with negligible labeling in the sn-1 and sn-2 positions. Nascent FA incorporation into TAG was kinetically independent of de novo DAG synthesis. TAG molecular species analysis indicated that the DAG used had the same composition as the endogenous bulk DAG derived from PC. Therefore the direct incorporation of nascent FA into TAG utilized a pre-existing DAG pool, not de novo synthesized DAG. In a later section below we use the relative rates of these three independent acyl transferase pathways to construct a more quantitative model of TAG synthesis in soybeans. This model takes account of PC acyl editing, which has not be previously included in models of oilseed TAG metabolism. 103 At least two systems of DAG sn-3 acylation producing TAG can be idenfified The first flux, as mentioned above, involves 11% of the total FA labeling that is directly incorporated into the sn-3 position of TAG. Other nascent FA are incorporated into glycerol lipids through acyl editing of PC, and de novo DAG synthesis which constitute approximately 57% and 17% of the total fatty acid labeling, respectively. The remaining 15% of acyl label unaccounted in the above three glycerolipid classes is present largely in minor phospholipid species. As TAG constitutes approximately 93% of the endogenous acyl lipid mass, eventually about 31% of the nascent fatty acids must end up in the sn—3 position of TAG. Thus about 20% of the nascent fatty acids incorporated into glycerolipids other than TAG must eventually move through into the sn-3 position of TAG to add to the 11% that are immediately incorporated. This represents a second, distinctive flux of FA into the sn-3 position of TAG. When compared to endogenous TAG, the molecular species analysis of TAG from [14C]acetate labeling (Figure 3-60) also strongly suggests two pathways to TAG, of approximately equal contribution. We estimated that TAG labeled with nascent FA at the sn-3 position provides approximately 35% of total TAG synthesis and has a total FA composition of 39% S, 29% M, 25% D and 7% T, with the unlabeled DAG coming from the endogenous DAG pool (Figure 3-63) and sn-3 made up of 75% S and 25% M (Figure 3-3C). Since the endogenous TAG composition is 20% S, 27% M, 41% D and 12% T, it is likely that a separate contribution to TAG synthesis is required to provide the remaining 65% of TAG. 104 This separate TAG synthesis must produce TAG of the composition 10% S, 26% M, 49% D and 15% T to balance out the endogenous composition. It is clear that this second TAG synthesis component is very different, containing reduced saturates and much higher PUFA than the initial [14C]acetate labeled TAG. The DAG moiety presumably comes from the bulk unsaturated DAG/PC pool, with a selectivity for sn-3 acylation for PUFA. The molecular species analysis of [‘4C]glycerol labeled TAG (Figure 3-8C) is also consistent with two pathways to TAG. We can group the molecular species of TAG into three groups; (1) high specific activity species SSS, SSM, SSD, SST, SMM, MMM (about 16% of TAG endogenous mass); (2) low specific activity species MTT, DDD, DDT, DTT (about 20% of endogenous TAG mass) (and possibly STT, MDD and MDT), and (3) the other species, of intermediate specific activity (about 64% of endogenous TAG mass). Clearly, rapidly labeled (i.e. high specific activity) molecular species are dominated by high levels of S and M, consistent with the nascent fatty acid labeling, while slowly labeled species have high levels of D and T. Time is required for further desaturation of the PC species derived from de novo synthesized DAG, to eventually produce DAG species with both positions occupied by PU FA. These can then be utilized by the TAG synthesis system that acylates the sn-3 position with PUFA, to produce TAG molecular species with all three positions occupied by PUFA. A PUFA selective DGAT or PDAT may be involved in this system. A DAG/DAG transacylase [122] has been proposed for the synthesis of TAG. Unsaturated FA would be transferred from the sn-2 position of one DAG 105 molecule to the sn-3 position of another, producing TAG and a sn-1-acyl-MAG. Re-acylation of MAG by MAGAT would bring about a significant enrichment of label in the sn-2 position of DAG. This is not observed. Therefore this pathway is not a major route of TAG synthesis in developing soybean embryos. Interconversion of DAG and PC and the supply of DAG for TAG synthesis It is clear from previous studies on oilseed metabolism that DAG and PC can be inter-converted. This is one way in which PUFA can be enriched for TAG biosynthesis. This inter-conversion is usually postulated to result from a highly reversible CPT [117, 118, 173]. It has been suggested that the PC and DAG pools are in equilibrium. The kinetics of glycerol backbone labeling by soybean embryos shown in Figure 3-7 are similar to those observed for detached linseed cotyledons [127], and provides the strongest evidence that the majority of TAG synthesis is from bulk DAG, via retro-conversion from PC, and not de novo synthesized DAG. If TAG and PC synthesis competed for the de novo DAG pool, with no opportunity for PC conversion back to DAG (Figure 3-9, Model A), TAG should label up 20x more rapidly than PC, to generate the desired ratio of end products. However, the converse is observed: PC labels up about 10x more rapidly than TAG (Figure 3-7A). The observation that the DAG pool used for sn-3 acylation with [14C]acetate labeled nascent acyl groups closely resembles the endogenous DAG and PC molecular species profiles (Figure 3-68), and not that of the de novo DAG pool (Figure 3-88), is also consistent with DAG first moving to PC, then back to DAG. 106 Model A G3P —>__) DAG __> TAG l PC M_od__el B G3P—> _._)DAG __) TAG ll Model B, variant 1 G3P 9 DAG __) TAG ll PC(active) $ PC(bulk) M_o_delc —> <— <— DAG 33" 9D“; —> PC 9 (oil synth) PTAG Model C, variant 1 G3P —_>_;.DAG _> PC _> DAG (oil synth) TTAG Model C, variant 2 cap ‘3‘ DAG -—> PC —> DAG ‘9 (active) (oil synth) —>TAG PC (bulk) Mggg! C, variant 3 G3P —> DAG —> PC —> DAG -> (active) (active) -> TAG PC (bulk) DAG (bulk) 107 Figure 3-9 Representations of mathematical models for TAG synthesis. Mathematical models tested against the kinetic labeling data by running simulations. Descriptions of the simulations are in Appendix B. As mentioned in the text the mathematical model that most fits the acetate and glycerol labeling is Model C, variant 3. The required DAG-PC inter-conversion might be supplied by different mechanisms, each of which may display explicit kinetic consequences. The simplest would be where both PC and DAG are the bulk biosynthetic pools (Figure 3-9, Model B). However, the DAG pool could be split into de novo and oil synthesis (oil synth) pools. Such a general mechanism (Figure 3-9, Model C) might be simplified if the reversible reactions were replaced by unidirectional reactions (Figure 3-9, Model 0, variant 1). Pool filling simulations for Models B and C, and several variants, were run to see how closely they would conform to the experimental kinetics of glycerol backbone labeling shown in Figure 3-7A. The variants included splitting PC and DAG pools into two pools, which would consist of a small biosynthetically active pool in kinetic equilibrium with a large bulk pool, for example (Figure 3-9 Model B, variant 1). Details of these simulations are provided in the Appendix B. Conditions were set to allow a simulation of glycerol backbone labeling from DAG that produced continuous DAG labeling but where PC labeling crossed over DAG labeling at about 30 minutes as in the experimental results (Figure 3-7A). The backbone labeling models were then tested against the acyl chain labeling results (Figure 3-2) to see if, when PC was labeled through acyl editing, how quickly would the acyl label move from PC to DAG. If the equilibration of [14C]acetate acyl label at the sn-1 and sn-2 positions of DAG is from PC conversion to DAG, in 30 minutes of assay no more than about 10% of the labeled acyl groups that accumulated in DAG can originate from PC (which is mostly sn-2 labeled through acyl editing, Figure 3-4A-B). However, the total label in PC is 3 times that of DAG, so only 108 about 3% of the total labeled acyl groups in PC originating from acyl editing can move through to DAG in 30 minutes. Simulations of acyl labeling with Model B failed to meet this criterion. In a simulation of Model B, variant 1 with the acyl editing input through PC(bulk), the movement of labeled acyl groups from acyl editing into DAG was still far too fast. By contrast, models of type C were generally much more successful at describing the DAG-PC precursor-product glycerol labeling kinetics and yet having slow movement of acyl editing labeling of PC move through into DAG. A observation in setting up the model C concerns the possibility of splitting the PC pool into a small, active pool and a large, bulk pool (Figure 3-9, Model C, variant 2). If this is done, then acyl editing must be largely a function of the bulk pool, not the active pool. Although several variants of Model C gave good simulations, one problem was that they predicted movement of glycerol labeling into TAG that was much slower than observed. This could be remedied by splitting the DAG(oil synth) pool into a small active and a large bulk component (Figure 3-9 Model C, variant 3). The de novo DAG pool must be small because it is compositionally distinct from the bulk DAG pool. In addition, in leaves, the endogenous DAG is very small, constituting less than 1% of the PC (Mhaske V. and Pollard M., unpublished data). This DAG presumably encompasses the de novo synthesized DAG pool we propose in Model C. In summary, the kinetics of [“C]glycerol labeling suggest that the major flux of glycerol-3-P through de novo synthesis into DAG is for PC synthesis, with almost none of the de novo synthesized DAG being channeled directly to TAG. The conversion of de novo 109 synthesized DAG to PC is fairly rapid, but residence in the large PC pool is much longer to allow for further desaturation, before conversion back to DAG for TAG synthesis. The de novo synthesized DAG and the DAG pool destined for oil biosynthesis are kinetically distinct and not in rapid equilibrium. A flux model for glycerolipid synthesis in developing soybean embryos Our analysis of glycerolipid acyl group and glycerol backbone labeling allows us to generate a model of the flux of acyl groups during oil synthesis (Figure 3-10). The individual elements of this model have already been discussed in the three preceding sections. A detailed step-by-step logic of model construction, along with its implicit assumptions and calculations, is presented in Appendix B. The flux model structure is based on the kinetic model C, variant 3 (Figure 3-9) which fit the [“Clacetate and [“C]glycerol labeling data. The flux model tracks 100 moles of fatty acids synthesized in the plastid through lipid metabolism over a small time increment. The fluxes are determined based off of initial rates of nascent FA incorporation into extra-plastidic glycerolipids (Figure 3- 2), their composition and molecular species (Figures 3-3, 3-4, 3-5, 3-6), along with determinations of the acyl groups used by de novo glycerolipid synthesis (Figure 3-8), and utilizes endogenous lipid compositions (Table 3-1, Figure 3-1) and mass balance to predict fluxes. Assuming steady state metabolism the end point reached in the developing soybean embryo accumulating oil will be 93 moles of acyl groups in TAG, 1.4 moles in DAG, 3.6 moles in PC and 2 moles in 110 DAG DA G3P 2 i, \i’ 63%4 i 67 .__..._..PG 63.4 . ,. int: 62 . .. j PA a (de ”OI/O) ‘. . . (Oil synth) 7" (Oil synth) . 59 pc (bulk) _ . PDAT 0-20 ' (57) l __ . E l ' FFA] Acyl editing ' Acyl-.CoA __ cycle I 89 I, \ “425427, 0 20 I11 ,. II “x. \ LPC - I (s <—=-' 100 mol FA: ‘- r \ PUFA selective 0W I .\ '16:0,18:0,18:1 ‘ ~ ~ ___________ ' I \ Plastid -..// I Channeled saturates : """""" ' or selective DGAT l l ___________ .. Figure 3-10 Soybean TAG synthesis flux model. The model assumptions and flux calculations are in Appendix B. The model follows 100 moles of newly synthesized FA in the plastid through the major fluxes of lipid metabolism such that final accumulation of FA are TAG 93 moles, PC 3.6 moles, DAG 1.4 moles, other lipids 2 moles. Other membrane lipids in soybean embryos are mostly made up of PE and PI. The arrows demonstrate the flux reactions with the major metabolite pools highlighted in boxes. All of the arrows represent steady-state fluxes, the black arrows are the route that nascent FA take for incorporation into glycerolipids as measured by [14C]acetate labeling. Dashed arrows represent uncertainty for that reaction. The numbers next to flux arrows represent steady-state fluxes or flux regions as calculated in Appendix B and utilizing mass balance for unknowns. The numbers in parentheses within the substrate pools represent the major initial accumulation of nascent FA in glycerolipids PC, DAG and TAG accumulate ~85% of initial FA. The * represents small fluxes of acyl groups to other lipids which is s 2% of the total steady-state acyl groups. ~15% of nascent FA initially accumulate in lipids other than PC, DAG and TAG. 111 other lipids (based on the values in Table 3-1). The model shows the fluxes, or expected flux ranges, for both the newly synthesized fatty acids and for the total flux. It should be noted that in terms of an activity (expressed in moles fatty acid accumulated unit time'1 unit fresh weight") the relative rate of lipid synthesis in soybean is slow (about 5 fold less) compared to other oilseeds such as safflower and rapeseed. Thus the model may or may not be a quantitative or qualitative representation of other oilseeds which accumulate palmitate, stearate, oleate, linoleate and linolenate. The model has a single PC pool. Although it is possible to consider organelles and specific domains of the endomembrane system giving rise to distinct sub-pools of PC, to date we have no kinetic or metabolic evidence to support such a conclusion. By contrast, our analysis strongly suggests three DAG pools. Pools for the immediate product of de novo synthesis of DAG and for TAG synthesis each contain only a small mole fraction of the total DAG, while the bulk DAG pool may be associated with the oil body fraction. We cannot rule out the possibility that a very small fraction of the de novo DAG pool is utilized directly for TAG synthesis by the traditional Kennedy pathway, as indicated by the dotted arrow. Likewise, the reversibility of PC-DAG inter-conversions is not clear: back reactions are indicated as dotted lines. An important aspect of PC metabolism is the acyl editing cycle. This is the largest acyl flux in the cytosol. Our analysis cannot place an exact value on this total flux, but an analysis of saturated fatty acid mass balance suggests a lower limit of over twice the nascent fatty acid incorporation into PC (about 125 moles time increment“), while 112 an upper limit may be over seven fold (about 427 moles time increment"). In considering an acyl editing cycle previously, from the data obtained from pea leaves, we considered the possibility that there was a “channeled” mole of mole exchange of nascent fatty acids with acyl groups released by acyl editing. However, if the acyl editing cycle flux is significantly larger than the incorporation of nascent acyl groups into PC via a PC:Iyso-PC acyl editing cycle, then the simplest way to depict the process is via a bulk acyl-CoA pool where acyl groups released by acyl editing and from de novo fatty acid synthesis mix, as shown in Figure 3-10. A surprise from the labeling was that we could make a kinetic distinction for two TAG synthesis systems. TAG synthesis will require 31 nmoles of acylation at the sn-3 position. About 11 moles are provided by the immediate incorporation of nascent FA, with a high preference for saturates. Analysis of the molecular species of [14C]FA labeled TAG suggests that this direct acylation utilizes the bulk DAG pool, not the de novo synthesized DAG pool. The high sn-3 [”Clsaturated FA labeling of bulk DAG suggest that either a specific pool of acyl- CoAs high in nascent saturates is delivered to a DGAT enzyme, or that the DGAT has a strong selectively for saturates from the bulk acyl group pool (Figure 3-10). At least two mechanisms, both of which are drawn into Figure 10, may provide the remaining 20 mol FA required for TAG synthesis, including essentially all the tri-PUFA TAG molecular species. A PUFA selective DGAT2 may utilize the bulk acyl pool or a PDAT reaction may transfer sn-2 PUFA from PC