"1‘. U ‘1. «~21..- ,, 1 ”o. :g‘flfi ‘ fiiig‘ This is to certify that the dissertation entitled Understanding Gene Expression and Metabolic Controls on Fatty Acid Biosynthesis in Plants. presented by Gustavo Bonaventure has been accepted towards fulfillment of the requirements for the Ph.D. /degc% in Genetics V/ Mr“ Major Jaw-I 6-1194”? Date rofessor s Signature MSU is an Affirmative Action/Equal Opportunity Institution Michigan State University 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 APR 2 1 2006 to k: p“'l WQ—t Em ; 30%5‘ Li 6/01 c:/CIRC/DateDue.p65-p.15 UNDERSTANDING GENE EXPRESSION AND METABOLIC CONTROLS ON FATTY ACID BIOSYNTHESIS IN PLANTS By Gustavo Bonaventure A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Genetics Graduate Program 2004 ABSTRACT UNDERSTANDING GENE EXPRESSION AND METABOLIC CONTROLS ON FATTY ACID BIOSYNTHESIS IN PLANTS. By Gustavo Bonaventure Plant fatty acids represent an important renewable source of highly reduced carbon chains that can be used for purposes ranging from food production to industrial feedstocks. So far, many attempts to engineer fatty acid production in plants have been based on current knowledge of the fatty acid metabolic pathway and the structural genes involved in it. However, plant tissues have proven to be not easily amenable to fatty acid engineering by overexpression of single FAS enzymes. These experiments lead to the conclusion that identification of regulatory factors limiting fatty acid production in plant tissues is crucial for manipulation of this metabolic pathway. At present, these factors have not been identified. Therefore, one major goal of this study has been to understand the signals and mechanisms that control fatty acid biosynthesis and the expression of genes for this primary metabolic pathway in plants. The results presented in this study demonstrate that acyl carrier protein (ACP) genes in Arabidopsis are under multiple levels of controls and suggest that some signals known to activate fatty acid gene expression in non-plant organisms have been conserved in plants (e. g., transcriptional activation by growth/cell cycle) whereas others differ (e.g., transcriptional activation by feedback mechanisms). Light has an important role on expression of ACP genes, increasing ACP4 mRNA levels and promoting the association of ACP transcripts with polyribosomes. Thus, this study demonstrates that different ACP genes are regulated by specific signals depending on the tissue and its fatty acid requirements. In addition, this thesis has explored the production of saturated fatty acids and their partitioning between diverse cellular processes by focusing on the characterization of an Arabidopsis acyl-ACP thioesterase mutant (fatb-ko). Analysis of this mutant has revealed the central role of FATB in the production of saturated fatty acids in all tissues. Moreover, redistribution of these molecules between different biosynthetic pathways in fatb-ko suggests that cells prioritize the synthesis of critical components for growth. In addition, slower growth and production of non-viable seeds in the mutant demonstrate an essential role of saturated fatty acids in plant growth and seed development. Isotope labeling experiments and western blots were used to investigate fatty acid metabolism and protein expression in fatb-ko leaves. The results indicate that fatb-ko leaves increase the rate of fatty acid synthesis by 40 % compared to wild type. The mutant also increases by 70% the initial rate of fatty acid breakdown. Western blot analysis reveal that BCCP, ACPs and 18:0-ACP desaturase levels are increased by 1.5- to 2-fold in mutant leaves compared to wild type. Thus, plant cells appear to have mechanisms that sense and respond to subnormal levels of saturated fatty acids in the cytoplasm. These mechanisms probably coordinate the requirements for lipid synthesis in the cytosol and fatty acid production inside the plastids. Moreover, the same mechanisms may also activate expression of nuclear genes for fatty acid synthesis. ACKNOWLEDGMENTS Many people have contributed to this research and my graduate education. My thanks to John Ohlrogge, my advisor, for giving me the opportunity to do my graduate studies in his lab and for his support over the years. This work would not have been possible without his guidance, the excellent working environment that his lab is, and the freedom I found to follow my ideas. My committee members, David Amosti, Christoph Benning, and Steve Triezenberg provided invaluable assistance, ideas and feedback over the years. Amazingly, although they have extremely busy schedules they have always been willing to meet with me, as well as with dozens of other grad students, to share their time and knowledge. Members of the Ohlrogge lab taught me many things in my time here. Special thanks to Mike Pollard for teaching me how to work with lipids on the bench and sharing his time and knowledge with me. Special thanks to David Schultz, Xiaoming B30 and Abe Koo and also to Fred, Quino, Katrin, Joerg, Guiyun, Ajay, Sari, Vandana, Naima, Troy, Linda, Anne, Jay, Phill, Greg, Yonghua, Sergei, etc. Jeannine Lee from the Genetics Graduate program has also been of invaluable assistance over the years. My thanks to Helmut Bertrand for being an excellent Director of the Genetics Program and for helping me and my wife to establish at Michigan State University. Alicia Pastor- Lecha and Ewa Danielewicz from the MSU Microscopy Facility provided me with invaluable help with the Electron and Laser Confocal microscopes. Finally, my thanks to my wife, Veronica, for all her love and for giving me total support and cheering me up during the tough periods of research when nothing seems to work , and also for teaching me to be more patient and tolerant in both science and life. To my daughter Marina for iv her unconditional love and teaching me that parenthood is totally compatible with science and sometimes even inspiring. Without my parents and sisters none of my achievements would have been possible. To all my friends at MSU that made all this possible: Rodrigo, Maite, Hector, Lizzy, Gustavo, Claribel, Chris, Xiao, Emilio, Francisco, Setsuko, and many many others. TABLE OF CONTENTS LIST OF FIGURES .......................................................................................................... IX LIST OF TABLES ............................................................................................................ XI CHAPTER 1 ....................................................................................................................... 1 INTRODUCTION .............................................................................................................. 1 Objectives ................................................................................................................... 6 Understanding regulation of F AS genes in Arabidopsis ............................................. 6 Transcriptional regulation of genes for fatty acid synthesis .................................... 7 Transcriptional regulation of F AS genes in microorganisms .................................. 7 Transcriptional regulation of FAS genes in animals ................................................ 9 Transcriptional regulation of FAS genes in plants ................................................ 10 Summary of Chapter 2 .............................................................................................. 12 Understanding regulation of fatty acid metabolism and FAS gene expression in a fatb mutant of Arabidopsis ........................................................................................ 13 Fatty acid synthesis in plants ................................................................................. 15 The role of fatty acids in the synthesis of different cellular components .............. 22 Membrane glycerolipids ........................................................................................ 22 Other fatty acid-derived products .......................................................................... 24 Regulation of fatty acid biosynthesis in plants ...................................................... 25 Regulation by light ................................................................................................. 25 Regulation by feedback inhibition ......................................................................... 26 Regulation by demand ........................................................................................... 27 LITERATURE CITED ................................................................................................. 28 CHAPTER 2 ..................................................................................................................... 33 DIFFERENTIAL REGULATION OF MESSENGER RNA LEVELS OF ACP ISOFORMS IN ARABIDOPSIS ...................................................................................... 33 ABSTRACT .................................................................................................................. 33 INTRODUCTION ........................................................................................................ 34 RESULTS ..................................................................................................................... 37 ACP4 mRNA Levels are Increased by Light ............................................................ 37 Polyribosomal Association of ACP mRNAs is Increased by Light ......................... 40 Polyribosomal association of ACP mRNAs in developing seeds of wild type B.napus and transgenics over—expressing MCTE ..................................................... 42 Conserved Motifs occur in the 5’ Leader Region of ACP mRNAs .......................... 43 The 5’ Leader Sequences of ACP] and ACP2 Increase Reporter Gene Expression 46 ACP mRNA Levels are Affected by a Sucrose-derived Signal and/or Growth Control in Cell Suspension Cultures ......................................................................... 49 vi ACP2 mRNA Levels are Affected by a Sucrose-derived Signal and/or Growth Control in planta ....................................................................................................... 52 Transcriptional regulation of Arabidopsis ACP2 gene by growth/cell cycle signals 54 Deletion analysis of Arabidopsis ACP2 promoter in tobacco BY-2 cells ................ 59 Conserved elements in Arabidopsis ACP promoters ................................................ 62 Changes in mRNA expression by inhibition of fatty acid synthesis in Arabidopsis cells ........................................................................................................................... 65 FAS genes do not respond to inhibition of fatty acid biosynthesis in Arabidopsis cells ........................................................................................................................... 67 DISCUSSION ............................................................................................................... 69 MATERIALS AND METHODS .................................................................................. 78 LITERATURE CITED ................................................................................................. 87 CHAPTER 3 ..................................................................................................................... 94 DISRUPTION OF THE FA TB GENE IN ARABIDOPSIS DEMONSTRATES AN ESSENTIAL ROLE OF SATURATED FATTY ACIDS IN PLANT GROWTH .......... 94 ABSTRACT .................................................................................................................. 94 INTRODUCTION ........................................................................................................ 95 RESULTS ................................................................................................................... 100 Mutant Isolation and Complementation Analysis ................................................... 100 FATB mRNA Expression Analysis ........................................................................ 102 FATB is Essential for Normal Seedling Growth .................................................... 103 Leaf and chloroplast morphology of wild type Arabidopsis and fab-k0 ................ 106 F ATB is Essential for Normal Seed Morphology and Germination ....................... 109 Fatty Acid Composition of fatb-ko Tissues. ........................................................... 110 Fatty Acid Composition of Individual Leaf Glycerolipids. .................................... 112 Acyl-ACP Thioesterase Activity ............................................................................ 112 Total Palmitate Content in Arabidopsis Leaf Tissue .............................................. 114 Leaf and Stem Surface Wax Analysis .................................................................... 115 Sphingoid Base Analysis ........................................................................................ 116 Leaf and stem cutin analysis ................................................................................... 117 fatb-ko act] double mutant ..................................................................................... 119 DISCUSSION ............................................................................................................. 121 Seedling Growth and Seed Development ............................................................... 123 Reduced Export of Palmitate in fatb-ko plants and Other Sources of Palmitate and Stearate in the Cell .................................................................................................. 125 Partition of Palmitate and Stearate to Non—glycerolipid Products. ......................... 127 Conclusions ............................................................................................................. 131 MATERIALS AND METHODS ................................................................................ 133 LITERATURE CITED ............................................................................................... 142 CHAPTER 4 ................................................................................................................... 146 vii DISRUPTION OF THE FA TB GENE INCREASES THE RATE OF FATTY ACID BIOSYNTHESIS,FATTY ACID AND LIPID TURNOVER, AND F AS PROTEIN EXPRESSION IN LEAVES. .......................................................................................... 146 ABSTRACT ................................................................................................................ 146 INTRODUCTION ...................................................................................................... 147 RESULTS ............... . .................................................................................................... l 5 1 Rate of fatty acid biosynthesis in leaves of wild type Arabidopsis and fatb-ko. 151 Rate of fatty acid turnover in wild type and fatb-ko leaves .................................... 153 Analysis of polar lipid classes in wild type and fatb-ko leaves .............................. 154 Analysis of radiolabeled C16 and C18 fatty acids in polar lipids of wild type and fatb- k0 leaves .................................................................................................................. 157 Mass distribution of C132C13 and ClozClg molecular species of polar lipids in wild type and fatb-ko leaves ............................................................................................ 160 Immunoblot analysis of wild type and fatb-ko leaves. ........................................... 164 DISCUSSION ............................................................................................................. 166 Increased rates of fatty acid biosynthesis and turnover in fatb-ko .......................... 168 Turnover of Polar Lipids in wild type and fatb-ko ................................................. 173 Conclusions and Future Perspectives ...................................................................... 174 MATERIALS AND METHODS ................................................................................ 175 LITERATURE CITED ............................................................................................... 181 CHAPTER 5 ................................................................................................................... 185 FINAL CONCLUSIONS, PERSPECTIVES AND FUTURE DIRECTIONS ............... 185 Conclusions and future directions on FAS gene expression in Arabidopsis .......... 185 Major conclusions of this thesis research related to F AS gene expression are: .. 185 Future directions .................................................................................................. 188 Conclusions and future directions on fatty acid metabolism and its regulation in fatb- k0. ............................................................................................................................ 192 Major conclusions of this thesis research related to the role of saturated fatty acid are: ........................................................................................................................... 192 Future directions .................................................................................................. 194 LITERATURE CITED ............................................................................................... 197 viii LIST OF FIGURES FIGURE 1. FATTY ACIDS EXPORTED FROM PLASTIDS HAVE DIVERSE ROLES IN PLANT CELLS. .......................................................................................................................... 3 FIGURE 2. SIMPLIFIED REPRESENTATION OF THE FATTY ACID BIOSYNTHESIS PATHWAY IN ARABIDOPSIS (ACCORDING TO OHLROGGE AND BROWSE, 1995). .............................. 16 FIGURE 3. SCHEMATIC REPRESENTATION OF THE ENZYMATIC REACTION CATA LYZED BY ACYL-ACP THIOESTERASES IN PLANTS. ..................................................................... 18 FIGURE 4. SIMPLIFIED SCHEME OF THE GLYCEROLIPID BIOSYNTHETIC PATHWAY IN ARABIDOPSIS (ADAPTED FROM BROWSE AND SOMERVILLE, 1991). ........................... 20 FIGURE 5. LIGHT-MEDIATED INDUCTION OF ACP4 mRNA LEVELS IN ARABIDOPSIS LEAF TISSUE. ....................................................................................................................... 39 FIGURE 6. LIGHT AFFECTS THE POLYRIBOSOME ASSOCIATION OF ACP MRNAS .............. 41 FIGURE 7. POLYSOME DISTRIBUTION OF ACP MRNAS IN WILD TYPE AND TRANSGENIC (MCTE) B. NAPUS DEVELOPING SEEDS ....................................................................... 43 FIGURE 8. CONSTRUCTS USED FOR ARABIDOPSIS TRANSFORMATION. ............................. 46 FIGURE 9. REPORTER GENE EXPRESSION IN DIFFERENT TISSUES OF ARABIDOPSIS TRANSGENIC PLANTS. ................................................................................................. 48 FIGURE 10. DIFFERENTIAL REGULATION OF ACP MRNA LEVELS IN STARVED ARABIDOPSIS CELLS ................................................................................................... 51 FIGURE 11. DIFFERENTIAL REGULATION OF ACP MRNA LEVELS IN ARABIDOPSIS CELLS GROWN IN THE PRESENCE OF SUCROSE. ...................................................................... 53 FIGURE 12. IN PLANTA REGULATION OF ACP2 MRNA BY CARBON .......................... 55 FIGURE 13. REGULATION OF ARABIDOPSIS ACP2 PROMOTER BY CARBON IN ARABIDOPSIS CELL SUSPENSION CULTURE ................................................................. 56 FIGURE 14. REGULATION OF ACP2 PROMOTER BY DIFFERENT CARBON SOURCES IN ARABIDOPSIS CELL SUSPENSION CULTURES ................................................. 58 FIGURE 15. ARABIDOPSIS ACP2 PROMOTER DELETION ANALYSIS IN TOBACCO BY-2 CELLS ............................................................................................ 61 FIGURE 16. CONSERVED ELEMENTS IN ARABIDOPSIS ACP PROMOTERS. ......................... 63 ix FIGURE 17. Inhibition of fatty acid synthesis by cerulenin in Arabidopsis cells .......... 68 FIGURE 18. SCHEMATIC REPRESENTATION OF THE ENZYMATIC REACTION CATALYZED BY ACYL-ACP THIOESTEASES (FAT). ............................................................................. 96 FIGURE 19. STRUCTURE OF THE ARABIDOPSIS FA TB GENE CARRYING THE T-DNA INSERTION. ............................................................................................................... 102 FIGURE 20. GROWTH AND MORPHOLOGY OF ARABIDOPSIS WILD-TYPE AND FA TB—KO PLANTS AND SEEDS ........................................................................... 105 FIGURE 21. GROWTH CURVES OF ARABIDOPSIS WILD-TYPE AND FA TB-KO PLANTS. ....... 107 FIGURE 22. MICROSCOPIC ANALYSIS OF LEAF SECTION AND CHLOROPLAST STRUCTURE OF WILD TYPE AND FA TB-KO SEEDLINGS ....................................................... 108 FIGURE 23. LEAF AND STEM CUTIN COMPOSITION OF WILD TYPE ARABIDOPSIS AND FA TB- KO(MOL%)... ....120 FIGURE 24. GROWTH AND MORPHOLOGY OF ARABIDOPSIS WILD-TYPE, FA TB-KO, ACT] AND FA TB-KO ACT] PLANTS. ..................................................................................... 122 FIGURE 25. SIMPLIFIED SCHEME OF PREDICTED FLUXES OF C16 AND C18 FATTY ACIDS IN MEMBRANE LEAF GLYCEROLIPIDS OF ARABIDOPSIS WILD TYPE, FA TB-KO AND FA TB-KO A('TI MUTANTS ......................................................................................................... 128 FIGURE 26. FATTY ACID TURNOVER IN WILD TYPE ( o ) AND FA TB-KO( o ) LEAVES. ...... 154 FIGURE 27. REDISTRIBUTION OF RADIOACTIVITY AMONG GLYCEROLIPID CLASSES OF WILD TYPE ( o ) AND FA TB-K0( o ) LEAVES ....................................................... 156 FIGURE 28. REDISTRIBUTION OF RADIOACTIVITY IN C”, FATTY ACIDS OF MEMBRANE GLYCEROLIPID CLASSES FROM WILD TYPE ( o ) AND FA TB-K0( o ) LEAVES. . . . . . . . 1 59 FIGURE 29. REDISTRIBUTION OF C18: ng AND szcjg MOLECULAR SPECIES OF POLAR LIPIDS IN WILD TYPE AND FA TB-KO ........................................................................... 162 FIGURE 30. SCHEME OF MASS LOSS FROM LABELED FATTY ACIDS IN LEAVES OF WILD TYPE AND FATB-K0..... .................................................................................................... 165 FIGURE 31. IMMUNOBLOT ANALYSIS OF BCCP, ACP AND 18:0-ACP DESATURASE IN LEAF TISSUE OF WILD TYPE AND FA TB—KO ARABIDOPSIS. .................................................. 167 FIGURE 32. COMPARATIVE SCHEME OF FATTY ACID METABOLISM IN WILD TYPE AND FA TB- K0 ............................................................................................................................ 170 LIST OF TABLES TABLE 1. FUNCTIONS OF LIPID MOLECULES IN HIGHER PLANTS ......................................... 2 TABLE 2. PROXIMAL UPSTREAM SEQUENCES OF ACP GENES IN DIFFERENT PLANT SPECIES ................................................................................................... 45 TABLE 3 . CHANGES IN mRNA ABUNDANCE OF GENES INVOLVED IN LIPID METABOLISM AFTER INHIBITION OF FATTY ACID SYNTHESIS IN ARABIDOPSIS CELLS. ...................... 70 TABLE 4. EXAMPLE OF TRANSCRIPTS THAT INCREASE MORE THAN 2 FOLD AFTER 6 HOURS OF CERULENIN TREATMENT. ....................................................................................... 71 TABLE 5. EXAMPLE OF TRANSCRIPTS THAT DECREASE MORE THAN 2-FOLD AFTER 6 HOURS OF CERULENIN TREATMENT. ....................................................................................... 72 TABLE 6 RELATIVE AMOUNTS OF FATB MRNA IN LEAF TISSUE OF WILD-TYPE ARABIDOPSIS AND FA TB-KO MUTANT ...................................................................... 103 TABLE 7. ARABIDOPSIS WILD-TYPE AND FA TB-KO BOLTING TIME AND GERMINATION RATES. ..................................................................................................................... 1 10 TABLE 8. FATTY ACID COMPOSITION OF WILD-TYPE AND FA TB-KO ARABIDOPSIS TISSUES (MOL%). .................................................................................................................. 1 1 1 TABLE 9. FATTY ACID COMPOSITION OF LEAF GLYCEROLIPIDS OF ARABIDOPSIS WILD- TYPE (WS) AND FA TB-KO MUTANT (MOL%). ............................................................ 113 TABLE 10. MAJOR COMPONENTS OF EPICUTICULAR LEAF WAXES FROFROM WILD-TYPE ARABIDOPSIS AND FA TB-KO MUTANT (MOL%). ....................................................... 116 TABLE 11. SPHINGOID BASE CONTENT OF LEAF TISSUE FROM WILD-TYPE ARABIDOPSIS AND FA TB-KO PLANTS (MOL%) ................................................................................. 118 TABLE 12. LEAF FATTY ACID COMPOSITION OF WILD TYPE ARABIDOPSIS (COLUMBIA), ACT] AND FA TB-KO A ("T1 DOUBLE MUTANT (MOL%) ................................................. 123 TABLE 13. RATES OF FATTY ACID SYNTHESIS (FAS) IN LEAVES OF ARABIDOPSIS WILD TYPE AND FA TB-KO MEASURED BY DIFFERENT RADIOLABELED SUBSTRATES ............ 152 xi CHAPTER 1 INTRODUCTION ”It takes a membrane to make sense out of disorder in biology. When the earth came alive, it began constructing its own membrane, for the general purpose of editing the sun. You have to be able to catch energy and hold it, storing precisely the needed amount and releasing it in measured shares. A cell does this, and so do the organelles inside. Each assemblage is poised in the flow of solar energy, tapping off energy from metabolic surrogates of the sun. To stay alive, you have to be able to hold out against equilibrium, maintain imbalance, bank against entropy, and you can only transact this business with membranes in our kind ofworld Lewis Thomas. Plant lipids are extremely diverse in structure and function (Table 1) and constitute the products of several distinct biosynthetic pathways. These molecules serve several essential functions in all organisms but their most critical role for life is the formation of cellular membranes. These membrane structures form the major barriers that delineate cells and their compartments and allow essential processes to occur (e.g., anabolic and catabolic reactions, transport by vesiculation and permeation, photon capture) (Rausch and Bucher, 2002; Bowsher and Tobin, 2001; Nebenfuhr and Staehelin, 2001). Among the many types of plant lipids, glycerolipids are usually the most abundant in plant cells (Ohlrogge and Browse, 1995). These lipids are the main constituent of cellular membranes and are derived from the glycerol and fatty acid biosynthetic pathways (Somerville et al., 2000). The fatty acid synthesis (FAS) pathway is a primary metabolic pathway, because it is found in every cell of the plant and is essential for growth. In addition to supplying fatty acids for the synthesis of membrane lipids, this pathway provides substrates for several other essential cellular processes in plants. For example, synthesis of sphingolipids, epicuticular waxes and cutin in epidermal cells and triacylglycerols in seeds. Moreover fatty acids are substrates for acylation reactions (e. g., proteins, sterols) and also precursors of signaling molecules (Figure 1). Table 1. Functions of lipid molecules in higher plants Function Lipid types Membrane components Glycerolipids, Sphingolipids, Sterols Storage compounds Triacylglycerols, Waxes Compounds active in electron transfer Chlorophylls, Ubiquinone, Plastoquinone Photoprotection and free radical protection Carotenoids, Tocopherols Surface protection Cutin, Suberin, Waxes, Triterpenes Protein modification 14:0, 16:0, Farnesyl and GeranylGeranyl Signaling pyrophosphate, Phosphatidylinositol, Ceramides, Dolichol Jasmonate, Diacylglycerol, ABA, GA, Brassinosteroids As a result of the dependence of multiple cellular processes on the production of fatty acids, the synthesis of these molecules is exquisitely controlled to balance supply and demand for acyl chains. For plant cells, this means matching the level of fatty acid synthesis to membrane biogenesis and to multiple cellular pathways depending on the Figure 1. Fatty acids exported from plastids have diverse roles in plant cells. In mesophyll cells fatty acids are used primarily for membrane glycerolipid synthesis. In epidermal cells the bulk of fatty acids goes into waxes and cutin. In seeds, triacylglycerols accumulate most of the exported fatty acids. In most cells saturated fatty acids are precursors of ceramides and substrates of acylation reactions. La) Aw_E._mEQ£ 303$ 04.... :55 . .. wan mEQ=owE=Qm =3 AIDE can BEEEQU .55 R .2255 wOXflg wEEBSoEm ’ acaLnEoE Tm _. Eon >33 353mm I 5395 92 -0‘ £85562} Eo< Baa”— /i one Eon >33 8538:: Emmi .8030 Figure l tissue, growth rate, stage of development and time of the day. For example, leaves increase by more than 10-fold the production of fatty acids during the day and some seeds accumulate 60 % of their weight in oil (Ohlrogge and Jaworski, 1997). Thus, a critical biochemical aspect of fatty acid biosynthesis in plants is to understand how cells adjust the production of fatty acids according to different demands and how they partition these molecules into several cellular pathways. Completion of the Arabidopsis genome sequence has accelerated the identification of structural enzymes involved in the synthesis of fatty acids and glycerolipids, and most of them have been either experimentally described or have strong candidates (Beisson et al., 2003; Mekhedov et al., 2000). By contrast, advances in the identification of regulatory factors and mechanisms of control for this metabolic pathway have been slower. Thus, a number of major question in the synthesis of fatty acids in plants remain unanswered: > How are fatty acid and lipid biosynthetic genes regulated? > Do global trans-acting factors exist that simultaneously control the expression of most