to DAG, generating TAG. The LPC produced by a PDAT reaction will add an 113 incremental flux to the PC acyl editing cycle (Figure 3-10). It is noteworthy that the stereochemical analysis of TAG from soybean oil shows that the sn-1 and sn- 3 positions have quite similar acyl compositions. Such observations have lead to speculations that the enzymes responsible for TAG synthesis are quite non- specific. However, at least in soybean, there is the coincidental summation of two routes with distinct specificities to give the overall composition. PC has long been considered the site of FA desaturation. In developing soybeans acyl editing, DAG-PC-DAG interconversion and a specific TAG synthesis system all contribute to the biosynthesis of TAG containing PUFA at higher levels than found in PC. Linking the Model to Questions of Biochemistry and Cell Biology Acyl Editing: The LPCAT reactions which are hypothesized to produce the acyl editing measured in pea leaves [162] and soybean embryos was first suggested as a mechanism to allow enrichment of PUFA for TAG synthesis in oilseeds. Subsequently, an in vitro mechanism, acyl exchange, a reversal of the LPCAT reaction, was proposed as a mechanism for acyl editing. It was demonstrated qualitatively by feeding free fatty acids to seeds [169]. In this study acyl editing is demonstrated as a major flux for endogenously synthesized fatty acids. Our previous identification of acyl editing in pea leaves and now in developing soybeans suggests that PC acyl editing may be a ubiquitous part of plant lipid metabolism. TAG synthesis in the cell must accommodate acyl editing, and indeed TAG composition may be significantly modified by its action. However, 114 there are likely to be more fundamental reasons for acyl editing mechanisms. First, it may be important in membrane dynamics and homeostasis. Control of the low level of lyso-PC may be critically important due to its detergent effects on membranes. Second, during periods of high fatty acid synthesis it ensures that levels of saturated-oleoyl and dioleoyl molecular species of membrane phospholipids (which may affect membrane fluidity, especially in the cold [174]) are kept low. Third, a high cycling rate allows for rapid metabolic changes in times of stress. Thus there is an advantage of adaptability for membrane biogenesis. Fourth, cycles in biology add robustness to metabolic networks. Fifth, PC is appears to be the acyl flux bank of the plant cell, rather than other phospholipids such as PE or Pl, which have more important and specific functions. The idea that a PC:Iyso-PC cycle is the principle acyl acceptor system in the cytosol of the plant cell is reinforced by the metabolic analysis we performed on the fatB mutant, where additional acyl flux was directed into PC and PC acyl cycling increased [161]. And finally, acyl editing may allow the removal of oxidized acyl groups in membrane lipids. ln seed, as in leaf [162], PC editing is an order of magnitude greater than PE editing. Also, seed acyl editing of PC at the sn-2 position is particularly dominant over that at the sn-1 position. This coincides with in vitro measurements of acyl exchange in developing seeds, where over 90% of the measured exchange is at the sn-2 position [143]. Assays of lyso-PL acyltransferase activity in microsomes developing soybean cotyledons with exogenous acceptor show a higher Vmax for LPEAT than LPCAT, but with 115 endogenous lyso-Iipid acceptors the LPCAT activity was significantly higher [175]. Thus further experiments are required to reveal whether lyso—PC generation is by way of a phospholipase A and/or a transacylase (eg. acyl exchange by a reversal of LPCAT) mechanism. DAG-PC Inter-Conversion and Localization of DAG pools: In leaf tissue DAG is a very minor glycerolipid. From this observation alone it may be inferred that CPT operates in the direction DAG —+ PC, with very little reverse reaction. As DAG is a component of lipid signaling Cascade [3], cellular control of DAG is likely to be tight. Thus our finding that there are at least two, and probably three, kinetically distinct DAG pools is hardly surprising. CPT would certainly be associated with the de novo DAG synthesis pool. Whether the reverse CPT reaction would also produce the DAG pool required for TAG synthesis is questionable. If so, it may be another gene product or post-translationally modified enzyme. In order for CPT to provide a flux running in the reverse direction from that associated with de novo glycerolipid synthesis would require a CPT with a locally different equilibrium position and /or a DAG removal mechanism such a phase partitioning to another membrane domain. It is also possible that a phospholipase C like activity provides the function of generating a bulk DAG pool for TAG synthesis from bulk PC. Thus there is no requirement for a rapidly reversible DAG-PC inter-conversion. The identification of distinct kinetic DAG pools raises the question of localization. In a comparative study of CPT and DGAT in microsomes from 116 several different seeds [16] found little DAG selectivity for either enzyme. They concluded that the exclusion of unusual fatty acids from membrane lipids was not achieved on the basis of CPT or DGAT specificities and postulated distinctly separate pools of DAG as an explanation. The concept of different endomembrane domains is well known. However, the location of the de novo DAG pool is unclear. This DAG pool may feed into a small, discrete PC pool, or into the bulk PC pool. If the former is the case, then TAG synthesis may occur from the bulk pool (which may itself have a more complex structure than we can observe here). However, if the latter is the case, then TAG synthesis will occur from a sub-domain. Or both DAG pools could each represent separate ER sub- domains, connected by the bulk PC pool. However, the bulk DAG pool is most likely associated with the oil bodies. It is a reasonable hypothesis that DAG is also phase partitioned along with TAG into the oil bodies during their biogenesis. TAG Synthesis: Despite the identification of three classes of genes that encode for enzymes of TAG synthesis in plants, and the observation that DGAT1 and DGAT2 proteins from tung seeds, when tagged, localize to different ER domains in tobacco and onion [133], the identification of kinetically distinct TAG synthesis routes with a preference for saturated or polyunsaturated FA was unexpected. It raises questions about acyltransferase genes, enzyme specificity and acyl-CoA channeling. Considering the synthesis of TAG highly enriched with saturates at the sn- 3 position, this is presumably a DGAT activity, and not PDAT, as it utilizes 117 nascent fatty acids immediately. Furthermore, PDAT would transfer unsaturates from the sn-2 position of PC. However, DGAT1 does not seem strong candidates for this activity. The dgat1 mutant of Arabidopsis has a slight increase in total saturates in TAG, with the fraction in the sn-3 position decreasing only slightly [176]. DGAT2 has been implicated in the production of TAG containing high amounts of novel unsaturated fatty acids, and on this basis also does not seem to be a good candidate. Perhaps another DGAT is responsible? This DGAT may have a very strong selectivity for saturates if it utilizes the mixed acyl- CoA pool. A caveat concerns acyl group channeling via ACBP. The numerous acyl-CoA binding proteins (ACBP) in plants [108-111, 113] may be involved in shuttling acyl groups to different fates. Selective transfer of newly synthesized saturated acyl groups to DGAT via a specific ACBP may have produced the observed kinetics. Considering the synthesis of TAG highly enriched with PUFA at the sn-3 position, including tri-PUFA TAG species, DGAT and PDAT mechanisms may be invoked. Both are shown in Figure 3-10. TAG synthesis via various transacylases of the PDAT family would transfer a sn-2 PUFA from PC to sn-3 DAG producing TAG and lyso-PC [119, 120, 177]. Thus PDAT would leave a footprint indistinguishable from sn-2 acyl editing. However, it cannot be the major producer of lyso—PC for acyl editing because it would not allow recycling of FA to feed de novo glycerolipid synthesis. For a putative DGAT to be involved, it would either have to have a strong preference for PUFA acyl-CoAs, or such molecules would have to be preferentially delivered to the site of the enzyme(s). It is 118 noteworthy that highly sn-2 labeled PC from acyl editing is not immediately available for conversion to DAG and then to TAG, otherwise TAG would contain more sn-Z label at early time points. A simple explanation for this is that the bulk PC pool that fills from acyl editing, creating a kinetic delay in its utilization. 3.5 Conclusion Through in vivo labeling experiments we have shown that an acyl editing cycle and generation of multiple sub pools of DAG are important for the synthesis of TAG in developing soybeans. Similar research in the leaves of peas suggests that acyl editing may be a ubiquitous part of plant acyl lipid metabolism. The individual enzymes/genes that are involved in these processes are unknown and thus represent an important area for further research on plant oil synthesis. As there are many unknown genes in Arabidopsis annotated as acyl transferases [134], further analysis of the acyl editing mechanism to determine if lipases or transacylases are involved may be very useful for final identification of the enzymes involved. Ultimate identification of acyl editing enzymes may require analysis of mutants. The model of newly synthesized acyl group fluxes may be useful in the interpretation of changes in lipid quantities or FA compositions identified during mutant screens or in targeted knockouts. It is also possible that not all the gene products/routes in the metabolism scheme have been identified. In this context, the use of quantitative metabolic analysis taking into account acyl editing with mutant lines may provide further clues to new genes with overlapping roles in the processes. The fact that soybean oil is a major worldwide source of 119 vegetable oil, with genetically engineered oil compositions in commercial production and under development, adds a biotechnology dimension to this study. Our identification of the major flux reactions of acyl groups from synthesis in the plastid to accumulation in TAG may allow a more directed approach toward identifying enzymes that might be useful in the engineering of TAG. The enzymatic reactions involved in acyl editing may also be important for transferring unusual FA from their site of synthesis on sn-2 PC to the three backbone locations of TAG. The synthesis of TAG and phospholipids must be intricately coordinated as both products require the synthesis of DAG. Production of different lipids from DAG may be controlled by utilizing multiple DAG pools in different locations. Identification of the sites and enzymes in each location may allow more efficient engineering of novel lipid metabolizing enzymes to their site of action in oilseed crop plants. Thus, a better understanding of the pathways of TAG biosynthesis including acyl edting and DAG production in developing soybeans may aid future efforts to engineer soybean oil with improved or novel properties. 3.6 Materials and methods Plant material-Developing soybean (Glycine max) embryos were harvested from plants grown in the greenhouse at ~24° C, supplemented with lights to maintain a 15 hr day. Embryos were harvested and cultured in media containing the carbohydrates, amino acids, inorganic salts and light conditions required for embryo growth to mimic in planta development as demonstrated 120 previously [178]. Harvested embryos (~21 mg dry weight/embryo for [“Clacetate labeling and ~15 mg dry weight/embryo for [14C]glycerol labeling) were equilibrated in culture conditions for 2.5 days before starting labeling experiments. RadiochemicaIs—[1-“C]Acetic acid, sodium salt (specific activity 50 mCi/mmol), and [“C(U)]glycerol (specific activity 150 mCi/mmol) were from American Radiolabeled Chemicals, Inc. (St. Louis, MO). [ 1“C]acetate and [14C]glycerol Iabeling- Pre-cultured embryos were moved to a single beaker containing fresh culture media plus radioactive substrate [“C]acetate (1 mM) or [14C]glycerol (0.5 mM) to start the labeling reaction. The media volume was just enough to cover all the embryos and was gently shaken in a water bath at 27° C under 30-40 pmol rn'2 s" of white light. At each time point the labeling reaction was quenched by transfering three embryos (four during [14C]glycerol labeling) to 6 ml of 85° C isopropanol for 10 min. General methods- Lipids were extracted from hot isopropanol quenched tissue with hexane/isopropanol [155] after homogenization with a mortar and pestle. Radioactivity in the total lipid samples, eluted lipids or organic and aqueous phases recovered from transmethylation was quantified by liquid scintillation counting (Beckman Instrument lnc., Fullerton, CA), while radioactivity on TLC plates was visualized and quantified by electronic radiography (Packard Instrument Co., Meriden, CT). AgNO3-TLC plates were prepared by impregnating Partisil® K6 silica gel 60 A TLC plates (Whatman, Maidstone, UK) with 10% AgN03 in acetonitrile (wlv), drying in air and activating at 110 °C for 5 min. 121 Unlabeled fatty acid methyl esters (FAMEs) were quantified by gas chromatography (GC) using a flame ionization detection and a 08-23 capillary column (30 m length x 0.25 mm inner diameter, 0.25 p film thickness; J&W). For preparative TLC all solvents contained 0.01 % butylated hydroxyl-toluene antioxidant. Polar lipid class analysis- polar lipid separation by TLC, recovery of polar lipids from TLC plates, PC molecular species separation, and PC FA positional localization were done as reported previously [162]. Neutral lipid class analysis-DAG and TAG were isolated by first acetylating an aliquot of the total lipids in acetic anhydridelpyridine, (3:2, v/v) and then separating the 1,2-diacyl-3-acetyl-glycerols (acetylated DAG) from TAG by silica TLC developed in hexane/diethyl ether/acetic acid (70:30:1, v/v/v). Individual lipids were eluted from TLC silica by chloroforrnlmethanol (4:1, v/v). Methanol and 0.88% aq. KCl were added to give chloroform/methanol/water ratios of 2:1:1, (v/v/v), resulting in a phase separation. The aqueous phase was back extracted with chloroform and lipids were recovered from the combined chloroform phases. Radiolabeled acyl group analysis-FAM E were prepared from total lipid extracts by heating to 80° C in 5% H2804 in methanol for 60 min. The organic soluble FAME were separated from the aqueous soluble compounds by the addition of water and hexane. FAME were prepared from purified individual lipids by base-catalyzed transmethylation [158]. For [“Clglycerol labeled lipids the proportion of label in the acyl groups versus the backbone/head group was determined by transmethylation [158] and scintillation counting of the separated 122 organic and aqueous phases. Aqueous phase radioactivity from PC was determined to be in the glycerol backbone and not the choline moiety previously [162]. [14C]FAME compositions were determined by separation based on the number of double bonds by AgNOa-TLC, the plates being developed to % height with hexane/diethyl ether (1 :1, vlv), then fully with hexane/diethyl ether (9:1, vlv). Saturated FAME were further separated by removal from AgNO3-TLC with CHClalMeOH (2:1, v/v) and separated on KC18 reversed phase silica gel 60 A TLC plates (Whatman, Maidstone, UK) by development in acetonitrile/methanol/water (65/35/0.5, vlv/v). Molecular species separation of radiolabeled DAG and TA G- Acetylated DAG isolated from TLC was separated into molecular species based on the number of double bonds by AgNOa-TLC [159], using a triple development (‘/2 then ’/4 development in chloroform/methanol (96:4, vlv), then fully in chloroform/methanol (99:1, vlv)). When necessary the saturate/diene and monoene/monoene molecular species were eluted from the silica [159] together and further separated on 15% AgNOa-TLC plates at -20° C with the same triple development as above. TAG molecular species were separated by AgNOa-TLC in two batches both at -20° C. The more saturated species (0-4 double bonds) were separated by triple development (first 60% then 75% development in chloroform/methanol (96:4, v/v) then full in chloroform/methanol (98:2, vlv)). The more unsaturated species (5-9 double bonds) were separated by triple development (60%, 75%, then full development in chloroform/methanol 94:6, v/v). When necessary the saturate/monoene/monoene and 123 saturate/saturate/diene molecular species were eluted together and separated by AgN03 TLC at -20° C by triple development in chloroform/methanol (99:1, vlv). The proportion of radioactivity in each molecular species band for DAG and TAG was determined by electronic autoradiography, then each band was eluted, the recovered lipids transmethylated, and the [‘4C]FAME analyzed by AgNOa-TLC, as described above. Regiochemistry of radiolabeled DAG and TAG-Collected acetylated DAG and TAG were digested with porcine pancreatic lipase (sigma) [159]. The digestion products were separated on silica TLC developed with hexane/diethyl ether/acetic acid (70:30:2, v/v/v). The amount of radioactivity in MAG, DAG, free fatty acids, and TAG was quantified by electrionic autoradiography, then each band was eluted, the recovered lipids transmethylated, and the [14C]FAME analyzed by AgNOa-TLC, as described above. Endogenous molecular species compositions- Total lipids from four soybean embryos (approx 13-14 mg dry weight), cultured under identical conditions to those used for labeling, were extracted according to Hara and Radin (1978). TAG concentrations were measured directly using electrospray ionization mass-spectrometry (ESI-MS). DAG and PC were converted to their sn-1,2-diacyl-3-acetyl-glycerol (ac-TAG) derivatives by digestion with pancreatic lipase C (for PC) followed by acetylation prior to ESI-MS analysis. ESI-MS in positive ion mode was performed by direct infusion with a Shimadzu (Columbia, MD) SlL-5000 autosampler into a Waters (Milford, MA) Quattro micro mass spectrometer. 10pl of sample in toluene was introduced to the electrospray 124 source by flow injection analysis conveyed by a 65:32:3 chloroform:methanol:100mM ammonium acetate solution at a flow rate of 0.1 ml/min. The capillary, extractor, and cone voltages were 3.2 W, 2.0 V, and 40 V respectively. The source and desolvation temperatures were 110 and 350 °C, respectively. The desolvation gas flow rate was 400 l/hr. Mass spectra were collected for 2 min; the m/z range scanned in the MS measurements was from 500 to 1000 (1 sec/scan) and in the MS2 measurements from 20 to the mass of the parent ion. Collision-induced dissociation used argon as the collision gas (2 x 10'3 mbar) with the collision energy set at 22 eV. Mass spectra data was acquired with MassLynx 4.0 software; [T AG-NH4]+ ion peaks were smoothed and integrated using QuanLynx software. In order to correct for the effect of the number of acyl chain carbons and double bonds on the signal strength [179], TAG standards with varying acyl chain length and number of double bonds were analyzed at different concentrations. After correcting for natural isotope abundance effects, the ion peak intensities for each TAG standard were normalized to the internal standard (10uM triheptadecanoin). The normalized peak intensity was plotted against TAG concentration. The slope of this standard curve was determined for each TAG standard. Multiple linear regression was then used to create a correction function relating the slope of the standard curve to the number of acyl chain carbons and double bonds. To determine the concentration of soybean TAG or ac-TAG molecular species, the ion peak intensities were deisotoped and corrected for natural 125 isotope abundance and then normalized to the 10uM triheptadecanoin internal standard. The correction function for acyl chain carbons and double bonds was then applied to determine the absolute concentration of each TAG species. When necessary, ESI-MS2 was used to determine the relative abundance of isomers with the same m/z value. The moles embryo" of each molecular species measured for TAG, DAG, PC are in Appendix B Tables 6-1, 6-2, 6-3, respectively. 3.7 Footnotes Abbreviations: ACP, acyl carrier protein; CPT, CDP-choline:1,2-diacyl-sn-glycerol cholinephosphotransferase ; DAG, sn-1,2-diacylglycerol; DGAT, diacylglycerolzacyl-CoA sn-3 acyltransferase; ER, endoplasmic reticulum; FA, fatty acid; G3P, glycerol-3-phosphate; LPCAT, lyso-phosphatidylcholine:acyl- CoA acyltransferase; PC, phosphatidylcholine; PDAT, (phospholipid:diacylglycerol acyltransferase); PE, phosphatidylethanolamine; PUFA, polyunsaturated fatty acid; TAG, triacylglycerol. 126 4 CONCLUSION The two pathway hypothesis of eukaryotic and prokaryotic glycerolipid assembly (Figure 1-1) was first proposed over 25 years ago based on radiotracer studies of plant tissue and isolated organelles [71, 84]. The age of genetics and molecular biology further confirmed this seminal work [85]. Since that time identification of PC acyl exchange reactions [140], hydrolytic turnover of PC acyl groups [74], channeling of nascent FA from the plastid into PC [79], and the mixing of nascent FA with previously synthesized FA during PC synthesis [149] have suggested that the steps of the eukaryotic pathway may be further refined. A re-examination of plant glycerolipid synthesis requires that new investigations circumvent the limitations of previous studies. Therefore, very rapid (as short as 1 min) in vivo kinetic labeling studies were conducted with [14C]acetate and [14C]glycerol. Controls with [1401002 and media containing additional carbon sources mimicking in planta conditions were utilized to limit changes in metabolism due to wounding or utilization of exogenously applied metabolic tracer compounds, respectively. Additionally, analysis included not only precursor-product relationships between labeled lipid classes, but also analysis of the regiochemistry and molecular species of lipids from the initial incorporation of labeled nascent FA and glycerol backbones into eukaryotic glycerolipids. 4.1 Acyl editing conclusions 127 Expanding pea leaves and developing soybean embryos were utilized for in vivo of labeling of the initial events of eukaryotic membrane and storage lipid synthesis. To ensure the in vivo relevance of the metabolic labeling data from exogenously applied [14C]acetate or [14C]glycerol to excised pea leaves, intact planted pea seedlings growing in pots were also labeled with 1“00;. The results with the 1“C02 precursor confirmed the conclusions based on exogenously applied [14C]acetate and [“Clglycerol metabolic tracers. Likewise, developing soybean embryos were pre-cultured for three days in media that allows excised embryos to mimic in planta growth. Trace amounts of [14C]acetate or [“Clglycerol were added to the culture media for labeling experiments which allowed exogenous tracers to be applied without wounding the embryos or changing the carbon sources used for heterotrophic growth and represent methodological improvements over most previous studies that utilized seed slices and a single radiolabeled carbon source. These controls allow us to be more confident that the labeling results reflect the reactions of central lipid metabolism that occur in viva. Kinetic analysis of rapid [14C]acetate and [“nglycerol incorporation into lipids allowed precursor-product relationships to be established for newly synthesized FA and glycerol backbone incorporation into eukaryotic glycerolipids, respectively. Stereochemical analysis of [“Cjacetate labeled FA in different lipids indicated that different mechanisms are employed for incorporation of nascent FA into PC, DAG and TAG. Molecular species separation of labeled lipids allowed determination of: (1) pre-existing FA that occur in the same molecule as 128 the [14C]acetate labeled newly synthesized FA; (2) the FA esterifed to [14C]glycerol during de novo glycerolipid synthesis. Finally, endogenous compositions of different lipid molecular species were determined utilizing ESI- MS/MS techniques which demonstrated differences between the steady state molecular species and those of de novo synthesized lipids. Together these analyses of de novo synthesized glycerolipids and incorporation of newly synthesized FA into glycerolipids were used to develop a revised model of eukaryotic membrane and storage lipid synthesis in plant leaves and seeds. The pathway of plant eukaryotic membrane lipid synthesis shown in most textbooks and reviews [34, 85] predicts that newly synthesized acyl groups exported from the plastid are first esterified to G3P producing initial PA molecular species of 18:1/18:1 and 16:0/18:1. PA is then converted to PC the major site of acyl desaturation, and that desaturated PC may be in equilibrium with DAG to produce other desaturated membrane or storage lipids (Figures 1-1 and 2-9 model 1). However, rapid labeling of the initial steps of eukaryotic lipid synthesis in pea leaves indicated that this simple model is incorrect and the major flux of newly synthesized FA into eukaryotic glycerolipids is not by G3P acylation but through acyl editing of PC, producing initial molecular species of PC containing one newly synthesized FA and one previously synthesized FA. Acyl editing involves a cycle of deacylation and reacylation of PC that by it self does not lead to net glycerolipid synthesis and produces an acyl group and lyso-PC as intermediates. Two models of acyl editing were possible in pea leaves (Figure 2- 9, model 2 and 3). (1) Plastid FA export is tied to acyl editing in a mole for mole 129 fashion, with newly synthesized FA directly esterified to the lyso-PC intermediate and the released acyl groups utilized for de novo glycerolipid synthesis (Figure 2- 9 model 2). (2) The acyl editing cycle is at a much higher rate than nascent FA synthesis such that a mixed acyl group pool is mostly FA recycled from acyl editing, and nascent FA in the mixed pool are rapidly incorporated into PC through the high flux of acyl editing (Figure 2-9 model 3). Labeling of initial glycerolipid synthesis in developing soybean embryos also indicated acyl editing as the major pathway for incorporation of newly synthesized FA into extra- plastidial glycerolipids. However, only the high acyl editing flux pathway is consistent with the soybean labeling results (Figure 3—10). Therefore, incorporation of nascent FA into glycerolipids through a high flux of acyl editing is probably the more accurate model in pea leaves and other plants. Research in this thesis demonstrates that incorporation of newly synthesized FA into extra-plastidial glycerolipids occurs through acyl editing in leaves and developing embryos from peas and soybeans, respectively. Pea and soybean are both “18:3” plants [82] and are both in the legume family Fabaceae of the order Fabales. Similar PC molecular species labeling results have also been observed in the leaves of the “16:3” plant Brassica napus [149] (order Capparales, family Brassicaceae) suggesting that acyl editing may be a ubiquitous part of plant leaf lipid metabolism. Additionally, lysophosphatidylcholine acyltransferase enzymatic activity has been found in a wide variety of plant leaf [91, 94, 143, 145] and oilseed tissues [130, 140, 142, 130 144, 163] further suggesting that acyl editing may be a central part of plant lipid metabolism. In soybean embryos the labeling experiments and endogenous PC, DAG and TAG molecular species compositions suggested that TAG was produced primarily from a PC derived DAG moiety generated by the removal of the phosphocholine headgroup (Figure 3-10). A traditional Kennedy pathway involving three consecutive backbone acylations was either very minor or nonexistent. The utilization of the PC DAG moiety for TAG synthesis has been suggested by in vitro assays of safflower [118] and sunflower [173] that demonstrated the reversibility of CDP-choline:diacylglycerol cholinephosphotransferase. In vivo pulse-chase backbone labeling demonstrated the utilization of PC backbone for TAG synthesis in linseed [117, 127] and safflower [168]. In each of these studies it is assumed that the DAG pool is in equilibrium with the DAG moiety of PC and that this equilibrated DAG pool is utilized for TAG synthesis. However, in the soybean embryo labeling reported in chapter three the positional localization of nascent [14C]FA incorporation into PC (>86% sn-2), DAG (equal sn-1 and sn-2) and TAG (>98% sn-3) suggested that newly synthesized DAG and acyl edited PC are not in rapid equilibrium and neither is rapidly used for TAG synthesis. Safflower (Carthamus tincton’us L.) oil contains over 75% PUFA and avocado (Persea amen’cana) oil less than 15%. The major initial product of [“Clglycerol labeling in safflower cotyledons was PC yet it was TAG in avocado mesocarp [168]. Thus, it has been suggested that utilization of the PC-DAG 131 moiety for TAG synthesis may only take place in oilseeds that utilize PC for FA modification and accumulate large amount of desaturated or unusual FA in TAG. However, the unusual FA petroselinic acid (18:1cisA6) produced in coriander and carrot endosperm [152] and the unusual FA 16:1-A6 produced in Thunbergia alata [153] are both synthesized in the plastid, yet still flux through PC prior to accumulation into TAG. If the PC-DAG moiety containing the unusual FA is used for TAG synthesis it would imply that FA modification on PC is not. required for conversion of PC to TAG. Additionally, in Brassica napus which contains 30% PUFA in seed TAG, four hour labeling of developing embryos with acetate and glycerol led the authors to suggest the Kennedy pathway is the major route to TAG synthesis [180, 181]. lt is clear that multiple pathways may be involved in TAG synthesis in different plant species and that utilization of the PC-DAG moiety for TAG is a mechanism that may or may not be used by various oilseed plants. However, the variability in the techniques of labeling studies in different plants (variability of oil tissue and developmental stage, wounding effects of excised tissue, different labeling time periods, FA or backbone labeling only, labeling with or without stereochemical and/or molecular species analyses) makes determinations of the sequence of reactions involved in the TAG synthesis pathway difficult to deduce from most previous studies of oil biosynthesis. The soybean inital glycerolipid synthesis studies presented here demonstrate that very short time point labeling kinetics and analyses of both the acyl and backbone components of TAG synthesis through regiochemistry and molecular species analyses are essential to generate an accurate model of TAG 132 synthesis. The soybean TAG synthesis model indicates that the PC-DAG moiety utilized by TAG synthesis is not the same as the DAG produced by de novo glycerolipid synthesis and therefore a direct Kennedy pathway may not be required in oilseeds (Figure 3-10). Whether a direct Kennedy pathway is a more major TAG biosynthesis route in other oilseeds will require additional in depth investigations of TAG synthesis in different plant species. 