genes? > What signals control the promoter activity of these genes? > How is the production of fatty acids controlled in plants? ‘9 What signals are sensed to adjust fatty acid synthesis rates to different cellular demands of fatty acids? > How is the production of fatty acids in the plastids coordinated with their utilization outside this organelle? \\ r Do signals communicate between the cytosol and the plastid fatty acid synthesis pathway? Objectives Based on these questions, the main objectives of the present work are to understand the signals that control the rate of fatty acid synthesis in plants, the expression of genes for this primary pathway and the production and partition of exported fatty acids from plastids. Understanding regulation of FAS genes in Arabidopsis TO begin to answer some of these important aspects of plant gene regulation and metabolism, Chapter 2 presents data on the signals and mechanisms that regulate the expression of Arabidopsis ACP isoforms. Why do we study ACP gene expression? First, these proteins are in the core of the fatty acids synthesis machinery and play a central role during acyl chain synthesis. Due to this pivotal role, it is likely that signals and mechanisms that affect ACP expression will also affect the expression of other fatty acid genes. Second, ACP mRNA and protein levels are among the most abundant of the fatty acid pathway, which in general are of low abundance. Third, ACP expression studies such as tissue specific expression and promoter analysis have been previously reported and provide partial but important information about regulatory signals and mechanisms (Baerson et al., 1994; Hlousek-Radojcic et al., 1992; Battey and Ohlrogge, 1990; Hannapel and Ohlrogge, 1988). Transcriptional regulation of genes for fatty acid synthesis Clues to the signals and mechanisms that activate plant FAS gene expression can be obtained from known signals and mechanisms in non-plant organisms, if some of these signals and mechanisms are conserved between distantly related living organisms. Considering the universal role of lipids in cell structure and regulatory processes it may be possible that fatty acid production and its genetic regulation Show, in some aspects, close resemblance in divergent organisms. For the purpose of reviewing how FAS genes are controlled in non-plant organisms, a short overview of transcriptional regulation of FAS genes in these organisms is given below. In addition, a brief summary of previous studies that analyzed different aspects of FAS gene expression in plants is presented. Transcriptional regulation of FAS genes in microorganisms The genes for fatty acid metabolism in E. coli are scattered about the genome with only two clusters, the minimal accBC and the fab cluster (Cronan and Rock, 1996). All genes are under regulation of F adR, a protein with dual functions that represses genes involved in fatty acid degradation and activates genes involved in fatty acid synthesis. In the absence of long chain acyl-COA, FadR directly binds to specific DNA sequences to simultaneously activate transcription of biosynthetic genes and repress catabolic genes. Thus, during exponential growth or in the absence of exogenous fatty acids, the pool of acyl-COA remains very small and FadR activates biosynthetic genes and Shuts off catabolic gene expression. When growth slows down or exogenous fatty acids are abundant, the acyl-COA levels build up and FadR de-activates expression of FAS genes and de-represses catabolic genes. Thus, in these organisms, FAS genes are almost exclusively controlled by growth or exogenous fatty acids and respond to the demand for new membranes (Cronan and Rock, 1996). In B. subtilis the genes involved in fatty acid metabolism are located mainly in gene clusters (e. g., the fabHAF operon) and are coordinately regulated by FapR. This regulator is a transcription factor that represses the expression of lipid metabolic genes by binding to a consensus sequence contained in their promoter regions (Schujman et al., 2003). Thus, expression of fatty acid gene clusters is de-repressed when cells are in the exponential phase of growth. Interestingly enough, addition of inhibitors of fatty acid biosynthesis (e.g., cerulenin) can also de-repress these genes, suggesting that feedback mechanisms of gene expression exist in this organism. During inhibition of FAS by cerulenin, the intracellular concentration of malonyl-COA is increased (Heath and Rock, 1995) and it is proposed that intracellular levels of malonyl-COA regulates FapR activity. Thus, if the rate of fatty acid synthesis falls below the normal levels (as a result of slow growth or inhibition), a transient increase in the intracellular concentration of malonyl- COA relieves FapR mediated repression of lipid biosynthetic genes (Schujman et al., 2003). In yeast, the synthesis of fatty acids is catalyzed by two multifunctional enzymes, acetyl COA carboxylase (ACC) and fatty acid synthase (FAS). ACC is a tetramer of identical subunits encoded by the FAS3 gene (Schuller et al., 1992). FAS consists of two multifunctional proteins. a and [3, which are encoded by two unlinked genes, FAS] and FAS2, respectively (Chirala, 1992). FAS], FASZ and FAS3 genes are coordinately regulated by growth and exogenous fatty acids. Moreover, when FASZ is over-expressed, expression of both FASl and FAS3 is also increased, suggesting that the FAS genes are coordinately regulated in yeast. The conserved GCCAAA element (UASfas) is present in all three FAS promoters and Specifically enhances the transcription of FAS genes. In addition, FAS] and FASZ share an UAS.no element that is common to genes involved in phospholipid biosynthesis. UAS,-no is a positive regulator of gene expression and is required for efficient expression of these genes. UASfas and UASmo act synergistically for Optimal expression of the FAS genes (Chirala, 1992; Schuller et al., 1992). In summary, microorganisms possess global trans-acting factors that bind to consensus sequences conserved in most of their FAS gene promoters and coordinately control the expression of most of these genes. Transcriptional regulation of FAS genes in animals Animals obtain fatty acids almost exclusively from their diet, and they have developed mechanisms that sense lipidic molecules such as cholesterol, fatty acids, fat-soluble vitamins and other lipids present in their food. These molecules or their derivatives bind to receptors in the nucleus or cytoplasm that directly or indirectly activate gene expression (Chawla et al., 2001 ). For example, one important group of receptors is the PPAR family (peroxisome proliferator-activated receptor), that is activated by polyunsaturated fatty acids and eicosanoids. PPARS belong to the nuclear receptor superfamily of transcription factors that contain a large COOH-terminal region with a ligand binding domain. Upon ligand binding, these transcription factors form heterodimers with the retinoid X receptor (RXR) and bind to specific DNA sequences found in their target promoters known as hormone response elements (HRE) (Chawla et al., 2001). Among their functions, PPARS can up- regulate proteins that increase metabolism and transport of fatty acids into the peroxisome and promote fat storage in the liver (Chawla et al., 2001). A second group of regulatory factors of fatty acid metabolism in animals is the sterol regulatory element binding protein (SREBP) family. These transcription factors are located on ER membranes in an inactive form, and are released by proteolysis in order to enter the nucleus. Proteolysis of SREBPS is inhibited by sterols and polyunsaturated fatty acids in mammals, providing a feedback regulatory mechanism for lipid synthesis (Dobrosotskaya et al., 2002). SREBPS interact with an escort protein, SCAP (SREBP cleavage-activating protein) that serves as sensors of sterols, fatty acids or phospholipids in animal cells (Seegmiller et al., 2002). Transcriptional regulation of FAS genes in plants 10 Similarly to other higher eukaryotes, multiple cis elements and trans factors are expected to be involved in regulation of plant fatty acid gene expression. However, in contrast to animals, which obtain fatty acids mainly from their diets, plants depend on their own production of fatty acids to subsist. In addition, the plant fatty acid machinery resembles that from bacteria and is plastid localized, although almost entirely nuclear encoded. Thus, regulatory mechanisms of fatty acid gene expression that communicate between plastids and nucleus are probably present in plant cells (Jarvis, 2001). The essential differences in the structure of fatty acid synthesis between animals and plants are most likely reflected in the mechanisms and Signals by which fatty acid and lipid genes are regulated in these organisms. For example, no obvious homologues of animal PPARS and SREBPS are evident in the Arabidopsis genome. One group of plant fatty acid genes that has received special attention is the ACP gene group. Baerson and Lamppa (1993) fused an approximate 900 bp fragment upstream of the Arabidopsis ACP2 gene to B-glucuronidase (GUS) and used the construct to generate transgenic tobacco plants. GUS expression was detected during seed development, in young leaves, in leaf epidermis, and flowers. In a second study, the same authors analyzed a series of six deletions of the same promoter in tobacco transgenic plants (Baerson et al., 1994). The results revealed distinct regions of the promoter involved in vegetative and reproductive development. A —320 to —236 bp region was important for expression of the reporter gene in leaves, whereas it did not alter expression in seeds and flowers. Seed expression was reduced when a —235 to ~55 bp region was deleted. This 11 region was also essential for expression in flowers. Thus, transcriptional regulation of the Arabidopsis ACP2 gene in different tissues involves different regions of the promoter and presumably tissue specific factors. Similarly, a deletion analysis of the Arabidopsis enoyl-ACP reductase promoter showed that three domains of the promoter were important for differential tissue expression. First, seed expression was unchanged by deletion to —47 bp of the transcription start Site, suggesting that seed Specific cis elements are located close to the transcription starting site. Second, removal of an intron in the 5’UTR resulted in increased expression in roots, suggesting the presence of negative regulatory elements in this region. A third region was important for high expression in young leaf tissue. Other studies reported the light-dependent expression of FAD7 (plastidial desaturase) (Nishiuchi et al., 1995), cold-dependent expression of FAD8 (also a plastidial desaturase) (Vijayan and Browse, 2002), and low-phosphate induction of SQDI and SQD2 (Yu et al., 2002; Essigmann et al., 1998). Thus, a subset of genes involved in lipid metabolisms respond to specific Signals depending on their distinct functions. In addition to specific environmental signals, the fatty acid and lipid genes that have been analyzed so far Show a common denominator, high levels of expression in rapidly expanding tissue where cell division is active (e.g., apical meristems), and also in developing seeds and flowers, where lipid accumulation occurs. Summary of Chapter 2 l2 In Chapter 2, we tested signals known to activate fatty acid and lipid gene expression in non-plant organisms for their ability to activate fatty acid genes in plant cells (e.g., growth-dependent expression, inhibition of fatty acid synthesis, sugar-sensing mechanisms). Results suggested that some of these signals but not all differentially increase expression of some ACP isoforms. Thus, it appears that some regulatory Signals have been conserved throughout evolution between plants, animals and lower organisms. For example, similarly to bacteria and yeast, ACP2 expression appears to respond to growth/cell division Signals, whereas inhibition of fatty acid synthesis has no effect on gene expression linked to fatty acid metabolism in plants. Analysis of the Arabidopsis ACP2 promoter demonstrated that activation by growth signals acts at the transcriptional level and identified promoter regions important for regulation and expression of ACP2. Light, which stimulates growth and chloroplast development has an important role on ACP gene expression at the levels of ACP4 mRNA abundance and association of ACP transcripts with polysomes. We also demonstrated that the 5’UTR of ACP transcripts increases gene expression and determines tissue specific expression. In summary, Chapter 2 discloses multiple levels of control on ACP gene expression and provides the basis for further exploration of gene regulation for fatty acid metabolism and also identification of regulatory factors for these genes. Understanding regulation of fatty acid metabolism and FAS gene expression in a fatb mutant of Arabidopsis 13 Chapter 3 takes a different approach and centers on the characterization of an Arabidopsis FA TB knock-out mutant (fatb—k0). FATB belongs to the acyl-ACP thioesterase family of proteins and alteration of its function by either mutation or over- expression has proved to be instructive in disclosing mechanisms of regulation of fatty acid metabolism (Eccleston and Ohlrogge, 1998; Ohlrogge et al., 1995). These observations together with our interest in understanding production of saturated fatty acids in plants and their role in development, prompted us to characterize the fatb-k0 mutant. Thus, Chapter 3 describes the central role of FATB in producing saturated fatty acids and the essential role of these molecules in plant growth and development. In addition, analysis of saturated fatty acid derivatives provides evidence for the involvement of F ATB in supplying fatty acids to distinct biosynthetic pathways and how cells control the economy of saturated fatty acids. These results provide clues to understand the effects of reduced saturated fatty acids on plant growth and set the basis for the experiments describe in Chapter 4. Chapter 4 illustrates the importance of altering acyl-ACP thioesterase expression in disclosing regulatory signals and mechanisms for regulation of fatty acid metabolism. fatb-k0 leaves respond to low levels of saturated fatty acids by increasing the rate of fatty acid synthesis by 40 % compared to wild type. Moreover, western blot analysis revealed that BCCP, ACPS and 18:0-ACP desaturase levels are increased by 1.5- to 2-fold in mutant leaves compared to wild type. Thus, higher rates of fatty acid synthesis are achieved at least in part by increasing fatty acid gene expression. In summary, plant cells appear to have mechanisms that sense and respond to subnormal levels of saturated fatty l4 acids in the cytoplasm. These mechanisms probably coordinate the requirements for lipid synthesis in the cytosol and fatty acid production inside the plastids. Moreover, the same mechanisms may also activate expression of nuclear fatty acid genes. With the purpose of introducing the role of acyl-ACP thioesterases in fatty acid synthesis in plants together with the role of fatty acids in the synthesis of different cellular components and the current knowledge in regulation of fatty acid synthesis, a brief overview ofthese topics is given below. Fatty acid synthesis in plants Plants are fundamentally different from other eukaryotes in the molecular organization of the enzymes of fatty acids biosynthesis (FAS). The individual enzymes of this primary pathway are dissociable soluble components located in the stroma of the plastids whereas in other eukaryotes fatty acid synthesis is catalyzed by multifunctional polypeptide complexes located in the cytosol (Somerville et al., 2000; Schuller et al., 1992; Chirala, 1992). The central carbon donor for fatty acid biosynthesis is the malonyl-COA produced by acetyl-COA carboxylase (ACCase). In dicot plants, plastidic ACCase is a mutisubunit enzyme composed of four dissociated polypeptides (biotin carboxyl carrier protein [BCCP], biotin carboxylase [BC], and a- and B-carboxyl transferases (CT). BCCP, BC and d-CT subunits are nuclear encoded whereas B-CT is plastome encoded (Sasaki et al., 1993). Before entering the fatty acid biosynthetic pathway, malonyl-COA is transferred 15 from coenzyme-A (COA) to an acyl carrier protein (ACP) by malonyl-CoAzACP transacylase. From this point on, all the reactions of the pathway involve ACP (Figure 2). After being transferred to ACP, the malonyl group enters a series of reactions that result 9° co cu,—c\ 2 w“ G) Acetyl-CoA acetyl-COA ' Carboxylase ADP. Pl @ o“c- ea -—c”0 (.__)® o“c-— ea —c"0 Condensation '0’ 2 ‘S‘ACP '0’ ’ ‘s-COA co,7 MalonyI-ACP Malonyl-CoA o ll CHJ—c- cH,-c —S-ACP l1 © 0 ,— o 3-ketobutyryl-AC P co Condensation 01‘3" c“,— c»;- £_S.Acp 3 @ NAOPH . II‘ o Butyryl-AC P NADP' _ _ _ " _ . a <5 cu, c s ACP @ Reduction 0 of 3-keto 3-ketoacyl-AC P mow ‘ Reduction group 070’9 M0,,“ , H. ' of double 9 continues bond cu3—c- cu,-c —$-ACP on o 3-hydroxybutyryl-ACP org-cu- CH- 3_ s. 1‘ c P Trans-Z-butenoyl-ACP n,o Dehydration Figure 2. Simplified representation of the fatty acid biosynthesis pathway in Arabidopsis (according to Ohlrogge and Browse, 1995). 1, Acetyl-COA carboxylase; 2, Malonyl-CoAzACP transacylase; 3, 3-Ketoacyl-ACP synthase (KAS); 4, 3-Ketoacyl-ACP reductase; 5, 3-Hydroxyacyl-ACP dehydratase; 6, 2,3-trans-enoyl-ACP reductase. l6 in the formation of a carbon-carbon bond and in the release of C02. These reactions are catalyzed by plastidial fatty acid synthase, which in plants is also a multisubunit enzyme composed of 3-ketoacyl-ACP synthases I, II and III, 3-ketoacyl-ACP reductase, 3- ketoacyl-ACP dehydrase, and enoyl-ACP reductase (Figure 2) (Somerville et al., 2000). All of fatty acid synthase subunits are nuclear encoded. The FAS pathway produces saturated fatty acids, however more than 75 % of plant fatty acids are unsaturated. The first double bond is introduced by a soluble 18:0-ACP desaturase in the plastid. Additional double bonds in fatty acids are incorporated by membrane-bound desaturases that act once the fatty acid has been incorporated into glycerol (Browse and Somerville, 1991). The elongation of fatty acids is terminated when the acyl group is removed from ACP. This reaction can be catalyzed by acyl-ACP thioesterases that hydrolyze the acyl-ACPS to release free fatty acids and ACPS or alternatively, the acyl group can be transferred to g1ycerol-3-phosphate or monoacylglycerol-3-phosphate by the action of acyl-ACP acyltransferases (Figure 3) (Somerville et al., 2000). More than half of the fatty acids produced in the plastids flow through acyl-ACP thioesterases and are exported from this organelle in the form of acyl-COAS. These molecules are later used for glycerolipid synthesis at the level of the endoplasmic reticulum (ER). Based on its Similarity to the animal pathway for glycerolipid synthesis it is named Eukaryotic Pathway (Figure 4). The fraction of fatty acids that remains in the plastid is used for glycerolipid synthesis within the plastid and based on its similarity to the bacterial pathway it is named the Prokaryotic Pathway (Figure 4). Some of the lipids synthesized by the Eukaryotic l7 Figure 3. Schematic representation of the enzymatic reaction catalyzed by acyl- ACP thioesterases in plants. Acyl-ACP thioesterases or FAT (fatty acid thioesterases) enzymes catalyze the hydrolysis of the thioester bond between acyl carrier proteins (ACP) and fatty acids to release free ACPS and free fatty acids inside the plastids. Fatty acid are then exported from the plastid as acyl-COA molecules. 18 .825 gamma 9 / e noses at 4.8.924. 98 “Gage m . a m «606.24. 92 m9??? .m a . E m 4.8.3:: E: most: m / a, El 38.285 ao<._e< 19 Figure 3 Figure 4. Simplified scheme of the glycerolipid biosynthetic pathway in Arabidopsis (adapted from Browse and Somerville, 1991). Fatty acid synthesis (FAS) takes place inside the plastid and its acyl-ACP products (16:0-ACP, 18:0-ACP and 18:1-ACP) are substrates of either acyl-ACP thioesterases (FAT) or acyl-ACP acyltranferases (ACT). By the action of the latter, acyl groups enter the Prokaryotic Pathway of lipid synthesis inside the plastids to produce MGDG, DGDG, PG and SL. By the action of acyl-ACP thioesterases, acyl groups leave the plastid as acyl-COA molecules and become substrates of the Eukaryotic pathway of lipid synthesis in the endoplasmic reticulum (ER). Fatty acids can return to the plastid in the DAG moiety of PC and are used for the synthesis of MGDG, DGDG and SL within the plastid. Abbreviations: PC, phosphatidylcholine; PE, phosphatidylethanolamine; MGDG, monogalactosyldiacylglycerol; DGDG, digalactosyldiacylglycerol; PG, phosphatidylglycerol; PI, phosphatidylinositol; PS, phosphatidylserine; SL, sulfolipid; DAG, diacylglycerol; PA, phosphatidic acid; LPA, lysophosphatidic acid. 20 c9353.— Ototmnao—onA—v oflwejd .05: ”5.. one" 05— nuow nump CHOP Hump Ono, Hump Hump num— HI. »» II» A _ 86: «no. as: «a. «no. «a. «”2 [IL It‘ll-ll. / 86: . NHG— «Amp IrIIL O2 2 . ‘3 1 ACP2+3 <1) i 0124812 time (hours) Figure 5. Light-mediated induction of ACP4 mRNA levels in Arabidopsis leaf tissue. (A) Northern blot analyses of two week-old Arabidopsis plants dark treated for 24 hours and reilluminated with fluorescent white light for the times indicated. Each lane of the blot contains 5 ug of total RNA from leaf tissue. The eIF 4A probe was used as a loading control. (B) Densitometry scanning of the blots in A. The signal intensities at each time were normalized with the corresponding eIF4A Signal and expressed relative to time 0 (set arbitrarily to one). (Caspar et al., 1993). In contrast to ACP4 transcript levels, the signal presented by a probe corresponding to the ACP2 coding sequence did not Show intensity variations in the conditions tested (Figure 5). This result agrees with a previous study that Showed that the Arabidopsis Acl].2 gene promoter (ACP2) does not confer light responsiveness to a reporter gene in transgenic 39 tobacco seedlings (Baerson et al., 1993). Because the ACP2 and ACP3 isoforms are 82% Similar at the nucleotide level, we assumed that the signal conferred by the ACP2 fragment represented the abundance of both, ACP2 and ACP3 transcripts (ACP2+3 in Figure 5). This assumption is also based on the observation that both proteins are expressed in similar amounts in Arabidopsis leaf tissue (Hlousek-Radojcic et al., 1992) and that the corresponding promoters confer similar levels of expression to a reporter gene in Arabidopsis leaves (Baerson et al., 1998). Polyribosomal Association of ACP mRNAs is Increased by Light Previous studies suggested that the expression of ACPS could be regulated at the level of translation. First, transgenic Brassica napus plants over-expressing a 12:0-ACP thioesterase Show an increase of approximately 2-fold in the ACP protein level with no Significant changes in the corresponding messenger abundance (Eccleston et al., 1998). Second, Hannapel et al (1988) reported that during seed development in soybean, the relative abundance of ACP and lectin mRNAs is at most 28-fold different whereas the corresponding protein levels differ at least ZOO-fold, suggesting a differential translational efficiency between these two mRNAs. Based on these observations, we compared the polyribosomal distribution of ACP mRNAs with other transcripts highly expressed in leaf tissue of Arabidopsis plants. In addition, the fact that Ii ght has a major effect on the polyribosomal association of several messengers RNA encoding plastidic proteins (Berry et al., 1990; Dickey et al., 1998) 40 prompted us to investigate the influence of light on the association of ACP mRNAs with ribosomes. Therefore, we analyzed Arabidopsis leaf tissue from plants either dark treated for 24 hours or reilluminated for 12 hours with Light Dark 0/o suc .fi-x-J‘f’m /’//’—’l ACP2+3 a I 9 u I I - FEDA ' jggfifi§$§“ .5 %_I_ i “W e. H j t?!!""”” Figure 6. Light affects the polyribosome association of ACP mRNAs. Leaf tissue extracts from plants grown in the dark for 24 h (Dark) or reilluminated for 12 h after a 12 h dark period (Light) were Spun in 15 to 60 % sucrose gradients. Sixty aliquots were collected at identical volume intervals and absorbance at 260 nm measured. Total RNA was extracted from 15 fractions of the gradients, loaded onto an agarose gel and analyzed by Northern blot. The blots were hybridized with probes corresponding to the sequences indicated. Fractions 7 to 15 correspond to the polyribosomal fractions (black line). white fluorescent light. The results demonstrated that the messengers for ACP2+3 and ACP4 were associated with polyribosomes (fractions 10 to 15 of the gradient) after 12 hours of illumination (Figure 6). In contrast, after 24 hours in the dark, the transcripts for the ACP isoforms appeared in the upper fractions (3 to 6) of the gradient, corresponding to low molecular weight polyribosomes or ribosome-free mRNAs (Figure 6). The polyribosomal distribution of the ferredoxin A (FEDA) mRNA was Similar to that of ACP mRNAs in both conditions (Figure 6). The eukaryotic initiation factor 4A (eIF4A) 41 transcript also showed association with polyribosomes in the light, but in contrast to ACP and FEDA mRNAs, a significant proportion of the eIF4A transcript was still bound to polyribosomes after 24 hours in the dark (Figure 6). In summary, these observations indicate that the transcripts encoding for ACP isoforms are associated with polyribosomes in Arabidopsis leaf tissue in a light-dependent manner similar to FEDA but differ Si gnificantly with respect to eIF4A mRNAs in the dark. Polyribosomal association of ACP mRNAs in developing seeds of wild type B.napus and transgenics over-expressing MCTE To evaluate if translation of ACP mRNAs was altered in developing seeds of transgenic Brassica napus over-expressing a medium-chain thioesterase (MCTE) compared to wild type, the polyribosomal distribution of ACP mRNAs was analyzed in seeds of both plant classes. Flowers from wild type plants and transgenic B. napus (event 198, Voelker et al., 1992) were tagged and developing seeds harvested at 21, 28 and 35 days after flowering. These time points correspond to mid-stage development of B.napus seeds, the period of active Oil biosynthesis and accumulation. The distribution profile of ACP transcripts on polysomes was Similar between B. napus wild type and transgenic seeds during the three developmental stages analyzed (Figure 7). These results suggested that translation of ACP mRNAs in seeds was not affected by over-expression of MCTE and mechanisms affecting protein stability were more likely explanations for higher FAS protein levels in transgenic seeds (Eccleston and Ohlrogge, 42 1998). In addition, association of ACP transcripts with high molecular weight polysomes indicated that ACP mRNAs were efficiently translated in seeds. MCTE Wild type Polysomes Polysomes o/o suc A///l 28DAF ,_ abun- 21w — Full.“ noon-a. Figure 7. Polysome distribution of ACP mRNAs in wild type and transgenic (MCTE) B. napus developing seeds. Developing seeds from wild type B. napus and transgenic overexpressing MCTE were harvested at 21, 28 and 31 days after flowering (DAF). Whole seeds were lised and extracts loaded onto 15-60 % sucrose gradients. Fractions containing polysomes were collected from the gradients after centrifugation and total RNA extracted and resolved by electrophoresis. After transfer, the blots were probed with a seed Specific ACP cDNA (Bn-ACP 28f10). Conserved Motifs occur in the 5’ Leader Region of ACP mRNAs The examination of Arabidopsis genomic and cDNA ACP sequences has revealed two unusual motifs in the 5’ leader region of ACP mRNAs (Table 2 and Ohlrogge et al., 1991). One feature is the presence of short sequences rich in pyrimidines. The high cytosine and thymidine content is uncommon, and leader sequences of plant mRNAs tend 43 to be rich in adenines and thymidines (Joshi, 1987). A second conserved motif is an element composed of seven nucleotides CTCCGCC (Table 2). In addition to Arabidopsis sequences, these features are also conserved in the 5’ leader region of ACP messengers from several diverse species (Table 2). The importance of 5’untranslated regions (UTRS) as essential regulators of gene expression in plants has been described (Dickey et al., 1998; Bolle et al., 1996). To analyze the participation of the 5’ leader sequences of the ACP mRNAs in the regulation of ACP expression, we generated independent Arabidopsis transgenic lines in which the 5’ UTR sequences of the Arabidopsis ACPl (lines atL-acpl) and ACP2 (lines atL—acp2) mRNAs were fused to the luciferase reporter gene (LUC) under the control of the CaMV35S promoter (Figure 8). To preserve the endogenous translation initiation site a short portion of the corresponding ACP coding sequence was also included (see Figure 8 for details). The 5’ leader regions of ACPl and ACP2 mRNAs present sequence variations but both have conserved CU rich regions and include the heptanucleotide motif CTCCGCC (Table 2). However, ACPI and ACP2 transcripts contain distinct translation initiation sites (Figure 8). The AUG context in the Acl].1 gene (ACPl) matches with one of the most common translation initiation sequences found in plants, AAACAAU_GGC whereas the translation initiation site in the Ac11.2 gene (ACP2), CUUCUAUQGC is found in a smaller number of plant genes (Joshi et al., 1997). This suggests that the efficiency of AUG recognition by the translation machinery might also influence ACP expression. A third line of transgenic plants carrying the LUC gene fused to a deleted version of the ACPl leader region (line atL-dell) was also generated. In this case, most 44 Table 2. Proximal upstream sequences of ACP genes in different plant species. A. thaliana ACPl: pt_c_tt_tgtacaCTCCGCCctctctccccatctctttcgacagatctcttctctctctc gtgtttcacgaaaca atg Athaliana ACP2 : mmCTCCGCCmatgtgaflgatgtggacgattcam atg Athaliana ACP3: actgtttctcatctcttcgt_chTCCGCCMQaamachgagctcitacgattcattcgttct atg Athaliana ACP4 : ccgaagataggcctgaatctccgagaacaaacCTCCGCCacaaaaacagaagacttgttgcttgcgtatcataat cgacgccgtatctctacttgctcgaacaaacccaaaaagatacatatcaagagattaaaccttatccaactaagagaagccatttttattttttttgggtctctgag ttgtgtattgagcttcatctccttcaa atg Athaliana ACPS: acaaaatagtaattcacggtc_cttgaacaATCCGCCaMgatcagatcgata atg B. campestris ACP-SQ: atcacgfltfigtacaCTCCGCCatctctctctctctcgagcagatctctctcgggaatatcgaca atg B. campestris ACP-sfl: caagctaccatgggacatcachtacaCTCCGCCatctctctccathtctctcgtgaataacgaaa atg B. napus ACP: atcacggmgtacaCTCCGCCatctctctctccttcgagcacagatctctctcgtgaatatcgaca atg B. napus ACP-28f10: acgctctgtacaCTCCGCCatctctctccattctctctcgtgagtaacgaca atg B. rapa ACP: gacatcacgflgtacaCTCCGCCatctctctctctctcgagcagatctctctcgggaatatcgaca atg C. glauca ACP: gcggccgctngTCCGTthatttcmgmccctctctcaaaflagatctctctctctcgctttgtatct atg Csativum ACP: tatcgtcacactctttgtgctCTCCGCCgtgtmgatcaataaacttttctcagatctaaactctatctatcaa atg C. lanceolata AC Pl-3: cCTCCGTCgtflaMagctaccaa atg C. lanceolata ACP l -l: cacggc_tCTCTCGCthatttgctcgctccctccctccctcccccatca atg Spinach ACPII: atctctctcctctttctctCTCCGCCacattcatt_ctga%actttctctcccctctctctccgcttcttca atg Spinach ACPI : aacttaataflactcaggataagcttctcactctctctctctctctctcttactacc atg Barley ACP2: ccgccgcCTCCACCgccgcCTCCACCgcccaccgcgccggcctctccccctgtcccgtctccccc atg Barley ACPl: aggaccagccggcctcttcccaccgcccccaaattctaccgagcagCTCAGCngccaaccc atg A sequence of arbitrary length upstream the of translation initiation site was selected. The sequence (CTCCGCC) and its derivatives are indicated in uppercase. The OT rich sequences are underlined. The ATG start codon is separated by one space at the right end of the sequences. A.= Arabidopsis, B.= Brassica, C. = Coriandrum 45 A) \ I ‘I (L) NPTH +1 \\\\\\\\1 GUS (R) m} Bl B) ///’ “4L- pCEtACPI IacctcIzaactctttgtacactccgccctctctccccatctctttcgacagatctcttctctctctcgtgtttcacgaaacaatggcgactcaattcagcgccatg pCaAC P22 acctcgagactgtttctctatctctttgtcttctccgcctcctccgatctcactccgatctctctacgattcattcttctatggcttccattgccatg pCadell I acctcgagctctttgtacaaacaatggcgactcaattcagcgccatg Figure 8. Constructs used for Arabidopsis transformation. (A) Scheme of the T-DNA constructs carrying the LUC and GUS genes. Black box: T- DNA borders (L) left and (R) right ends, Open box: nopaline synthase 3’ polyadenylation Signal, NPTII: neomycin phosphotransferase coding sequence, black triangle: CaMV35S promoter, LUC: firefly luciferase coding sequence, GUS: B-glucuronidase coding sequence, B: BamHI, P: PstI, Bl: BglII, +1 denotes the transcription initiation Site. (B) Sequences of the 5’ UTRS derived from Arabidopsis ACPl (pCaACPl), ACP2 (pCaACP2) and the deleted version of ACP] (pCadell) used in the three independent constructs. The ACP translation initiation codon is underlined. The displayed sequences start at the predicted transcription initiation site and therefore the leaders generated by these constructs have 8 additional bases from the vector sequence (underlined). of the 5’ UTR was removed, conserving only the first nucleotides towards the 5’ end and the ACPI translation initiation Site (see Figure 8 for sequence details). As an internal control for position effect and copy number, all the T-DNA constructs also carried a cassette expressing the B-glucuronidase (GUS) gene under the CaMV35S promoter (Figure 8). The 5’ Leader Sequences of ACP] and ACP2 Increase Reporter Gene Expression 46 We analyzed the LUC and GUS specific activities in expanding leaves of two week-old transgenic plants. The atL-acpl and atL-acp2 transgenic lines Showed 10- and 20-fold higher LUC/GUS ratios respectively compared to the atL-dell lines after 24 hours in the dark (Figure 9A). From these results we conclude that the presence of the ACP 5’ UTRS is essential for high expression of ACP] and ACP2 in Arabidopsis leaves. Based on the polyribosomal distribution of the ACP transcripts (Figure 6) we also asked whether the ACP 5’ UTRS could have a light regulatory role similar to other UTRS from plastid proteins (Dickey et al., 1998; Bolle et al., 1994). After 6 hours of re-illumination, the activity ratio of the reporter genes in the atL-acpl and atL-acp2 lines differed approximately 20- and 30-fold with respect to atL-dell lines (Figure 9A). The additional increase upon reillumination in the LUC/GUS ratios suggests that light-enhanced expression of the ACP-LUC constructs is at least partly mediated by the ACP 5’UTR. Arabidopsis developing seeds are green and posses photosynthetic capacity. However, they are considered heterotrophic as imported sucrose is their major carbon source. As shown in Figure 9B, the LUC/GUS ratios of specific activities in mid-stage developing seeds differed from the ratios observed in leaves. The activity ratio in atL-acpl transgenic lines was approximately 2.5-fold higher than the ratio in atL-acp2 plants and l3-fold higher than the atL-dell transgenic lines (Figure 9B). Thus, in contrast to leaf tissue, the 5’ UTR of ACP] mRNA appeared to confer a preferential expression of the LUC gene in 47 A) Leaves B) Seeds C) Roots LUC/GUS Ratio I atL-dell Cl atL-acpl I atL-acp2 Figure 9. Reporter gene expression in different tissues of Arabidopsis transgenic plants. atL-dell. atL-acpl and atL-acp2 indicate the transgenic lines transformed with pCadell, pCaAC P1 and pCaAC P2 respectively (see Figure 8). (A) Transgenic plants were dark treated for 24 h (Dark) or reilluminated for 6h (Light) with white fluorescent light. LUC and GUS specific activities were measured in leaves of 10 independent transgenic lines for atL-dell, atL-acpl and atL-acp2. LUC/GUS ratios are the average of the 10 individual ratios from each transgenic line. The atL-acpl and atL-acp2 LUC/GUS ratios are expressed with respect to atL-dell LUC/GUS ratio (set arbitrary to one). The bars denote the standard deviation of the average. (B) LUC and GUS Specific activities were measured in roots Of 10 independent lines for atL-dell, atL-acpl and atL-acp2. LUC/GUS ratios were calculated and represented as in A. (C) Reporter gene activity was measured in pools of developing seeds from 10 independent lines for atL-dell, atL-acpl and atL-acp2. The LUC/ GUS ratios were calculated and represented as in A. developing seeds compared to the 5’UTR of ACP2 mRNA. It is noteworthy that the level of reporter gene expression in atL-acpl lines relative to control lines was similar in leaves and seeds (between 10- and 20-fold) (Figure 9A and B). 48 The expression of the reporter genes was also evaluated in root tissue from 2 week-old transgenic plants. In contrast to leaf and seed tissue, the influence of the AC P1 and ACP2 5’UTRS was less pronounced and for atL-acpl plants the ratio of reporter gene activity was Similar to the ratio in atL-dell plants (Figure 9C). This result suggests that the 5’ leader of the ACPI mRNA does not affect the expression of the LUC mRNA in roots. For atL-acp2 plants, the LUC/GUS ratio was 4-fold higher than the ratio in atL-dell plants (Figure 9C). Similar to leaf tissue, the influence on LUC expression of the ACP2 leader was higher than the effect produced by the ACPI leader on the same reporter gene. However, the relative ratios of reporter gene activity conferred by the ACP2 leader was between 7 and lO-fold lower in roots compared to leaves (Figure 9A and C). ACP mRNA Levels are Affected by a Sucrose-derived Signal and/or Growth Control in Cell Suspension Cultures In bacteria and yeast the expression of several genes involved in fatty acid and lipid synthesis is tightly coupled to growth (Jiang et al., 1994; Carman et al., 1999). Cells growing in the presence of a carbon source such as sucrose Show high rates of transcription of these genes (Carman et al., 1999). In addition to their role as energy sources, sugars have been demonstrated to control the expression of plant genes involved in diverse processes such as starch metabolism (Nakamura et a1, 1991), storage protein accumulation (Hattori et a1, 1990) and lipid degradation (Graham et al, 1994). Based on these observations, we asked whether the expression of some of the fatty acid synthesis genes in plants might also be under metabolic and/or growth control by sugars. For this purpose we analyzed the mRNA levels of ACP2, ACP3, ACP4 and the endoplasmic 49 reticulum (ER) associated delta-12-desaturase (F AD2) in Arabidopsis mesophyll-derived cell suspension cultures. In order to distinguish between the messengers corresponding to ACP2 and ACP3 in this experiment, we synthesized a messenger specific probe based on the 3’ UTR sequences of these genes. The transcript levels of ACP2 and ACP3 isoforms declined approximately 2-fold after 48 hours of starvation (Figure 10). In contrast, the ACP4 and FAD2 transcript levels showed no significant variation during this 48 hours period (Figure 10). Thus, the absence of sucrose in the media and/or the reduced growth rate preferentially affected the expression of the ACP2 and ACP3 mRNAs. In the same experiment the transcript levels of the ribulose-1,5- biphosphate-carboxylase 1A small subunit (RBCSlA) increased approximately 2-fold after 12 hours of starvation and slightly declined afterwards (Figure 10). This result was in agreement with the fact that the gene for the small subunit of Rubisco is activated by light and repressed by sugars (Terzaghi et al., 1995). To evaluate the effect of addition of sucrose to starved cells in the absence of light, the cells were grown in the dark and in the presence of 5 8 mM sucrose. The transcript levels of ACP2 and ACP3 increased approximately 2.5-fold after 24 hours (Figure 11). In contrast, if the cells were kept in an osmotic control media in the dark the level of the same messenger RNA remained steady (Figure 11). Interestingly, we observed that the incubation of starved cells in the dark and 58 mM sucrose had a negative effect on the ACP4 mRNA levels, decreasing its abundance more than 2.5-fold. Conversely, no variations in the levels of the same mRNA were observed With. the osmotic control media in the dark (Figure 11). Thus, both the absence of light and the presence of 58 mM sucrose were responsible for the down-regulation of the ACP4 50 A) B) Time (hours) 06122448 3 i C‘. . ‘2 ACP3 1" ‘2’ ‘ ‘ 3 2 ,. . I j; , RBCSIA ACP4 “pull. 5 1 . _ ,. E , _~. . 01, A ACP4 FAD2 35* BF 3" ‘1‘ I“ is l \ j ~ .. g FAD2 RBCSIA Q.... “0 ACP2and3 eIF4A] a q as M n 4 Time (hours) Figure 10. Differential regulation of ACP mRNA levels in starved Arabidopsis cells. (A) Cells were starved for 48 hours in the presence of light and total RNA was extracted at the times indicated. Each lane of the Northern blots contains 2.5 ug of total RNA. The blots were hybridized with probes corresponding to the sequences indicated. The eIF4A probe was used as a loading control. (B) The Signals in A were quantified by scanning densitometry and the values are represented as in Figure 5. transcript levels. The same mRNA profile was observed for RBCSIA and Similar mechanisms of regulation might be operating for both genes (Figure 11). In the case of FAD2 mRNAs, no differences in the relative levels of the corresponding messenger were found in the conditions tested (Figure 10 and 11). Thus, in contrast to ACP, neither starvation nor light altered FAD2 mRNA abundance in liquid cell culture. To investigate whether sugars could have a direct role in the regulation of ACP transcript levels, we examined if ACP expression could be altered by uncoupling growth from the 51 presence of sugars in the media. For this purpose, cells were starved in an osmotic control media (58 mM mannitol) for 48 hours and subsequently transferred into media containing either 58 mM 3-O-methyl-glucopyranose (3-OMG) or 2 mM 2-deoxy-glucose (2-d-Glc). The effect of these two glucose analogs on gene expression has been previously studied in Arabidopsis cell-suspension cultures (Fujiki et al., 2000). The first glucose analog is taken up by cells but not phosphorylated whereas the second is phosphorylated but not further metabolized (Dixon et al., 1979). The presence of 3-OMG and 2-d-Glc in the media neither showed a positive effect on ACP2+3 transcript levels nor did it decrease the ACP4 mRNA level after 24 hours (data not shown). These results suggested that cell growth and/or a metabolic Signal generated by sucrose or downstream of glucose were necessary to control ACP mRNA levels in Arabidopsis cell liquid culture. ACP2 mRNA Levels are Affected by a Sucrose-derived Signal and/or Growth Control in planta To evaluate if mechanisms that regulate ACP2 transcript levels in Arabidopsis cell cultures also operate in intact plants, Arabidopsis seedlings were grown in liquid media in the presence or absence of carbon. The results in Figure 12 indicated that ACP2 mRNA levels responded to carbon signals in intact plants similarly to cell suspension cultures. Therefore, the underlying mechanisms were most likely physiological mechanisms of regulation of ACP2 expression in plants rather than mechanisms acting only in cell suspension cultures. 52 A) Dark/Sucrose (hours) Dark/mannitol (hours) ACP2 " “CU - '- ACP3 a. , ACP4 a ? 21' ‘ M M ii a M FAD2 r m A . RBCSIA . M u - a . “ fl ” I" ~ '6 ~ eIF4A 9' 1‘" B) 8 g 3 AC P2 and 3 5 l B a i E g 1 FAD2 E ACP4 g 0. RBCSIA 0 3 6 12 24 Time (hours) Figure 11. Differential regulation of ACP mRNA levels in Arabidopsis cells grown in the presence of sucrose. (A) Cells were starved for 48 hours in the light and subsequently transferred to either sucrose containing media (dark/sucrose) or an osmotic control media (dark/mannitol) and incubated in the dark. Total RNA was extracted at the times indicated. Each lane of the Northern blots contains 2.5 ug of total RNA. The blots were hybridized with probes corresponding to the sequences indicated and the eIF4A probe was used as a loading control. (B) The blots corresponding to the dark/sucrose treatment were quantified by scanning densitometry and the values are represented as in Figure 5. 53 Transcriptional regulation of Arabidopsis ACP2 gene by growth/cell cycle signals The increase in ACP2 mRNA abundance after carbon induction (Figures 10, l 1, 12) may be brought about by transcriptional or post-transcriptional mechanisms (e.g., mRNA stability). Therefore, to investigate which mechanisms are involved in ACP2 gene expression, a 1.5 Kbp DNA region upstream of this gene was fused to luciferase (LUC) reporter and transformed into Arabidopsis. In order to eliminate the effect on gene expression of the ACP2 5’ leader (Figures 8 and 9), this region was not included in the construct (see materials and methods). As a control, a 2 Kbp DNA region upstream of the eIF 4A1 gene was fused to LUC. The steady-state levels of eIF4Al mRNA remained constant regardless of the presence or absence of carbon in the media (Figures 10 and 11). In addition, all constructs carried a second reporter gene (GUS) under the constitutive CaMV35S promoter to normalize LUC activity for T-DNA copy number, position effect and in this case also for differential transcriptional activity in cells with variant energetic status (presence/absence of carbon sources). Since Arabidopsis cells grown in liquid culture are recalcitrant to Agrobacteria-mediated transformation, transgenic cell suspension cultures were induced from transgenic Arabidopsis plants. Stable transformation was preferred for these experiments because problems associated with transient transformation (e.g., loss of vector and reporter gene activity) were avoided. This is particularly important when performing experiments that 54 - Sucrose + Sucrose Hours: 0 12 24 48 12 24 At-ACPZ rRNA Figure 12. In planta regulation of ACP2 mRNA by carbon. Arabidopsis seedlings were grown in liquid media for 2 weeks in the presence of 58 mM sucrose and light. Subsequently, seedling were transferred to media containing 58 mM mannitol and incubated for 2 days in the light (-Sucrose). Samples were taken at 0, 12, 24 and 48 hours and total RNA extracted. After 2 days of starvation seedlings were transferred to media containing 58 mM sucrose for 1 day (+Sucrose). Samples were taken at 12 and 24 h after induction and total RNA isolated. RNA was resolved by electrophoresis and blotted onto nylon membranes. Filters were hybridized with an Arabidopsis ACP2 cDNA probe (At-ACP2). Ribosomal RNA (rRN A) in the gel was visualized by ethidium bromide staining. are carried out during several days. In addition, insertion Of the transgene in the genome provides a better platform for regulation of ACP2 promoter activity. Interaction of foreign DNA with nucleosomes in the chromosome context Should facilitate the organization of the promoter region (e.g., nucleosome position, binding of factors) to better reflect the actual promoter structure and regulation. 55 SUC ACP2:LUC STV-48h SUC-24 h SUC eIF4A] :LUC STV-48h SUC-24 h Fold Induction of LUC activity Figure 13. Regulation of Arabidopsis ACP2 promoter by carbon in Arabidopsis cell suspension culture. Arabidopsis cell suspension cultures were generated from leaf tissue of transgenic plants carrying 1.5 Kb of ACP2 promoter (ACP21LUC) and 2 kB of eIF4A] promoter (eIF4AlzLUC). Vectors also carried the GUS reporter gene under a constitutive promoter (CaMV35S:GUS) to control for T-DNA position and copy number. Three independent cell liquid cultures for each construct were generated from three independent transgenic plants. Cells were maintained in media supplemented with 58 mM sucrose (SUC). At the beginning of the experiment, cells were starved in media containing 58 mM mannitol for 2 days (STV-48h). After this period cells were transfered to sucrose containing media for 1 day (SUC- 24h). Cell samples were taken at the beginning and end of each period and LUC and GUS specific activities assayed in triplicate. LUC activity was normalized with GUS activity (LUC/GUS) and expressed as fold induction of LUC activity. LUC/GUS ratios were set to l for STV-48 to Simplify the interpretation of the results. *’*--‘ The results in Figure 13 indicated that the LUC/GUS ratio of specific activities was reduced by 2-fold after 2 days of starvation and increased by more than 3-fold when sucrose was added back. Changes in LUC expression in transgenic Arabidopsis cells were Similar to those in ACP2 mRNA levels under the same growth conditions (Figures 10 and 11). Moreover, no Significant differences in LUC/GUS ratio of specific activities were observed when LUC was under regulation of eIF4Al promoter. These results agreed with previous experiments in which eIF4A] transcript levels did not change with carbon availability (Figures 10 and 11). Thus, a 1.5 Kb DNA fragment upstream of the ACP2 gene promoter was sufficient to confer carbon regulation of gene expression. These results indicated that changes in ACP2 mRNA abundance were the result of transcriptional activation/de-repression by carbon rather than increased transcript stability. Other carbon sources were tested for their ability to induce LUC expression in Arabidopsis cells. In this regard, glucose and fructose either alone or together were capable to activate LUC expression in cells carrying ACP22LUC construct (Figure 14). These results demonstrated that metabolizable carbon sources other than sucrose were also capable to regulate ACP2 gene promoter. To test if carbon sources different from sugars could also increase LUC expression, Arabidopsis cells were incubated with 3 mM sodium acetate after starvation. Acetate is rapidly taken up and used as substrate for fatty acid biosynthesis, however it is not a good source of carbon to sustain Arabidopsis cell growth (data not Shown). Nevertheless, acetate regulates gene expression of genes 57 involved in various cellular processes (Sheen,l990). The results in Figure 14 indicated that acetate failed to induce LUC expression after starvation and suggested that growth Fold Induction of LUC activity [0 _. 'N o.) .Is mannitol sucrose fructose glucose deoxy 2-glucose 3-O-M-glucose acetate sucrose + cerulenin llllllll Figure 14. Regulation of ACP2 promoter by different carbon sources in Arabidopsis cell suspension cultures. Transgenic Arabidopsis cells expressing ACP22LUC and CaMV3SS:GUS constructs were starved for 2 days in mannitol containing media and then induced with different carbon sources for 1 day. After this period, LUC and GUS Specific activities were assayed in triplicate. LUC and GUS specific activities assayed in triplicate. LUC activity was normalized with GUS activity (LUC/GUS) and expressed as fold induction Of LUC activity. LUC/GUS ratios were set to 1 for mannitol treatment to Simplify the interpretation Of the results. Sucrose, glucose and fructose were at 58 mM, 2-deoxy- glucose (2-deoxy glc) at 5 mM, 3—O-methyl glucose (3-O—methyl glc) at 58 mM, acetate at 3 mM, cerulenin at 0.01 mM . Bars denote Standard deviations of the average. and not the sole presence of carbon sources was necessary to regulate ACP2 gene promoter. To further investigate the relationships between ACP2 gene transcription, carbon supply and growth, two different strategies were followed. First, cells were incubated with sugar analogs (3-O-methyl-glucose and 2-deoxy-glucose) that are not metabolized by cells yet sensed as carbon and therefore provide a sugar-sensing signal (Fujiki et al., 2000). Second, cells were incubated in the presence of sucrose plus 10 uM cerulenin. The rationale for the latter experiment was to provide carbon for growth but at the same time to halt growth/cell division by limiting supply of fatty acids. The results in Figure 14 indicated that both treatments, sugar analogs and sucrose plus cerulenin, failed to induce ACP2:LUC expression after starvation. These data suggested that growth and/or cell division is a requisite for regulation of transcription by the ACP2 gene promoter. Possibly, transcription of ACP2 gene may be regulated by cell-cycle signals. In agreement with this hypothesis, microarray profile analysis of synchronized Arabidopsis cell cultures demonstrated that ACP2 mRNA expression is cell-cycle coordinated (Menges et al., 2002). Deletion analysis of Arabidopsis ACP2 promoter in tobacco BY-2 cells To identify discrete DNA elements in the Arabidopsis ACP2 promoter responsible for transcriptional regulation by grth and/or cell cycle, a promoter deletion analysis was 59 conducted. A promoter deletion series of the Arabidopsis ACP2 promoter was fused to the luciferase reporter gene and the constructs used to transform tobacco BY-2 cells. Tobacco cells, instead of Arabidopsis, were chosen for this experiment because the former cells have the advantage of being transforrnable by Agrobacteria and consequently more constructs could be conveniently analyzed. Nonetheless, this system may have disadvantages. First, mechanisms existent in Arabidopsis to activate ACP2 gene expression could be different from those in tobacco cells. Second, factors in tobacco cells may not recognize elements in the Arabidopsis ACP2 promoter. The results in Figure 15 demonstrated that LUC expression driven by the 1.5 Kb ACP2 promoter responded to carbon Signals in tobacco BY-2 cells, similar to Arabidopsis cells. Furthermore, the increase and decrease in LUC expression by carbon availability were in the same range as observed in Arabidopsis cells (Figure 13 and 14). Thus, these data indicated that mechanisms for ACP2 gene expression were conserved between tobacco BY-2 cells and Arabidopsis cells and also that tobacco factors can recognize and regulate Arabidopsis ACP2 promoter by growth/cell cycle signals. The deletion series of the ACP2 promoter disclosed the presence of both positive and negative domains for gene expression in the Arabidopsis ACP2 promoter (Figure 15). Removal of the —1,500/—750 fragment reduced LUC expression by 6-fold, indicating the presence of positive elements in this promoter area. Although overall LUC expression was affected, the response to carbon was unaltered, suggesting that growth/cell cycle responsive 60 LUC activity - o 00 -1573 +1 Suc LUC Stv-48h - Sue-24h Suc LUC Stv-48h Sue-24h Suc LUC Stv-48h Suc-24h Suc LUC Stv-48h Sue-24h Figure 15. Arabidopsis ACP2 promoter deletion analysis in tobacco BY-2 cells. Tobacco BY-2 cells were transformed via Agrobacteria with a binary vector containing a deletion series of the Arabidopsis ACP2 promoter fused to luciferase (LUC). Vector also carried GUS fused to CaMV35S. Three individual calli were used to generate three independent cell suspension cultures for each construct. Cells were maintained in media with 58 mM sucrose (Suc). At the beginning of the experiment cells were starved in media containing 58 mM mannitol for 2 days (Stv- 48h). After this period cells were transferred to sucrose containing media for 1 day (Sue-24h). Cell samples were taken at the beginning and end of each period and LUC and GUS specific activities assayed in triplicate. Luc activity was normalized by GUS activity and expressed as LUC activity (units of light emitted). elements were still present in the 750 bp promoter fragment. Interestingly, deletion Of an additional 250 bp (-750 to —500) reconstituted LUC expression to levels Similar to those Obtained with the 1.5 Kb promoter fragment (Figure 15). This result indicated that 61 negative elements of expression were localized in the —750/—500 region. LUC expression still responded to carbon signals and therefore growth/cell cycle elements were present in the —500 promoter fragment. Further deletion of the promoter (-500 to —250) decreased by 10-fold the overall expression of LUC, indicating the presence of positive elements in the —500/-250 fragment. Induction of LUC activity by carbon was in the same range as with the other constructs (~ 2-3 fold) indicating the presence of growth/cell cycle responsive elements within the —250 promoter fragment (Figure 15). Conserved elements in Arabidopsis ACP promoters In addition to the conserved elements found in the 5’ UTRS (Table 2), computer alignments of Arabidopsis ACP genes (plastidial isoforms) disclosed the presence of two additional conserved elements (Figure 16). First, an 11 bp element was present in the promoter regions of the five ACP genes (consensus sequence: CCTGCATCTCC). This element was neither present in ACP genes of other plant species nor in other Arabidopsis FAS and lipid biosynthetic genes (Beisson et al., 2003). Therefore, it may be that the 11 bp element is Specific for Arabidopsis ACP promoters. Nevertheless, it has to be considered that information on ACP promoter sequences from different plant Species is still limited. Second, a 6 bp element (GCCAAA) was found at —200 bp in ACP2 and ACP3 promoter regions but not in the other Arabidopsis ACP promoters (Figure 16). Interestingly, the same element is present in FASl and FASZ genes of yeast that are regulated by growth (Chirala, 1992; Schuller et al., 1992) as well as in several Arabidopsis genes which are cell-cycle regulated (e. g., At-CDC6, A t-MCM3) (Stevens et Figure 16. Conserved elements in Arabidopsis ACP promoters. Two conserved elements in the Arabidopsis ACP genes were identified by computer alignments. An 11 bp element (highlighted in gray) was found in the promoter regions of the five Arabidopsis ACP genes. The consensus sequence for this element is indicated at the bottom of the figure together with its location (relative to the start codon [capital letters]) in the different ACP genes. Interestingly, the 11 bp element in the ACPS gene is located in the coding sequence. A 6 bp element GCCAAA (underlined) was found only in ACP2 and ACP3 promoters. This element overlaps with the 11 bp element and this observation may further suggest that these two elements are important for ACP gene expression. 63 Iputer ms of red at apital me is only i this gene v ACP1 ttgtcCCTGGTTCTCCgactgagagaagcagccatgatcttagtaaaccttgaggagaaag atatatagaaacttaaccaaaaaacttcttcttgctcttccctcttatggtgactagtatt gtgtttcacgaaacaATG ACP2 ttccaaCCTGCATCTfiQQAAAgcacccaactccacctgacttgtctgtgtctgttgcgttca ctttcattggaatctcagattttttattt‘ttttgtttgttttggttgttgcgatttgtgactataaaacctctcc cacttggttcttcactctcactgtttctcatctcttcgtcttctccgcctccttcaatctcactccgatctc tctacgattcattcgttctfi AC P3 tttgttatttctctaCCTGCATCT_(3—__CCAAA tcacccacctcacctgacttgtctgtctgaattct ctcattggaatctcagaaaagttttttttttttttttgtttaggttaaaatactataaaattaaaaataagtct cccacttggattcttcactgtgtctcacatctctttgtcttctccgcctcctccgatctcactccgatctc tctacgattcattcttc tATG AC P4 gtgccgaagataggCCTGAATCTCCgagaacaaacctccgccacaaaaacagaagactt gttgcttgcgtatcataatacgacgccgtatctctacttgctcgaacaaacccaaaaagatacatat caagagattaaaccttatccaactaagagaagccatttttattttttttgggtctctgagttgtgtattga gcttcatctccttcaafl ACPS ggatggtacttagaatagattttccaaaatgatggcagaatagaacgtggctctataaatacataaa tcccagcagtgtttgccatcagctacaaaatagtaattcacqctccttaaacaatccgccatctctct cgatcagatcgataATG gcgacaagtttctgcCCTCCATCTCC Consensus sequence (location) CCTGCATCTCC ACP1 (-615) CCTGGTTCTCC 1 1 bp element ACP2 (-21 l) CCTGCATCTCC ACP3 (-202) CCTGCATCTCC ACP4 (-188) CCTGAATCTCC ACPS (+17) CCTCCATCTCC Figure 16 64 al., 2002; de Jager et al., 2001). Although ACP2 and ACP3 appear to derive from a recent gene duplication (these isoforms are > 80 % identical. in their amino acid sequences and the respective genes lay one next to the other in the Arabidopsis genome), the divergent amino acid sequences of the other ACP isofomis (60-70 % identity) indicate that they diverged long ago. Therefore, the 11 bp element present in the promoter region of the five ACP isoforms represents a DNA motif that has been conserved for a long period throughout evolution. Thus, this element is a good candidate for a regulatory element of ACP expression. Because the presence of the 6 bp element in the ACP2 and ACP3 promoters is the result of a more “recent” gene duplication event, its conservation in these two promoters may only reflect its short existence in evolutionary terms. However, the presence of the same element in other Arabidopsis cell-cycle regulated genes suggests that the 6 bp element is also a good candidate for ACP gene regulation. Changes in mRNA expression by inhibition of fatty acid synthesis in Arabidopsis cells To investigate Whether feedback mechanisms for FAS gene regulation Similar to those in bacteria (Schujman et al., 2003) also exist in plant cells, Affymetrix GeneChip analysis was used to follow changes in gene expression after inhibition of fatty acid synthesis in Arabidopsis cells. Gene expression analysis of cell suspension cultures presents several advantages over analysis of whole tissues. First, cell cultures are homogeneous systems as opposed to tissues which are composed of different cell types. Second, growth 65 conditions can be rapidly changed and cells more readily take up components from the media. Finally, cells Share the same environment and therefore they should perceive Similar signals. Cerulenin, a potent inhibitor of fatty acid synthesis in plants was utilized in the experiment. This drug crosses cell membranes rapidly and specifically inhibits condensing enzyme KAS I (B-ketoacyl-ACP synthase) (Schneider and Cassagne,l995). In order to evaluate the concentration of cerulenin required to inhibit fatty acid synthesis significantly but with minimal cytotoxic effects, cells were incubated with 0, 10, 30 and 100 uM cerulenin for 0. 