4.2 Additional significance of acyl editing and PC derived TAG synthesis The process of acyl editing described in this thesis has possible significance for understanding additional areas of basic plant lipid biology. The mechanism of fatty acid export from the plastid for utilization by the eukaryotic pathway acyltransferases has yet to be elucidated. Experiments with isolated pea chloroplasts indicate that newly synthesized FA are channeled into PC, possibly on the outer chloroplast envelope [79]. Isolated chloroplasts cannot synthesize PC de novo; therefore acyl editing is very likely involved in incorporation of newly synthesized FA into PC at the plastid envelope. Lateral diffusion of PC through the PLAM [94, 95] to the cellular endomembrane system would allow a mechanism of plastid FA export that would not require acyl-CoA to pass through the cytosol. Acyl editing may also be involved in ensuring that newly synthesized glycerolipids do not disturb the integrity of cellular membranes. PUFA containing lipids are required for proper membrane physiology and function, especially at low temperatures [174, 182]. Desaturation of 18:1 to 18:2 and 18:3 lagged 133 significantly behind the rate of 18:1 production in pea and soybean (Figures 2-2 and 3-3, respectively). However, acyl editing produced initial molecular species of PC with mixtures of previously synthesized PUFA with newly synthesized saturated or monounsaturated FA (Figures 2-5 and 3-5). The mixed FA molecular species may benefit the plant by not allowing accumulation of relatively saturated lipid molecular species produced from only newly synthesized FA (e.g. 16:0/18:1, 18:1/18:1). Without acyl editing, plant tissue that is rapidly producing lipids (e.g. expanding leaves or developing oilseeds) may accumulate relatively saturated molecular species in biosynthetic regions of the ER membrane that could affect the membrane physical properties and function. The supply of FA has been shown to be one of the limiting factors of TAG accumulation in oilseeds [183]. High levels of externally supplied FA produces a feedback control on fatty acid synthesis through modulation of plastidic ACCase activity [58]. In soybean embryos most newly synthesized FA are initially incorporated into glycerolipids through acyl editing (Figure 3-10). Acyl editing enzymes may be able to regulate or sense free FA levels which could be involved in the signaling to modify ACCase activity. In plant species that accumulate unusual FA in TAG the amount of unusual FA in membrane lipids is typically very minor. However, in transgenic plants producing unusual FA the membrane lipids can sometimes accumulate more unusual FA than TAG [128]. An acyl editing scheme that selectively removes modified FA from PC for TAG synthesis and incorporates newly synthesized FA into PC to be further modified may be involved in plant species that naturally produce unusual FA. Likewise, in 134 transgenic plants that express the biosynthetic enzymes for unusual FA but do not express unusual FA selective acyl editing enzymes, the unusual FA may not be efficiently removed from the site of synthesis on PC. A similar mechanism selective for PUFA may be involved in plant species that accumulate large amounts of PUFA in TAG. The kinetic modeling of [14C]acetate and [14C]glycerol labeling of soybeans (Figures 3-9 and 3-10) suggested at least three DAG sub-pools that are kinetically distinct. The individual DAG sub-pools may suggest separate sub- cellular localizations that are related to separate functions. Soybean TAG synthesis primarily utilized the bulk DAG produced from bulk PC. The PC headgroup removal to generate the bulk DAG may be localized to an ER region associated with a developing oil body for TAG synthesis. A possible function of utilizing bulk PC in a separate location for TAG synthesis may be to ensure the use of desaturated PC (and thus DAG) and not de novo synthesized DAG/PC or acyl edited PC which would contain at least one acyl group that was not desaturated or otherwise modified. Utilization of PC that diffuses laterally through the membrane from the site of desaturation to the proximity of the oil body may be a mechanism to separate the steady-state desaturated bulk PClDAG for polyunsaturated TAG synthesis without the need for a Kennedy pathway that utilizes a high PUFA acyl-CoA pool or enzymes that are selective for PUFA. 4.3 Future research on acyl editing and TAG synthesis pathways 135 The next logical step in the investigation of lipid synthesis through acyl editing is to identify the enzymes/genes responsible for the deacylation and reacylation of both the sn-1 and sn-2 position of PC. The acyl editing model presented (Figure 3-10) utilizes a sn-1/sn-2 lyso-PC intermediate. Therefore an LPCAT enzyme is probably involved in the lyso-PC reacylation and possibly the PC deacylation by the reverse action of LPCAT [140]. Arabidopsis thaliana contains 11 gene candidates for lyso-phospholipid acyltransferases [134], and recently two Arabidopsis genes have been expressed in yeast and the activities demonstrated to be broad specificity lyso-phospholipid acyltransferases with the highest activity for lyso—PC [184]. Alternatively phospholipases may be involved in the deacylation of PC of which Arabidopsis has at least one PLA1 and six PLA2 gene candidates [134]. Assuming the knockout of acyl editing enzymes are not lethal, the redundancy in Arabidopsis suggests that a fonrvard genetic screen is unlikely to yield which gene products are involved in acyl editing. A reverse genetic approach may be more suitable; however redundancy may also be a problem and may require testing many combinations of gene knockouts. Additionally, the labeling experiments and analysis methods presented in this thesis to investigate acyl editing are laborious and time consuming and thus are not suitable for screening of many different plants. A possible preliminary labeling experiment to analyze reverse genetic candidates would be to rapidly pulse label fatty acid synthesis with [“Clacetate and measure the ratio of [“ClDAG to [“0ch. If acyl editing is reduced in the mutants the initial labeling ratio should 136 increase. Further analysis will be required to determine if a change in DAG and PC labeling is actually due to a change in acyl editing. A possible non metabolic labeling based analysis method to analyze reverse genetic acyl editing mutant candidates would be to use mass spectrometry to quantify lipid molecular species. If acyl editing is knocked out or significantly reduced by a mutation then more nascent FA might enter de novo glycerolipid synthesis and thus initially produce larger amounts of 18:1/18:1 & 16:0/18:1molecular species in PA, DAG and PC. Additionally, levels of lipid molecular species that have nascent FA/previously synthesized FA composition (e.g. 18:1/18:3) produced from acyl editing might be reduced. However, desaturation of the initially produced molecular species would presumably not be affected and any change in molecular species abundances may be small and transient. Therefore to have the highest probability of success, tissue that is rapidly producing lipids such as young expanding leaves should be utilized. Since there are many possible gene candidates for acyl editing enzymes further analysis of the mechanism of acyl editing to determine whether lipases or transacylases are involved could narrow down the number of possibilities. In this regard incorporation of oxygen-18 into the carbonyl oxygen of newly synthesized FA through [180]acetate or H2180 labeling and measuring 180 content with mass spectrometry would allow for measurements of FA hydrolysis. Each time a FA is hydrolyzed by a lipase and re-esterified there is 50% exchange of the carbonyl oxygen with the cellular H20 [74]. If acyl editing proceeds by way of a lipase then there would be different amounts of 18O in the acyl-CoA pool and PC. However, if 137 acyl groups are removed from PC by a transacylation reaction such as reverse LPCAT then 180 content of the acyl-CoA pool and PC should be the same. Similar experiments utilizing [‘3O]phosphate may be able to determine if PC headgroup removal to produced DAG for TAG synthesis is by way of a phospholipase D and PA phosphatase (as suggested for eukaryotic galactolipid DAG production [89, 90]) or reverse action of cholinephosphotransferase [16, 117,118]. Genetic approaches to identify the enzymes of acyl editing would be most successful in the model species Arabidopsis thaliana, which has a fully sequenced genome and many genetic/molecular biology tools available. However, acyl editing was demonstrated in pea and soybean and not Arabidopsis. Thus identification of the acyl editing enzymes from pea and soybean by enzyme purification may be a logical approach. In this regard LPCAT activity has been measured in microsomes of many different leaf [91, 94, 143, 145] and oilseed tissue [130, 140, 142, 144, 163], which suggest that at least partial purification of enzymatic activity is possible. Preliminary experiments suggest that an acyl editing activity is associated with isolated pea leaf chloroplasts (Appendix C). Interestingly in five minutes of [14C]acetate labeling of pea chloroplasts only nascent [14C]18:1 was incorporated into PC and was located ~93% at the sn—2 position of PC. Similarly most in vitro LPCAT activities measured in microsomes were also mostly sn—2 specific. Therefore the substantial sn-1 acyl editing activity identified by in vivo labeling in pea leaves (Figure 2-6) may be a soluble activity that is lost by isolation of chloroplasts or 138 microsomes alone. If an enzyme is able to be purified the amino acid sequence may be identified by Edman degradation or mass spectrometry based sequencing. Once the protein/gene sequence is known genetic and molecular biology techniques could be utilized to confirm its function by analysis of acyl editing in knockout/knockdown mutants and over expressers. The models of pea leaf and soybean acyl editing presented describe an acyl editing cycle involving nascent saturated and monounsaturated FA incorporation into the sn-1 and sn-2 position of PC. Some evidence in chapter three suggests that within the acyl editing cycle and glycerolipid synthesis there could be different mechanisms involved in saturate and 18:1 incorporation into glycerolipids which should be studied further. Soybean saturate labeling increased with time in proportion to 18:1 in PC, while saturate labeling was constant and in very different proportions to 18:1 in DAG and TAG (Figure 3-7B). Additionally, the primarily sn-2 and 18:1 specific acyl editing in pea chloroplasts (Appendix C) suggest that separate enzymatic activities and possibly locations are involved with incorporation of nascent 18:1 into PC as compared to nascent saturates in pea leaves. The high sn-3 saturate labeling of nascent FA incorporation into TAG of developing soybeans may suggest that there is a specific channeling of saturated FA to a DGAT. Release of newly synthesized saturated FA and 18:1 from ACP at the termination of fatty acid synthesis is carried out by two separate thioesterases, FATB and FATA respectively. The Arabidopsis fatb mutant which reduces the amount of saturated FA exported from the plastid [161, 185, 186] and increases total 18:1 plastid export and 139 breakdown may be useful for analysis of acyl editing mechanisms. If there is a saturate specific acyl editing enzyme that is required for nascent saturated FA incorporation into glycerolipids then a mutant should have a change in lipid composition similar to fatb, and the double mutant lipid composition should be similar to the single mutants. If the excess 18:1 exported from the plastid in the . fatb mutant is incorporated into glycerolipids through acyl editing then the rate of nascent FA incorporation into PC should increase. If acyl editing is required for turnover and breakdown of the excess 18:1 in the fatb mutant, then B-oxidation in an acyl editing and fatb double mutant might be reduced as compared to the fatb mutant alone. Additionally, a fatb and 18:1 acyl editing mutant may accumulate excess 18:1 containing PC molecular species as compared to wild type. Soybean embryo fatty acid synthesis produced more newly synthesized [“Clacetate labeled 18:0 than 16:0. However, the endogenous FA mass composition contains more 16:0 than 18:0. The newly synthesized [“C]18:0 that is initially more abundant than newly synthesized [“C]16:0 in soybean glycerolipids may be a transient pool of 18:0 that is used for synthesis of other products such as sphingolipids, cutin, or wax. Alternatively, 18:0 may be desaturated while bound to PC or within the recycled acyl-CoA pool. The plastidic stearoyl-ACP desaturase is considered the only known 18:0 desaturase in plants [34]. Identification of an extra-plastidic stearate desaturase would be novel to plant lipid metabolism. Further [14C]acetate soybean labeling with much longer time points and/or pulse-chase [“Cjacetate labeling experiments may be 140 able to reveal the down steam products of the initially high 18:0 production for additional investigation. The identification of multiple sub-pools and DAG that are not in equilibrium suggests that the enzymes producing each sub-pool are not in the same cellular location. Even though we could not distinguish multiple pools of PC kinetically, acyl edting may be at a separate location than de novo PC synthesis. Cell biology localization techniques such as green fluorescent protein (GFP) fusions of the enzymes for de novo glycerolipid synthesis, acyl editing, and TAG synthesis could be employed to confirm the separate localization hypothesis. An interesting facet of localization studies would be whether different isoforms of enzymes localize to the same sub-cellular location. For instance if plastid associated acyl editing is only sn-2 specific then a sn-2 LPCAT may be localized to the PLAM region, but an sn-1 LPCAT that utilizes saturates may be soluble or localized to an acyl-editing domain of the ER. A hypothesis as to why many transgenic plants expressing FA modification enzymes do not accumulate the same quantities of unusual FA in TAG as do the native species is that unknown factors channel the unusual FA from site of synthesis to incorporation into TAG. The GFP fusions used for localization studies could also be used for co- lmmunoprecipitation studies to identify proteins interacting with the PC acyl editing enzymes or the TAG synthesis DGATs and PDATs. Identification of the enzymes required for incorporation of modified FA into TAG and the ability to express them in the correct location to be efficiently utilized will greatly increase 141 our ability to generate transgenic crops that accumulate unusual FA of interest in TAG. This dissertation has demonstrated that FA incorporation into membrane lipids and TAG involves large fluxes of acyl editing in and out of PC in pea leaves and developing soybeans. The relation of acyl editing to previous studies suggest that acyl editing may be involved in many different plant species and further research may imply that it is a central part of plant glycerolipid metabolism. The acyl editing models presented update the simple de novo glycerolipid synthesis model found in most lipid biosynthesis reviews and textbooks. Further research of the mechanisms of acyl editing and the enzymes/genes involved will deepen the understanding of plant lipid synthesis and possibly increase our ability to modify the incorporation of FA into plant lipids to suit our needs for food, fuel and industrial applications. 142 5 APPENDIX A: CHAPTER 2 SUPPLEMENT NOTE 1 Discussion: 18: 1/18:1 and 16:0/18:1 Are Not Initial Molecular Species of Eukaryotic Glycerolipids Rapidly expanding pea leaves at day 8 and day 9 produce about 2.2 and 0.6 pmol FAlhr/g fresh wt. respectively in the light, and contain 20-25 pmol FAlg fresh wt. Although endogenous 16:0/18:1 and 18:1/18:1 molecular species of PA and PC can arise by non-biosynthetic mechanisms, if we treat them as essentially all biosynthetic, produced by de novo synthesis from glycerol-3-P, and estimating their PA and PC levels at 0.002 and 0.2 mole % respectively, the residence time for any labeled FA passing through these molecular species pools is calculated as 0.7-3 sec and 1.1-5 minutes respectively. Thus, a pool of 16:0/18:1 and 18:1/18:1 molecular species of PC should have been easily detected by our analysis. 143 SUPPLEMENTAL FIGURES Figure 5-1 Supplemental Figures S1A-S1M: Lipidomics of pea leaves and acetate labeled pea leaves. Pea leaf lipid samples were analyzed by ESI-MS/MS by the Kansas Lipidomics Research Center. Extraction of lipids was conducted by their standard Arabidopsis leaf protocol (Web site: h_ttp:/lwww.k-state.ed u/ligid/lipidomics/leaf- extraction.html). Data is represented as mol percent of total lipids measured. Error bars are standard error, n = 5. Black bars: 8 day old pea leaves out and quenched immediately. White bars: 8 day old pea leaves out into strips and labeled with 1 mM cold (‘20) acetate for 5 min and quenched immediately. Mock labeling conditions were the same as for [“Clacetate labeling. A-B: The sum of total molecular species for each lipid class measured. C-M: Individual molecular species for each lipid class measured. A: Digalactosyldiacylglycerol (DGDG), Monogalactosyldiacylglycerol (MGDG), PC, PE, PG, phosphatidylinositol (Pl). B: phosphatidylserine (PS), PA, Lyso-PC, Lyso-PE, Lyso-PG. C: DGDG. D: MGDG. E: PC. F: PE. G: PG. H: Pl. l: PS. J: PA. K: Lyso—PC. L: Lyso-PE. M: Lyso-PG. 144 SlA: Total molecular species by lipid class 45 40- 35- \c 301 ,°_ 25- ° 201 E 15. 10- 5. 0. 1:1-Ila: U H "‘ a. m 2 P" DGDG MGDG Lipid class 145 SIB: Total molecular species by lipid class LOO—_— 0.75-i 0.50 mol % 0.25 0.00 < 9.. (I) Q-l LySOPC I LysoPE I LysoPG Lipid Class 146 SIC: DGDG umuwm Lawn umuwm . 9mm H6”. anon muwm v6». muem GEM Huvm Nuvm muvm vuvm muvm envm Molecular Species 147 SID: MGDG muwm Sam mumm . Guam dwm Nuwm anon anon muwm Guem ten Nuvm muvm vuvm muvm oxen 42 + 40 38 1 5- 1.0. 0 SJ 0 0 e\e 2:: Molecular Species 148 SIE: PC «an Q3. 339 mucv Numm muwm 3am muwm can». den N6». nuem 96m mnwm can :3“ Nuvm muvm 93% can». Molecular Species 149 SIF:PE I d C I d u d I d 5 0 5 0 5 0 5 0 5 0 2 2 l l 1 1 0 0 0 0 o\o 7:: NA? m£§ v36 fiév 95v flan flan 35m flan Efim den fiém 96m fibn flan 96m Cvm fivm mgfi VGfi Molecular Species 150 SIG: PG 32:1 32:0 34:4 34:3 34:2 34:1 34:0 Molecular Species 151 SIH: Pl BCleS Molecular Sp 152 mol % 0.225 811: PS 0.200' 0.175- 0.150d 0.125-i 0.100q 0.075' 0.050- 0.025d 0.000- -.II _II. III VMNfimeMNfingMN—VMNFVMNF'IMN vvvveweeeewweowmeoceccnumuvv MMMMMMMMMMMMMMMMVVVV‘G'VI'V'VV'V Molecular Species 153 SlJ: PA 0.14 0.12- 0.10' /o c 0.08- E 0.06] 0.044 0.02- 34 1 36:6 Molecular Species 154 S] K: Lyso-PC 16:1 16:0 18:3 18:2 13:1 18:0 Molecular Species 155 SlL: Lyso-PE 0.025 0.020' 0.015% mol (yo 0.010- 0.005- 0.000 u 16:1 16:0 18:3 18:2 18:1 Molecular Species 156 SIM: Lsyo-PG 0.005 0.004- o 0.003' I 0.002- (LOOP—Ill 0.000- T. 16:1 16:0 18:3 18:2 18:1 Molecular Species [1101 /o 157 Figure 5-2 Estimates of single and dual acyl-labeled PC molecular species for Model 3 (Figure 2-9). Calculation of PC molecular species containing one labeled acyl group (PC[FA*FA]) or two labeled acyl groups (PC[FA*FA*]), based on x moles of labeled fatty acid (FA*) and y moles of unlabeled fatty acid (FA) distributed randomly between sn-1 and sn-2 positions. The distribution is calculated as PC[FA*FA]/(PC[FA*FA*] + PC[FA*FA]). The small amount of labeled saturates largely at the sn-1 position (Figure 2-6) will not greatly skew these results. 100 40 o. 8 5 60~ '5'. o 40 § ’I § ‘IE 5.5 § 5!; ‘IIE = 8 I: 3‘ 1:5 1:5 *‘ 1:5 Ills Ill15:0 2 20 . \ ’l D ’l- ’I- \ /I- ’II- ('2’ .§ 5' R 5'? filé S 4% 5"2 914-0 0 4 .5 1.: at I}: A}: .5 1!: all: ‘ a 13:0 TAG 20:0 TAG 18:0 TAG 16:0 PC 18:0 PC 16:0 11 12.0 Alpha oxidation substrate Figure 6-2 Alpha oxidation of [“Clacetate labeled saturated FA from TAG and PC. Individual saturated FA acids from TAG and PC were subject to chemical alpha oxidation. Each reaction is a mixture of FA from 20, 30, 120 & 360 min acetate labeling. The reaction produced a series of FA each shortened by 10 from the carboxyl end. Results are expressed as relative specific activity to that of the full length reaction substrate. The X-axis groups each individual substrate reaction products with decreasing chain length from left to right. Individual FA products bar shading is in the key. Method: Individual saturated FAME from TAG and PC were collected by first AgNOa then reversed phase TLC (as in methods) and saponified to FA. 100 ug of non labeled carrier FA was added to each reaction with the respective FA chain length. The reaction was started by adding 10 mg of finely ground KMnO4 and 0.5 ml acetone to each dried FA sample and heated to 60° C. After six hours remaining KMnO4 was reduced in the cooled sample by adding 1 ml of 0.5M FeSO4 in 0.5 N H2804. FA were extracted with hexane and FAME generated by reaction in 5% sulfuric acid in MeOH at 50° C for 30 min. FAME extracted by hexane were split into two aliquots, mass quantified by GC-FID, and radioactivity quantified by reversed phase TLC and electronic autoradiography. 160 B (D '- o S l- a) — P . r022 0 l- I- S ‘3 20:0): 2 a) 2 LIJ _2 < . .: 220 a) 1". ‘- ' .1" 88 SM 'More saturated"More PUFA' (C) Acetylation PLC digestion - (A) Polar lipids - (B) Neutral lipids m Figure 6-3 TLC image examples and flow chart for analysis of radiolabeled glycerolipids. A-E. Examples of the major TLC systems used during analysis of radiolabeled glycerolipid classes, molecular species, and FA composition. Molecular species are represented as a combination of two or three FA, total saturates, S; monoenes (18:1), M; dienes (18:2), D; trienes (18:3), T. A, Neutral lipid class TLC, pictured 6 min [14C]acetate. B, Polar lipid class TLC, pictured 6 min [“Clacetate. C, TAG molecular species argentation TLC, pictured 30 min [“Clglycerol. D, Molecular species separation of PC and DAG as 1,2-diacyl-3- acetyl-glycerols, pictured 6 min [14C]acetate PC. D, FAME argentation TLC, pictured FAMEs from different 6 min [“Clacetate PC molecular species. 161 A: 2 min 4o 35~ ' ".1301 |15.525- r" 320- E15- $10. .. n 0---r'.-r-r-L. . . . 8855§§585t ISDMID C:10min 40 35« 1 11.1301 E25: _ 5'20- - E151 3210. 5. cull-.11.. .. . aacsgssast IS UM ID B: 6 min °/. [14C]FAME 4O 351 301 25- 20‘ 15- 10‘ IS UM ID D: 30 min 1%. [14C]FAME 40 35- 30- 25- 20- 15- 10- 5. 0 . IS EIM ID Figure 6-4 [“Clacetate labeled PC molecular species time course. A-D, molecular species are represented as pairs of FA, with bar heights represent % of species among the total. Bar shading represent % of [“ClFA among species. In species with only one color it means that only that FA is labeled the other FA within the molecular species is unlabeled. Bar shading: S, total saturates, black bars; M, monoenes (18:1), white bars; D, dienes (18:2), gray bars. A-D Time of [14C]acetate labeling: A, 2 min; B, 6 min; c, 10 min; 0, 30 min. 162 A; 6 min Figure 6-5 [“Clacetate labeled DAG molecular :2 species time course. 30 - A-C, molecular species are u 25 . - represented as pairs of FA, with 3 20 , _ ID bar heights represent % of g P species among the total. Bar 15 _, UM . 14 IS shadlng represent% of[ C]FA 1° ' __ among species. In species with 5 ' only one color it means that 0 - r . . . . . only that FA is labeled the other ‘0 D '- 2 2 0 t- 0 '- [: FA in the molecular species is co co (0 2 o O "’ 5 2 unlabeled. Bar shading: 8, total B' 10 min saturates, black bars; M, ' monoenes (18:1), white bars; D, 40 dienes (18:2), gray bars. A-C, 35 . Time of [“Clacetate labeling: A, 30 , 6 min; B, 10 min; C, 30 min. 25 - n- g 20 . ' - '0 °‘° 15 - GM 10 . IS 5 ‘ ll 0 T I I I l I T assassras: C: 30 min 40 35 - 30 - 25 - F g 20 ‘ _ r .0 z! 15 , DM 10 1 IS 5 1 0 " r r T T T r 8 8 I7: 5 E ‘2’ "' 8 E l: 163 NGmm 20 ID 16" UM . IS 312 P =3 8- 4. 0- , wEDPEDPDFEEOPOPtDPt m w o o o wflgwfiggameEEEEESQO BHOmm 20 ID 16‘ DM 012- IS 1’ 1- :2 a- 4. 0'1 wEOFEDHOFtEOFDFtDPt m . o o o $$8e5558em§§§§2280e CHMmm 25 20- 9 2 0154 g 3 3 5 $1015; 5 a: s 5. 55 225 I 0_ ‘gsfiss ‘ A" I wEDPEDFDPEEDPOFtDFt m o o o gflgwéégfimwgéégzzgoo 164 Figure 6-6 [14C]acetate labeled TAG molecular species time course. A-C, molecular species are represented as three of FA, with bar heights represent % of species among the total. A-B, bar shading represent % of [“C]FA among species. In species with only one color it means that only that FA is labeled the other FA in the molecular species is unlabeled. Bar shading: S, total saturates, black bars; M, monoenes (18:1), white bars; D, dienes (18:2), gray bars. A 6 min labeling; B, 10 min labeling. C, Only the total [“C] label per molecular species is reported. A: PC 40 35 - B 30 - .5 20 _ 5 I6 min (i) 15 . ‘ § § 010 mln °\o 10 . s s E I30 min 5 . s E s uESl-MS/MS 0 q 5 \ \ SS SM SD ST MM MD MT DD DT ‘lT Molecular species B: DAG 35 30 - E 25 - g 20 ‘ g ‘ I6 min 3 15 I g s E ‘ 010 min ,3 10 ' E E S s s .30 min - 5 3 s s E nESI-MS/MS o a i h 5 > 5 SS SM SD ST MM MD MT DD DT TT Molecular species Figure 6-7 [“Clglycerol labeled molecular species time course of PC and DAG. A-B, molecular species are represented as a combination of two FA. Total saturates, S; monoenes (18:1), M; dienes (18:2), D; trienes (18:3), T. [“C]glycerol labeled molecular species are backbone label only (acyl chain labeling hase been subtracted). Bar shading represents labeling time: 6 min, black bars; 10 min, white bars; 30 min, gray bars. The endogenous mass molecular species compositions are from ESI-MS/MS measurements, hatched bars. A, PC. B, DAG. 165 45 40 1 ~ § 35 J § - n 30 V § I30 min TAG E. 25 . g Q ‘ C16 min DAG g 20 - % % § I30 min DAG 15 § % $ US min PC § g $ .30 min PC 1° ‘ Q S S V § § § § 5- n t % |t 0 _ \\ .lk § & Figure 6-8 Fatty acid composition of [14C]glycerol labeled TAG, DAG and PC molecular species. The fatty acid composition of acyl groups esterified during de novo glycerolipid synthesis with glycerol-3-phosphate acylation. Values are calculated from the [14C]glycerol backbone labeled molecular species in figures 70 (TAG) and S7 (PC and DAG). Fatty acids are represented as number of double bonds: saturates, S; monoenes, M; dienes, D; trienes, T. Bar shading: 30 min TAG, black bars; 6 min DAG, white bars; 30 min DAG, light gray bars; 6 min PC, hatched bars; 30 min PC, dark gray bars. 166 TAG Molecular Species Enomoles TAG/embryo PPI 4.19 1 0.17 PPL 12.92 :I: 0.35 PPO 4.52 :t 0.35 PPS 0.66 :l: 0.27 PLL 8.03 :l: 0.28 PLI 26.86 :t 1.23 PLL 38.60 t 1.11 POI 15.78 :I: 0.45 POL 37.72 1: 2.13 PSI 2.16 :l: 0.12 POO 13.70 :I: 0.26 PM 9.26 :l: 0.18 PSO 6.02 i 0.78 Lll 14.10 i 0.77 LLI 34.12 :l: 1.82 OH 7.40 d: 0.39 OLT 39.06 :l: 1.75 LLL 28.94 :t 1.29 Sll 1.33 i: 0.05 OLL 43.15 :I: 1.75 OOl 12.88 :I: 0.52 SLI 7.69 :l: 0.31 OOL 28.44 :I: 2.06 OLL 10.49 i: 0.76 SOI 5.27 :l: 0.38 SOL 16.20 :1: 0.76 000 11.95 :l: 0.56 SSI 0.79 :l: 0.03 800 8.23 :l: 0.40 SSL 2.97 :l: 0.14 SSO 3.35 :l: 0.21 ALL 2.43 :l: 0.09 AOl 2.85 :l: 0.11 AOL 3.70 :l: 0.20 Table 6-1 TAG Molecular Species Composition of Soybean Embryos Individual TAG molecular species of developing soybean embryos were quantified using ESI-MS and ESI-M82. Values represent nanomoles of TAG per embryo (1: standard deviation). Descriptions of the various molecular species do not imply any specific stereochemical arrangement of the different acyl chains. Abbreviations for fatty acids: P = palmitate (16:0), S = stearate (18:0), O = oleate (18:1), L = linoleate (18:2), l = linolenate (18:3) and A = arachidonate (20:0) 167 DAG Molecular Sgecies picomoles DAG/embryo PP 19.75 :I: 2.94 Pl 384.64 :I: 22.84 PL 1629.70 :1: 146.31 PO 878.23 :I: 80.62 PS 80.59 d: 11.76 M 125.62 :I: 10.41 Ll 751.03 :I: 39.99 OI 537.02 :I: 30.00 LL 1364.04 :l: 76.21 SI 130.86 :I: 8.31 OL 1884.44 :1: 119.75 SL 692.70 :I: 43.88 00 1239.94 :l: 78.56 $0 594.25 :I: 46.74 SS 105.53 :I: 11.42 Table 6-2 DAG Molecular Species Composition of Soybean Embryos DAG molecular species of developing soybean embryos were converted to their sn-1,2-diacyly-3-acetyl-glycerol derivatives and then quantified using ESI-MS and ESI-M82. Values represent picomoles of DAG per embryo (i standard deviation). Descriptions of the various molecular species do not imply any specific stereochemical arrangement of the different acyl chains. Abbreviations for fatty acids: P = palmitate (16:0), 8 = stearate (18:0), O = oleate (18:1), L = linoleate (18:2), and l = linolenate (18:3). 168 PC Molecular Species gicomoles Pgembrvo PP 952.25 :1: 118.71 Pl 832.32 :I: 97.16 PL 3827.04 :1: 456.43 PO 2141.75 1245.10 PS 524.49 :I: 76.10 M 523.51 :I: 98.42 LI 1836.70 :l: 218.43 OI 1244.15 :l: 128.81 LL 3483.64 :t 360.67 SI 358.86 :I: 37.87 0L 4360.23 :l: 460.21 SL 1912.57 :I: 186.51 00 2964.48 :l: 289.10 SO 1730.68 :l: 165.72 SS 259.02 :I: 43.78 Table 6-3 PC Molecular Species Composition of Soybean Embryos DAG molecular species of developing soybean embryos were converted to their sn-1,2-diacyly-3-acetyI-glycerol derivatives and then quantified using ESI-MS and ESI-M82. Values represent picomoles of PC per embryo (:l: standard deviation). Descriptions of the various molecular species do not imply any specific stereochemical arrangement of the different acyl chains. Abbreviations for fatty acids: P = palmitate (16:0), 8 = stearate (18:0), 0 = oleate (18:1), L = linoleate (18:2), and l = linolenate (18:3). 169 6.2 Supplement: Kinetic modeling of pool filling To choose between alternative models of TAG synthesis which take into account various options for DAG-PC interconversions simulations of kinetic labeling were undertaken (Figure 3-9, and6-9). The expectation was not that any of the more complex models would exactly match the actual kinetics measured (Figures 3-2, 3-7). Instead the modeling is used to formulate the best descriptions of acyl metabolism. In such modeling of this in vivo data there is a list of implicit assumptions: (1) The cells are in a uniform physiological state. (2) Although exogenous “non-physiological” substrates may penetrate differently, may perturb metabolism slightly differently, they effectively report the bulk tissue metabolism. (3) In vivo variance, including that generated by using different experiments used for acetate and glycerol labeling is small. (4) We assume that there is no futile cycling of acyl groups, that TAG is a static end product, and that different lipid molecular species have similar time constants for pool filling. (5) Minor acyl fluxes, sUch as to sphingolipids, plastid lipids, PE etc. are ignored. The modeling is set up with 1000 moles FA in total PC and 400 moles in total DAG (based off of Table 3-1), which are invariant in size over the hour period the simulation is run. The inputs are either 100 moles/hr of FA into the 170 system into the DAG(input) pool, or 300 moles via acyl editing into a bulk or small active PC pool. The output is 100 moles/hr of FA into TAG. A simple Excel spreadsheet is used to run the simulation, using 0.01 hour increments, that is, with 1 mole FA input per time increment when modeling for de novo DAG input. The models simulated are based off those represented in Figure 3-9. In the case of the simplest model tested, model B (Figure 6-9a), the rate constant for DAG to TAG conversion k1 will be 0.25 hr", to allow a fixed rate of TAG synthesis of 100 moles/hr (400 moles DAG x 0.25 hr"). The rate constants for the DAG to PC (kg) and PC to DAG conversions (k3) can be set at any value, but because the DAG and PC pool sizes must be kept constant, the ratio of k2:k3 = 2.5. The values of kg and k3 were then varied to allow a simulation of glycerol backbone labeling from DAG that produced continuous DAG labeling but where PC labeling crossed over DAG labeling at about 30 minutes as in the experimental results (Figure 3- 7A). This occurred when the forward and back reactions for the DAG to PC interconversion were 20 times the rate of TAG synthesis (Figure 6-9a). These parameters were then used to test what would be predicted for an acyl editing input (Figure 6-9b). At 30 minutes about 21% of the PC would be converted to DAG. This is inconsistent with experimental observations that show about 3% of the PC have actually moved from PC to DAG. Thus model B fails to predict the slow movement of labeled PC to DAG. If at 0.5 hr 21 % PC was converted to DAG that would change the composition and stereochemistry of the DAG product dramatically. 171 Additional pool filling simulations for Models B and C (including variants) were run to see how closely they would conform to the experimental kinetics of glycerol backbone labeling, of TAG labeling, and of acyl editing. Some of these are shown in Figure 6—9. Variants on model B suffered the same problem as the simplest form of model B (too much back conversion of acyl labeled PC to DAG at early time points). By contrast, variants on model C was generally much more successful at describing the DAG-PC precursor-product glycerol labeling kinetics and yet having slow movement of acyl editing labeling of PC move through into DAG. Model C variants have as a common theme at least two kinetically distinct DAG pools, DAG(in) for the immediate products of de novo DAG synthesis, and DAG(out), which supplies TAG synthesis. The DAG input pool is kept small (2% of PC), in keeping with the observation of very small DAG pools observed in leaf, because the de novo DAG pool must be compositionally distinct from the bulk DAG pool, and because glycerol labeled DAG moves through to PC very significantly within 30 minutes (PC pool filling to TAG takes 10 hours, so if 15 min pool filling of DAG occurs, it will be 1/40 of PC). To fit for the most simplified version of Model C (—> DAG -—> PC —> DAG —>TAG) gives are rather flattened DAG filling curve (6- 90) compared to Figure 3-7, although this can be attenuated by adding a direct low flux loop from DAG(in) to DAG(out). Thus Model C, variant 1 seems unlikely. Splitting the PC pool into a small, active pool and a large, bulk pool Model C, variant 2 (Figure 6-9d) greatly alleviates the problem of slow DAG pool filling 172 at later times seen in Model C, variant 1. Keeping the active PC pool at 2% of the total PC, and therefore at the same size as DAG(in), requires that the interchange between the active and bulk PC pools be 3 times the rate of TAG synthesis. If this is done, then acyl editing must be largely a function of the bulk pool (Figure 6-9e), not the active pool (Figure 6-9f). In proposing two PC pools, it is not clear whether the desaturases act on PC during its residence in the small, active PC pool, or the bulk PC pool, or both. One problem with simulations of Model C, variants 1 (Figure 6-9c) and 2 (Figure 6-9d), was that they predicted movement of glycerol labeling into TAG that was much slower than observed (Figure 3-7A). This could be remedied by splitting the DAG(out) pool into a small active and a large bulk component (Figure 6-9g, Model C, variant 3). Of course, with multiple TAG synthesis routes, this partitioning of DAG(out) may still be a simplification. More generally, the splitting of the intermediate pools of PC and DAG allows for a large degree of flexibility in modeling TAG synthesis, with many combinations of pool sizes and exchange rates. The exercise of modeling initial kinetics has lead to Model C, variant 3 as the preferred model, giving a reasonable though not an exact simulation. However, many more complex variants are possible and cannot be adequately distinguished. Clearly, additional methods, such as longer term and pulse-chase labeling, will be required to refine the model. 