2 and 6 hours. Labeled acetate was added to the cultures and after the incubation period, cells were quenched with hot isopropanol and lipids extracted. Analysis of labeled fatty acids demonstrated that after 2 hours the inhibition of F AS was 65%, 82% and 94% for 10, 30 and 100 uM cerulenin, respectively, compared to the control (Figure 17). After 6 hours, inhibition of FAS in the presence of cerulenin remained Similar to values measured at 2 hours (70%, 84% and 96% for 10, 30 and 100 uM cerulenin respectively) (Figure 17). These data indicated that the effect of cerulenin on F AS persisted after 6 hours of treatment. In addition, the constancy in the percentage of FAS inhibition throughout the experiment suggested no major cytotoxic effects by cerulenin on Arabidopsis cells during this period. Based on these results, a concentration of 10 uM of cerulenin was chosen for gene expression experiments. This low concentration of inhibitor is enough to substantially reduce fatty acid synthesis (70%) and yet provide small amounts of fatty acids that would minimize cytotoxic effects and stress responses. 66 FAS genes do not respond to inhibition of fatty acid biosynthesis in Arabidopsis cells Affymetrix gene chips (8K) representing approximately one third of the Arabidopsis genome were used to evaluate changes in gene expression afier inhibition of fatty acid synthesis. Two flasks with 100 mL of cells growing at exponential phase were incubated in the presence of 10 BM cerulenin for 6 hours. Samples (20 mL) were taken at 0, 2 and 6 hours after drug treatment from each of the flasks (two‘biological replications per time point). Total RNA was isolated from each sample and used to synthesize cDNA with poly(T) primers fused to the T7 promoter. Subsequently, T7:cDNAs were used as templates for cRNA synthesis by in vitro transcription in the presence of biotinylated- ATP. Finally, biotinylated-cRNA was used for GeneChip hybridization reactions. The results in Table 3 indicated that mRNA levels from some genes involved in fatty acid and lipid metabolism changed slightly after 2 and 6 hours of inhibition of fatty acid synthesis. For example, the mRNA levels corresponding to the biotin containing subunit (BCCP) of acetyl-COA carboxylase increased ~ 2-fold after 6 h of cerulenin treatment. A few other mRNAs showed slight increases (e.g., 1.4-fold for ACP2 and cytosolic acetyl-COA carboxylase). The standard deviations of these values were within 5 % of the average value and therefore the fold increases were reliable. In contrast, the mRNA levels of other lipid related genes were reduced by cerulenin treatment. Transcripts for acyl- ACP 67 140000 120000 100000 80000 DPM 60000 40000 ~ 20000 0 , fItfiM 10 UM , 112130 uM C1100 uM 0 hours 2 hours 6 hours Figure 17. Inhibition of fatty acid synthesis by cerulenin in Arabidopsis cells. Arabidopsis cells were incubated with 0, 10, 30 and 100 uM cerulenin for 0, 2 and 6 hours. Labeled acetate (0.025 mCi) was added to the cultures. After the incubation periods cells were quenched with hot isopropanol and lipids extracted. Radiolabeled fatty acids were transmethylated and separated by thin layer chromatography. Quantitation of bands corresponding to fatty acids was performed by scanning in an Instant-Imager and scintillation counting. Results are expressed in desintegrations per minute (DPM). thioesterases declined approximately 1.6-fold after 6 h of drug treatment (Table 3). Thus, although small differences could be seen in the mRNA levels of some fatty acid synthesis genes, mRNA levels for most FAS and lipid synthesis genes were not significantly altered by inhibition of fatty acid biosynthesis. 68 Significant changes in mRNA levels were observed for genes related to cellular processes such as stress response, metabolism, drug resistance and several regulatory proteins such as transcription factors and kinases. Approximately 90 genes presented increases in transcript abundance of more than 2-fold after 6 h of cerulenin treatment (Table 4). Changes in the transcript levels of genes related to stress response, metabolism and drug resistance most likely reflected the stress conditions imposed by inhibition of fatty acid synthesis. Approximately 110 genes presented lower mRNAs levels than control cells after 6 h of cerulenin treatment (Table 5). DISCUSSION One objective addressed by the current work is the nature of the signals that give rise to changes in the expression of genes involved in fatty acid synthesis. Although a limited number of previous studies suggest a minor role of light on the expression of these genes (Battey et al., 1990; Baerson et al., 1993, Scherer et al., 1987), the mutually dependent relation between cthTOplast biogenesis and de novo production of glycerolipids envisage a closer connection between light and at least some of the genes for this pathway. In this report we demonstrated that light affects the expression of mRNAs encoding acyl carrier protein isoforms from Arabidopsis. First, we demonstrated that the levels of the messenger for ACP4 increased 4- to 5-fold after light treatment of dark grown plants. The similarity in the kinetics of mRNA induction between ACP4 and FEDA in Arabidopsis suggests Similar mechanisms of activation for both genes. Nevertheless, it remains to be examined whether the increase of 69 Table 3 . Changes in mRNA abundance of genes involved in lipid metabolism after inhibition of fatty acid synthesis in Arabidopsis cells. descriptions Fold chan e at 2 h at 6 h AT5G16390 acetyl-COA carboxylase (BCCP subunit) 1.7 i 2.1 AT4G38570 putative phosphatidylinositol synthase 1.7 r 2.1 AT4G22340 CDP-diacylglycerol synthetase-like protein 1.5 2.0 AT1G3617O acetyl-COA carboxylase subunit 1.2 1.5 AT1G54580 acyl carrier protein isoform 2 1.0 1.4 AT1G54580 acyl carrier protein isoform 2 1.2 1.4 AT3G51840 Short-chain acyl COA oxidase 1.2 1.2 AT3G12120 delta-12 desaturase (Fad2) -2.1 1.2 AT4G30950 omega-6 fatty acid desaturase (fad6) 1.0 1.2 DL4775C thioesterase like protein 1.0 1.1 glycerol-3-phosphate acyltransferase (ACTI) 1.0 1.1 AT1G01480 acetyl-COA carboxylase (ACC2) 1.0 1.1 ATZGZS710 biotin holocarboxylase synthetase -1.1 1.1 ATSGOSSSO omega-3 fatty acid desaturase (fad8) 1.0 1.1 AT2G44620 acyl carrier protein precursor (mitochondrial) -1.0 1.1 AT2G32260 putative phospholipid cytidylyltransferase 1.0 1.1 AT4G23850 acyl-COA synthetase-like protein 1.3 1.1 AT1G67730 b—keto acyl reductase (glossy8) 1.3 1.1 AT4G13840 fatty acid elongase-like protein (cer2-like) 2.0 1.1 DL4770C thioesterase like protein 1.6 1.1 AT5G35360 biotin carboxylase subunit (CAC2) -l .3 1.1 AT4020870 fatty acid hydroxylase-like protein 1.0 1.1 ATZG31360 delta 9 desaturase 1.0 1.1 AT2G33150 3-ketoacyl-COA thiolase -1.3 1. 1 AT3G11170 omega-3 fatty acid desaturase 1.0 1.1 AT4G34510 putative ketoacyl-COA synthase -l.3 1.1 AT1G76490 3-hydroxy-3-methylglutaryl COA reductase -8.3 1.1 AT3G1 1 170 fatty acid desaturase -1.9 - l .l AT4G34250 fatty acid elongase-like protein -1.0 -1.1 DL4435W triacylglycerol lipase like protein - l .2 - l .1 AT1G68530 very-long-chain fatty acid condensing enzyme CUTl -l .0 -1.1 AT4G25050 acyl carrier-like protein (ACP-4) 1.0 -1.2 *Fold change: fold increase in signal intensity compared to time o h 70 Table 4. Example of transcripts that increase more than 2 fold after 6 hours of cerulenin treatment. descriptions Fold chan e at 2 h at 6 h T020048 jasmonate inducible protein isolog 9.0 11.2 AT5G13930 chalcone synthase (CHS) 1.9 10.8 AT2G34660 multidrug resistance-associated protein 2 7.5 9.3 AT4GI9030 nodulin 26 like protein 14.2 8.7 AT4G12360 putative lipid transfer protein 6.5 5.4 AT1G17740 Phosphoglycerate dehydrogenase 2.6 5.3 AT1G14900 high mobility group protein a (HMGa) 1.9 5.1 AtMYB42 R2R3-MYB transcription factor. 9.0 5.0 AT2G33810 squamosa promoter binding protein-like 3. 4.5 4.7 AT5G18170 glutamate dehydrogenase 1 (GDHl) 2.4 4.6 ATl 610460 germin-like protein (GLP7) 4.7 4.6 AT4G21880 membrane-associated salt-inducible protein 4.0 4.6 AT5G08640 flavonol synthase 1.6 4.2 AT2G16590 putative protein 24.7 4.1 AT3G22490 LEA D34 protein homologue typel. 2.6 4.0 AT1G78820 receptor like ser/threo kinase ARK3 3.7 4.0 AT2G47000 multi drug resistance proteins 4.4 3.8 AT2G01830 putative histidine kinase (CREl) 3.0 3.6 F13P17.32 cytochrome P450 homolog 2.4 3.5 F16N3.5 prolamin box binding factor (PBF) 3.7 3.4 AT1G09530 phytochrome-associated protein 3 (PAP3). 2.0 3.3 AT1G19050 ARR7 mRNA for response regulator 7 3.8 3.2 AT2G1 1620 putative 3.3 3.2 AT2G18650 RING zinc finger protein 2.8 3.1 *Fold change: fold increase in signal intensity compared to time o h 71 Table 5. Example of transcripts that decrease more than 2-fold after 6 hours of cerulenin treatment. 72 descriptions Fold change at 2 h at 6 h ATSGS756O xyloglucan endotransglycosylase (TCH4) -3.45 -8.85 AT4G27280 calcineurin B-like protein 3 -4.05 -8.45 AT3G47380 putative -2.35 -6.95 AT2G19800 putative -0.55 -6.40 AT1G32170 xyloglucan endotransglycosylase-related protein -3.55 -6.35 AT4GZ7450 glutamine-dependent asparagine synthetase -1.20 -5.75 late embryogenesis abundant protein homolog -2.90 -5.65 AT2G37170 aquaporin -4. 50 -5 .40 AT4G23550 putative protein -1 .60 -5.40 AT3G47340 glutamine-dependent asparagine synthetase -3.55 -5.40 AT2G33830 putative auxin-repressed protein -3.25 -5.30 AT1G131 10 cytochrome P450 -0.15 -5.20 AT2G44380 putative CHP-rich zinc finger protein -0.40 -5.05 AT2G28630 putative fatty acid elongase -1.60 -5.00 class 1 non-symbiotic hemoglobin -l.15 -5.00 AT5G59520 zinc transporter (ZIP2) -4.50 -4.95 AT4G02380 putative -2.80 -4.95 AT5G24090 acidic endochitinase gene -3.55 -4.95 class IV chitinase -3.70 -4.90 AT3G57700 protein kinase-like protein -3.85 -4.85 DL4265W membrane transporter like protein -5. 15 -4.80 A_TM017A05.3 putative -0.75 -4.80 AT2G44840 ethylene response element binding protein -4.50 -4.75 AT4G08950 putative phi-l-like phosphate-induced protein -2.20 -4.65 AT4G22590 trehalose-6-phosphate phosphatase-like protein -2.50 -4.65 *Fold change: fold decreased in signal intensity compared to time o h the ACP4 transcript levels is direct due to light (e.g. via phytochromes) or indirect via cell growth. The analyses of 1 Kb of genomic sequence upstream of the Acll.4 gene disclosed the presence of several GATA-like motifs. The GATA (or I) boxes are regulatory elements that are functionally important in many light-regulated promoters. The core element is defined as GATAA and related GATA motifs with variable flanking sequences are found in several promoters (Terzaghi et al., 1995). For instance, the le upstream region of the ferrredoxin-A gene contains seven copies of related GATAA elements. Although the actual transcription initiation site for the ACP4 gene has not been mapped yet, we localized the GATA-like motifs respective to the AUG initiation codon of ACP4. Thus, the upstream region of this gene presents two AAGATAA elements at — 715 and -812, one AGATAA at —505 and three GATAA at —70, -526 and —908 base pairs respective to the translation initiation codon. The presence of these elements in the upstream region of the gene for ACP4 suggests the direct role of light on the transcription of this gene. In contrast, only one and two copies (expected by chance) of these elements are present in the upstream regions (1 Kb) of AclI.2 and AclI.3 genes respectively. This observation is consistent with results that demonstrate that light has no impact on mRNA levels of ACP2 and most likely ACP3 (Figure 5 and Baerson et al., 1993). A second aspect of the influence of light on ACP gene expression was observed at the level of ribosome association. A general effect of light on polysome formation has been previously described in several studies (Giles et al., 1977; Mosinger et al., 1983). These works Show that the proportion of ribosomes present as polyribosomes increases substantially in response to light. Moreover, continuous far—red (FR) light mediates a 73 strong increase in the relative level of polysomes, demonstrating the participation of phytochromes in the response. Despite this general effect on translation, there is evidence that particular mRNAs are less affected by the decrease in the translation rate in the absence of light. For example, Berry et al (1990) showed that in Amaranth cotyledons the expression of the small and large subunits of Rubisco are not found associated with polyribosomes in cotyledons of dark grown plants but rapidly recruited onto them upon illumination. In contrast, mRNAs encoding non-light regulated proteins are associated with polysomes regardless of the light-dark treatment. Our results (Figure 6) demonstrated that the transcripts for the ACP isoforms tested and FEDA largely dissociated from polysomes during the dark period in contrast to the non-plastidic eIF 4A mRNA, which remained associated with high molecular weight polyribosomes in the dark. Thus, these observations reinforce the notion that cytoplasmic mRNAs are differentially associated with polyribosomes during the light-dark period (Berry et al., 1990; Petracek et al., 1997). In chloroplasts of Chlamydomones and plants, both translation and mRNA stability are enhanced by nuclear encoded factors that bind to the 5’ UTR of several chloroplast- encoded RNAS (Danon et al., 1991; McCormac et al., 2000). Likewise, several studies confirmed the major role of the 5’UTR in translation and mRNA stability of cytoplasmic mRNAs (Dickey et al., 1998; Lukaszewicz et al., 1998). A previous report demonstrated that these regions can also be important for transcription activation (Bolle et al., 1994 and 1996). Analyses of the 5’ leader sequences of several genes encoding for thylakoid proteins in spinach, disclosed the presence of CT-rich regions, designated CT-leader 74 boxes (CT-LB and CT-B) (Bolle et al., 1994). The deletion of these elements from the PsaF (subunit III of photosystem I) and PetH (plastocyanin) upstream gene sequences severely reduces the transcription of a reporter gene in transgenic tobacco (Bolle et al., 1994). It is noteworthy that the CT-LB box is also present in the 5’UTR of the spinach ACP II (Bolle et al., 1996). The presence of several CT-rich elements and the CTCCGCC box in the ACP 5’UTRS of several diverse plant species (Table 2) suggests that these motifs might be important for ACP gene expression. The results presented in Figure 9 provide convincing evidence that these regions have a positive effect on reporter gene expression in transgenic Arabidopsis plants. Moreover, this effect was enhanced by light, suggesting that the ACP 5’UTRs may contain light-responsive elements that increase gene expression post-transcriptionally and/or transcriptionally in the context of the CaMV3SS promoter (Figure 9). We also demonstrated that the 5’ UTRS of Arabidopsis ACP1 and ACP2 conferred differential reporter gene expression in a tissue specific manner (Figure 9). At this point it is not possible to distinguish between a transcriptional, translational or mRNA stability mechanism to explain these differences in reporter gene activity. Gallie et al (1992) showed that 5’ leader length alters the expression of a reporter gene in carrot protoplasts. In particular, translation is influenced by leader length, and a 74 base leader construct is expressed 6-fold more highly than a 29 base leader construct. Conversely, Bolle et al (1994) found that in transgenic tobacco, the expression of a reporter gene under CaMV35S promoter is not altered when the 5’ leader of PsaF (188 bases) is added to the 5’ UTR (24 bases) of the native reporter gene mRNA. Based on 75 these results, it is possible that part of the differences observed in the LUC/GUS ratios in transgenic Arabidopsis tissues could be attributed to differences in the length of the ACP and control leaders (Figure 8 and 9). However, the tissue specific differences in reporter gene activity conferred by the three constructs analyzed, that ranged from almost no effect in roots to 10 to 30-fold increase in leaves and seeds, supports a sequence- dependent enhancement by the ACP 5’ UTRS. The results presented in Figures 10 and 11 for Arabidopsis mesophyll-derived cell culture indicate that the messengers for ACP2, ACP3 and ACP4 are under growth and/or metabolic control. The induction of the ACP2 and ACP3 transcript levels in the presence of 58 mM sucrose together with the inability of sugar analogs to reproduce this effect demonstrates that cell growth and/or a signal generated by sucrose or downstream of glucose is required for up-regulation of these messenger RNAS. In contrast to the response of these messenger RNAS, the ACP4 mRNA levels remained steady in starved cells and in the presence of 3-O-methyl-glucopyranose whereas they were down- regulated in the presence of 58 mM sucrose in the dark (Figure 10 and 11). The observation that mRNA levels for the ER localized F AD2 desaturase Showed no variation under the conditions analyzed indicates that the response of genes encoding plastid components of fatty acid synthesis may differ from those controlling later steps such as cytosolic desaturation. Regulatory mechanisms of ACP2 gene expression also responded to carbon signals in planta, suggesting that these mechanisms were physiologically relevant and not the 76 consequence of cell culture induction. Analysis of the ACP2 promoter demonstrated that carbon-mediated regulation of ACP2 gene expression was transcriptional and that promoter elements responsible for this regulation were located within —250 bp from the transcription initiation site. Comparison of Arabidopsis ACP promoters identified a GCCAAA element located at —200 bp of ACP2 and 3 genes. This element was also found in yeast FAS genes (Chirala, 1992; Schuller et al., 1992) and Arabidopis cell-cycle regulated genes (Stevens et al., 2002; de Jager et al., 2001). Therefore, it is a good candidate element for growth/cell cycle regulation of A CPZ gene expression. Analysis of gene expression after inhibition of fatty acid synthesis suggested that, in contrast to bacteria, fatty acid and lipid synthesis genes in plants did not respond to feedback mechanisms of gene expression after blocking this primary pathway. Thus, mechanisms other than regulation of FAS mRNA abundance may be involved in counteracting the inhibitory effect of cerulenin in plant cells (e.g., post-translational modifications of gene expression, metabolic regulation of FAS enzymes). In summary, results of this study demonstrate that expression of genes involved in plant fatty acid synthesis is under multiple levels of control. Light clearly has a major impact on ACP4 mRNA levels, but this impact does not extend to the other ACP isoforms examined here. These different responses to light on ACP are probably transcriptional, based on the presence of GATA boxes in the ACP4 but not other ACP upstream regions. Light control is also exerted post-trancriptionally because all ACP isoforms are more highly associated with polysomes in the light than the dark. Experiments with suspension 77 cultures indicate that differences in expression patterns of ACP isoforms are also observed in their response to sugar supplements. A common element in both the leaf and tissue culture experiments is that ACP4, the most leaf-specific isoform, behaves in a manner Similar to mRNA for genes involved in photosynthesis (FEDA or RBCSIA). In contrast, expression of the other ACP isoforms may be most responsive to demands for fatty acid synthesis brought about by enhanced growth. This level of complexity demonstrated by multiple ACP genes under different controls may reflect the need of plant cells to tightly regulate both the amount and the cellular destination of fatty acids produced in plastids (Ohlrogge and Browse, 1995) and to match the supply of fatty acids to different tissue and environmental demands. MATERIALS AND METHODS Nomenclature There are five genes for plastidial ACP in the Arabidopsis thaliana genome (Mekhedov et al., 2000). In this study we report results for Ac11.1, AclI.2, AclI.3 and AclI.4 genes, corresponding to the gene products ACP1, 2, 3 and 4 respectively. Acl].1 (GenBank accession number X13708) was previously described in Post-Beittenmiller et al., 1989 and Ac112.2 and AclI.3 (GenBank accession numbers X57698 and X57699 respectively) in Lamppa et al., 1991. We propose that the ACP major leaf isoform (Shintani, 1996) gene be named AclI.4 (isoform ACP4, At4g25050 or F13M23.190) and that Ac11.5 (ACPS) referred to GenBank accession number: A_TM021B04.6 or AF 007271. 78 Plant Growth and Lighting conditions Wild type Arabidopsis thaliana plants (ecotype Columbia) were grown on soil at 20°C in a 12:12 hours lightzdark photoperiod (80-100 umol m'2 s"). For the analyses of light- mediated induction of mRNA levels, 2 week-old plants were left in the dark for 24 hours and reilluminated with white fluorescent light (80-100 umol m'2 s”) for the times indicated in Figure 5. Leaf tissue was harvested and immediately frozen in liquid nitrogen prior to RNA extraction. For polyribosomal association analyses, 2 week-old plants were dark incubated for 12 hours and then incubated in the dark for an additional 12 hours or illuminated for 12 hours. Leaf tissue was harvested at the end of each period and processed as indicated above. Constructs and Plant Transformation Sequences corresponding to the 5’ untranslated regions of ACP1 (GenBank accession number X13708) and ACP2 (X57698) mRNAs were obtained by PCR amplification of Arabidopsis (ecotype Columbia) genomic DNA using the primers: 5 ’CATGCTCGAGCTCTTTGTACACTCCGCCCT3’ and 5 ’- CATGCCATGGCGCTGAATTGAGTCGCCAT-3’ for ACP1 (gene AclI.1); 5’- CATGCTCGAGACTGTTTCTCTATCTCTTTG-3’ and 5’- CGCGCCATGGCAATGGAAGCCATAGAAGAAT-3’ for ACP2 (gene AclI.2). The PCR amplification products were digested with 20101 and Ncol and cloned in frame to the 79 firefly luciferase gene (LUC) in pUC-LUC-BT2 (WeiBhaar et al., 1991) to give pTACPl and pTAC2 respectively. A truncated version of the ACP1 5’ leader sequence was generated by in vitro annealing of two synthetic oligonucleotides: 5’- TCGAGCTCTTTGTACAAACAATGGCGACTCAATTCAGCGC-3’ and 5’- CATGGCGCTGAATTGAGTCGCCATTGTTTGTACAAAGAGC-3’. The double- stranded fragment was cloned in frame to the LUC coding region in pUC-LUC-BT2 to give pTdell. All constructs were confirmed by sequencing. The vector pTACPl was digested with BamHI and the vectors pTACP2 and pTdell with BglII. The resulting fragments were cloned independently in the BamHI site of the binary vector pCAMBIA- 2201 (Hajdukiewicz et al., 1994) to generate pCaACPl, pCaACP2 and pCadell respectively (Figure 8). The orientation of the inserts was confirmed by restriction mapping using Pstl. Agrobacterium tumefaciens strain C58Cl (pM90) (Koncz et al., 1986) was transformed with the pCaACPI, pCaACP2 and pCadell binary vectors by the freeze/thaw transformation protocol (An, 1987). The leaf-vacuum infiltration method was used for Arabidopsis transformation (Bechtold et al., 1993). The transgenic lines transformed with pCaACPl , pCaACP2 and pCadell were named atL-acpl, atL-acp2 and atL-dell respectively. A 1.5 Kb DNA fragment upstream of the Arabidopsis ACP2 gene was amplified by PCR using the following primers: fwd, 5’ GACTGGATCCTGCATTAGCTGGTAGTTCAG 3’ and rvs, 5’ GACTCTCGAGTGAGAAACAGTGAGAGTGAA 3’. The resulting PCR fragment was digested with BamHI and Xhol and ligated into the pUC-LUC-BT2 vector upstream of the luciferase coding sequence (Td-2-ACP2-LUC). A 2 Kbp upstream region 80 corresponding to the Arabidopsis eIF 4A1 gene was amplified by PCR using the following primers: fwd, 5’ GACTGGATCCGAAAACGAACCTCAGTTCAG 3’ rvs, 5’ GACTCTCGAGGAGAGACTGGTGGAATATCC 3’. The resulting PCR fragment was digested with BamHI and Xhol and ligated into the pUC-LUC-BT2 vector upstream of the luciferase coding sequence (Td-2-eIF4A1-LUC). The constructs were digested with Bglll and independently subcloned into the BamHI site of pCAMBIA-2201 binary vector. The orientation of the insert was analyzed by digestion with Pst-I and sequencing. For construction of the ACP2 promoter deletion series, the vector Td-2-ACP2-LUC carrying 1.5 Kbp upstream region of the ACP2 gene was used as template for PCR. A combination of forward primers with the same reverse primer (same sequence as rvs primer described above) was used to generate the deletions. The forward primers were as follows: D750, 5’ GACTGGATCCGTAGTA AACCACTATTGAAA 3’; D500 5’ GACTGGATCCCACATAGTAAATTGATGCGT 3’; D250 5’ GACTGGATCCGACGTAGTAGAATTTACAAA 3 ’; D l 00 5’ GACTGGATCCTGACTTGTCTGTGTCTGTTG 3’. The PCR products were digested with BamHI and Xhol and ligated into Td-2 vector. The different constructs were digested with BglII and ligated into the BamHI Site of pCAMBIA 2201 as indicated above. All constructs were confirmed by restriction mapping and sequencing. Selection of Transgenic Plants and Tissue Collection 81 Seeds (T1) from transformed Arabidopsis plants were surface-sterilized and sown on seed germination media in the presence of 50 ug/mL kanamycin. After one week, resistant seedlings (T1 plants) were transplanted individually to soil and grown at 20°C in a 18:6 hours lightzdark period until they set seeds. For reporter gene analyses in leaf tissue, seeds from T2 plants were selected in germination media under 50 ug/mL kanamycin and transferred to soil. Plants from 10 independent transgenic lines for each construct (see Figure 9) were grown in a 12:12 hours lightzdark cycle for 2 weeks. The last day the plants were kept in the dark for 24 hours and the leaf tissue from half of the plants per line was collected and immediately frozen in liquid nitrogen for subsequent enzyme analyses. Leaf tissue from the remaining plants was harvested as above after 6 h of reillumination with white fluorescent light (80-100 umol m"2 s"). For reporter gene analyses in developing seeds, T2 seedlings were selected as above and transferred to soil. Ten independent lines per construct (see Figure 9) were grown in a 16:8 hours lightzdark cycle. Flowers from the primary stem were tagged in the morning every day during one week. Flower anthesis was considered as day zero in our experiment, and mid-stage siliques (8 to 10 days after anthesis) were dissected and the seeds immediately frozen in liquid nitrogen for subsequent enzyme analyses. Cell Suspension and Root Liquid Cultures Arabidopsis (ecotype Columbia) suspension cultures (Axelos et al., 1992) were a gift from N. Raikhel laboratory (Michigan State University, DOE Plant Research Laboratory). Suspension cultures were maintained in a 12:12 hour lightzdark photoperiod 82 at 22°C on a rotary Shaker (120 rpm) in CSM-sue media (0.32% (w/v) Gamborgs with minimal organics (Sigma G5893), 58 mM sucrose, 0.05% (w/v) MES, 1.1 pg mL'l 2,4D, pH 5.7). For the analyses of mRNA levels in different growth conditions, 100 mL cell cultures were grown for 3 days in CSM-sue in the presence of white light (100 umol m’2 3"). After the third day, the cells were pelleted at 700 rpm and washed twice with CSM- ma (0.32% (w/v) Gamborgs with minimal organics (Sigma G5893), 58 mM mannitol, 0.05% (w/v) MES, 1.1 ug mL" 2,4D; pH 5.7). Cells were finally resuspended in 150 mL of CSM-ma and incubated in the presence of light for 2 days (starved cells). An aliquot of 10 mL was taken at the times indicated in Figure 10 and the cell pellet stored at -80°C until RNA extraction. The remaining 100 mL of the culture was pelleted as before, resuspended in either 50 mL of CSM-sue or 50 mL of CSM-ma and incubated in the dark for 24 hours. An aliquot of 10 mL was taken at the times indicated in Figure 11 and the cell pellet stored at —80°C until RNA extraction. In the experiment with glucose analogs, cells were starved with CSM-man for 2 days and subsequently transferred to CSM supplemented with either 58 mM of 3-O-methyl-D-glucopyranose (Sigma) or 0.2 mM of 2-deoxy-D-glucose (Sigma). Cells were incubated in the dark and an aliquot of 10 mL was taken at 0, 3, 6, 12 and 24 hours and the cell pellet stored at -80°C until RNA extraction. Induction of Arabidopsis cell suspension culture from leaf tissue was performed as described by Axelos et al (1992). Tobacco bright yellow-2 cells (BY-2) were transformed via A grobacterium tumefaciens strain LBA4404 as described by BaO et al (2002). 