173 Figure 6-9 Kinetic modeling of pool filling simulations Graphs of simulations of pool filling, either from de novo DAG acyl labeling (it will be the same for glycerol backbone de novo labeling), or from acyl editing inputs. The pool filling represents the labeling experiments, where nascent, labeled acyl groups are available from fatty acid synthesis for glycerolipid synthesis. Figure 6-9a: Pool filling from de novo DAG synthesis using model B. Figure 6-9b: Pool filling from acyl editing using model B. Figure 6-9c: Pool filling from de novo DAG synthesis using model C, variant 1. Figure 6-9d: Pool filling from de novo DAG synthesis using model C, variant 2. Figure 6-9e: Pool filling from acyl editing via the bulk PC pool using model C, vaflant2. Figure 6-9f: Pool filling from acyl editing via the active PC pool using model C, vanant2. Figure 6—99: Pool filling from de novo DAG synthesis using model C, variant 3. Figure 6-9a: Pool filling from de novo DAG synthesis using model B. 70 O DAG 60 ‘ I PC 50 D TAG 40 30 Glycerolipids 0 0.2 0.4 0.6 0.8 1 1.2 Time (Hours) 174 Figure 6-9b: Pool filling from acyl editing using model B 250 IPC 200 ‘ ODAG chAG in 150 4 :9 .9 73 8 100 - 2 O (tutu. («“61“ 50 - “(titut‘ulultltufl ( «th‘“ («“th “(tum)“ O lllllIi1iilllllllllltl‘l‘l‘l(l(l(l(l(l(l(lIIlllllllllllllllllllllllllllllllIlllllllllllllllllIIIIIIIIIIIIIIIIIIIIIIIIll 0 0.2 0.4 0.6 0.8 1 Time (Hours) Figure 6-9c: Pool filling from de novo DAG synthesis using model C, varient 1. 1.2 80 IPC 70 j chAG 60 . ODAG 50 . 40 ~ uruuuuuuuuuuuututu) u 30* Glycerolipids 20* 10~ kIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIllllllllllllIIIIIIIIIIIIIIIIIIIIIIIIIIIIII 0 0.2 0.4 0.6 0.8 1 0 Time (Hours) 175 1.2 Figure 6-9d: Pool filling from de novo DAG synthesis using model C, variant 2. 70 60 _ DTAG ODAG 50 - IPC {’3’ 40 J '5. «t = ( (““1 g “(““muuuut .Z" 30 ~ (“(((l((( 0 20 . 10 - o 7— 0 0.2 0.4 0.6 0.8 1 1.2 Time (Hours) Figure 6-9e: Pool filling from acyl editing via the bulk 350 PC pool using model C, variant 2. 300 ‘ DTAG IPC 250 ‘ ODAG 8 9 £150 - o 100 - 50 - O G ((((((((((((((((((((((((((((((((((((((((((((((((((((((((((((((((_(_(_(.(_(_((_((((((((((((((((((((((((((((O 0 0.2 0.4 0.6 0.8 1 1.2 Time (Hours) 176 Figure 6-9f: Pool filling from acyl editing via the active 250 PC pool using model C, variant 2. CITAG 200 ‘ IPC ODAG fi 150 - .9 9 £100 (9 (tutu tttttt(((““““«““ 50 ‘ (tttttttttt «(t N“ ((((““ “mum“ «M‘m‘u ;_ O (((«flflfigfifiu I -~ , 0 0.2 0.4 0.6 0.8 1 1.2 Time (Hours) Fi ure 6-9 : Pool fillin from de novo DAG s nthesis 9 9 9 Y usin model C, variant 3. 70 9 60 . DTAG ODAG 50 1 IPC 8 3:140 ‘ E (tumult. 8 30 q ((t(((((((((((( . O ,. J. tutu“ 20 ‘ («((“w 10 - 0 thkk "lluu'lnu'""flu""unlulll'l'lllllllllllllllllllllllllllllllllIIIIIIIII 0 0.2 0.4 0.6 0.8 1 1.2 Time (Hours) 177 6.3 Discussion of soybean TAG synthesis flux model Figure 3-10 Model assumptions The following assumptions used to generate a model of acyl group fluxes in soybean embryos are based on the endogenous lipid compositions and the [“C]acetate/[“C]glycerol labeling results and conclusions. 1) De novo synthesis 3 of DAG and acyl editing of PC are metabolically distinct processes. 2) Nascent FA directly enter PC through sn-1/sn-2 acyl editing. 3) Nascent FA can mix with recycled FA released from acyl editing for de novo glycerolipid synthesis. 4) The major flux of de novo synthesized DAG is for PC synthesis. 5) TAG synthesis utilizes the bulk DAG pool (not de novo DAG) generated from bulk PC. 6) The bulk DAG generating reaction is in a different location or selective such that de novo synthesized PC and acyl edited PC are not rapidly used for TAG synthesis. 7) In a soybean embryo rapidly accumulating oil 90% of FAS is ultimately used for TAG synthesis and most of the PClDAG produced is turned over for TAG synthesis. 8) Newly synthesized FA are incorporated into PC:DAGzTAG at relative initial rates of approximately 4:1 :1, respectively. Thus the initial FAS product incorporation into glycerolipids: 60% PC, 15% DAG, 15% TAG (total 90%). 178 Table 6-4 TAG synthesis model: nascent FA flux value calculations Total S (% 14C) M (% 14C) Nascent Nascent [“C]FA s Flux M Flux PC 57 5 95 2.9 54.1 DAG 17 29 71 4.9 12.1 TAG 1 1 75 25 8.2 2.3 Other 15 29 71 4.4 10.6 [14C]acetate labeling values used to calculate the flux of newly synthesized saturated (S, 16:0, 18:0) and monounsaturated (M, 18:1) FA through the lipid metabolism model towards TAG. Total [“C]FA values are based on the fact that 85% of newly synthesized FA are incorporated into PC, DAG and TAG at a ratio of 10:3:2. The remaining 15% most likely is incorporated into other membrane lipids. Nascent S & M flux values are The model assumes that 93% of FA will ultimately flux into TAG to generate the endogenous lipid composition. Supplement: steflw step logic in flpx moc_Lel construction Model development assumes 100 moles of FAS and utilizes mass balance along with the initial nascent FA incorporation rates and compositions (Table 6- 4). Newly synthesized FA are incorporated into PC:DAGzTAG at relative initial rates of 10:3:2, respectively. Thus of initial FAS product incorporation into glycerolipids: 57% PC, 17% DAG, 11% TAG (total 85%). If 93 moles of FA are ultimately incorporated into TAG (based on Table 1, and that PC represents ~1/2 of soybean membrane lipids) and the remaining 7 moles for membrane lipid synthesis then 31 and 3.5 moles of G3P are required for acyl group esterification, respectively. De novo synthesis of 34.5 mol of total glycerolipids would require 69 mol of FA to flux through PA and DAG into the sn-1/sn-2 positions of membrane lipids and TAG. The remaining 31 moles from FAS will ultimately be esterifed to 179 the sn-3 position of TAG. If 32 moles of nascent [“C]FA are used by de novo glycerolipid synthesis (Table 6-4, DAG + other lipids), then an additional 37 moles of recycled [‘ZC]FA are required to balance de novo synthesis. Presumably this is provided by the acyl chains released by acyl editing to allow 57 mol of nascent [‘4C]FA incorporation into PC through Lyso-PC acylation. However, 57 mol of PC acyl editing will not provide enough saturates to sustain) de novo glycerolipid synthesis. Saturates make up ~29% of acyl groups entering de novo glycerolipid synthesis, as estimated from [“C]glycerol labeling (Figure 6- 8). Thus total saturates would comprise 20 moles of the 69 moles acylated to G3P. Newly synthesized FA would contribute 9.3 moles of saturates (Table 6-4), requiring 10.7 moles of recycled saturates. However, 57 mol of PC acyl editing with equal sn-1/sn-2 contribution could only provide up to 7.6 mol of saturated FA (PC 26.5% S x 28.5 mol sn-1 FA release). Additionally, nascent FA incorporation into PC was 14% sn-1 and 86% sn-2. Saturates are predominately at the sn-1 position, therefore if the labeling reflects the positional release of FA during acyl editing much less saturates would be released and recycled into de novo synthesis. Therefore the flux of acyl editing cannot be one acyl group released from PC for each nascent acyl group incorporated, and must be 1.4 times that required for incorporation of nascent FA into PC with the possibility of being much greater. The excess flux of recycled acyl groups produced during acyl editing may be re-incorporated into PC through the acyl editing cycle (Figure 3- 10), with the entire process reducing the saturated concentration of PC and redistributing FA between the sn-1/sn-2 positions. 180 It is very unlikely that nascent acyl groups are channeled directly into Lyso-PC that is under a high flux of reacylation with recycled acyl groups, but more likely the nascent acyl groups mix with the recycled acyl groups of the rapid acyl editing cycle (Figure 3-10). Initial PC and DAG contain different [14C]FA incorporation rates and compositions indicating there is either different mixed pools of new and recycled acyl groups feeding de novo synthesis and acyl editing or that the acyl editing flux is selective and fast enough that over three times as much nascent FA enter PC than DAG (Figure 3-2) through de novo synthesis from the same mixed pool. To determine the relative flux of acyl groups around the acyl editing cycle we consider how many more FA must be removed from the mixed acyl pool by acyl editing compared de novo synthesis to produce the different [”C]FA labeling amounts and stereochemistry of PC and DAG. To simplify the compositional and stereochemical differences of acylation we consider the relative flux of [“C]M into the sn-2 position of both lipids, assuming initial labeling of PC to be 100% sn-2 M and DAG to be 50% sn-2 M. In this scenario 54.1 moles nascent [“C]M enter sn-2 PC and 11.4 moles enter de novo glycerolipid synthesis (de novo DAG + other membrane lipids, Table 6-4). Therefore the flux of nascent [“CJM into sn-2 PC is a maximal 4.8 times faster than esterification of sn-2 G3P, any greater flux would lower the amount of nascent acyl groups entering de novo glycerolipid synthesis. However, if there are any enzymatic selectivites of LPAT or acyl editing enzymes there is a possibility that the flux could be lower. 181 TAG synthesis will require 31 nmoles of sn-3 acylation. About 11 moles of this flux is provided by the immediate incorporation of nascent FA, with a high preference for saturates (Table 64). Analysis of the molecular species of [“ClFA labeled TAG suggests that this direct acylation utilizes the bulk DAG pool, not the de novo synthesized DAG pool, and provides about 35% of the molecular species in endogenous TAG. The other 65% of TAG synthesis would be compensated by utilizing the recycled acyl-CoA pool from acyl edting and also PDAT reactions. 182 7 APPENDIX C: PC ACYL EDITING IN ISOLATED PEA CHLOROPLASTS Introduction Isolated pea chloroplasts have been shown to channel newly synthesized FA into PC, even though chloroplasts cannot synthesized PC de novo [79]. To determine if acyl editing is involved in this channeling of newly synthesized FA into PC, isolated pea leaf chloroplasts were labeled with [“C]acetate. The [“CJFA labeled PC was analysed for labeled FA composition, regiochemistry, and molecular species. Results and Discussion After 5 min of [14C]acetate labeling of isolated pea chloroplasts over half the radioactivity was in free fatty acids and acyl CoA as reported previously [79]. PC contained the most labeled FA out of all labeled polar lipids and was entirely [“C]18:1. The stereochemical analysis indicated that ~93% of labeled FA were localized to the sn-2 position. Separation of labeled molecular species (Figure 7- 1) demonstrated that nascent [14C]FA were incorporated next to unlabeled saturates, mono-, di- and tri-unsaturates. Together these results are very similar to the acyl editing measured by in vivo labeling of pea leaves, with the exception of a highly sn-2 specific FA incorporation. 183 60 1 IPC 50 I PG 40 i 30 J 20 L 10 % 14C of total molecular species 0 . SS SD ST SM MM MD MT DD DT TT Figure 7-1 [“C]acetate labeled PC and PG molecular species from isolated chloroplasts. Isolated chloroplasts were labeled with [14C]acetate for five minutes. Molecular species of PC and PG were separated by argentation TLC. Bar heights represent percent of labeled species among each lipid. Molecular species are represented as a pair of FA. S, saturates; M, monounsaturates; D, diunsaturates, triunsaturates. Black bars, PC. Gray bars, PG. 184 The radiolabeled FA composition and molecular species of the prokaryotic lipid PG were analyzed to ensure that the PC labeling results were from acyl editing and not due to a transfer of the PC phosphocholine headgroup to a prokaryotic pathway synthesized DAG. PG from five minute [“C]acetate labeled pea leaf chloroplast contained approximately equal amounts of saturated and monounsaturated radiolabeled FA. The molecular species labeled were predominantly fully saturated and saturated/monounsatured (Figure 7-1). Together the PG labeling is the expected result of a de novo synthesized prokaryotic lipid. The difference in PG and PC labeling confirms that incorporation of newly synthesized FA into PC of isolated chloroplasts is not from prokaryotic pathway produced DAG. Therefore, incorporation of newly synthesized FA into PC of isolated pea chloroplasts is through acyl editing of the sn-2 position. In pea leaves PC sn-2 acyl-editing may be located within the outer chloroplast envelope or associated with the chloroplast through the PLAM region. The lack of PC sn-1acyl editing in isolated chloroplasts compared to intact leaves may be because this process takes place within a different cellular membrane, involves soluble enzymes, or the activity may be disrupted by chloroplast isolation. Methods Pea plants were grown, chloroplasts isolated, labeling conditions and lipid extraction were as done previously [79]. Labeling was started by adding 5 pCi of 50 mCi/mmol [“C]acetate resuspened in labeling media to 0.4 mg chlorophyll of chloroplasts with a final volume of 0.2 ml. Chloroplasts were incubated at room 185 temperature under ~180 mol of photons/m2/s of white light, with gentle shaking for 5 min. The labeling reaction was stopped by adding 2 ml of chloroform/methanol (1 :1, v/v) to start the lipid extraction. Lipid separation, radioactive fatty acid composition, regiochemistry and molecular species were determined as done previously [162]. 186 10. 11. 12. 13. 14. REFERENCES Chapman, K.D., Occurrence, metabolism, and prospective functions of N- acylethanolamines in plants. Progress in Lipid Research, 2004. 43(4): p. 302-327. Cheong, J.J. and Y.D. Choi, Signaling pathways for the biosynthesis and action of jasmonates. Journal of Plant Biology, 2007. 50(2): p. 122-131. Wang, X.M., Lipid signaling. Current Opinion in Plant Biology, 2004. 7(3): p. 329-336. Wang, X.M., et al., Signaling functions of phosphatidic acid. Progress in Lipid Research, 2006. 45(3): p. 250-278. Shah, J., Lipids, lipases, and lipid-modifying enzymes in plant disease resistance. Annual Review of Phytopathology, 2005. 43: p. 229-260. Nawrath, C., Unraveling the complex network of cuticular structure and function. Current Opinion in Plant Biology, 2006. 9(3): p. 281-287. Franke, R. and L. Schreiber, Suben'n - a biopolyester forming apoplastic plant interfaces. Current Opinion in Plant Biology, 2007. 10(3): p. 252-259. Kolattukudy, P.E., Structure, biosynthesis, and biodegradation of cutin and suberin, Annual Review of Plant Physiology and Plant Molecular Biology, 1981. 32: p. 539-567. Koch, K. and W. Barthlott, Plant epicuticular waxes: Chemistry, form, self- assembly and function. Natural Product Communications, 2006. 1(11): p. 1067-1 072. Kunst, L. and AL. Samuels, Biosynthesis and secretion of plant cuticular wax. Progress in Lipid Research, 2003. 42(1): p. 51-80. Long, L.M., et al., The maize epicuticular wax layer provides UV protection. Functional Plant Biology, 2003. 30(1): p. 75-81. Dyer, J.M., et al., High-value oils from plants. The Plant Journal, 2008. 54(4): p. 640-655. Durrett, T.P., C. Benning, and J. Ohlrogge, Plant triacylglycerols as feedstocks for the production of biofuels. The Plant Journal, 2008. 54(4): p. 593-607. Guschina, LA. and J.L. Harwood, Lipids and lipid metabolism in eukaryotic algae. Progress in Lipid Research, 2006. 45(2): p. 160-186. 187 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. McKeon, T.A., J.T. Lin, and A.E. Stafford, Biosynthesis of ricinoleate in castor oil, in Chemicals Via Higher Plant Bioengineering. 1999, Kluwer Academic l Plenum Publ: New York. p. 37-47. Vogel, G. and J. Browse, Cholinephosphotransferase and diacylglycerol acyltransferase - Substrate specificities at a key branch point in seed lipid metabolism. Plant Physiology, 1996. 110(3): p. 923-931. Jaworski, J. and EB. Cahoon, Industrial oils from transgenic plants. Current Opinion in Plant Biology, 2003. 6(2): p. 178-184. Drexler, H., et al., Metabolic engineering of fatty acids for breeding of new oilseed crops: strategies, problems and first results. Journal of Plant PhysioIOQY. 2003. 160(7): p. 779-802. Abbadi, A., et al., Biosynthesis of very-Iong-chain polyunsaturated fatty acids in transgenic oilseeds: Constraints on their accumulation. Plant Cell, 2004. 16(10): p. 2734-2748. Graham, I.A., T. Larson, and J.A. Napier, Rational metabolic engineering of transgenic plants for biosynthesis of omega-3 polyunsaturates. Current Opinion in Biotechnology, 2007. 18(2): p. 142-147. Napier, J.A., Transgenic plants as a source of fish oils: healthy, sustainable and GM. Journal of the Science of Food and Agriculture, 2007. 87(1): p. 8-12. Napier, J.A., The production of unusual fatty acids in transgenic plants. Annual Review of Plant Biology, 2007. 58: p. 295-319. Broun, P. and C. Somerville, Accumulation of ricinoleic, lesquerolic, and densipolic acids in seeds of transgenic Arabidopsis plants that express a fatty acyl hydroxylase cDNA from castor bean. Plant Physiology, 1997. 113(3): p. 933-942. Lee, M., et al., Identification of non-heme diiron proteins that catalyze triple bond and epoxy group formation. Science, 1998. 280(5365): p. 915- 91 8. Cahoon, E.B., et al., Biosynthetic origin of conjugated double bonds: Production of fatty acid components of high-value drying oils in transgenic soybean embryos. Proceedings of the National Academy of Sciences of the United States of America, 1999. 96(22): p. 12935-12940. Smith, MA, et al., Heterologous expression of a fatty acid hydroxylase gene in developing seeds of Arabidopsis thaliana. Planta, 2003. 217(3): p. 507-51 6. 188 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. Cahoon, E.B., J. Shanklin, and J.B. Ohlrogge, Expression of a coriander desaturase results in petroselinic acid production in transgenic tobacco. Proceedings of the National Academy of Sciences of the United States of America, 1992. 89(23): p. 11184-11188. Ohlrogge, J.B., D.N. Kuhn, and PK. Stumpf, Subcellular-localization of acyl canier protein in leaf protoplasts of spinacia-oleracea. Proceedings of the National Academy of Sciences of the United States of America, 1979. 76(3): p. 1194-1198. Gueguen, V., et al., Fatty acid and lipoic acid biosynthesis in higher plant mitochondria. Journal of Biological Chemistry, 2000. 275(7): p. 5016-5025. Schwender, J. and J.B. Ohlrogge, Probing in vivo metabolism by stable isotope labeling of storage lipids and proteins in developing Brassica napus embryos. Plant Physiology, 2002. 130(1): p. 347-361. Molina, L, et al., The lipid polyester composition of Arabidopsis thaliana and Brassica napus seeds. Phytochemistry, 2006. 67(23): p. 2597-2610. von Wettstein-Knowles, P., Analyses of bariey spike mutant waxes identify alkanes, cyclopropanes and internally branched alkanes with dominating isomers at carbon 9. Plant Journal, 2007. 49(2): p. 250-264. Postbeittenmiller, D., G. Roughan, and J.B. Ohlrogge, Regulation of plant fatty-acid biosynthesis - analysis of acyl-coenzyme a and acyl-acyl carrier protein substrate pools in spinach and pea-chloroplasts. Plant Physiology, 1992. 100(2): p. 923-930. Ohlrogge, J. and J. Browse, Lipid Biosynthesis. Plant Cell, 1995. 7(7): p. 957-970. McKeon, TA. and PK. Stumpf, Purification and characterization of the stearoyl-acyl canier protein desaturase and the acyl-acyl canier protein thioesterase from maturing seeds of safflower. Journal of Biological Chemistry, 1982. 257(20): p. 2141-2147. Sperling, P. and E. Heinz, Isomeric sn-1-octadecenyl and sn-2- octadecenyl analogs of lysophosphatidylcholine as substrates for acylation and desaturation by plant microsomal-membranes. European Journal of Biochemistry, 1993. 213(3): p. 965-971 . Hanrvood, J.L., Recent advances in the biosynthesis of plant fatty acids. Biochimica Et Biophysica Acta-Lipids and Lipid Metabolism, 1996. 1301(1- 2): p. 7-56. 189 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. Los, DA. and N. Murata, Structure and expression of fatty acid desaturases. Biochimica Et Biophysica Acta-Lipids and Lipid Metabolism, 1998. 1394(1): p. 3-15. Tocher, D.R., M.J. Leaver, and PA Hodgson, Recent advances in the biochemistry and molecular biology of fatty acyl desaturases. Progress in Lipid Research, 1998. 37(2-3): p. 73-1 17. Hlousek-Radojcic, A., et al., Fatty acid elongation is independent of acyl- coenzyme A synthetase activities in leek and Brassica napus. Plant Physiology. 1998. 116(1): p. 251 -258. Fehling, E. and KD. Mukherjee, Acyl-CoA elongase from a higher plant (Lunaria annua): metabolic intermediates of very-long-chain acyl-CoA products and substrate specificity. Biochimica et Biophysica Acta (BBA) - Lipids and Lipid Metabolism, 1991. 1082(3): p. 239-246. Ohlrogge, J.B. and JG. Jaworski, Regulation of fatty acid synthesis. Annual Review of Plant Physiology and Plant Molecular Biology, 1997. 48: p. 109-1 36. Konishi, T., et al., Acetyl-CoA carboxylase in higher plants: Most plants other than gramineae have both the prokaryotic and the eukaryotic forms of this enzyme. Plant and Cell Physiology, 1996. 37(2): p. 117-122. Schulte, W., et al., Multi-functional acetyI-CoA carboxylase from Brassica napus is encoded by a multi-gene family: Indication for plastidic localization of at least one isoform. Proceedings of the National Academy of Sciences of the United States of America, 1997. 94(7): p. 3465-3470. Cronan, J.E. and G.L. Waldrop, Multi-subunit acetyl-CoA carboxylases. Progress in Lipid Research, 2002. 41(5): p. 407-435. Sasaki, Y., et al., Chloroplast-encoded protein as a subunit of acetyl-coA carboxylase in pea plant. Journal of Biological Chemistry, 1993. 268(33): p. 25118-25123. Sasaki, Y., et al., Chloroplast RNA editing required for functional acetyl- CoA carboxylase in plants. Journal of Biological Chemistry, 2001. 276(6): p. 3937-3940. Shorrosh, B.S., et al., Structural-analysis, plastid localization, and expression of the biotin carboxylase subunit of acetyl-coenzyme-A carboxylase from tobacco. Plant Physiology, 1995. 108(2): p. 805-812. Shorrosh, B.S., et al., The pea chloroplast membrane-associated protein, IEP96, is a subunit of acetyl-CoA carboxylase. Plant Journal, 1996. 10(2): p. 261 -268. 190 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. Choi, J.K., et al., Molecular-cloning and characterization of the cDNA coding for the biotin-containing subunit of the chloroplastic acetyl- coenzyme—A carboxylase. Plant Physiology, 1995. 109(2): p. 619-625. Roesler, K.R., et al., Co-purification, co-immunoprecipitation, and coordinate expression of acetyl-coenzyme A carboxylase activity, biotin carboxylase, and biotin carboxyl carrier protein of higher plants. Planta, 1996. 198(4): p. 517-525. Ke, J.S., et al., Coordinate regulation of the nuclear and plastidic genes coding for the subunits of the heteromeric acetyl-coenzyme a carboxylase. Plant Physiology, 2000. 1 22(4): p. 1057-1071. Madoka, Y., et al., Chloroplast transformation with modified ach operon increases acetyl-CoA carboxylase and causes extension of leaf longevity and increase in seed yield in tobacco. Plant and Cell Physiology, 2002. 43(12): p. 1518-1525. Thelen, J.J. and J.B. Ohlrogge, The multisubunit acetyl-CoA carboxylase is strongly associated with the chloroplast envelope through non-ionic interactions to the carboxyltransferase subunits. Archives of Biochemistry and Biophysics, 2002. 400(2): p. 245-257. Ruuska, S.A., et al., Contrapuntal networks of gene expression during Arabidopsis seed filling. Plant Cell, 2002. 14(6): p. 1191-1206. Eccleston, V.S. and J.B. Ohlrogge, Expression of lauroyl-acyl carrier protein thioesterase in Brassica napus seeds induces pathways for both fatty acid oxidation and biosynthesis and implies a set point for triacylglycerol accumulation. Plant Cell, 1998. 10(4): p. 613-621. Hunter, SC. and J.B. Ohlrogge, Regulation of spinach chloroplast acetyl- CoA carboxylase. Archives of Biochemistry and Biophysics, 1998. 359(2): p. 170-178. Shintani, D.K. and J.B. Ohlrogge, Feedback inhibition of fatty-acid synthesis in tobacco suspension cells. Plant Journal, 1995. 7(4): p. 577- 587. Browse, J., P.G. Roughan, and CR. Slack, Light control of fatty-acid synthesis and diumal fluctuations of fatty-acid composition in leaves. Biochemical Journal, 1981. 196(1): p. 347-354. Sasaki, Y., A. Kozaki, and M. Hatano, Link between light and fatty acid synthesis: Thioredoxin-Iinked reductive activation of plastidic acetyl-CoA carboxylase. Proceedings of the National Academy of Sciences of the United States of America, 1997. 94(20): p. 11096-11 101. 191 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. Kozaki, A. and Y. Sasaki, Light-dependent changes in redox status of the plastidic acetyl-CoA carboxylase and its regulatory component. Biochemical Journal, 1999. 339: p. 541-546. Kozaki, A., et al., Recombinant carboxyltransferase responsive to redox of pea plastidic acetyl-CoA carboxylase. Journal of Biological Chemistry, 2000. 275(14): p. 10702-10708. Kozaki, A.K., K. Mayumi, and Y. Sasaki, ThioI-disulfide exchange between nuclear-encoded and chloroplast-encoded subunits of pea acetyl-CoA carboxylase. Journal of Biological Chemistry, 2001. 276(43): p. 39919- 39925. Savage, L.J. and J.B. Ohlrogge, Phosphorylation of pea chloroplast acetyl-CoA carboxylase. Plant Journal, 1999. 18(5): p. 521-527. Shintani, D., et al., Antisense expression and overexpression of biotin carboxylase in tobacco leaves. Plant Physiology, 1997. 114(3): p. 881- 886. Thelen, J.J. and J.B. Ohlrogge, Both antisense and sense expression of biotin carboxyl carrier protein isoform 2 inactivates the plastid acetyl- coenzyme A carboxylase in Arabidopsis thaliana. Plant Journal, 2002. 32(4): p. 419-431. Roesler, K., et al., Targeting of the Arabidopsis homomeric acetyl- coenzyme A carboxylase to plastids of rapeseeds. Plant Physiology, 1997. 113(1): p. 75-81. Kornberg, A. and WE. Pricer, Enzymatic esterification of alpha- glycerophosphate by long chain fatty acids. Journal of Biological Chemistry, 1953. 204(1): p. 345-357. Kennedy, E.P. and SB. Weiss, Function of cytidine coenzymes in the biosynthesis of phospholipides Journal of Biological Chemistry, 1956. 222(1): p. 193-213. Weiss, S.B., E.P. Kennedy, and J.Y. Kiyasu, Enzymatic synthesis of triglycerides. Journal of Biological Chemistry, 1960. 235(1): p. 40-44. Roughan, PG. and CR. Slack, Cellular-organization of glycerolipid metabolism. Annual Review of Plant Physiology and Plant Molecular Biol09Y. 1982. 33: p. 97-132. Browse, J., et al., Fluxes through the prokaryotic and eukaryotic pathways of lipid-synthesis in the 16:3 plant Arabidopsis-thaliana. Biochemical Journal, 1986. 235(1): p. 25-31. 192 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. Marechal, E., et al., Lipid synthesis and metabolism in the plastid envelope. Physiologia Plantarum, 1997. 100(1): p. 65-77. Pollard, M. and J. Ohlrogge, Testing models of fatty acid transfer and lipid synthesis in spinach leaf using in vivo oxygen-18 labeling. Plant Physiology, 1999. 121(4): p. 1217-1226. Jones, A., H.M. Davies, and TA Voelker, Palmitoyl-acyl carrier protein (A CP) thioesterase and the evolutionary origin of plant acyl-A CP thioesterases. Plant Cell, 1995. 7(3): p. 359-371. Donnann, P., T.A. Voelker, and J.B. Ohlrogge, Cloning and expression in escherichia-coli of a novel thioesterase from arabidopsis-thaliana specific for long-chain acyl-acyl carrier proteins. Archives of Biochemistry and Biophysics, 1995. 316(1): p. 612-618. Andrews, J. and K. Keegstra, Acyl-coA synthetase is located in the outer- membrane and acyl-coA thioesterase in the inner membrane of pea chloroplast envelopes. Plant Physiology, 1983. 72(3): p. 735-740. Block, MA, et al., The acyl-00A synthetase and acyl-coA thioesterase are located on the outer and inner membrane of the chloroplast envelope, respectively. Febs Letters, 1983. 153(2): p. 377-381. Koo, A.J.K., J.B. Ohlrogge, and M. Pollard, On the export of fatty acids from the chloroplast. Journal of Biological Chemistry, 2004. 279(16): p. 16101-16110. Jouhet, J., E. Marechal, and MA. Block, Glycerolipid transfer for the building of membranes in plant cells. Progress in Lipid Research, 2007. 46(1): p. 37-55. Benning, C., 00. Xu, and K. Awai, Non-vesicular and vesicular lipid trafficking involving plastids. Current Opinion in Plant Biology, 2006. 9(3): p. 241-247. Mongrand, S., et al., The C-16 : 3/C-18 : 3 fatty acid balance in photosynthetic tissues from 468 plant species. Phytochemistry, 1998. 49(4): p. 1049-1064. Browse, J., et al., Mutants of arabidopsis deficient in the synthesis of alpha-linolenate - biochemical and genetic-characterization of the endoplasmic-reticulum linoleoyl desaturase. Journal of Biological Chemistry, 1993. 268(22): p. 16345-16351 . Roughan, P.G., R. Holland, and CR. Slack, The role of chloroplasts and microsomal fractions in polar-lipid synthesis from [acetate-1-C-14] by cell- 193 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. free preparations from spinach (Spinacia-oleracea) leaves. Biochemical Journal, 1980. 188(1): p. 17-24. Browse, J. and C. Somerville, Glycerolipid synthesis - biochemistry and regulation. Annual Review of Plant Physiology and Plant Molecular Biology, 1991. 42: p. 467-506. Slack, C.R., P.G. Roughan, and N. Balasingham, Labeling studies in vivo on metabolism of acyl and glycerol moieties of glycerolipids in developing maize leaf. Biochemical Journal, 1977. 162(2): p. 289-296. Slack, CR. and PG. Roughan, Kinetics of incorporation invivo of [C- 14 ]acetate and [carbon-G14] dioxide into fatty-acids of glycerolipids in developing leaves. Biochemical Journal, 1975. 152(2): p. 217-228. Kunst, L., J. Browse, and C. Somerville, Altered regulation of lipid biosynthesis in a mutant of Arabidopsis deficient in chloroplast glycerol-3- phosphate acyltransferase activity. Proceedings of the National Academy of Sciences of the United States of America, 1988. 85(12): p. 4143-4147. Andersson, M.X., J.M. Kjellberg, and AS. Sandelius, The involvement of cytosolic lipases in converting phosphatidyl choline to substrate for galactolipid synthesis in the chloroplast envelope. Biochimica Et Biophysica Acta-Molecular and Cell Biology of Lipids, 2004. 1684(1-3): p. 46-53. Awai, K., et al., A phosphatidic acid-binding protein of the chloroplast inner envelope membrane involved in lipid trafficking. Proceedings of the National Academy of Sciences of the United States of America, 2006. 103(28): p. 10817-10822. Bessoule, J.J., E. Testet, and C. Cassagne, Synthesis of phosphatidylcholine in the chloroplast envelope after import of lysophosphatidylcholine from endoplasmic-reticulum membranes. European Journal of Biochemistry, 1995. 228(2): p. 490-497. Mongrand, S., J.J. Bessoule, and C. Cassagne, A re-examination in vivo of the phosphatidylcholine-galactolipid metabolic relationship during plant lipid biosynthesis. Biochemical Journal, 1997. 327: p. 853-858. Mongrand, S., C. Cassagne, and J.J. Bessoule, Import of lyso- phosphatidylcholine into chloroplasts likely at the origin of eukaryotic plastidial lipids. Plant Physiology. 2000. 122(3): p. 845-852. Kjellberg, J.M., et al., Acyl-CoA dependent acylation of phospholipids in the chloroplast envelope. Biochimica Et Biophysica Acta-Molecular and Cell Biology of Lipids, 2000. 1485(2-3): p. 100-110. 194 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. Andersson, M.X., M. Goksor, and AS. Sandelius, Optical manipulation reveals strong attracting forces at membrane contact sites between endoplasmic reticulum and chloroplasts. Journal of Biological Chemistry, 2007. 282(2): p. 1170-1174. Vance, J.E., Phospholipid-synthesis in a membrane-fraction associated with mitochondria. Journal of Biological Chemistry, 1990. 265(13): p. 7248-7256. Gaigg, 8., et al., Characterization of a microsomal subfraction associated with mitochondria of the yeast, Saccharomyces—cere visiae - involvement in synthesis and import of phospholipids into mitochondria. Biochimica Et Biophysica Acta-Biomembranes, 1995. 1234(2): p. 214-220. Jouhet, J., et al., Phosphate deprivation induces transfer of DGDG galactolipid from chloroplast to mitochondria. Journal of Cell Biology, 2004. 167(5): p. 863-874. Eaton, S., K. Bartlett, and M. Pourfarzam, Mammalian mitochondrial beta- oxidation. Biochemical Journal, 1996. 320: p. 345-357. lchihara, K., K. Yamane, and E. Hirano, Acyl-CoA synthetase in oilseeds: Fatty acid structural requirements for activity and selectivity. Plant and Cell Physiology, 1997. 38(6): p. 717-724. Frentzen, M., et al., lntraorganelle localization and substrate specificities of the mitochondrial acyl-coA-sn-glycerol-3-phosphate o-acyltransferase and acyl-coA-1-acyl-sn-glycerol-3-phosphate o-acyltransferase from potato-tubers and pea leaves. European Journal of Biochemistry, 1990. 187(2): p. 395-402. Gerhardt, B., Fatty-acid degradation in plants. Progress in Lipid Research, 1992. 