83 For root liquid culture, approximately 50 seeds from ten independent transgenic lines (T2 plants) per construct (see legend Figure 9) were surface sterilized and incubated in root liquid media (0.43% (w/v) MS salts, 0.05% (w/v), MES, 0.1% (v/v) 1000X B5 vitamin stock, 50 ug mL‘I kanamycin, pH 5.7) at 4°C for two days. The seeds were then incubated at 22°C in continuous fluorescent white light on a rotary shaker (120 rpm) for 2 weeks. Root tissue from approximately 20 seedlings per transgenic line was collected under the dissecting microscope and frozen in liquid nitrogen for subsequent enzyme analyses. RNA Extraction and Northern Analyses Total RNA was isolated from plant tissue and cell pellets by standard phenol/chloroform extraction and lithium chloride precipitation (Sambrook et al., 1989). The RNA was fractionated in formaldehyde-agarose gels and blotted onto Hybond-N filters (Amersham-Pharmacia Biotech, NJ). In all cases hybridization reactions were done at 42°C in 50% formamide with a-[32P]-dCTP radiolabeled fragments corresponding to ACP2 (X57698), ACP3 (X57699), ACP4 (RXW18), FEDA (M35868), eIF4A (X65052), FAD2 (L26296) and RBCSIA (X13611). After autoradiography, the relative mRNA levels from the northern blots were quantified by scanning densitometry (Molecular Dynamics). Messenger Specific probes for ACP2 and ACP3 transcripts were generated by PCR amplification of Arabidopsis (ecotype Columbia) genomic DNA corresponding to the 3’UTR of these transcripts. The primers used were: 5’- TGAAAAGGCCAAGTAGAAT-3’ and 5’-GTCAGATACAAGCCTTGTA-3’ for 84 ACP2; 5’-GGAAAAGGCCAAGTAGAAA-3’ and 5’-CAAGCCTTGTAATAATTATC- 3’ for ACP3. Polyribosome Analyses The protocol for polyribosome isolation was adapted from Petracek et al (1997) with the following modifications: approximately 1 g of Arabidopsis leaf tissue or B.napus developing seeds were homogenized in 4 mL of U buffer (200 mM Tris-HCl , pH 8.5, 50 mM potassium chloride, 25 mM magnesium chloride, 2 mM EGTA, 100 ug/mL heparin, 2% polyoxyethylene, and 1% deoxycholic acid), centrifuged at 13,000 g for 15 min at 4°C and the complete supernatant (4 mL) was loaded onto a 30 mL linear sucrose gradient (15 to 60%). After gradient centrifugation, fifteen 2 mL fractions were collected by dripping directly into 2 mL of phenol-chloroform, 50 uL of 10% SDS, 40 uL Of 0.5M EDTA, and 10 uL of 100 mM aurin-tricarboxylic acid (Sigma). The final RNA pellets were washed with 70% ethanol, dried and resuspended in 10 LIL water, 20 uL forrnamide, 6.5 uL formaldehyde, and 3.5 uL MOPS buffer (0.2 M MOPS, pH 7, 50 mM sodium acetate, 5 mM EDTA). Samples were heated for 15 min at 68°C and loaded onto a formaldehyde-agarose gel. Northern analyses were performed as described above. For gradient UV profile analyses, two gradients were spun in parallel and one used to measure UV absorbance. Fifty 100 uL samples were taken at identical volume intervals and the absorbance measured at 260 nm. Tissue Extraction and Enzyme Assays in Transgenic Plants and Cells. 85 Leaf tissue or cell pellets (~ 0.25 g) were ground with a mortar and pestle and resuspended in either 300 LIL of GUS buffer (200 mM Trizma-HCL, pH 7.0, 10 mM 13- mercaptoethanol, 10 mM EDTA, 0.01% (w/v) SDS, and 0.1% (v/V) Triton X-100) or 300 uL of LUC buffer (100 mM potassium phosphate, pH 7.8, 1 mM EDTA, 7 mM 13- mercaptoethanol, 10 % (v/v) glycerol) and further homogenized on ice. Root tissue (~ 0.2 g) was homogenized and resuspended in IOOuL of either GUS buffer or LUC buffer. Developing seeds (~ 0.1 g) were homogenized in either 200 LIL of GUS buffer or LUC buffer. In all cases, cell debris was removed by centrifugation for 15 min at 13,000g at 4°C. The supernatant was transferred to a clean tube and kept on ice. For GUS activity, 50 uL of the sample supernatant were diluted in 400 LIL of GUS buffer and mixed with 50 uL of 10 mM 4-methylumbelliferyl-b-D-glucuronide (Sigma). The reaction was incubated at 37°C and a 100 uL aliquot was diluted in 0.2 M NazCO3 at time 0 and every 15 min for 60 minutes. UV fluorescence was measured at 365-mn excitation and 455-nm emission with a Hitachi F -2000 fluorometer. GUS Specific activity is expressed as nmol methylumbelliferone produced min'1 mg"1 protein. For LUC analyses, a 25 uL aliquot Of the sample was mixed with 100 uL of LUC reagent (20 mM Tricine, pH 7.8, 5 mM MgC12, 0.1 mM EDTA, 3.3 mM DTT, 270 uM COA, 500 uM luciferin (Promega), 500 uM ATP). The linear range of the reaction was determined by mixing increasing extract volumes with 100 LIL Of LUC reagent. Luminescence was measured (3 s delay, 10 5 integration time) in a Tumer TD-20e luminometer. LUC Specific activity is expressed as light units/mg protein. In all cases the GUS and LUC specific activities and LUC/GUS ratios were measured in triplicate and independently for each transgenic line. Protein 86 concentration was determined using a protein assay kit (Bio-Rad) with BSA (Sigma) as the standard. Determination of FAS inhibition by cerulenin in Arabidopsis cell cultures. Cells were incubated in the presence of 0, 10, 30 and 100 uM of cerulenin and 25 uCi of [l-MC] sodium acetate for 0, 2 and 6 hours. Total lipids were extracted by hexane- isopropanol method and fatty acids transmethylated with 5% (v/v) sulfuric acid in methanol for 1 hour at 80° C. FAMES were separated by thin layer chromatography (TLC) with 90/10 (v/v) hexane/diethyl-ether. Radiolabeled bands corresponding to fatty acids were quantitated by scanning in an Instant Irnager (Packard, Meriden, CT) and scintillation counting (Beckman Instruments, Fullerton, CA). Affymetrix GeneChip analysis. All the experimental protocols were performed according to the Affymetrix GeneChip® Expression Analysis Manual (http://www.affymetrix.com). Data analysis was performed with GeneChip® Data Mining Tool Software (Affymetrix, Inc; Santa Clara, CA). LITERATURE CITED An G (1987) Binary Ti-Vectors for plant transformation and promoter analyses. Method Enzymol 153: 292-305 Axelos M, Curie C, Mazzolini L, Bardet C, Lescure B (1992) A protocol for transient gene expression in Arabidopsis thaliana protoplasts isolated from cell suspension cultures. Plant Physiol Biochem 30: 123-128 87 Axelos M, Curie C, Mazzolini L, Bardet C, Lescure B (1992) A protocol for transient gene expression in Arabidopsis thaliana protoplasts isolated from cell suspension cultures. Plant Physiol Biochem, 30: 123-128. Baerson SC, Vander Heiden MG, Lamppa GK (1998) Overlapping yet distinct promoter activities of two Arabidopsis genes coding for nearly identical isoforms of the acyl carrier protein. Int J Plant Sci 159: 533-538 Baerson SR, Lamppa GK (1993) Developmental regulation of an acyl carrier protein gene promoter in vegetative and reproductive tissues. Plant Mol Biol 22: 255-67 Bao X, Focke M, Pollard M, Ohlrogge J (2000) Understanding in vivo carbon precursor supply for fatty acid synthesis in leaf tissue. Plant J 22: 39-50 Bao X, Katz S, Pollard M, Ohlrogge J (2002) Carbocyclic fatty acids in plants: biochemical and molecular genetic characterization of cyclopropane fatty acid synthesis of Sterculiafoetida. Proc Natl Acad Sci U S A, 99: 7172-7. Battey JF and Ohlrogge JB (1990) Evolutionary and tissue specific control of expression of multiple acyl carrier protein isoforms in plants and bacteria. Planta 180: 352-360 Bechtold N, Ellis J, Pelletier G (1993) In-planta Agrobacterium-mediated gene-transfer by infiltration of adult Arabidopsis thaliana plants. Cr Acad Sci III-Vie 316: 1194-1199 Berry JO, Breiding DE, Klessig DF (1990) Light-mediated control of translational initiation of ribulose-l, 5-bisphosphate carboxylase in amaranth cotyledons. Plant Cell 2: 795-803 Bolle C, Herrmann RG, Oelmuller R (1996) Different sequences for 5'-untranslated leaders of nuclear genes for plastid proteins affect the expression of the beta- glucuronidase gene. Plant Mol Biol 32: 861-8 Bolle C, Sopory S, Lubberstedt T, Herrmann RG, Oelmuller R (1994) Segments encoding 5'-untranslated leaders of genes for thylakoid proteins contain cis-elements essential for transcription. Plant J 6: 513-523 Browse J, Roughan PG, Slack CR (1981) Light control of fatty acid synthesis and diurnal fluctuations of fatty acid composition in leaves. Biochem J 196:347-54 Brun T, Roche E, Kim KH, Prentki M (1993) Glucose regulates acetyl-COA carboxylase gene expression in a pancreatic beta-cell line (INS-1). J Biol Chem 268: 18905-11 88 Carman GM, Henry SA (1999) Phospholipid biosynthesis in the yeast Saccharomyces cerevisiae and interrelationship with other metabolic processes. Prog Lipid Res 38: 361- 399 Caspar T, Quail PH (1993) Promoter and leader regions involved in the expression of the Arabidopsis ferredoxin A gene. Plant J 3: 161-74 Chory J, Susek RE (1994) Light signal transduction and the control of seedling development. In E Meyerowitz, C Somerville, eds, Arabidopsis, Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY, 579-614 Danon A, Mayfield SP (1991) Light regulated translational activators: identification of chloroplast gene Specific mRNA binding proteins. EMBO J 10: 3993-4001 de Jager SM, Menges M, Bauer UM, Murray JAH (2001) Arabidopsis E2F1 binds a sequence present in the promoter of S phase regulated gene AtCDC6 and is member of a multigene family with differential activities. Plant Mol Biol, 47: 555-568. de Silva J, Loader NM, Jarman C, Windust JHC, Hughes SG, Safford R (1990) The isolation and sequence analysis ot two seed-expressed acyl carrier protein genes from Brassica napus. Plant Mol Biol, 14: 537-548. Dickey LF, Petracek ME, Nguyen TT, Hansen ER, Thompson WF (1998) Light regulation of Fed-1 mRNA requires an element in the 5' untranslated region and correlates with differential polyribosome association. Plant Cell 3: 475-84 Dixon M, Webb EC (1979) Enzymes (London: Longman): 248-251 Eccleston VS, Ohlrogge JB (1998) 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 10: 613- 22 Fujiki Y, Ito M, Nishida I, Watanabe A (2000) Multiple signalling pathways in gene expression during sugar starvation. Pharmacological analysis of din gene expression in suspension—cultured cells of Arabidopsis. Plant Physiol 124: 1139-1147 Gallie DR, Walbot V (1992) Identification of the motifs within the tobacco mosaic-virus 5'-leader responsible for enhancing translation. Nucleic Acids Res 20: 4631-4638 Giles AB, Grierson D and Smith H (1977) In vitro translation of messenger RNA from developing bean leaves. Evidence for the existence of stored messenger RNA and its light induced mobilization into polyribosomes. Planta 136: 31-36 Graham IA, Denby KJ, Leaver CJ (1994) Carbon catabolite repression regulates glyoxylate cycle gene-expression in cucumber. Plant Cell 6: 761-772 89 Hajdukiewicz P, Svab Z and Maliga P ( 1994) The small, versatile pPZP family of Agrobacterium binary vectors for plant transformation. Plant Mol Biol 25: 989-994 Hannapel DJ, Ohlrogge JB (1988) Regulation of acyl carrier protein messenger RNA levels during seed and leaf development. Plant Physiol 86: 1174-1178 Hattori T, Nakagawa S, Nakamura K (1990) High level expression of tuberous root storage protein genes of sweet potato in stems of plantlets grown in vitro on sucrose medium. Plant Mol Biol 14: 595-604 Hlousek-Radojcic A, Post-Beittenmiller D, Ohlrogge JB (1992) Expression of constitutive and tissue Specific acyl carrier protein isoforms in Arabidopsis. Plant Physiol 98: 206-214 Hunter SC, Ohlrogge JB (1998) Regulation of spinach chloroplast acetyl—COA carboxylase. Arch Biochem Biophys 359: 170-8 Jiang P, Cronan JE (1994) Inhibition of fatty acid synthesis in Eschericia coli in the absence of phospholipid synthesis and release of inhibition by thioesterase action. J Bacteriol 176: 2814-2821 Joshi CP (1987) An inspection of the domain between putative TATA box and translation start site in 79 plant genes. Nucleic Acids Res 15: 6643-53 Joshi CP, Zhou H, Huang XQ, Chiang VL (1997) Context sequences of translation initiation codon in plants. Plant Mol Biol 35: 993-1001 Kasemir H (1979) Control of chloroplast formation by light. Cell Biol Int Rep 3: 197- 214 Ke J, Wen TN, Nikolau BJ, Wurtele ES (2000) Coordinate regulation of the nuclear and plastidic genes coding for the subunits of the heteromeric acetyl-coenzyme A carboxylase. Plant Physiol 122: 1057-71 Koch KE (1996) Carbohydrate-modulated gene expression in plants. Annu Rev Plant Physiol Plant Mol Biol 47: 509-540 Koncz C, Schell J (1986) The promoter of TL-DNA gene 5 controls the tissue-specific expression of chimeric genes carried by a novel type of Agrobacterium binary vector. Mol Gen Genet 204: 383-396 Lamppa G, Collen J (1991) Analysis of two linked genes coding for the acyl carrier protein (ACP) from Arabidopsis thaliana (columbia). Plant Mol Biol 16: 469-474 90 Lukaszewicz M, Jerouville B, Boutry M (1998) Signs of translational regulation within the transcript leader of a plant plasma membrane H+-ATPase gene. Plant J 14: 413-423 Mekhedov S, Martinez de Ilarduya O, Ohlrogge J (2000) Toward a functional catalog of the plant genome. A survey of genes for lipid biosynthesis. Plant Physiol 122: 389-401 Menges M, Hennig L, Gruissem W, Murray JA (2002) Cell cycle-regulated gene expression in Arabidopsis. J Biol Chem, 277: 41987-2002. Mosinger E and Schopfer P (1983) Polysome assembly and RNA synthesis during pytochrome mediated photomorphogenesis in mustard cotyledons. Planta 158: 501-511 Nakamura K, Ohto MA, Yoshida N, Nakamura K (1991) Sucrose induced accumulation of B-amylase occurs concomitant with the accumulation of starch and sporamin in leaf petiole cuttings of sweet potato. Plant Physiol 96: 902-909 Nishiuchi T, Nakamura T, Abe T, Kodama H, Nishimura M, Iba K (1995) Tissue- specific and light responsive regulation of the promoter region of Arabidopsis thaliana chloroplast w-3 fatty acid desaturase gene (FAD7). Plant Mol Biol 29: 599-609 Ohlrogge J, Browse J (1995) Lipid biosynthesis. Plant Cell 7: 957-70 Ohlrogge JB, Browse J, Somerville CR (1991) The genetics of plant lipids. Biochim Biophys Acta 1082: 1-26 Ohlrogge JB, Jaworski JG (1997) Regulation of fatty acid synthesis. Annu Rev Plant Physiol Plant Mol Biol 48: 109-136 Petracek ME, Dickey LF, Huber SC, Thompson WF (1997) Light-regulated changes in abundance and polyribosome association of ferredoxin mRNA are dependent on photosynthesis. Plant Cell 12: 2291-300 Post-Beittenmiller MA, Hlousek-Radojcic A, Ohlrogge JB (1989) DNA sequence of a genomic clone encoding an Arabidopsis acyl carrier protein (ACP). Nucleic Acid Res 17: 1777 Sambrook J, Fritsch EF, Maniatis T (1989) Molecular Cloning, a Laboratory Manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY Sasaki Y, Kozaki A, Hatano M (1997) Link between light and fatty acid synthesis: thioredoxin-linked reductive activation of plastidic acetyl-COA carboxylase. Proc Natl Acad Sci 94: 11096-101 Savage LJ, Ohlrogge JB (1999) Phosphorylation of pea chloroplast acetyl-COA carboxylase. Plant J 18: 521-527 91 Scherer DE, Knauf VC (1987) Isolation of a cDNA clone for the acyl carrier protein-I of spinach. Plant Mol Biol 9: 127-134 Schneider F, Cassagne C (1995) Specific inhibition of plant fatty acid elongation by a long-chain cerulenin analogue. Eur J Biochem, 228: 704-9. Schujman GE, Paoletti L, Grossman AD, de Mendoza D (2003) FapR, a bacterial transcription factor involved in global regulation of membrane lipid biosynthesis. Dev Cell, 4: 663-72. Schuller HJ, Hahn A, Troster F, Schutz A, Schweizer E (1992) Coordinate genetic control of yeast fatty acid synthase genes FASl and FASZ by an upstream activation site common to genes involved in membrane lipid biosynthesis. EMBO J 11: 107-114 Sheen J (1990) Metabolic repression of transcription in higher plants. Plant Cell, 10:1027-38. Shintani DK (1996) How plants manage their fatty assets: A study into the organization and regulation of the plant fatty acid biosynthetic pathway. PhD thesis. Michigan State University, East Lansing Somerville C, Browse J, Jaworski JG, Ohlrogge J (2000) Lipids. In B Buchanan, W Gruissem, R Jones, eds, American Society of Plant Physiologists, Rockville, Maryland, 456-527 Staub JM, Maliga P (1994) Translation of psbA mRNA is regulated by light via the 5'- untranslated region in tobacco plastids. Plant J 4: 547-53 Stevens R, Mariconti L, Rossignol P, Perennes C, Cella R, Bergounioux C (2002) Two E2F sites in the Arabidopsis MCM3 promoter have different roles in cell cycle activation and meristematic expression. J Biol Chem, 277: 32978-84. Subrahmanyam S, Chirala S (1992) Coordinated regulation and inositol-mediated and fatty acid mediated repression of fatty acid synthase genes in Saccharomyces cerevisiae. Proc Natl Acad Sci, 89: 10232-10236. Susek RE, Ausubel FM, Chory J (1993) Signal transduction mutants of Arabidopsis uncouple nuclear CAB and RBCS gene expression from chloroplast development. Cell 74: 787-799 Terzaghi WB, Cashmore AR (1995) Light-regulated transcription. Annu Rev Plant Phys 46: 445-474 Voelker TA, Worrell AC, Anderson L, Bleibaum J, Fan C, Hawkins DJ, Radke SE, Davies HM (1992) Fatty acid biosynthesis redirected to medium chains in transgenic oilseed plants. Science, 257: 72-4. 92 Weber N, Taylor DC, Underhill EW (1992) Biosynthesis of storage lipids in plant cell and embryo cultures. Adv Biochem Eng Biotechnol 45: 99-131 Weiflhaar B, Block A, Armstrong GA, Herrmann A, Schulze-Lefert P, Hahlbrock K (1991) Regulatory elements required for light-mediated expression of the petroselinum crispum chalcone synthase gene. Syrnp Soc Exp Biol 45: 191-210 93 CHAP TER 3 DISRUPTION OF THE FA TB GENE IN ARABIDOPSIS DEMONSTRATES AN ESSENTIAL ROLE OF SATURATED FATTY ACIDS IN PLANT GROWTH ABSTRACT Acyl-ACP thioesterases determine the amount and type of fatty acids that are exported from the plastids. In order to better understand the role of the FATB class of acyl-ACP thioesterases we identified an Arabidopsis mutant with a T-DNA insertion in the FA TB gene. Palmitate (16:0) content of glycerolipids of the mutant was reduced 42% in leaves, 56% in flowers, 48% in roots and 56% in seeds. In addition, stereate (18:0) was reduced by 50% in leaves and 30% in seeds. The grth rate was reduced in the mutant resulting in 50 % less fresh weight at 4 weeks compared to wild-type plants. Furthermore, mutant plants produced seeds with low viability and altered morphology. Analysis of individual glycerolipids revealed that the fatty acid composition of prokaryotic plastid lipids was largely unaltered whereas the impact on eukaryotic lipids varied but was particularly severe for phosphatidylcholine (PC) with more than a 4-fold reduction of 16:0 and 10- fold of 18:0 levels. The total wax load of fatb-k0 plants was reduced 20 % in leaves and 50% in stems, implicating FATB in the supply of saturated fatty acids for wax biosynthesis. Analysis of C13 Sphingoid bases derived from 16:0 indicated that, despite a 50 % reduction in exported 16:0, the mutant cells maintained wild-type levels of Sphingoid bases presumably at the expense of other cell components. Cutin composition of the mutant was reduced by more than 85 % in CH, derivatives, however its total cutin load was unaltered compared to wild type. The growth retardation caused by the fatb 94 mutation is enhanced in a fatb-k0 act] double mutant where saturated fatty acid content is further reduced. Together, the results demonstrate the in vivo role of FATB as a major determinant of saturated fatty acid synthesis and the essential role of saturates for the biosynthesis and/or regulation of cellular components critical for plant growth and seed development. INTRODUCTION In plants, de novo fatty acid synthesis in plastids can be terminated either by the action of plastidial acyltransferases that transfer the acyl group of acyl-ACP to produce glycerolipids within the plastid (prokaryotic pathway) or by acyl-ACP thioesterases (FAT) that release free fatty acids and ACP. After export from the plastid, free fatty acids are re-esterified to coenzyme-A (COA) to form the cytosolic acyl-COA pool, which is primarily used for glycerolipid biosynthesis at the endoplasmic reticulum (ER) (eukaryotic pathway) (Figure 4) (Browse and Somerville, 1991). In Arabidopsis leaves, 18:1 and 16:0 are the major products of plastid fatty acid synthesis and approximately 60 % of these products are exported to the cytosol as free fatty acids. In other tissues or plant Species, flux through the acyl-ACP thioesterase to the eukaryotic pathway is more predominant with 90 % or greater contribution. Therefore, thioesterases play an essential role in the partitioning of de novo synthesized fatty acids between the prokaryotic and eukaryotic pathways. Moreover, thioesterase substrate specificity determines the chain length and saturation level of fatty acids exported from the plastid (Pollard et al., 1991). Based on 95 Figure 18 Schematic representation of the enzymatic reaction catalyzed by acyl- ACP thioesteases (FAT). Acyl-ACP thioesterases or FAT enzymes (fatty acid thioesterase) hydrolyze the thioester bond between acyl canier proteins (ACP) and fatty acids to release free ACP and free fatty acids inside the plastid. Free fatty acids are then exported as acyl-COA molecules. In most plants the are two classes of FAT enzymes, A and B. The FATA class has higher in vitro specificity for unsaturated fatty acids (18:1) whereas FATB has more activity towards saturated fatty acids (14:0, 16:0 and 18:0). The X axis of the histograms represent the fatty acid attached to ACP to form the acyl-ACP substrate. 96 3: E: 3: a: i/ a \ \ililli / 3 anon—>0 gums—n— @ / 3 8 4 a "8:4 2: _« .5 «a. 33:8 2.3 c: m ._. 0.4). However, since approximately half of the homozygous fatb-k0 plants were lost during germination (see below) an expected ratio of 2.5:1 (resistantzsusceptible) better fit the observed ratio (x2 = 0.18, P > 0.6). One hundred and ten individuals of these 280 resistant plants were grown to full maturity and of these, 25 had a slow-growth phenotype (see below). Again, the observed phenotypic segregation ratio agreed with the expected 2:0.5 ratio (x2 = 0.2, P > 0.6) of segregation for a single T-DNA insertion when considering the lower germination rate of the mutant. A subset of the 110 BASTA resistant plants were randomly selected and subjected to PCR and GC analysis to determine the genotype and the fatty acid composition, respectively. All plants with wild-type visual phenotype and fatty acid composition were heterozygous for the FA TB T-DNA insertion, whereas plants with mutant visual phenotype and fatty acid composition were homozygous for the same insertion (data not shown). The confirmation that the visual phenotype and the genetic (T-DNA) and the biochemical (fatty acid composition) markers all co-segregated suggested that the T- DNA insertion in the FA TB gene was responsible for both the observed growth and fatty acid phenotypes. In order to rule out the possibility that a second-site mutation closely linked to the FA T B T-DNA insertion was responsible for the phenotype, the wild-type FATB cDNA was inserted under the control of the constitutive CaMV35S promoter and expressed in homozygous fatb-k0 plants. Transgenic lines resistant to both hygromycin-B (transgene T-DNA) and BASTA (knockout T-DNA) were indistinguishable from wild- type and showed normal growth and biochemical characteristics (see below). Therefore, 101 J L-202 (__ T-DNA (5.5 Kb) LB ‘ 5’.. cttgcaccctga attatatgctc.. .3’ Figure 19. Structure of the Arabidopsis FA TB gene carrying the T-DNA insertion. A, The FA TB gene is composed of 5 exons and the T-DNA is located in the second intron. The arrows above the gene scheme represent the primers used for the initial screening of thefatb-knockout plants (F w: forward, Rv: reverse, JL-202: left border of T- DNA). The arrowheads underneath exons 2 and 3 represent the primers used for mRNA quantification by real-time PCR. B, Sequence of a portion of intron 2 encompassing the Site of integration of the T-DNA (arrow) we conclude that disruption of the FA TB gene is responsible for the phenotypes Observed in the mutant plants. FATB mRN A Expression Analysis The T-DNA was located in the second intron of the FA TB gene and therefore it was possible that the cells could correctly Splice out at least a fraction of the precursor FATB RNA to yield mature mRNA. Indeed, by using RT-PCR with a set Of primers that spanned the second intron of the FA TB gene (see Figure 19), small amounts of correctly 102 spliced mRNA were detected (data not shown). To determine the extent of gene disruption, the F ATB transcript levels were quantified by real-time PCR in wild-type and mutant leaf tissue and found to be more than ISO-fold lower in mutant than in Wild-type (Table 6). Therefore, although PCR could detect correctly spliced FATB mRNA, this transcript represented less than 0.7 % of wild-type levels. Furthermore, western blot analysis developed with anti-Arabidopsis FATB antibodies did not detect FATB protein in the insertion mutant (data not shown). These results indicated that the T-DNA insertion generated an essentially complete knock-out mutant. Table 6 Relative amounts of FATB mRNA in leaf tissue of wild-type Arabidopsis and fatb-k0 mutant FATB mRNA EIF4A1* mRNA ratio FATB/eIF4A] C, value 2Ct Ct value 2Cf WT (Ws) 23.5 i 0.4 1.2 107 21.4 a 0.36 2.8 106 0.24 fatb-k0 (het) 24.3 3: 0.1 1.0 107 21.2 i 0.15 2.3 106 0.23 fatb-k0 (horn) 30.4 a. 0.3 1.4 10" 21.2 a 0.18 2.4 106 0.0016 *The eukaryotic protein synthesis initiation factor A1 (eIF4Al) mRNA was used as an internal control since the levels of this transcript did not differ in leaf tissue of wild-type and fatb-k0 plants. Het and Hom: heterozygous and homozygous for T-DNA insertion respectively. FATB is Essential for Normal Seedling Growth The first visual characteristic offatb-knockout plants was their Size compared to wild- type (Figure 20A and B). The rosettes offatb-ko plants were approximately half the diameter of wild-type rosettes during the first weeks of growth at 22°C. In addition, the bolting time was delayed in the mutant. More than 90% of the wild-type plants bolted after 4 weeks under our growing conditions, whereas development was delayed in fatb-k0 such that only after the Sixth week did more than 90% offatb-ko plants bolt (Table 7). 103 The morphology of the different organs from mutant plants was unchanged compared to wild-type. However the stems of the mutant elongated more slowly than wild-type stems. AS shown in Figure 21A and B, a plot of the fresh weight of the aerial parts of wild-type and fatb-k0 mutant plants indicated that during the first 4 weeks after germination the plants grew at a constant rate. However, the rate was slower for the mutant. Results in Figure 21B (log scale) indicated that wild-type plants increased their fresh weight 10.6- fold (i 0.4) per week, whereas for the fatb-k0 plants the increase was only 8.8-fold (i 0.6). This 17 % reduction in growth rate led to a reduction of more than 50 % in the fresh weight of the mutant after 4 weeks (Figure 20A). Growth of both wild-type and mutant plants slowed after the fourth week, but more so for the former, such that by the sixth week the fatb-knockout plants differed in size and fresh weight by about 25% (Figure 2 1 A). During the growing phase the percent ratio of dry to fresh weight remained at ~9 % while during the drying period it increased to ~14 % for both wild-type and mutant plants. Taken together, the similar morphology but different growth rates suggested that differences in wild-type and mutant plants were the consequence of a reduced grth rate and not altered development of the mutant. Wild-type and fatb-k0 plants were grown in the presence of l % sucrose on either plates or liquid culture to determine if the normal growth rate could be recovered. Sucrose availability did not eliminate the slower growth rate of the mutant (data not shown) suggesting that photosynthetic capacity or carbon limitations of the fatb-knockout plants were not causes of the reduced growth rate. 104 Figure 20. Growth and morphology of Arabidopsis wild-type and fatb-k0 plants and seeds. A, Four weeks old wild-type (left) and fatb-k0 (right) plants. B, Two weeks Old wild-type (left) and fatb-k0 (right) plants. C, wild-type Arabidopsis seed. D, E and F, wild-type-like, intermediate deformed and very deformed seeds from fatb-k0 plants respectively. 105 Plant grth rate is modified by temperature and part of this effect may be associated with variations in the physical properties of cell membranes. To test the possibility that the growth retardation of the fatb-k0 mutant was a function of temperature, wild-type and fatb-k0 plants were grown for 2 weeks at 22°C and then transferred to three different temperatures (16°C, 22°C and 36°C). The fatb-k0 plants Showed the same percent reduction (~50 %) of fresh weight per seedling compared to wild-type at the three different temperatures. Therefore, the slower relative growth of the mutant plants was not altered within the range of temperatures used. In addition we tested wether adding exogenous saturated fatty acids by either spraying plants or supplementing seedlings grown in liquid culture could overcome the slower growth of the fatb-k0 mutant. These procedures were not sufficient to chemically complement the fatb-k0 phenotype. In fact, addition of higher amounts of exogenous fatty acids showed deleterious effects on plant growth (data not Shown). Leaf and chloroplast morphology of wild type Arabidopsis and fab-k0 The leaves offatb-ko average 50 % of the area and fresh weight of wild type leaves (Figure 20 and data not shown). These differences may be brought about by reduced cell Size or less number of cells per leaf. In addition, leaves are composed of different cell types (e.g., mesophyll, epidermal) and some of those might be differentially affected by the genetic lesion. We analyzed optical sections of leaf tissue from wild type and mutant plants by laser confocal microscopy (LCM) and leaf epidermal cells by scanning electron microscopy (SEM) (Figure 22A to D). The structures of leaves and epidermis together 106 if A :1 3-5; Wild type '6 l s 31 m E 2.5.. a 2 l "' 1 51 ~ on ‘1 . fatb-k0 0.5' O " ' r +—#+-—-~*—+~*l 4' t t ~ws~r—+—~t 1 2 3 4 5 6 7 8 weeks after germination 10 Wild type B 0.1 0.01 grams/seedling (log scale) 0.001 weeks after germination Figure 21. Growth curves of Arabidopsis wild-type and fatb-k0 plants. A, The fresh weight of aerial parts of Arabidopsis plants was measured on a weekly basis for a period of 8 weeks. Each time point is the average value of at least 7 individual plants. B, Logarithm of the fresh weight versus time. with Sizes of epidermal and mesophyll (palisade and spongy) cells were similar between both plant classes (Figure 22A to D). Thus, these observations indicated that the smaller size offatb-ko leaves compared to wild type was the result of fewer cells per leaf rather than smaller cell Sizes. These results contrasted with those Obtained with the Arabidopsis fab2 mutant, in which decrease in leaf size was the consequence of reduced 107 Figure 22. Microscopic analysis of leaf section and chloroplast structure of wild type and fatb-k0 seedlings. A—B, Scanning electron micrographs of wild type (A) and fatb-k0 (B) leaf epidermal cells. C-D, leaf sections of wild type (C) and fatb-k0 (D). E-F, Transmission electron microscopy (TEM) of wild type (E) and fatb-k0 (F) leaf chloroplasts. G-H, TEM of thylakoid membranes from wild type (G) and fatb-k0 (H) leaf chloroplasts. 108 expansion of leaf cells (Lightner et al., 1994). Growth of fatb-k0 in the presence of sucrose did not revert its growth phenotype, suggesting that photosynthetic capabilities of the mutant did not differ substantially from wild type plants. Analysis of leaf chloroplast ultrastructure by transmission electron microscopy (TEM) Showed no obvious differences between organelles of wild type and fatb-k0 (Figure 22E and F). The size and number of chloroplasts per cell of the mutant were Similar to wild type. Moreover, structure and number of thylakoid membranes were also Similar between the two plant classes (Figure 22G and H). In summary, these observations agreed with similar photosynthetic capabilities between wild type and mutant plants. FATB is Essential for Normal Seed Morphology and Germination Germination of seeds produced by fatb—k0 plants was reduced by approximately 50 % either on soil or 1 % sucrose (Table 7). Close examination of mature seeds produced by the fatb-k0 mutant revealed a continuous range of deformity in seed morphology, with wild-type-like seeds in one extreme to very deformed seeds (approximate fi'equency of 20 %) in the other (Figure 20C-F and Table 7). The germination rate of very deformed seeds was only 16 %. These observations suggested that some stage of seed or embryo development may be substantially affected in the mutant. By analyzing developing siliques it was not evident that deformed seeds were located in Specific segments of this organ. Upon surface sterilization, some mutant seeds also lost the seed coat, suggesting 109 Table 7. Arabidopsis wild-type and fatb-k0 bolting time and germination rates. Bolting time* (%) Time WT (WS) fatb-k0 4th week 92.3 27.6 2 5‘h week 7.7 72.4 Germination rate (%) WT (WS) fatb-k0 Soil 96.7 45 .6 1% sucrose 94.4 45.8 Deformed seeds in total seeds (%) WT (WS) fatb-k0 0.8 21.7 Germination on soil of deformed seeds (%) fatb-k0 16 * 18:6 h lightzdark; 22 C that the structure of this tissue could be altered. Scanning electron microscopy (SEM) analysis of the seed coat from mutant plants with wild-type-like or intermediate morphology did not Show any obvious structural differences compared to wild-type seeds (Figure 20C-F). Many deformed seeds from the mutant displayed a shriveled seed coat (Figure 20F). Fatty Acid Composition of fatb-k0 Tissues. 110 _-.I_IE| 1.3;..1‘ AS indicated in Table 8, palmitate (16:0) in homozygous fatb-k0 plants was reduced 42 % in leaves, 56% in flowers, 48% in roots and 56% in seeds compared to wild-type. Stearate (18:0) decreased almost 50 % in leaves and 30 % in seeds, with negligible changes in flowers and roots. The fatb-k0 plants also showed an increase of 150-200 % in oleate (18:1) and 40-60 % in linoleate (18:2) in leaves, flowers and roots. Linolenate (18:3) declined by 15-20 % in leaves, flowers and roots. Seed unsaturated fatty acids were less affected. Together, these results demonstrate the in vivo role of FATB as a major determinant of 16:0 in all the tissues analyzed and also indicate that FATB contributes to the level of 18:0 in leaves and seeds. fatb-k0 plants transformed with the wild-type FATB cDNA under the CaMV35S constitutive promoter had a fatty acid composition very similar to wild type, confirming that the FATB cDNA complemented the biochemical phenotype of the mutant (Table 8). Table 8. Fatty acid composition of wild-type and fatb-k0 Arabidopsis tissues (mol%). 16:0 16: 1(9) 1621(3) 16:3 18:0 18:1(9) 18:2 18:3 Other Leaftissue WT (WS) 17.5 0.6 3.7 11.8 1.5 2.6 14.4 47.7 fatb-k0 10.1 1.3 3.4 12.4 0.8 8.2 23.2 40.1 3SS-FATB** 15.5 0.6 4.0 11.9 1.3 4.6 16.8 45.0 Flowers WT (WS) 27.4 1.7 1.2 2.4 3.2 30.6 31.3 fatb-k0 12.0 4.6 2.6 2.6 8.5 41.9 24.7 Roots WT (WS) 25.3 0.9 6.6 3.6 28.8 21.3 13.2 fatb-k0 13.4 1.1 6.0 9.2 42.5 17.4 10.1 Seeds WT (WS) 8.3 0.3 3.5 13.0 27.5 18.6 28.6 fatb-k0 3.6 0.4 2.4 11.9 33.1 16.2 32.1 * Average of at least 3 samples. Standard deviation values are not shown and represent less than 5% of the average values. ** fatb-k0 plants transformed with CaMV35S-FATB cDNA. Ill Fatty Acid Composition of Individual Leaf Glycerolipids. The fatty acid composition of individual glycerolipids from homozygous fatb-k0 and wild-type leaves are presented in Table 9. Palmitate reductions occurred mainly in extra- plastid lipids. While phosphatidylethanolamine (PE), phosphatidylserine (PS) and phosphatidylinisitol (PI) had an approximately 50 % reduction in 16:0 compared to wild type, in phosphatidylcholine (PC) the reduction was almost 80 %. The palmitate levels in plastid lipids were less affected with a significant reduction (40 %) only in sulfoquinovosyldiacylglycerol (SQDG). All of the extra-plastid glycerolipids except PI had reduced 18:0. Again PC was the most affected with a 10-fold reduction in 18:0. The characteristic changes in unsaturated fatty acids seen in Table 8 for total leaf lipids, namely increases in 18:1 and 18:2, and a decrease in 18:3, were most pronounced for the phospholipids and SQDG. The data in Table 9 also indicated that despite the changes in fatty acid composition there were no major differences in the relative proportions of leaf glycerolipids between wild—type and fatb-k0 plants. And finally, the total amount of fatty acid methyl esters produced by acid-catalyzed transmethylation of Arabidopsis leaves was the same for the wild-type (11.5 umoleS/gfw) and fatb-k0 (11.6 umoles/gfw), indicating that the fatb-k0 did not affect net fatty acid accumulation per fresh weight. AcyI-ACP Thioesterase Activity To determine if any compensatory changes occurred in acyl-ACP thioesterase activity, mutant and wild-type plants were assayed for hydrolysis of 18:1-ACP and 16:0-ACP. The FA TA gene product has an acyl specificity 18:1 >> 18:0 >> 16:0 while the FA TB 112 Table 9. Fatty acid composition of leaf glycerolipids of Arabidopsis wild-type (WS) and fatb-k0 mutant (mol%). % total 16:0 1621(3) 16:3 18:0 18:1(9) 18:1(11) 18:2 18:3 PC WT (WS) 11.2 i 0.3 21.1 2.4 6.1 - 34.7 35.5 fatb-k0 13.6 e 0.7 4.5 0.2 17.8 1.3 46.1 30.1 PE WT (WS) 10.0 i 0.2 29.0 2.1 2.8 - 35.1 30.8 fatb-Ito 7.9 :r 0.6 11.6 1.1 11.3 1.2 50.8 24.9 PS WT (WS) 0.9 i 0.1 30.8 6.8 2.9 27.1 32.3 fatb-k0 0.8 i 0.1 17.2 2.9 10.2 44.3 25.2 PI WT (WS) 5.5 e 0.1 40.0 2.0 2.2 25.3 30.3 fatb-k0 4.4 i 0.1 21.2 1.8 9.2 36.9 29.3 PG WT (WS) 12.0 i 0.3 29.2 22.0 1.1 4.5 - 10.8 32.1 fatb-k0 12.6 i 0.3 24.3 20.7 - 8.5 0.9 16.9 27.6 SQDG WT (WS) 4.3 i 0.1 34.6 1.5 3.6 0.7 21.5 37.8 fatb-k0 4.2 i 0.6 20.4 0.5 10.5 1.2 36.4 30.8 DGDG WT (WS) 20.2 e 0.3 12.8 4.2 1.0 1.1 - 6.7 73.9 fatb-k0 19.4 e 0.1 12.0 5.4 0.5 1.2 0.5 7.9 70.5 MGDG WT (WS) 35.5 e 0.1 1.6 38.1 0.8 3.0 53.4 fatb-k0 36.7 i 0.2 1.3 40.3 1.6 3.2 50.7 gene product has a specificity 16:0 > 18:1 > 18:0 (Doermann et al., 1995; Salas and Ohlrogge, 2002). The activity from the FA TA gene product dominates acyl-ACP thioesterase activity measurements made With crude extracts. OleoyI-ACP hydrolytic 113 activity in leaf extracts of wild-type and mutant plants was similar (~l25 pmol*min' 1"‘mg") and hydrolytic activity on 16:0-ACP was close to background levels and therefore difficult to quantify. These results indicate that measurable acyl-ACP hydrolytic activity does not change in mutant leaf extracts compared to wild-type and therefore endogenous levels of FATA activity are not up-regulated in the mutant. Total Palmitate Content in Arabidopsis Leaf Tissue Most acid or base-catalyzed transmethylation methods for fatty acid analysis efficiently convert O-acyl groups such as found in glycerolipids to fatty acid methyl esters. However, N-acyl groups such as found in Sphingolipids react only very slowly using such methods. In order to evaluate the total palmitic and stearic content (0- and N-linked) of Arabidopsis cells, a strong alkaline hydrolysis on total leaf tissue, on lipids extracted with multiple chloroform methanol extractions, and on the solvent-extracted residue was performed. The total 16:0 content in leaves of wild-type plants was 1.87 i 0.02 umoles/gfw, whereas leaves offatb-ko plants contained 1.14 i 0.03 umoles/gfw. Thus, a 39 % reduction in total 16:0 was observed in the mutant, similar to the 42 % reduction of 16:0 in glycerolipids (Table 8). The amount of total stearic acid was also reduced by 50 % in leaf tissue, from 0.16 i 0.01 umoles/gfw in wild-type plants to 0.075 i 0.005 umole'S/gfw in the mutant. The analysis of extracted lipids and solvent-extracted residue indicated that almost all the 16:0 and 18:0 in the cells was co-extractable from leaf tissue, since the lipid fraction contained similar amounts of saturated fatty acids to the total tissue (data not shown). In contrast, the solvent-extracted residue which may contain acylated proteins and other insoluble lipids, contained 3 % of the total 16:0 and no 114 detectable 18:0. Similar reductions of 16:0 and 18:0 were observed in all the fractions analyzed, indicating that absence of FATB reduced saturated fatty acid levels in both organic soluble and insoluble components. Leaf and Stem Surface Wax Analysis The very long chain fatty acids (VLCFA) required for wax synthesis are produced by elongation of 16 and 18 carbon saturated fatty acids (Post-Beittenmiller, 1996). To determine if the 40-50 % reduction in saturated fatty acids influences total wax load or composition in the fatb-k0 mutant, leaf and stem epicuticular waxes were analyzed. The results in Table 10 indicated that at 5 weeks total wax load per fresh weight in leaves was reduced by 20 % in the fatb-k0 mutant. However, no novel components or substantial changes in the distribution of wax components were observed at this stage. A 20 % reduction in leaf wax load was consistently observed at different stages of plant development (data not Shown). Analysis of primary stems indicated a 50 % reduction in wax load per fresh weight in the fatb-k0 mutant compared to wild type (1.1 i 0.1 umol/gfw versus 2.2 i 0.3 umOl/gfw respectively), again without changes in the distribution of wax components. These data indicate that supply of saturated fatty acids by FATB is one factor limiting wax biosynthesis but that reduction of this supply does not result in the replacement of 16:0 by 18:1 or other precursors for surface wax structures. The role of FATB in supplying wax precursors was more evident in stems than in leaves as the former tissue accumulates higher amounts of epicuticular waxes. Similar tissue specific wax reductions are observed in most of the Arabidopsis wax biosynthetic mutants (eceriferum) in which reductions in stem waxes are larger than in 115 Table 10. Major components of epicuticular leaf waxes from wild—type ArabidJorsiS and fatb-k0 mutant (mol%). (carbon chain length) WT (WS) fatb-k0 Alkanes 27 1.45.0.1 l.2:l:0.05 29 21.93:].2 22.1:E0.l 30 1.23:0] 1.0:t0.03 31 39.4 i 1.1 40.5 :1: 0.5 32 0.9 i 0.1 0.9 i 0.03 33 11.7 i 0.4 12.7 :t 0.03 Free fatty acids 24 3.5 i 0.2 3.3 i 0.06 26 5.7 i 0.2 5.3 i 0.09 28 2.2 31:04 1.8101 30 0.7 d: 0.1 0.5 i 0.1 Primary alcohols 26 3.1 i 0.4 2.3 i 0.3 28 3.6i 0.5 3.63:0.3 30 2.2 i 0.1 2.2 i 0.02 32 2.3 3: 0.1 2.2 :1: 0.02 Total (pg/gm) 104.1 A 6.6 83.9 :1: 4.5 Total (umol/gfw) 0.24 :t 0.01 0.19 i 0.01 leaf waxes (Rashotte et al., 2001). Sphingoid Base Analysis Since sphingoid base and Sphingolipid synthesis are initiated by serine palmitoyltransferase (Lynch, 1993) an analysis of sphingoid bases was conducted. Leaves of wild-type (0.54 i 0.09 umoles/gfw) and fatb-k0 (0.50 i 0.08 umoles/gfw) plants did not differ significantly in the total amount of sphingoid bases. However, 116 differences could be observed in the relative abundance of the individual sphingoid bases (Table 11). Most significantly, trihydroxy-18:0 (118:0) increased by almost 4-fold in the total Sphingoid bases of the fatb-k0 mutant. The total sphingoid base composition was similar to that reported by Sperling et al. (1998): the most abundant sphingoid base in Arabidopsis leaf tissue was t18:1(8E), followed by t18:1(8Z). Monoglucosylceramide (MGC) is considered one of the most abundant Sphingolipids and therefore the sphingoid base composition in leaf MGC was analyzed. The base composition in MGC of wild-type and mutant plants was indistinguishable, and was Similar to that reported by Imai et al (2000) (Table 11). The difference in composition between total and MGC sphingoid base compositions indicates that MGC is not the predominant Sphingolipid in Arabidopsis leaves. As suggested by Imai et al. (2000), complex Sphingolipids such as phosphoinositolceramides could be more abundant than MGC in Arabidopsis leaf tissue. Sphingoid bases in extracted lipids and solvent-extracted residue were also analyzed. The latter fraction can contain sphingoid bases from highly glycosylated phophoinositolceramides and GPI moieties from GPI-anchored proteins. Again, no major changes in sphingoid base composition between Wild-type and fatb-k0 plants were observed except in trihydroxy 18:0 (Table l 1). Leaf and stem cutin analysis The monomers of cutin are synthesized from the COA esters of palmitic acid (16:0) and oleic acid (18:1), respectively. To determine if the reduction in saturated fatty acids in fatb-k0 influences cutin composition, leaf and stem cutin were analyzed. The results shown in Figure 23 indicate that 16 carbon derivatives of cutin were reduced by more 117 Table 11. Sphingoid base content of leaf tissue from wild-type Arabidopsis and fatb-k0 plants (mol%) 118:1(8E) t18:1(8Z) 118:0 d18:1(8E) d18:1(8Z) d18:2(4,8) Total Tissue WT (WS) 60 32 2 4 t 2 fatb-k0 57 32 8 2 l 1 Monoglucosylceramide WT (WS) 24 59 t 11 t 5 fatb-ko 25 59 t 1 0 2 5 Solvent Extracted Lipids WT (WS) 62 33 2 2 1 t fatb-k0 57 36 5 2 t t Solven t-Extracted Residue WT (WS) 85 7 3 t 3 l fatb-k0 76 7 13 t 3 1 Dihydroxy bases: 8-sphingenine (d18:1[8E or Z]), 4,8-sphingadienine (d18:2[4E,EZ or 4E,8Z]); trihydroxy bases: 4-hydroxysphinganine (t18:0) and 4-hydroxy-8-sphingenine (t18:1[8E or 2]). t, traces. than 85 % in fatb-k0 leaves and stems compared to wild type. These results demonstrated that FATB provides 16:0 for cutin biosynthesis in Arabidopsis. Although fatb-k0 has reduced 16 carbon derivatives in cutin the total cutin load was unaffected in the mutant (approximately 5 umol/g residue in leaves and stems of both plant classes). Thus, the mutant compensates for the decrease in CH, derivatives by producing more C18 derivatives (Figure 23). 118 fatb-k0 act] double mutant The lipid analysis demonstrated that despite the absence of FATB in the plastids, mutant plants contain approximately 50 % of the saturated fatty acids found in wild-type plants. What is the origin of these remaining saturated fatty acids? Palmitate and stereate, as 16:0-ACP and 18:0-ACP respectively, may be utilized directly by acyltransferases in the plastid for prokaryotic lipid synthesis (Somerville et al., 2000). In order to evaluate how much of the saturated fatty acids remaining in the fatb-k0 plants derive from acyl group fluxes through plastidial acyltransferases, a cross between fatb-k0 plants and the Arabidopsis act] (atsl) mutant was performed. The act] mutant has reduced plastidial glycerol-3-phosphate:acyl-ACP acyltransferase activity, the first step in the plastid pathway of glycerolipid biosynthesis. Although act] plants contain reduced amounts of 16:3 and higher amounts of 18:1, grth of this mutant is normal (Kunst et al., 1988). However, the fatb-k0 act] double mutant was severely impaired in growth when compared to wild-type,fatb—ko and act] plants (Figure 24). Leaf fatty acid analysis of the double mutant indicated that this tissue contained approximately 3-4 mol% 16:0, representing a 70 % reduction compared to wild type (Table 8 and 12). The levels of 18:1 in leaves were increased in the double mutant compared to fatb-k0 whereas 18:2 and 18:3 levels remained almost unchanged (Table 8 and 12). Interestingly, leaf levels of 18:0 in the double mutant were Similar to those in fatb-k0 leaves, indicating no further decrease in overall 18:0 by the second mutation. Analysis of individual lipid classes from leaf tissue of the double mutant revealed that the C H, acyl composition of extraplastidial glycerolipids and PG did not differ substantially from those in fatb-k0 leaves (data not 119 A Composition of leaf cutin f Hf _. 80 i .WS 1 70 . i 60 Ufatb-ko l 50 - .\° T: 40 E 30 .. 20 . ‘° I. ll 0 I: LLtLj: I I 7 I 7 I I 7 [069 761/0676: ’IKO’Q/OIIKO .90 0,0 760 4); [04) 6); 6’7 7 If ’Q/ o o ‘90 76106-13 ‘97 (5)0 ’39 4), 7 (20 B Composition of stem cutin 80 I“ — . .WS 70 l 60 - Qfatb-ko 50 . ‘ 7 "i’ o\° l a 40 l E l 30 20 A . l -. [1112.6 7 7 I I I 7 7 70%? *5? Yo o’e’e’Q 9 ’2, F0 761/ "a ’59, Yo 70 ’Q, ’0 6.0/302,3 ,2’962 <2 ’6‘. 7 76>. ’29 {O {O .0 .9 , ’59, A», ‘9'; 63.0 Figure 23. Leaf and stem cutin composition of wild type Arabidopsis and fatb-k0 (mol%). Cutin components were reduced with LiAlH4 and analyzed by GLC. A, cutin monomer composition of four-week old leaves from wild type and fatb-k0 plants. B, cutin monomer composition of six-week old stems from wild type and fatb-k0 plants. Bars denote standard deviations. 120 shown). Moreover, the relative abundance of each glycerolipid was similar. These results demonstrated that the second mutation (actI) affected primarily 16:0 accumulation in plastidial glycerolipids with minor effects on extraplastidial lipids compared to fatb-k0. In summary, double mutant plants blocked in both F ATB and ACTI were further reduced in saturated fatty acid content to levels ~ 30 % of wild-type plants. These plants displayed a smaller size and thus this more severe grth phenotype associated with a greater reduction in saturated fatty acids further demonstrates the essential role of saturated fatty acids in maintaining normal rates of plant growth. DISCUSSION Acyl-ACP thioesterases are responsible for the export from the plastid of fatty acids produced by the de novo fatty acid synthesis system. In this study an Arabidopsis insertion mutant of the FA TB gene was isolated,‘and we describe its effects on plant growth and on the production and utilization of saturated fatty acids. In a previous study, antisense of Arabidopsis FATB using the CaMV3SS promoter resulted in a substantial reduction of 16:0 only in flowers and seeds and minimal difference in leaves and roots and did not Show any visual phenotype (Doermann et al., 2000). The lack of anti-sense impact on leaves suggested either FATB was not the major controller of 16:0 in leaves or that the reduction of FATB mRNA was not sufficient to reduce 16:0 levels in certain tissues. The characterization of an Arabidopsis fatb-knockout mutant in this study clarifies that the second interpretation is correct and that FATB is a major control point 121 Figure 24. Growth and morphology of Arabidopsis wild-type,fatb-ko, act] and fatb- ko act] plants. Four weeks Old wild-type (WS) (bottom right), fatb-k0 (bottom left), act] (top right) and fatb-k0 act] double mutant (top left) plants. The size of act] mutants is similar to the size of wild type Columbia (Col) plants (not shown). The size of wild type plants derived from the cross between fatb-k0 (WS) and act] (C01) is intermediate to the parental ecotypes (not shown). for saturated fatty acid fluxes in all tissues. The more extensive biochemical phenotype and the reduced growth rate and seed viability observed in the mutant compared to antisense plants suggest that the FATB enzyme or its mRNA may be in large excess and 122 Table 12. Leaf fatty acid composition of wild type Arabidopsis (Columbia), act] and fatb-k0 act] double mutant (mol%) 16:0 16:1 16:1(3) 16:3 18:0 18:1(9) 18:2 18:3 Wild type 16.2 1.4 3.0 11.3 1.5 3.0 13.8 49.8 (Col) act] 14.3 1.1 3.1 0.4 1.6 11.9 24.1 43.1 fatb-k0 act] 3.1 1.2 3.2 0.5 0.8 23.4 24.8 41.5 difficult to reduce sufficiently by antisense methods to produce a growth phenotype. Seedling Growth and Seed Development A large number of mutants with diverse changes in fatty acid composition have been isolated in Arabidopsis (Wallis and Browse, 2002). Most of these are not readily distinguishable from wild-type plants when grown at standard temperatures (15°C-25°C). One of the few examples in which a mutation affecting fatty acid composition has consequences for plant growth is the fab? mutant in Arabidopsis, in which elevated levels of 18:0 in the membrane lipids results in dwarf plants (Lightner et al., 1994). However, the phenotype can be partially ameliorated by growing the mutant at high temperature, suggesting that membrane fluidity or a related physical property reduces growth of the mutant at 22°C. The Arabidopsis fatb-knockout is the first example of an Arabidopsis mutant with reduced levels of saturated fatty acids where a reduction in vegetative growth at standard growth conditions occurred. Low temperature did not alleviate nor did high temperature 123 exacerbate the Slow growth phenotype offatb-ko plants suggesting that effects other than changes in bulk membrane physical properties were limiting the growth of the mutant. Hence, in contrast with fab2 in which high 18:0 levels may disrupt the proper function of membranes, reduced saturate levels in fatb-k0 plants may be altering the biosynthesis and function of critical cell components. However, we can not rule out that lower amounts of saturates may be associated with more subtle changes in the physical properties of cellular membranes that could affect functions such as transport or vesicle formation. The fatb-k0 mutant with a ca. 50% reduction in palmitate also contrasts with the fab] mutant that is increased approximately 50% in palmitate but displays normal growth (Wu et al., 1994). The fatb-k0 mutant is also distinguished from other fatty acid mutants in its effect on seed development and germination (Figure 20C-F and Table 7). However, in contrast to the Slow growth phenotype that occurred in all seedlings, the penetrance of the seed phenotype was incomplete. At this stage it is not clear whether the seed defects are a consequence of alterations during specific seed developmental phases, or perhaps an indirect effect due to deficiencies in supply of nutrients from maternal tissues. Palmitoyl-ACP pools in the plastid are subject to three major reactions; acyltransfer to glycerol, elongation to 18:0-ACP or hydrolysis by FATB. Mutations that block or reduce all three of these fates are now available. In the fab] mutant (Wu et al., 1994), reduction in 16:0-ACP elongation results in increased 16:0 in both the prokaryotic and eukaryotic lipids suggesting that flux into both these pathways can be increased by increased availability of 16:0-ACP. Moreover, in both fatb-k0 and fab] leaves there is an increase in 16:1(9) levels, suggesting that 16:0-ACP pools are increased within chloroplasts of 124 these two mutants. In contrast, in the act] mutant the loss of the acyltransferase pathway results in increased 16:0-ACP elongation to 18:0 rather than increased flux to the eukaryotic path via the FATB thioesterase (Kunst et al., 1989). Similarly, in the fatb-k0, the reduction in flux via the thioesterase also primarily increased elongation to C18 rather than increased flux of C K, into prokaryotic lipids. These contrasting responses suggest that the elongation rate of 16:0-ACP is regulated primarily by availability of substrate but that the contributions of the FATB and acyltransferase reactions to 16:0 flux likely have additional levels of control. Reduced Export of Palmitate in fatb-k0 plants and Other Sources of Palmitate and Stearate in the Cell Lipid analysis (Table 8) demonstrated that despite the homozygous fatb-k0 insertion, mutant plants still produce approximately 50 % of the palmitate found in wild-type plants. How and where is the remaining palmitate balance in a plant cell produced? A Similar statement and question can be made for stearate. Palmitate is both an intermediate and an end product of de novo fatty acid synthesis in the plastid. Palmitate, as 16:0-ACP, may be utilized directly by acyltransferases in the plastid for prokaryotic lipid synthesis, in particular by the lyso-phosphatidic acid sn-2 acyltransferase (Frentzen et al., 1983). Alternatively, after hydrolysis by acyl-ACP thioesterases, free fatty acids including palmitic acid may be exported from the plastid, a proposed mechanism now substantiated by in vivo labeling (Pollard and Ohlrogge, 1999). And finally, plant mitochondria have the capacity for de novo synthesis of fatty acids (Wada et al., 1997) and, although considered a minor pathway this organelle could partially compensate for low 16:0 levels in the fatb mutant. Could (I) the transfer of palmitate from prokaryotic lipids to 125 eukaryotic lipids, (2) the FATA acyl-ACP thioesterase, and/or (3) mitochondrial fatty acid synthesis account for the remaining exported palmitate production? The cross offatb-ko with act] plants demonstrates that the prokaryotic pathway provides about 60 % of the saturated fatty acids in leaves offatb-ko. Approximately half of the saturates that are still produced in the double mutant can be attributed to plastidial PG (produced with prokaryotic character by an unknown path) and to FATA activity. The Arabidopsis FATA-encoded thioesterase has a small but measurable in vitro activity towards 16:0 and 18:0-ACP (about 2 % and 16 % of the activity towards 18:1-ACP, respectively) and our results suggest that these enzymes have in vivo hydrolytic activity towards 16:0- and 18:0-ACP, a conclusion also drawn by Nadev et al. (1992). The mutants studied here demonstrate that FATA, mitochondrial FAS or other sources of 16:0 are minor contributors to leaf saturated fatty acid flux in comparison to FATB and the plastid acyltransferases. As summarized in Figure 25, the results of this study allow a better estimate of the relative contributions of alternative pathways for saturated fatty acid supply in plants. In wild-type plants, the total C16 fatty acids incorporated into membrane glycerolipids is 33.3 units, which includes 16:1(3)-trans in PG and 16:2 and 16:3 in MGDG. About 23 units are used for prokaryotic lipid synthesis, assuming that all PG and MGDG are derived from this pathway, and that the proportion of 16:0-containing DGDG that is prokaryotic is one third, based on its sn-2 distribution (Kunst et al., 1989). Assuming that half the SQDG is of prokaryotic origin, of the remaining 10.3 units which are exported 126 from the plastid, about 2.5 units return to this organelle as SQDG and DGDG while the remaining 7.8 units are used for phospholipid synthesis (Figure 25). In the fatb-k0 line we do not know what proportion of the 16:0-containing DGDG and SQDG pools are now of prokaryotic origin, so estimates are ranges. About 2.6-4.2 units of palmitate (25-45 % of wild-type) are exported and 2.6 units (33 % of wild-type) are retained for phospholipid synthesis while 0-1.6 units (0-65 % of wild type) are returned for plastid lipid synthesis. Palmitate used for prokaryotic lipid synthesis is barely affected by the mutation, increasing from 23 to 23.6-25.6 units. In the fatb-k0 act] double mutant, about 3.5 units of palmitate (35 % of wild type) are exported and 2.5 units (33 % of wild type) are retained for phospholipid synthesis while 0-1 units (0-40 % wild type) are returned to the plastid. Partition of Palmitate and Stearate to Non-glycerolipid Products. Palmitate constitutes the primary saturated fatty acid exported from plastids and incorporated into membrane glycerolipids. In addition to glycerolipids, several other cellular components derived from the exported palmitate and/or stearate have essential structural and perhaps signaling roles for cell growth. First, sphingoid bases in plants are 18 carbon amino alcohols that are synthesized outside the plastid from palmitoyl-COA and serine by the action of serine palmitoyltransferases (Lynch, 1993). In addition, 16:0 127 Figure 25. Simplified scheme of predicted fluxes of C16 and C13 fatty acids in membrane leaf glycerolipids of Arabidopsis wild type, fatb-k0 and fatb-k0 act] mutants. The numbers in the figure represent average values of mol% units of C16 and C13 fatty acids accumulated in membrane glycerolipids. A range of values is given when alternative biosynthetic pathways are considered (see text for details). The average mol% units were calculated based on the mol% abundance of a particular fatty acid in a Specific lipid, the relative abundance of the lipid and the contribution of each pathway (prokaryotic and eukaryotic) to the synthesis of that lipid (Browse et al., 1986) (see Discussion for more details). Arrows represent the flux of acyl molecules through FATB and FATA (acyl-ACP thioesterases B and A respectively), KAS-II (3-ketoacyl-ACP synthase-II) and LPAAT (lyso-phosphatidate sn-2 acyltransferase) that are either used for lipid biosynthesis in the plastid (prokaryotic pathway) or exported from the same organelle. ACTl (glycerol-3-phosphate acyltransferase) transfers 18:1 to glycerol-3-P in the prokaryotic pathway. Question marks indicate that: l, the actual flux of 16:0 through FATA in wild type leaves is not known and an upper limit of 4 mol% units can be estimated based on fatb-k0 data (however, flux through FATA in fatb-k0 could be a compensatory mechanism); and 2, alternative pathways for biosynthesis of PG in the fatb-k0 act] double mutant can be considered. 