31(4): p. 417-446. Olsen, J.A. and KR. Lusk, Acyl-coA synthetase-activity associated with rapeseed lipid body membranes. Phytochemistry, 1994. 36(1): p. 7-9. Shockey, J.M., M.S. Fulda, and J.A. Browse, Arabidopsis contains nine long-chain acyl-coenzyme A synthetase genes that participate in fatty acid and glycerolipid metabolism. Plant Physiology, 2002. 129(4): p. 1710- 1 722. Fulda, M., et al., Two long-chain acyl-CoA synthetases from Arabidopsis thaliana involved in peroxisomal fatty acid beta-oxidation. Plant Journal, 2002. 32(1): p. 93-103. 195 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. Schnurr, J.A., et al., Fatty acid export from the chloroplast. Molecular characterization of a major plastidial acyl-coenzyme A synthetase from Arabidopsis. Plant Physiology, 2002. 129(4): p. 1700-1709. Schnurr, J., J. Shockey, and J. Browse, The acyl-CoA synthetase encoded by LACSZ is essential for normal cuticle development in Arabidopsis. Plant Cell, 2004. 16(3): p. 629-642. Burton, M., et al., Evolution of the acyl-CoA binding protein (A CBP). Biochemical Journal, 2005. 392: p. 299-307. Kojima, M., et al., The effects of down-regulating expression of Arabidopsis thaliana membrane-associated acyl-CoA binding protein 2 on acyl-lipid composition. Plant Science, 2007. 172(1): p. 36-44. Engeseth, N.J., et al., Characterization of an Acyl-CoA-binding protein from Arabidopsis thaliana. Archives of Biochemistry and Biophysics, 1996. 331(1): p. 55-62. Leung, K.C., et al., ACBP4 and ACBP5, novel Arabidopsis acyl-COA- binding proteins with kelch motifs that bind oleoyl-CoA. Plant Molecular Biology. 2004. 55(2): p. 297-309. Chye, M.L., B.Q. Huang, and S.Y. Zee, Isolation of a gene encoding Arabidopsis membrane-associated acyl-CoA binding protein and immunolocalization of its gene product. Plant Journal, 1999. 18(2): p. 205- 214. Li, H.Y. and ML. Chye, Membrane localization of Arabidopsis acyl-CoA binding protein ACBP2. Plant Molecular Biology, 2003. 51(4): p. 483-492. Leung, K.C., et al., Arabidopsis ACBP3 is an extracellularly targeted acyl- CoA-binding protein. Planta, 2006. 223(5): p. 871-881. Benning, C. and H. Ohta, Three enzyme systems for galactoglycerolipid biosynthesis are coordinately regulated in plants. Journal of Biological Chemistry, 2005. 280(4): p. 2397-2400. Voelker, T. and AT Kinney, Variations in the biosynthesis of seed-storage lipids. Annual Review of Plant Physiology and Plant Molecular Biology, 2001. 52: p. 335-361. Slack, C.R., et al., Some evidence for the reversibility of the "cholinephosphotransferase-catalysed reaction in developing linseed cotyledons in vivo. Biochimica Et Biophysica Acta, 1983. 754(1): p. 10-20. 196 118. 119. 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. Slack, C.R., et al., Some properties of cholinephosphotransferase from developing safflower cotyledons. Biochimica Et Biophysica Acta, 1985. 833(3): p. 438-448. Dahlqvist, A., et al., Phospholipid : diacylglycerol acyltransferase: An enzyme that catalyzes the acyl-CoA-independent formation of triacylglycerol in yeast and plants. Proceedings of the National Academy of Sciences of the United States of America, 2000. 97(12): p. 6487-6492. Stahl, U., et al., Cloning and functional characterization of a Phospholipid : Diacylglycerol acyltransferase from Arabidopsis. Plant Physiology. 2004. 135(3): p. 1324-1335. Mhaske, V., et al., Isolation and characterization of an Arabidopsis thaliana knockout line for phospholipid: diacylglycerol transacylase gene (At5g13640). Plant Physiology and Biochemistry, 2005. 43(4): p. 413-417. Stobart, K., et al., Triacylglycerols are synthesised and utilized by transacylation reactions in microsomal preparations of developing safflower (Carthamus tinctorius L) seeds. Planta, 1997. 203(1): p. 58-66. Zou, J.T., et al., The Arabidopsis thaliana TAG1 mutant has a mutation in a diacylglycerol acyltransferase gene. Plant Journal, 1999. 19(6): p. 645- 653. Zhang, F.Y., M.F. Yang, and Y.N. Xu, Silencing of DGAT1 in tobacco causes a reduction in seed oil content. Plant Science, 2005. 169(4): p. 689-694. Singh, S.P., et al., Metabolic engineering of new fatty acids in plants. Current Opinion in Plant Biology. 2005. 8(2): p. 197-203. Wilson, R.F., et al., Involvement of phospholipids in poly-unsaturated fatty- acid synthesis in developing soybean cotyledons. Plant Physiology, 1980. 66(4): p. 545-549. Slack, C.R., P.G. Roughan, and N. Balasingham, Labeling of glycerolipids in cotyledons of developing oilseeds by [1-C-14]acetate and [2-H- 3]glycerol. Biochemical Journal, 1978. 170(2): p. 421 -433. Millar, A.A., M.A. Smith, and L. Kunst, All fatty acids are not equal: discrimination in plant membrane lipids. Trends in Plant Science, 2000. 5(3): p. 95-101. Moreau, RA. and PK. Stumpf, Recent studies of the enzymic-synthesis of ricinoleic acid by developing castor beans. Plant Physiology, 1981. 67(4): p. 672-676. 197 130. 131. 132. 133. 134. 135. 136. 137. 138. 139. 140. Bafor, M., et al., Ricinoleic acid biosynthesis and triacylglycerol assembly in microsomal preparations from developing caster-bean (Ricinus- communis) endosperm. Biochemical Journal, 1991. 280: p. 507-514. Stahl, U., A. Banas, and S. Stymne, Plant microsomal phospholipid acyl hydrolases have selectivities for uncommon fatty-acids. Plant Physiology, 1995. 107(3): p. 953-962. Kroon, J.T.M., et al., Identification and functional expression of a type 2 acyl-CoA:diacylglycerol acyltransferase (DGA T2) in developing castor bean seeds which has high homology to the major triglyceride biosynthetic enzyme of fungi and animals. Phytochemistry, 2006. 67(23): p. 2541 - 2549. Shockey, J.M., et al., Tung tree DGA T1 and DGA T2 have nonredundant functions in triacylglycerol biosynthesis and are localized to different subdomains of the endoplasmic reticulum. Plant Cell, 2006. 18(9): p. 2294-2313. Beisson, F., et al., Arabidopsis genes involved in acyl lipid metabolism. A 2003 census of the candidates, a study of the distribution of expressed sequence tags in organs, and a Web-based database. Plant Physiology, 2003. 132(2): p. 681-697. Lands, W.E.M., Lipid metabolism. Annual Review of Biochemistry, 1965. 34: p. 31 3-&. Purdon, A.D., et al., Energy consumption by phospholipid metabolism in mammalian brain. Neurochemical Research, 2002. 27(12): p. 1641-1647. Purdon, AD. and SI. Rapoport, Energy requirements for two aspects of phospholipid metabolism in mammalian brain. Biochemical Journal, 1998. 335: p. 313-318. de Kroon, A., Metabolism of phosphatidylcholine and its implications for lipid acyl chain composition in Saccharomyces cerevisiae. Biochimica Et Biophysica Acta-Molecular and Cell Biology of Lipids, 2007. 1771(3): p. 343-352. Kol, MA, et al., Uptake and remodeling of exogenous phosphatidylethanolamine in E-coli. Biochimica Et Biophysica Acta- Molecular and Cell Biology of Lipids, 2004. 1636(2-3): p. 205-212. Stymne, S. and AK. Stobart, Evidence for the reversibility of the acyl-coA- lysophosphatidylcholine acyltransferase in microsomal preparations from developing safflower (Carthamus-tinctorius L) cotyledons and rat-liver. Biochemical Journal, 1984. 223(2): p. 305-314. 198 141. 142. 143. 144. 145. 146. 147. 148. 149. 150. 151. Stymne, 8., AK. Stobart, and G. Glad, The role of the acyl-coA pool in the synthesis of poly-unsaturated 18-carbon fatty-acids and triacylglycerol production in the microsomes of developing safflower seeds. Biochimica Et Biophysica Acta, 1983. 752(2): p. 198-208. Rochester, CF. and D.G. Bishop, The role of lysophosphatidylcholine in lipid-synthesis by developing sunflower (Helianthus-annuus L) seed microsomes. Archives of Biochemistry and Biophysics, 1984. 232(1): p. 249-258. Serghinicaid, H., et al., Oleoyl-phosphatidy/choline molecular-species desaturated in pea leaf microsomes - possible substrates of oleate- desaturase in other green leaves. Plant Science, 1988. 54(2): p. 93-101. Demand re, 0., et al., Phosphatidylcholine molecular-species formed by lysophosphatidylcholine acyltransferase from soya bean microsomes. Phytochemistry, 1994. 35(5): p. 1171 -1 175. Bertrams, M., K. Wrage, and E. Heinz, Lipid labeling in intact chloroplasts from exogenous nucleotide precursors. Zeitschrift Fur Naturforschung C-a Journal of Biosciences, 1981. 36(1-2): p. 62-70. Wang, X.M., Plant phospholipases. Annual Review of Plant Physiology and Plant Molecular Biology, 2001. 52: p. 211-+. Noiriel, A., et al., Expression in yeast of a novel phospholipase A1 cDNA from Arabidopsis thaliana. European Journal of Biochemistry, 2004. 271(18): p. 3752-3764. lshiguro, S., et al., The DEFECTIVE IN ANTHER DEHISCENCE1 gene encodes a novel phospholipase A1 catalyzing the initial step of jasmonic acid biosynthesis, which synchronizes pollen maturation, anther dehiscence, and flower opening in Arabidopsis. Plant Cell, 2001. 13(10): p. 2191-2209. Williams, JP, et al., The role of phosphatidylcholine in fatty acid exchange and desaturation in Brassica napus L. leaves. Biochemical Journal, 2000. 349: p. 127-133. Schwender, J., et al., Rubisco without the Calvin cycle improves the carbon efficiency of developing green seeds. Nature, 2004. 432(7018): p. 779-782. Ratcliffe, R.G. and Y. Shachar—Hill, Measuring multiple fluxes through plant metabolic networks. Plant Journal, 2006. 45(4): p. 490-511. 199 152. 153. 154. 155. 156. 157. 158. 159. 160. 161. 162. 163. Cahoon, EB. and J.B. Ohlrogge, Apparent role of phosphatidylcholine in the metabolism of petroselinic acid in developing umbelliferae endosperm. Plant Physiology, 1994. 104(3): p. 845-855. Schultz, OJ. and J.B. Ohlrogge, Biosynthesis of triacylglycerol in Thunbergia alata: Additional evidence for involvement of phosphatidylcholine in unusual monoenoic oil production. Plant Physiology and Biochemistry, 2000. 38(3): p. 169-175. Somerville, C. and J. Browse, Plant lipids - Metabolism, mutants, and membranes. Science, 1991. 252(5002): p. 80-87. Hara, A. and NS. Radin, Lipid extraction of tissues with a low-toxicity solvent. Analytical Biochemistry, 1978. 90(1): p. 420-426. Arnon, D. l., Copper enzymes in isolated chloroplasts - polyphenoloxidase in Beta-Vulgaris. Plant Physiology, 1949. 24(1): p. 1-15. Khan, M.U. and JP. Williams, lMproved thin-layer chromatographic method for separation of major phospholipids and glycolipids from plant lipid extracts and phosphatidyl glyceml and bis-(monoacylglyceryl) phosphate from animal lipid extracts. Journal of Chromatography, 1977. 140(2): p. 179-185. lchihara, K., et al., An improved method for rapid analysis of the fatty acids of glycerolipids. Lipids, 1996. 31(5): p. 535-539. Christie, W.W., Lipid Analysis: Isolation, Separation, Identification and Structural Analysis of Lipids. 3rd ed. 2003, Bridgwater, England: The Oily Press an imprint of PJ Barnes & Associates. Schwender, J., Y. Shachar—Hill, and J.B. Ohlrogge, Mitochondrial metabolism in developing embryos of Brassica napus. Journal of Biological Chemistry, 2006. 281(45): p. 34040-34047. Bonaventure, G., et al., Metabolic responses to the reduction in palmitate caused by disruption of the FA TB gene in Arabidopsis. Plant Physiology, 2004. 135(3): p. 1269-1279. Bates, P.D., J.B. Ohlrogge, and M. Pollard, Incorporation of Newly Synthesized Fatty Acids into Cytosolic Glycemlipids in Pea Leaves Occurs via Acyl Editing. Journal of Biological Chemistry, 2007. 282(43): p. 31206- 31 216. lchihara, K., et al., 1-Acylglycerophosphocholine O-acyltransferase in maturing safflower seeds. Planta, 1995. 1 96(3): p. 551 -557. 200 164. 165. 166. 167. 168. 169. 170. 171. 172. 173. 174. 175. Cases, 8., et al., Identification of a gene encoding an acyl CoA : diacylglycerol acyltransferase, a key enzyme in triacylglycerol synthesis. Proceedings of the National Academy of Sciences of the United States of America, 1998. 95(22): p. 13018-13023. Cases, 8., et al., Cloning of DGA T2, 3 second mammalian diacylglycerol acyltransferase, and related family members. Journal of Biological Chemistry, 2001. 276(42): p. 38870-38876. Lard izabal, K.D., et al., DGA T2 is a new diacylglycerol acyltransferase gene family - Purification, cloning, and expression in insect cells of two polypeptides from Mortierella ramanniana with diacylglycerol acyltransferase activity. Journal of Biological Chemistry, 2001. 276(42): p. 38862-38869. Saha, S., et al., Cytosolic triacylglycerol biosynthetic pathway in oilseeds. Molecular cloning and expression of peanut cytosolic diacylglycerol acyltransferase. Plant Physiology, 2006. 141(4): p. 1533-1543. Griffiths, G., S. Stymne, and AK. Stobart, Phosphatidylcholine and its relationship to triacylglycerol biosynthesis in oil-tissues. Phytochemistry, 1988. 27(7): p. 2089-2093. Griffiths, G., S. Stymne, and AK. Stobart, The utilization of fatty-acid substrates in triacylglycerol biosynthesis by tissue-slices of developing safflower (Carthamus-tinctorius L) and sunflower (Helianthus-annuus L) cotyledons. Planta, 1988. 173(3): p. 309-316. Allen, D.K., Ohlrogge, J.B., Shachar-Hill, Y., The role of light in soybean seed filling metabolism. Currently submitted to Plant Journal, 2008. Pollard, MR. and P.K. Stumpf, Long-chain (C-20 and C-22) fatty-acid biosynthesis in developing seeds of Tropaeolum-majus - an in vivo study. Plant Physiology, 1980. 66(4): p. 641-648. Pollard, MR. and P.K. Stumpf, Biosynthesis of C-20 and C-22 fatty-acids by developing seeds of Limnanthes-alba - chain elongation and delta-5 desaturation. Plant Physiology, 1980. 66(4): p. 649-655. Triki, S., C. Demandre, and P. Mazliak, Biosynthesis of triacylglycerols by developing sunflower seed microsomes. Phytochemistry, 1999. 52(1): p. 55-62. Wallis, J.G. and J. Browse, Mutants of Arabidopsis reveal many roles for membrane lipids. Progress in Lipid Research, 2002. 41(3): p. 254-278. Rajasekharan, R. and V. Nachiappan, Use of photoreactive substrates for characterization of Iysophosphatidate acyltransferases from developing 201 176. 177. 178. 179. 180. 181. 182. 183. 184. 185. soybean cotyledons. Archives of Biochemistry and Biophysics, 1994. 311(2): p. 389-394. Katavic, V., et al., Alteration of seed fatty-acid composition by an ethyl methanesulfonate-induced mutation in Arabidopsis-thaliana affecting diacylglycerol acyltransferase activity. Plant Physiology. 1995. 108(1): p. 399-409. Ghosal, A., et al., Saccharomyces cerevisiae phospholipid : diacylglycerol acyl transferase (PDA T) devoid of its membrane anchor region is a soluble and active enzyme retaining its substrate specificities. Biochimica Et Biophysica Acta-Molecular and Cell Biology of Lipids, 2007. 1771: p. 1457-1463. Allen, D.K., Y. Shachar—Hill, and J.B. Ohlrogge, Compartment-specific labeling information in C-13 metabolic flux analysis of plants. Phytochemistry, 2007. 68(16-18): p. 2197-2210. Han, XL. and R.W. Gross, Quantitative analysis and molecular species fingerprinting of triacylglyceride molecular species directly from lipid extracts of biological samples by electrospra y ionization tandem mass spectrometry. Analytical Biochemistry, 2001. 295(1): p. 88-100. Perry, H.J. and J.L. Harwood, Radiolabeling studies of acyl lipids in developing seeds of Brassica-napus - use of [1 -C-14]acetate precursor. Phytochemistry, 1993. 33(2): p. 329-333. Perry, H.J. and J.L. Harwood, Radiolabeling studies of acyl lipids in developing seeds of Brassica-napus .2. Use of [2-H3jglycerol precursor in radiolabeling studies of acyl lipids in developing seeds of Brassica-napus. Phytochemistry, 1993. 34(1): p. 69-73. Slack, CR. and PG. Roughan, Rapid temperature-induced changes in fatty-acid composition of certain lipids in developing linseed and soybean cotyledons. Biochemical Journal, 1978. 170(2): p. 437-439. Bao, X.M. and J. Ohlrogge, Supply of fatty acid is one limiting factor in the accumulation of triacylglycerol in developing embryos. Plant Physiology, 1999. 120(4): p. 1057-1062. Stahl, U., et al., A family of eukaryotic lysophospholipid acyltransferases with broad specificity. FEBS Lett, 2008. 582(2): p. 305-9. Salas, J.J. and J.B. Ohlrogge, Characterization of substrate specificity of plant FatA and FatB acyl-ACP thioesterases. Archives of Biochemistry and Biophysics, 2002. 403(1): p. 25-34. 202 186. 187. Bonaventure, G., et al., Disruption of the FA TB gene in Arabidopsis demonstrates an essential role of saturated fatty acids in plant growth. Plant Cell, 2003. 15(4): p. 1020-1033. Cahoon, E.B., et al., Engineering oilseeds for sustainable production of industrial and nutritional feedstocks: solving bottlenecks in fatty acid flux. Current Opinion in Plant Biology, 2007. 10(3): p. 236-244. 203 l l 336 II II LL95 129 3 a...“ h. A; .. 1 .at. w. ..