128 \ :66 3.88 J \ Baas / N 2525 j x .. J / x 9 \ nu<93 54.93 J H253 __-m(v_ 4 94m .243 Ba 4 BB .323 :95. 4 Ba 5% _ I _ - I _ - I _ _ 9n 3. .2 3.2 2. I. n m... S ... a: c « 8E5 \ 8.536 3. .SE \ . . .. Tm \ / 3.: Inn / 2 n: . 111 l 9.1. 1 1. J u . . m m J I . . B D At. . u a )3 ;. S m At lhIIIII. 1 «F111 11 .Vb. T1 1k IL Mr. . .II M.” c In. .hHm . II 0 .\ m C In . . In . D n1. 01 n . in _ _ o a 1 1 1w. n Inc IR. 1. a e mi la 10. i x s . 011. PL 10 129 Figure 25 or its hydroxylated derivatives can be N-linked to sphingoid bases to form ceramides and sphingolipids. Second, in epidermal cells, saturates (16:0 and 18:0) are precursors of the cutin/suberin monomers and wax components. (Post-Beittenmiller,1996). Third, myristoylation (14:0) and palmitoylation (16:0) of proteins is critical for the localization and regulation of protein activity (Yalovsky et al., 1999). In wild-type leaves the total 16:0 content is about 1.9 umoles/gfw. Deducting the contribution from prokaryotic lipid synthesis, about 1.2 umoles/gfw of 16:0 must be exported from the plastid to fuel glycerolipid synthesis. Sphingolipid base synthesis requires another 0.5 umoles/gfw of palmitate export. In addition, there are significant levels of 16:0 and 2-hydroxy-1620 N-acyl groups in sphingolipids, so the total flux of palmitate into Sphingolipids is actually higher. Thus Sphingolipid synthesis may consume 30-40 % of the total cytosolic palmitate pool. However, despite a reduction of about 50 % in extraplastidial 16:0 in the fatb mutant, the total amount per fresh weight of sphingoid bases in leaf tissue was similar to wild-type. One interpretation of the constancy of Sphingolipid production is that sphingoid base synthesis is tightly maintained at the expense of acyl composition changes in other glycerolipids. Furthermore, because Sphingolipids are essential lipids for cell growth (Wells and Lester, 1983) the slow grth of the fatb-k0 plants could result from a slower supply of the critical 16:0 component to sphingolipids. 130 Although in leaf mesophyll cells the major fraction of fatty acids is used for biosynthesis of membrane glycerolipids, in epidermal cells most newly produced fatty acids are directed toward the biosynthesis of cutin and epicuticular waxes. In Arabidopsis leaves, an epicuticular wax load of 0.2 pmoles/gfw represents only approximately 10 % of the total leaf pool of palmitate plus stearate, but within the epidermal cells the proportions will be much higher. Analysis of leaf and stem epicuticular waxes in the fatb-k0 mutant showed a 20 % and 50 % reduction in total wax load respectively, indicating that the FATB thioesterase is one source for production of wax precursors in epidermal cells. Moreover, a reduction of about 85 % in 16 carbon derivatives from cutin demonstrated that FATB supplies palmitic acid for cutin biosynthesis. In leaf tissue of the fatb mutant, all extra-plastidial phospholipids showed a reduction of approximately 50 % in the relative 16:0 levels (Table 9), except PC that showed a 78 % reduction. The fact that PC was most affected may suggest that this phospholipid is a key pool in the delivery or partition of palmitate in the cell, for example as an indirect donor of saturated groups to sphingolipid biosynthesis. In addition, the lO-fold reduction in the relative 18:0 levels also implicated that PC may play a role in 18:0 partitioning. PC plays a major role in the flux of glycerol and fatty acids during membrane glycerolipid biosynthesis and therefore an intriguing question is whether the larger changes in the fatty acid composition of PC could affect its function and be responsible, at least in part, for the slower growth rate of the fatb-knockout plants. Conclusions 131 The fatb-k0 line shows a reduction in saturated fatty acids exported to the cytosol, a 17 % reduction in the rate of growth, and altered seed morphology and germination. Although this study clearly demonstrates the requirement for the FATB gene and saturated fatty acids for normal rates of Arabidopsis growth and viable seed formation, the specific function(s) supplied by saturated fatty acids to sustain normal growth are still uncertain. Other than the reduction in saturated fatty acid content, and a decrease in wax load, alterations in glycerolipids and sphingoid base compositions were minor. Future work will focus on whether the growth rate of the mutant is linked to the biosynthesis of specific cellular components, subtle variations in membrane properties or changes in F AS/ lipid tumover-degradation rates or a combination of effects. The recent development of new isotope labeling techniques to investigate lipid synthesis and tumover-degradation will be critical to answer these questions (Pollard and Ohlrogge, 1999; Bao et al., 2000). However, it is important to note that a lack of change of a critical component for growth may point to its essential nature more than change. Thus, the slower production of an essential lipid in the fatb mutant could slow growth to a balance point between synthesis of that component and growth and therefore, no change in the level of the key component per plant weight would be expected in the mutant. For example, although we observed no overall reduction in sphingoid base accumulation, the essential nature of sphingolipid synthesis for growth in other systems (Wells and Lester, 1983) suggests compositional changes might not be expected. Similarly, if rates of protein acylation, cutin biosynthesis or synthesis of other saturate derived components are essential to growth, a biochemical phenotype in these components may not be observed. Thus, the isolation of suppressor mutants for the fatb-k0 phenotype may also provide insights into the underlying 132 mechanisms which connect the supply of saturated fatty acids to the biosynthesis and/or regulation of essential plant growth components. MATERIALS AND METHODS Plant Material and Growth Conditions Wild-type Arabidopsis thaliana and fatb-k0 mutant plants (ecotype Wassilewska) were grown at 80-100 umol m'2 s", 22°C under l8-h light/6-h dark photoperiod. Seeds were always estratified for three days at 4°C. Selection of T-DNA tagged plants was carried out by soaking the soil with 50 ug/mL of commercial BASTA (Finale) (AgrEvo, Montvale, NJ). Surface sterilized seeds of Arabidopsis were germinated on 0.8 % (w/v) agar solidified MS medium supplemented with l % (w/v) sucrose. For the experiments at different temperatures the plants were grown for 2 weeks at 22°C (16 hours light) and then transferred to 16°C and 36°C or kept at 22°C under identical lighting conditions. Arabidopsis fatb-k0 act] double mutant was generated by using Arabidopsis fatb-k0 plants as a pollen donor and Arabidopsis act] (Kunst et al., 1988) as pollen recipient. Seeds obtained from the crosses were sown on soil in the presence of 50 ug/mL BASTA in order to select for heterozygous act]/+ fatb—k0/+ seedlings. F2 seedlings from F1 heterozygous plants were screened by GC and PCR in order to detect homozygous, heterozygous and wild types for act] and fatb-k0 loci. Mutant Isolation 133 A transfer DNA (T-DNA, 5.5 Kb) insertion into the FATB gene was identified by screening pooled genomic DNA prepared from a T-DNA tagged Arabidopsis thaliana (ecotype Wassilewska) collection (Sussman et al., 2000). The gene-specific primers used for the screening of insertions into the FA TB gene were 5’- CTCATATCCACATATATCTCTCTCTCACC-3’ forward and 5’- CAAGCAAGCAAGGTGGTAGTAGCAGATAT-3’ reverse, and the T-DNA specific primer matching the left end of the T-DNA (IL-202) was 5’- CATTTTATAATAACGCTGCGGACATCTA C-3’. A FATB genomic fragment was labeled using random priming and used to detect specific PCR products by Southern blot hybridization (Sambrook et al., 1989). The T-DNA/FA TB junctions from both ends of the insertion were PCR amplified, subcloned and sequenced from both ends to determine the insertion point for the 5.5 kb T-DNA. Complementation Analysis The binary vector pBINAR-Hyg (Becker D, 1990) carrying the wild-type FATB cDNA was a gift from P. Doermann. The vector was used to transform Arabidopsis fatb- knockout plants by Agrobacterium vacuum infiltration (Bechtold et al., 1993). The transformation of homozygote plants for the FATB T-DNA insertion did not render transgenic seedlings and therefore heterozygote fatb-k0 plants were transformed. Transgenic seedlings were first selected on agar plates in the presence of 25 ug/mL of hygromycin-B. After 7 days, seedlings were transferred to soil pre-soaked with 50 ug/mL of BASTA in order to select against wild-type plants for the fatb-knockout T-DNA 134 insertion. DNA was extracted from hygromycin-B (transgene T-DNA) and BASTA (knock out T-DNA) resistant plants by using the Qiagen Plant DNA extraction kit (Chatsworth, CA). Homozygote fatb-k0 plants were identified by PCR using the same primers as for the PCR originally used to isolate the fatb-k0 mutant. Real-Time PCR Quantification of mRNAs Total RNA was prepared from leaf tissue of wild-type (Wassilewska) and fatb-k0 mutant Arabidopsis by using the Qiagen Plant RNA extraction kit (Chatsworth, CA) according to the instructions. A 5 ug aliquot was used as a template for cDNA synthesis employing the SuperScript First Strand Synthesis system and oligo dT primers (Stratagene). Specific primers for the second and third exons of the FATB gene were designed with Primer Express sofiware (PE Applied Biosystems). The sequences of forward and reverse primers are: 5 ’ -AATCATGTTAAGACTGCTGGATTGC-3 ’ and S ’- ATACCATTCTTTCCAGACTGACTGA-3’ respectively (Figure 19). Primers were verified by showing that the PCR reaction product produced a single band after agarose gel electrophoresis. Real Time Quantitative PCR analysis was performed according to the manufacturer’s instructions (PE Applied Biosystems). The reaction contained, in a final volume of 30 uL, 250 ng of reverse transcribed total RNA, 1.5 uM of the forward and reverse primers, and 2X SYBR Green PCR Master Mix. All reactions were performed in triplicate. The relative amounts of all mRNAs were calculating using the Comparative CT method as described in User Bulletin #2 (PE Applied Biosystems). Arabidopsis eukaryotic protein synthesis initiation factor 4A1 (eIF4A1) mRNA was used as an internal control for variations in amounts of mRNA. Levels of FATB mRNA were 135 normalized to eIF 4A1 mRNA levels and presented as a ratio between wild-type and fatb- ko mutant plants. The forward and reverse primers used to amplify eIF4A] mRNA were 5 ’ -CCAGAAGGCACACAGTTTGATGCA-3 ’ and 5 ’ - AGACTGAGCCTGTTGAATCACATC-3’ respectively. Fatty Acid Analysis of Glycerolipids from Different Tissues of Arabidopsis Approximately 0.1 g fresh weight (gfw) of tissue from 5-week old Arabidopsis plants was heated at 90°C for 1 hour in 0.3 mL of toluene and 1 mL of 10% (v/v) boron trichlodide/methanol (Sigma) with heptadecanoic acid (17:0) as an internal standard. After acidification with aqueous acetic acid fatty acid methyl esters were extracted two times with hexane and analyzed by gas chromatography with a flame ionization detector (GC-FID) on a DB-23 capillary column. Individual Glycerolipid Analysis One gfw of leaf tissue from 5-week old Arabidopsis plants was ground in liquid nitrogen. Lipids were extracted in hexane-isopropanol and glycerolipid classes were separated by thin layer chromatography (TLC) on K6 silica plates (Whatman Inc, Clifton, PA) impregnated with 0.15 M ammonium sulphate and activated for 3 hours at 110° C (Kahn and Williams, 1977). The TLC plates were developed three times with 912308 (vzvzv) acetoneztoluenezwater and lipids were detected after spraying with 0.2 % (w/v) 2’-7’- dichlorofluorescein/methanol and viewing under UV light. Standards were used to identify the different glycerolipid classes. Lipids were eluted from the silica with 136 chloroform-methanol and fatty acid methyl esters prepared and analyzed as described above. Leaf Epicuticular Waxes Approximately 3 gfw of leaf tissue from 5-week old Arabidopsis plants were used for epicuticular wax analysis. The tissue was dipped in chloroform for 30 seconds and then the following internal standards (IS) were added; n-octacosane at 20 pg per gr of fresh weight, docosanoic acid and l-tricosanol both at 10 pg per gr of fresh weight. All the compounds were purchased from Sigma. After evaporation of the chloroform under nitrogen the epicuticular waxes were silylated to convert free alcohols and carboxylic acids to their trimethylsilyl(TMS)-ethers and -esters respectively. The epicuticular waxes were heated at 110° C for 10 min in 100 uL of pyridine and 100 uL of N,O- bis(trimethylsilyl)trifluoroacetamide (Sigma). After cooling, the solvent was evaporated under nitrogen and the product was resuspended in 1:1 (volzvol) heptaneztoluene for GC analysis. GC conditions were the following: a HP-5 capillary column (30 m x 0.32 mm x 0.25 pm film thickness) with helium carrier gas at 2 ml/min was used; injection was in split mode; injector and FID-detector temperatures were set at 360°C; and the oven temperature was programmed at 150°C for 3 minutes, followed by a 10°C/min ramp to 350°C, and then held for an additional 20 min at 350°C. GC/MS analysis was also performed to identify components of the mixture. Sphingoid Base Analysis 137 Approximately 1 gfw of leaf tissue from 5-week old Arabidopsis plants was heated at 110°C for 24 hours with 4 mL of dioxane (Sigma) plus 3.5 mL of 10 % (w/v) aqueous Ba(OH)2 (Sigma) (Sperling et al., 1998). D-erythro-sphingosine (Matreya Inc, Pleasant Gap, PA) at 25 ug/gfw and heptadecanoic acid (Sigma) at 400 ug/gfw were added as internal standards. After saponification the sample was extracted with chloroform to obtain the sphingoid bases in the organic phase and fatty acids in the alkaline aqueous phase. Each phase was independently analyzed as described below. The chloroform fraction containing the sphingoid bases was back-extracted with an equal volume of 0.4 M aqueous HCl. The acid aqueous phase (containing the protonated sphingoid bases) was then titrated with KOH to pH = 10. The sphingoid bases were re- extracted into chloroform and after evaporation under nitrogen resuspended in 1 mL of methanol. The sphingoid bases were oxidized to their corresponding aldehydes by stirring the sample with 100 uL of 0.2 M sodium periodate (Sigma) at room temperature for 1 hour in the dark (Kojima et al., 1991). The aldehydes were recovered by hexane extraction and used directly for GC analysis. GC conditions were the following: a HP-5 capillary column (30 m x 0.32 mm x 0.25 mm film thickness) with helium carrier gas at 2 ml/min was used; injection was in split mode; injector and FID-detector temperatures were set at 250°C; and the oven temperature was programmed at 100°C for 3 minutes, followed by a 10°C/min ramp to 260°C, and then held for an additional 10 min at 260°C. GC/MS analysis was also performed to identify components of the mixture. 138 The basic aqueous phase was acidified with HCl to pH < 4 and extracted twice with hexane to recover fatty acids. Fatty acids were transmethylated and analyzed by GC using the same protocol as indicated above. Monoglucosylceramide (MGC) Analysis Approximately 10 g of Arabidopsis leaf tissue was quenched with 50 mL of hot isopropanol and ground in a polytron. The extract was filtered and the residue extracted with 25 mL of 2:1 (vzv) chloroformzmethanol and re-filtered. This residue was re- extracted with 25 mL of 1:2 (vzv) chloroformzmethanol and again filtered. All the three filtrates were combined and evaporated to dryness on a rotary evaporator. The lipid fraction was finally dissolved in 5 mL of chloroform (“Lipid Fraction” in Table 10). The solvent extracted residue dried under vacuum (“Solvent Extracted Residue” in Table 10). The lipid fraction was subjected to a partial base transmethylation to convert most of the O-acyl glycerolipids to fatty acid methyl esters under conditions which leave the N-acyl groups intact. This was achieved by vortexing the lipids (100 mg) with 2M KOH in methanol (0.6 mL) plus hexane (4 mL) for 2 min at room temperature. The reaction was quenched by adding 4 mL of 1M aqueous acetic acid. The hexane phase was removed and the acidified aqueous phase extracted with 2:] then 1:2 (vzv) chloroformzmethanol. The hexane and chlorofonn2methanol fractions were combined and evaporated to dryness under nitrogen. Lipids were analyzed by TLC as described for “Individual Glycerolipid Analysis” above. A band that co-migrated with beta-D-Glucosyl Ceramide standard (Matreya Inc, Pleasant Gap, PA) was eluted from the silica with 2:1 and 1:2 (vzv) chloroformzmethanol. The solvent was dried under nitrogen and the MGC sample was 139 cleaved by alkaline hydrolysis in 1 mL dioxane (Sigma) and 1 mL 10% (w/v) aqueous Ba(OH)2 (Sigma) for 24 hours at 110°C. Sphingoid bases were recovered and their aldehyde-derivatives analyzed as indicated above for total sphingoid base analysis. Cutin Analysis Approximately 20 g of leaf or stem tissue was ground in liquid nitrogen and extracted with 200 mL of isopropanol. The extract was filtered and the residue re-extracted with 200 mL 2/1 (v/v) chlorofomr/methanol. After filtering the residue was re-extracted with 200 mL 1/2 (v/v) chloroform/methanol and air-dried. An excess (2.5 times by weight) of LiAlH4 plus 6 mL of tetrahydrofuran were added to 100 mg of residue. Hydrolysis was at 80°C for 48 hours with periodic vortexing. After hydrolysis, the excess of LiAlH4 was decomposed by careful addition of ethyl-acetate to the reaction mixture. The mixture was acidified by the addition of 5 mL of water plus 0.8 mL of concentrated HCL. Cutin components were extracted two times with 6 mL of diethyl-ether. The ethereal solution was dried under nitrogen and the sample dissolved in 0.1 mL pyridine plus 0.1 mL of BSTFA (SIGMA). Silylation was at 110° C for 10 min. The excess of reagent was evaporated under nitrogen and the sample dissolved in 1/1 (v/v) heptane/toluene for GLC-MS analysis in an HP-5 capillary column with the oven temperature programmed at 90°C for 5 min, followed by 10°C/min ramp to 300°C, and then held for an additional 10 min at 300°C. Acyl-ACP Thioesterase Activities 140 Assays for 16:0-ACP and 18:1-ACP hydrolysis were performed according to Eccleston and Ohlrogge, 1996. Scanning electron microscopy Plants were harvested at 3 weeks of age and fixed in 3% glutaraldehyde in 0.1 M PIPES (pH 7.2) overnight at 4°C. Samples were washed three times with 0.1 M PIPES (pH 7.2), 10 min per wash, before dehydration through graded ethanol series (50, 60, 70, 80, 90, 100%). Dehydrated samples were transferred to a drying apparatus and critical point dried at C02. Dehydrated specimens were affixed to stubs with double-sided tape and coated with gold in argon. SEM pictures were taken at the Center for Advanced Microscopy at Michigan State University in a J EOL (Japan Electron Optics Laboratories) J SM-6400V scanning electron microscope. Transmission electron microscopy For TEM small leaf samples were cut from the middle of the oldest true leaves and immersed in 3%glutaraldehyde in 0.1 M PIPES buffer (pH 7.2). Samples were then fixed overnight at 4°C as described above. After three washes in 0.1 M PIPES buffer, the tissues were postfixed in 2% aqueous osmium tetroxide overnight at 4°C. Samples were again washed 3 times in 0.1M PIPES (pH 7.2) and then dehydrated through graded ethanol series as above. TEM samples were embedded in epoxy resin and thin sectioned before staining. Accession Numbers 141 The accession numbers for the Arabidopsis thaliana proteins described in Figure 25 and text are: acyl-ACP thioesterase B (FATB) (At1g08510); acyl-ACP thioesterase A (FATA) (At3g25110 and At4gl3050), glycerol-3-phosphate acyltransferase (ACTl) (At1g32200 ), 3-ketoacyl-ACP-synthase-II (KAS-II) ( Atlg74960), lyso-phosphatidate sn-2 acyltransferase (LPAAT) (At4g30580) and eIF4Al (At3gl3920). LITERATURE CITED Bao X., Focke M., Pollard M., Ohlrogge J. (2000) Understanding in vivo carbon precursor supply for fatty acid synthesis in leaf tissue. Plant J, 22: 39-50. Bechtold N., Ellis J., Pelletier G. (1993) In-planta Agrobacterium-mediated gene- transfer by infiltration of adult Arabidopsis thaliana plants. Cr Acad Sci III-Vie, 316: 1194-1199 Becker B. (1990) Binary vectors which allow the exchange of plant selectable markers and reporter genes. Nucleic Acids Res, 18: 203. Browse J., Warwick N., Somerville C.R., Slack CR. (1986) Fluxes through the prokaryotic and eukaryotic pathways of lipid synthesis in the '1613' plant Arabidopsis thaliana. Biochem J, 235: 25-31. Browse J., McConn M., James Jr D., Miquel M. (1993) Mutants of Arabidopsis deficient in the synthesis of alpha-linoleate. J Biol Chem, 268: 16345-16351. Browse J., Somerville C. (1991) Glycerolipid synthesis - biochemistry and regulation. Annu Rev Plant Physiol Plant Mol Biol, 42: 467-506 Buhr T, Sato S, Ebrahim F, Xing A, Zhou Y, Mathiesen M, Schweiger B, Kinney A, Staswick P, Clemente T (2002) Ribozyme termination of RNA transcripts down regulate seed fatty acid genes in transgenic soybean. Plant J, 30: 155-163. Doermann P., Voelker T.A., Ohlrogge J.B. (1995) Cloning and expression in Escherichia coli of a novel thioesterase from Arabidopsis thaliana specific for long-chain acyl-acyl carrier proteins. Arch Biochem Biophys, 316: 612-618. 142 Doermann P., Voelker T.A., Ohlrogge J.B. (2000) Accumulation of palmitate in Arabidopsis mediated by the acyl-acyl carrierprotein thioesterase FATBl. Plant Physiol, 123: 637-44. Eccleston V.S., Ohlrogge J.B. (1998) 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, 10: 613- 22. Frentzen M., Heinz E., McKeon T.A., Stumpf P.K. (1983) Specificities and selectivities of g1ycerol-3-phosphaste acyl-transferase and monoacylglycerol-3-phosphate acyltransferase from pea and spinach chloroplasts. Eur J Biochem, 129: 629-638. Imai H., Morimoto Y., Tamura K. (2000) Sphingoid base composition of monoglucosylceramide in Brassicaceae. J Plant Physiol, 157: 453-456. Kahn M.U., Williams J.P. (1977) Improved thin layer chromatographic method for the separation of major phospholipids and glycolipids from plant lipid extracts and phosphatidylglycerol and bis(monoacylglycerol) phosphate from animal lipid extracts. J Chromatogr, 140: 179-185. Kojima M., Ohnishi M., Ito S. (1991) Composition and molecular species of ceramide and cerebroside in scarlet runner beans (Phaseolus coccineus L.) and kidney beans (Phaseolus vulgaris L.). J Agric Food Chem, 39: 1709-1714. Kunst L., Browse J., Somerville C. (1989) A mutant of Arabidopsis deficient in desaturation of palmitic acid in leaf lipids. Plant Physiol, 90: 943-947. Kunst L., Browse J., Somerville C. (1988) Altered regulation of lipid biosynthesis in a mutant of Arabidopsis deficient in chloroplast glycerol-3-phosphate acyltransferase activity. Proc Nat Acad Sci USA, 85: 4143-4147 Lightner J., Wu J.R., Browse J. (1994) A mutant of Arabidopsis with increased levels of stearic—acid. Plant Physiol, 106: 1443-1451 Lynch D.V. (1993) Sphingolipids. In Lipid Metabolism in Plants, T.S. Moore, ed. (Boca Raton, FL: CRC Press), pp. 285-303 Mc Elhaney RN. (1989) The influence of membrane lipid composition and physical properties of membrane structure and function in Acholeplasma laidlawii. Crit Rev Microbiol, 17 : 1-32. Mekhedov S., de Ilarduya O.M., Ohlrogge J. (2000) Toward a fimctional catalog of the plant genome. A survey of genes for lipid biosynthesis. Plant Physiol, 122: 389-402. 143 Nadev Y., Wierzbicki A., Knowtoln S., Pierce J., Ripp K., Hitz W., Aegertcr M., Browse J. (1992) Alteration of the levels of palmitic and stearic acids in somatic soybean embryos by co-suppression of the acyl-ACP thioesterase. In Biochemistry and Molecular Biology of Membrane and Storage lipids in Plants, N. Murata and C. Somerville, ed (Rockville, MR: ASPP Press), pp. 61-65 Ohlrogge J., Browse J. (1995) Lipid biosynthesis. Plant Cell, 7: 957-970. Pollard M.R., Ohlrogge J.B. (1999) Testing models of fatty acid transfer and lipid synthesis in spinach leaf using in vivo oxygen-18 labeling. Plant Physiol, 121: 1217-1226 Pollard M.R., Anderson L., Fan C., Hawkins D.H., Davies H.M. (1991) A specific acyl-ACP thioesterase implicated in medium chain fatty acid production in immature cotyledons of Umbellularia californica. Arch Biochem Biophys, 284: 306-312. Post-Beittenmiller D. (1996) Biochemistry and molecular biology of wax production in plants. Annu Rev Plant Physiol Plant Mol Biol, 47 : 405-430. Rashotte AM, Jenks MA, Feldmann KA (2001) Cuticular waxes on eceriferum mutants of Arabidopsis thaliana. Phytochemistry, 57: 115-123. Roughan G., Nishida I. (1990) Concentrations of long chain acyl acyl carrier proteins during fatty acid synthesis by chloroplasts isolated from pea (Pisum sativum), safflower (Carthamus tinctoris) and amaranthus (Amaranthus lividus) leaves. Arch Biochem Biophys, 276: 38-46. Salas J.J., Ohlrogge J.B. (2002) Characterization of substrate specificity of plant FatA and fatb acyl-ACP thioesterases. Arch Biochem Biophys, 403: 25-34. 8011 J., Roughan G. (1982) Acyl Acyl carrier protein pool sizes during steady state fatty acid synthesis by isolated spinach chloroplasts. Febs Letters, 146: 189-192. Somerville C., Browse J., Jaworski J.G., Ohlrogge J. (2000) Lipids. In B Buchanan, W Gruissem, R Jones, eds, American Society of Plant Physiologists, Rockville, Maryland, 456-527 Sperling P., Zahringer U., Heinz E. (1998) A sphingolipid desaturase from higher plants-Identification of a new cytochrome b(5) fusion protein. J Biol Chem, 273: 28590-28596 Sussman M.R., Amasino R.M., Young J.C., Krysan P.J., Austin-Phillips S. (2000) The Arabidopsis knockout facility at the University of Wisconsin-Madison. Plant Physiol, 124: 1465-1467 144 Voelker T.A., Jones A., Cranmer A.M., Davies H.M., Knutzon D.S. ( 1997) Broad- range and binary-range acyl-acyl-carrier protein thioesterases suggest an alternative mechanism for medium-chain production in seeds. Plant Physiol, 114: 669-77 Wada H., Shintani D., Ohlrogge J. (1997) Why do mitochondria synthesize fatty acids? Evidence for involvement in lipoic acid production. Proc Natl Acad Sci U S A, 94: 1591- 1596 Walenga R.W., Lands W.E. (1975) Effectiveness of various unsaturated fatty acids in supporting growth and respiration in Saccharomyces cerevisiae. J Biol Chem, 250: 9121- 9129. Wallis J.G., Browse J. (2002) Mutants of Arabidopsis reveal many roles for membrane lipids. Prog Lipid Res, 41: 254-78 Wells G.B., Lester R.L. (1983) The isolation and characterization of a mutant strain of Saccharomyces cerevisiae that requires a long chain base for growth and for synthesis of phosphosphingolipids. J Biol Chem, 258: 10200-10203. Wilson RF, Marquardt TC, Novitzky WP, Burton JW, Wilcox JR, Kinney AJ, Dewey RE (2001) Metabolic mechanisms associated with alleles governing the 16:0 concentration of soybean oil. J Amer Oil Chem Soc, 78: 335-340. Wu J.R., James D.W., Dooner H.K., Browse J. (1994) A mutant of Arabidopsis deficient in the elongation of palmitic acid. Plant Physiol, 106: 143-150. Yalovsky S., Rodriguez-Concepcion M., Gruissem W. (1999) Lipid modification of proteins-slipping in and out of membranes. Trends Plant Sci, 4: 439-445. 145 CHAPTER 4 DISRUPTION OF THE FA TB GENE INCREASES THE RATE OF FATTY ACID BIOSYNTHESIS,FATTY ACID AND LIPID TURNOVER, AND FAS PROTEIN EXPRESSION IN LEAVES. ABSTRACT Disruption of the FATB gene in Arabidopsis (fatb-k0) results in ~ 60 % reduction in saturated fatty acid export from plastids and 17 % reduction in the growth rate of the mutant compared to wild type (Chapter 3 and Bonaventure et al., 2003). In this study we report that although fatb-k0 seedlings grow more slowly than wild type the rate of fatty acid synthesis (FAS) in leaves of the mutant increases by 40 %. Moreover, the protein levels of BCCP, ACPS and 18:0-ACP desaturase are increased by 1.5-2 fold in fatb-k0 leaves compared to wild type, suggesting concerted up-regulation of FAS protein expression. To maintain constant amounts of fatty acids in leaves, thereby counterbalancing their higher rate of production, the mutant also increases by 70 % the initial rate of fatty acid breakdown. However, although fatb-k0 leaves have higher rates of fatty acid synthesis and turnover, the composition of membrane lipids is similar to wild type. Thus, homeostatic mechanisms to preserve membrane compositions are functional and compensate for substantial changes in rates of fatty acid and glycerolipid metabolism in the mutant The results suggest that fatb-k0 cells attempt to increase saturated fatty acid production, however a surplus of fatty acids above amounts needed for leaf growth is created and leads to increased turnover. In the mutant, the surplus of fatty acids exported from leaf plastids (largely C18 unsaturated fatty acids) and 146 1: AF. I n I In'. .2 o 1 incorporated into polar lipids is mainly lost via degradation or interconversion of C182C18- PC species. In addition, Cu, fatty acids are also rapidly removed from this phospholipid in fatb-k0 leaves, suggesting that mechanisms are not present which can preferentially preserve the saturated fatty acids. INTRODUCTION In plants, the major pathway for the de novo fatty acid synthesis occurs in the plastid (Ohlrogge et al., 1979) and this organelle exports fatty acids to supply a diverse array of extraplastidial biosynthetic pathways and cellular processes in the cell (Kolattukudy, 2003; Yalovsky et al., 1999; Post-Beitenmiller, 1996; Lynch, 1993). Production of fatty acids to be exported from plastids depends on the activity of acyl-ACP thioesterases (F ATs) that hydrolize the acyl group from acyl-acyl carrier protein (acyl-ACP) to release free fatty acids and ACP (reviewed by Voelker et al., 1997). After export from the plastids, free fatty acids are re-esterified to coenzyme-A (CoA) to form the cytosolic acyl-COA pool (Pollard and Ohlrogge, 1999). In most plant cells, these molecules are primarily used for the biosynthesis of membrane glycerolipids at the level of the endoplasmic reticulum (ER) (Browse and Somerville, 1991). However, in other tissues exported acyl-COA molecules have different destinations. For example, in oilseeds, the major fraction of acyl groups is incorporated into triacylglycerols (TAG) that may constitute more than 90 % of total lipids. In epidermal cells, a significant fraction of acyl- CoA molecules is utilized for the synthesis of waxes and cutin (Kolattukudy, 2003; Post- Beittenmiller, 1996). Furthermore, all cells synthesize sphingolipids, and we recently 147 estimated that as much as 30-40 % of exported 16:0-COA is needed for sphingoid base synthesis in leaves (Bonaventure et al., 2003). Finally, exported 14:0-CoA and 16:0-COA participates in other acylation reactions, such as protein acylation (Yalovsky, 1999). Because acyl-ACP thioesterases terminate fatty acid synthesis and allow the export of fatty acids from plastids, these enzymes are critical regulators of cellular metabolism, controlling the partitioning of de nova-synthesized fatty acids between plastids and cytosol and therefore their supply to diverse cellular processes. Two classes of FAT enzymes has been described in most plants, namely FATA and FATB (Voelker et al., 1997). The FATA class has highest in vitro activity for 18:1-ACP and lower for saturated acyl-ACP substrates. In contrast, the FATB class prefers saturated acyl-ACP but also shows activity for unsaturated acyl-ACPS (Salas and Ohlrogge, 2002; Voelker et al., 1997; Doermann et al., 1995). The Arabidopsis genome encodes two FATA genes and a single FATB gene (Beisson et al., 2003; Mekhedov et al., 2000). We previously described the isolation and analysis of an Arabidopsis mutant disrupted in the FA TB gene (fatb-k0) (Bonaventure at al., 2003). In this mutant, the overall amounts of saturated fatty acids in different tissues are reduced by 40 to 50 % compared to wild type. This reduction occurs only in the cytosolic pool of saturated fatty acids, affecting the fatty acid composition of extraplastidial glycerolipids, synthesis of wax in leaves and stems, and the fatty acid composition of TAGS in seeds. However, although sphingolipid synthesis is initiated from 16:0-COA, no reductions were observed in total sphingoid base content, suggesting that plants may prioritize the synthesis of these essential lipids (Wells and Lester, 1.983). These observations also suggested the existence of a metabolic hierarchy 148 in the economy of saturated fatty acids in plant cells, which prioritizes synthesis of essential components for growth (Bonaventure et a] 2003). The reduction in the pool of cytosolic saturated fatty acids also slows down the growth of the mutant, resulting in seedlings approximately half the size of wild type seedlings (Bonaventure et al, 2003). Thus, reduction in the export of saturated fatty acids from plastids affects cellular processes that are critical for plant growth. Complete suppression or disruption of any of the core enzymes of FAS would be expected to reduce fatty acid synthesis and affect plant performance. In support of this, analysis of tobacco plants engineered to constitutively express an antisense transcript of the tobacco biotin carboxylase (BC) or the Arabidopsis biotin carboxylase carrier protein (BCCP), showed reductions of leaf fatty acids together with a stunted phenotype (Shintani et al, 1997). Growth phenotypes together with reduction in leaf lipids were also obtained with antisense constructs directed against stearoyl-ACP desaturase and ACP4, and with induced mutations in the fab2 and enoyl-ACP reductase genes (Branen et al., 2003; Mon at al., 2000; Lightner et al., 1994). It is apparent from these studies that even slight reductions in leaf fatty acids synthesis can have pleiotropic effects on plant growth and development. Some of these effects come from changes in chloroplast membrane structure and therefore loss of photosynthetic capability. By contrast, a wide range of mutations in plant fatty acid desaturases, acyltransferases and condensing enzymes demonstrate that the fatty acid composition of plant membranes can be altered considerably with no apparent phenotype under normal growth conditions (Wallis and Browse, 2002; Kunst et al., 1989; Wu et al., 1994). These mutants reveal the plasticity of 149 lipid metabolism in plants and the capacity of these organisms to adjust to alternative lipid compositions. Based on these observations, one possible mechanism responsible for the slower growth of fatb-k0 plants could be a reduced synthesis of fatty acids and therefore a decline in the rate of membrane lipid biosynthesis (Branen et al., 2003; Shintani et al., 1997). Alternatively, slower growth of the mutant may be the result of reduced synthesis of other critical components such as sphingolipids, cutin and waxes or lower rates of acylation reactions. In addition, lipid-derived signaling molecules that affect growth could be affected in the mutant (Nandi et al., 2003). When fatb-k0 plants are grown at a range of different temperatures, no suppression of the fatb-k0 grth phenotype is observed, suggesting that changes in the bulk properties of cellular membranes are not the main factor affecting plant performance (Bonaventure et al., 2003). However, more subtle changes in transport and vesiculation cannot be ruled out. Likewise, when the mutant is grown at different relative humidity levels and even in liquid culture, no reversion of the growth phenotype occurs, suggesting that increased water loss through reduced wax content is also not a major factor affecting growth. Similar results are obtained when fatb-k0 is supplemented with exogenous sugar and in this case its photosynthetic capability appears not to limit its growth (Bonaventure et al., 2003). In addition, the chlorophyll content and the ultrastructure of fatb-k0 chloroplasts are similar to wild type (Chapter 3, Figure 22) 150 g To further elucidate the role of saturated fatty acids in plant growth and to understand fatty acid partition and F AS regulation in fatb-k0, a series of isotope labeling experiments were conducted. Unexpectedly, the rates of both fatty acid synthesis and turnover were higher in fatb-k0 than wild type leaves. In addition, up-regulation of three proteins involved in FAS was detected, suggesting that higher F AS rates were achieved at least in part by higher FAS protein expression. Thus, fatb-k0 plants appear to induce a futile cycle of fatty acid production and degradation, perhaps as an attempt to increase saturated fatty acid synthesis. RESULTS Rate of fatty acid biosynthesis in leaves of wild type Arabidopsis and fatb-k0. The rate of fatty acid biosynthesis in leaves correlates mainly with the expansion rate of this organ, being higher in younger leaves and reflecting the demand for new membranes to sustain cell division and chloroplast biogenesis (Bao et al., 2000; Browse et al., 1981). It was previously reported that fab-k0 plants grow slowly resulting in a 50 % decline in their fresh weight compared to wild type (Bonaventure et al., 2003). To evaluate if the disruption of the FATB gene reduces the rate of fatty acid synthesis in leaves, incorporation of labeled precursors was used to assess the in vivo rate of this pathway in Arabidopsis wild type and fatb-k0. Because isotope incorporation rates can be influenced by internal pools which might differ between mutant and wild type (Nunn et al., 1977; Cronan et al., 1975), three different labeled substrates were tested. 151 aura-ruin” h. First, the rate of 3 H incorporation into fatty acids from 3HzO was measured in leaf tissue of wild type and fatb-k0. The results in Table 13 indicate that mutant leaves incorporated 3 H into fatty acids at a rate 40 % higher than wild type tissue. The incorporation of 3 H was linear for the first two hours of the assay and followed non-linear kinetics thereafter (data not shown). Similar to other systems, a lag phase of 10-15 min for 3 H incorporation into fatty acids was observed (data not shown and Browse et al., 1981). Second, leaf tissue of wild type Arabidopsis and fatb-k0 were incubated with exogenous [MC] acetate as the label substrate. Similar to the results obtained with radiolabeled water, mutant leaves incorporated 14C from acetate into fatty acids at a rate 30 % higher than wild type (Table 13). The kinetic of 14C acetate incorporation into fatty acids was linear for at least 1 hour (data not shown). Finally, intact wild type and mutant seedlings were labeled with 14C02. As shown in Table 13, fatb-k0 leaves incorporated 14C from C0; into fatty acids at a higher rate than wild type (ca. 50 %). In all the experiments, the chlorophyll content per gram fresh weight (gfw) in leaves of the two plant classes was within 5 % of one another (on average: 1.05 i 0.01 mg/gfw for wild type and 1.07 i 0.03 mg/gfw for fatb-k0). Table 13. Rates of fatty acid synthesis (FAS) in leaves of Arabidopsis wild type and fatb-k0 measured by different radiolabeled substrates Substrate Wild type* fatb-k0* IncreaseI (%) 3H20 2801 i 139 3878 i 145 40 14C02 42.0 i 1.6 63.6 a: 2.9 50 "‘C acetate 39.0 i 1.2 50.9 i 1.4 30 Units are in nmol of 3 H or 14C / h / mg Chl incorporated into fatty acids and values represent initial rates of fatty acid synthesis (see Materials and Methods). lPercent increase of fatb-k0 rates versus wild type rates. 152 gm In conclusion, by using three different radiolabeled substrates it was demonstrated that fatb-k0 leaves synthesized fatty acids at an average rate 40 % higher than wild type. Rate of fatty acid turnover in wild type and fatb-k0 leaves We previously reported that leaves from fatb-k0 plants contained the same amount of fatty acids per gram of fresh weight than wild type (Bonaventure et al., 2003). Therefore, if the FAS rate in fatb-k0 leaves is on average 40 % higher than wild type, the rate of fatty acid tumover must concomitantly increase to maintain fatty acid amounts constant in this tissue. Therefore, to test this conclusion fatty acid turnover was evaluated by labeling intact seedlings with a 30 min pulse of 14CO; and then determining the level of MC in fatty acids at different times up to 120 hours (Figure 26). During the first 6 hours of the chase period, there was net accumulation of '4C into fatty acids of both wild type and mutant (Figure 26). This net accumulation of label most likely reflected the use of 14C labeled carbohydrates as substrates for fatty acid synthesis (Bao et al., 2000). After 6 h of the initial l4CO; pulse, the amount of label in fatty acids from wild type and mutant began to decay (Figure 26). In the 6-120 h period, the radioactivity in fatty acids dropped by 70 % in fatb-k0 leaves compared to 47 % in wild type leaves (Figure 26). Therefore, wild type and mutant leaves have an average turnover rate of 9 % and 14 % per day respectively (55 % increase in fatb-k0). However, as the label in fatty acids decayed following exponential kinetics in the mutant, the average rates did not properly reflect the actual fatty acid turnover rates (Figure 26). Thus, if initial rates were considered instead (6-24 h period), wild type and mutant plants lost 153 1 0.39 and 0.67 nmol 14C/mg Chl/hour, respectively (Figure 26). These results demonstrate that the initial rate of fatty acid turnover in the mutant was approximately 70 % higher that in wild type. 80 nmoles 14C in fatty acids/mg Chl time (hours) Figure 26. Fatty acid turnover in wild type ( O ) and fatb-k0 ( O ) leaves. Intact plants were pulsed for 30 min with 2 mCi of 14C02 and fatty acids chased for a period of 5 days in a non-radioactive atmosphere at 22° C. At the times indicated leaf samples were harvested and the fatty acids purified and their radioactivity measured as described in “Materials and Methods”. Analysis of polar lipid classes in wild type and fatb-k0 leaves To investigate whether higher rates of fatty acid turnover in fatb-k0 leaves affected the tunover of particular membrane glycerolipids, we analyzed leaf glycerolipid classes after pulsing wild type and fiztb-ko seedlings with 14CO; and chasing lipids for different times 154 up to 120 h. The results shown in Figure 27 indicate that the distribution of radiolabeled plastidial and extraplastidial glycerolipids was similar between wild type and fatb-k0 leaves. The label in phosphatidylcholine (PC) from wild type and fatb-k0 declined by approximately 65 % in the 0-120 h period (Figure 27). This high rate of label disappearance in PC is explained by the donation of diacylglycerol moieties (DAG) from PC to galactolipids and sulfolipids in chloroplasts (Roughan, 1970). Phosphatidylethanolamine (PE) was the only glycerolipid analyzed with lower label in fatb-k0 than wild type leaves (Figure 27). In wild type tissue, PE accounted for approximately 7 to 9 % of labeled lipids throughout the 5 days period whereas in the mutant represented only 5 to 6 %. Moreover, these levels remained constant throughout the chase period (Figure 27). This observation agreed with the lower steady state amounts of PE observed in fatb-k0 leaves compared to wild type (Bonaventure et al., 2003). Lower incorporation of radioactivity in PE of the mutant was also observed at early time points during continuous labeling of leaves with [HQ-acetate (data not shown). These data suggest reduced synthesis rather than increased degradation of this phospholipid in fatb-k0 leaves. In the case of plastidial lipids, the percentage of label in monogalactosyldiacylglycerol (MGDG) and digalactosyldiacylglycerol (DGDG) increased by approximately 25 % and 155 Percentage of radioactivity Percentage ofradioactivity time (hours) Figure 27. Redistribution of radioactivity among glycerolipid classes of wild type ( O ) and fatb-k0 ( O ) leaves. Seedlings were pulsed with 14C02 and leaf glycerolipids separated by TLC and their activities quantitated as described in “Materials and Methods”. Abbreviations: PC, phophatidylcholine; PE, phosphatidylethanolamine; MGDG, monogalactosyldiacylglycerol; DGDG, digalactosyldiacylglycerol; PG, phosphatidylglycerol. 150 % respectively in both wild type and fatb-k0 during the chase period (Figure 27). These increments in labeled MGDG and DGDG at longer times represented the contribution of the eukaryotic pathway to the plastidial pool of these lipids (Browse et al., 156 1989). These data together with the decline of labeled PC are consistent with the parallel operation of the prokaryotic and eukaryotic pathways of lipid biosynthesis in wild type and fatb—k0 leaves. In contrast, the percentage of label in phosphatidyldiacylglycerol (PG), which is mainly produced in the plastid, declined in both plant classes by approximately 30 % (Figure 27). The kinetics of l4C labeling in the different lipid classes of wild type Arabidopsis leaves were in accordance with previous studies in which the same tissue was labeled with 14C- acetate (Wu et al., 1994; Browse et al., 1989; Kunst et al., 1989). In conclusion, the similarity in the distribution of label among lipid classes in wild type and fatb-k0 leaves indicate that despite higher rates of fatty acid turnover cells maintained a constant composition of polar lipids. Analysis of radiolabeled C16 and C13 fatty acids in polar lipids of wild type and fatb- ko leaves The fatty acid composition of leaf membrane glycerolipids is different between wild type andfatb-ko. In this regard, PE, PS and P1 in the mutant have reductions of approximately 50 % in their 16:0 content compared to wild type, whereas in PC the reduction is 80 % (Bonaventure et al., 2003). To understand how the lack of FATB activity affects fatty acid partitioning into leaf glycerolipids, we evaluated redistribution of labeled fatty acids in these lipids. Wild type and mutant seedlings were pulsed with 14CO; and radiolabeled fatty acids in polar lipids chased for different times up to 120 b (Figure 28). In order to obtain an accurate estimate of 14C18 and 14C”, levels, fatty acids were first hydrogenated 157 ~ I.‘ l; to remove double bonds and then separated by reverse phase-thin layer chromatography (TLC) according to their chain lengths. The results of the analysis revealed that in the mutant, redistribution of labeled C13 and Cu, fatty acid was similar to wild type in plastidial lipids but differed substantially in extraplastidial lipids (Figure 28). In PC of wild type leaves, labeled C16 fatty acids increased by 25 % in the 0-120 h period (from 20 % to 25 %) whereas in PC of the mutant they decreased by 60 % (from 15 % to 6 %) (Figure 28). These results indicated that in PC of fatb-k0 leaves, C16 fatty acids were first incorporated at 30 % lower levels than wild type and later removed to reach 75 % lower levels than wild type. Moreover, PC in the fatb-k0 lost almost 50 % of its labeled C16 fatty acids in 6 hours, suggesting a rapid removal of this fatty acid from PC (Figure 28). Therefore, the ~80 % reduction in the amount of C 10 in PC of fatb-k0 compared to wild type (Bonaventure et al., 2003), results from both reduced incorporation but mainly from higher removal of this fatty acid (Figure 28). Labeled C10 fatty acids in PE redistributed differently from those in PC. In both, wild type and fatb-k0 leaves their levels decreased after the initial incorporation (Figure 28). 158 MGDG k.- Percentage ome fatty acids s. Percentage 0me fatty acids 0 6 12 24 48 72 time (hours) 40 i 1 i r i r in r i I i r i 0 6 12 24 48 72 96 120 time (hours) Figure 28. Redistribution of radioactivity in C16 fatty acids of membrane glycerolipid classes from wild type ( O ) and fatb-k0 ( O ) leaves. Seedlings were pulsed with 14C02, lipid classes separated by TLC and their corresponding fatty acids transmethylated, hydrogenated and separated by reverse phase TLC. Abbreviations: PC, phosphatidylcholine; PE, phosphatidylethanolamine; MGDG, monogalactosyldiacylglycerol; DGDG, di galactosyldiacyl glycerol; PG, phosphatidyl glycerol. 159 Nevertheless, PE from mutant leaves incorporated lower levels of C u, fatty acids than wild type (Figure 28). Thus, in contrast to PC, the steady state differences in the C16 content of PE between wild type and fatb-k0 was primarily the result of lower incorporation instead of increased removal of this fatty acid (Bonaventure et al., 2003). In contrast to extraplastidial glycerolipids, the incorporation and redistribution of label in C18 and C16 fatty acids of plastidial glycerolipids was similar between wild type and fatb- ko (Figure 28). These observations agreed with previous studies in which no significant changes in the fatty acid composition of MGDG, DGDG and PG were found between wild type and the mutant (Bonaventure et al., 2003). These results also suggest that the ratio of degradation or interconversion to other components of C13 and Cu, fatty acids from MGDG, DGDG and PG was not altered in fatb-k0 compared to wild type. In both, MGDG and DGDG, the C16 fatty acid levels declined in accordance with their dependence on the prokaryotic and eukaryotic pathways for their synthesis (Figure 28). Moreover, the more pronounced decline of labeled C16 in DGDG reflected the higher dependence of this lipid on the eukaryotic pathway (Browse et al., 1986). In the case of phosphatidylglycerol (PG), the situation was different from the other major plastidial lipids and reflected its primary synthesis by the prokaryotic pathway (Browse and Somerville, 1991). The level of C16 fatty acids incorporated into this phospholipid at time 0 was 55 % and increased to more than 60 % at 120 h (Figure 28). These data suggest that Cu, fatty acids in PG were more stable than C18 fatty acids. Mass distribution of C133C13 and szClg molecular species of polar lipids in wild type and fatb-k0 leaves 160 _ .T'm'mf Arabidopsis membrane glycerolipids are composed of either C163C18 or C182C13 fatty acids in a 1:1 molar ratio. Whereas lipids derived from the prokaryotic pathway are predominantly C1(,:C13, those derived from the eukaryotic pathway are either C lg:C .8 or ClozCrx (Browse and Somerville, 1991). In order to better understand the fluxes of C13 and CH, fatty acids into lipids and their degradation or interconversion to other lipids, the decay in the mass of labeled fatty acids was calculated for the different molecular species of polar lipids (Figure 29). The results shown in Figure 29A indicate that in wild type leaves, PC lost 4.8 nmol l“C/mg, Ch] of c.8138 and 3.3 nmol 14C/mg (3111 of (3,613.8 during the 6-120 h time frame. In parallel, MGDG and DGDG together gained 3.5 nmol l4C/mg Chl of C13:C13 and lost 12.4 nmol l4C/mg Chl as C162C13 (Figure 29A). Thus, gain of Clnglg species in MGDG and DGDG could be accounted for by transfer of these molecules from PC. The remaining 1.3 nmol 14C/mg Chl of C132C18 lost from PC could either be degraded in the cytosol or interconverted into other lipid-derived components (e.g., membrane lipids). Thus, in wild type, most of C182C13 loss in PC (73 %) was via MGDG and DGDG biosynthesis in the plastid (Figure 30). The fate of lost szClg species in PC could be explained by trafficking of these molecules back to the plastid, degradation in the cytosol or interconversion into other molecules (Figure 30). 161 ”15"“ A. wild type I)’. fatb-I10 PC PC 20 [j C16:C18 10 0 __ 20 7 PE PE 10 20 i DGDG DGDG 10 nmol l4C lmg Chl 20 20- 10 0 .; N v o v- N 19 N60 NON .— Time (hours) Figure 29. Redistribution of C13: C13 and C16:C13 molecular species of polar lipids in wild type and fatb-k0. Absolute masses for radiolabeled C13: C13 and C162C13 species were calculated using total label in fatty acids, polar lipid abundance and percentage of radiolabeled C13 and C16 in the different polar lipids. Abbreviations: PC, phosphatidylcholine; PE, phosphatidylethanolamine; MGDG, monogalactosyldiacylglycerol; DGDG, digalactosyldiacylglycerol; PG, phosphatidyl glycerol. 162 In fatb-k0 leaves, PC lost 10.4 nmol MC/mg cm of cure18 and 2.4 nmol MC/mg of C1sz 18 in the 6-120 h period (Figure 29B). MGDG and DGDG together gained 3.1 nmol l4C/mg Chl of C182C18 and lost 21.9 nmol l4C/mg Chl of CmiClg (Figure 29B). Thus, from the total mass of labeled Clnglg lost from PC only 3.1 nmol 14C/mg Chl contributed to galactolipid synthesis whereas 7.3 nmol l4C/mg Chl were not accounted for. Thus, these results indicate that in contrast to wild type, the major fraction (70 %) of C132C13 lost from PC was not incorporated into galactolipids in the mutant. Importantly, this fraction represented 62 % of the label lost from extraplastidial lipids in fatb-k0 whereas only 18 % in wild type (Figure 30). The fate of this mass of labeled C182C13 together with lost 06sz from PC could be degradation in the cytosol, interconversion to other cellular components or trafficking to the plastid in case of the latter (Figure 30). A second important observation of these experiments was that disappearance of C162C13 species from MGDG, DGDG and PG represented the main loss of mass from labeled fatty acids in wild type and fatb-k0 leaves. In leaves of the former, total loss of labeled fatty acids was 23.4 nmol l4C/mg Chl in the 6-120 h period (Figures 1 and 5). Disappearance of C16:C13 species from MGDG, DGDG and PG accounted for 16.4 mnol l4C/mg Chl, and corresponded to ~70 % of total mass loss (Figure 30). In the mutant, the total loss of labeled fatty acids was 40.8 nmol l4C/mg Chl in the same period and in this case 29 nmol l4C/mg Chl were lost through plastidial Cloiclg species (~70 % of total mass loss) (Figure 26 and 5). These results indicate that a major fraction of labeled fatty 163 acids was eliminated by turnover of C162C18 Plastidial species and that the percentage of total label loss by this route was similar between wild type and fatb-k0 (~70 %). In summary, these results demonstrate that increased fatty acid synthesis in fatb-k0 was compensated by increased absolute rates of lipid turnover in a way that preserved the ratio of polar lipids. Furthermore, the results suggest that in the mutant the surplus of fatty acids exported from plastids and incorporated into polar lipids is lost at the level of PC via degradation or interconversion of this phospholipid into other components. In addition, this higher degradation/interconversion rate of PC in fatb-k0 leaves may explained why Cu, fatty acids are rapidly removed from this phospholipid and accumulate at levels 80 % lower than wild type, whereas in the rest of the extraplastdial lipids they accumulate at levels 50 % lower than wild type (Bonaventure et al., 2003). Immunoblot analysis of wild type and fatb-k0 leaves. To determine if the increase in the rate of fatty acid biosynthesis in fatb-k0 leaves was correlated with an up-regulation of FAS protein expression, specific antibodies against BCCP (biotin carboxylase carrier protein), acyl carrier proteins (ACP) and stearoyl-ACP desaturase were used for immunoblot analysis of protein extracts from wild type and fatb-k0 leaves. As shown in Figure 31 protein levels of the BCCP subunit of plastidic acetyl-COA carboxylase (ACCase) were increased by 1.5-fold in fatb-k0 leaves compared to wild type. Similarly to BCCP, the protein levels of ACPS and stearoyl-ACP desaturase were induced by approximately 2—fold in the mutant (Figure 31). Arabidopsis leaves 164 I express several isoforms of plastidic ACPS with ACP-4 as the most abundant in this tissue followed by ACP—2 and 3 (Hlousek-Radojcic et al.. 1992). The immunoblot results Total +42 wt +60fatb I 1 plastid Croicrs -l6.4 Wt -29fatb Degradation? : lnterconversion? 1 1‘ Cisicrs Crsicrs - 3.3 Wt - 1.3 wt 9.41am M l——_. C18:C18 C16:Cl8 - 2.4 wt - 0.3fatb - 1.8 atb Figure 30. Scheme of mass loss from labeled fatty acids in leaves of wild type and fatb-k0. The arrows depict the flux of either fatty acids (16:0/18:1) exported from the plastid or €111:ng and C 162C”; in the different lipids (gray: wild type (wt); black:fi1tb-ko (fatb)). The values inside the boxes are nmol l4C/mg Chl of C13: Cu; and ClézClg species. Initial positive values represent total label input at time 6 hours and negative values loss of labeled C13: C 13 and CmICIg species after 120 hours. Loss of label from galactolipids and PG most likely represent degradation. Loss of labeled C16:C13 from PC could not be assigned to a specific pathway, it may occur by conversion to galactolipids, degradation or interconversion to other cellular components. Similar fates can be assigned to lost ClgICm-PC. C lg:C jg-PE and C miClg-PE. Abbreviations: GaL: galactolipids (MGDG and DGDG); PC, phosphatidylcholine; PE, phosphatidylethanolamine; PG, phosphatidyl glycerol, FAS: fatty acid synthesis pathway. 165 indicated that most of ACP isoforms were up—regulated in leaves of the mutant (Figure 31). In contrast, a single band in the blot probed with anti-stearol-ACP desaturase antibody appeared up-regulated in fatb-k0 leaves, suggesting that one major leaf isoform of this enzyme was induced (Figure 31). Changes in the expression of some FAS proteins occur during leaf development, being higher in young leaves and declining after this tissue completes its expansion (J. Ohlrogge, unpublished results). Thus, to determine whether the increased levels of BCCP, ACPS and 18:0-ACP desaturase in fatb-k0 were the result of differences in the developmental stage of wild type and mutant leaves, immunoblot analysis was also performed on leaf extracts at different stages of development (2, 3 and 4 weeks old seedlings). The results showed a consistent increase (1 .3-2 fold) in the levels of the three FAS proteins in fatb-k0 leaves compared to wild type at the different stages of development (data not shown). DISCUSSION Acyl-ACP thioesterases (FAT) are responsible for the export of fatty acids from the plastid produced by the de novo fatty acid synthesis. In this study, isotope labeling experiments and western blots were performed to investigate changes in lipid metabolism and protein expresSion brought about by disruption of the FATB gene in Arabidopsis (Bonaventure et al., 2003). These experiments suggest that fatb-k0 alters its fatty acid and 166 sneer-‘91 glycerolipid metabolism together with FAS protein expression to compensate for the deficiency in saturated fatty acids. However, higher FAS rate in fatb-k0 increases the production and export from the plastids of 18:1 over 16:0 fatty acids (Bonaventure et al., fatb-k0 wild type A 3.75 7.5 15 30 3.75 7.5 15 30 (ug) BCCP -5 "— ‘- —-— -—-- ACPS M., 18:0-ACP , ....... Desaturase 2.5 B 2.0 Fold increase 1.5 (fatb—k0 vs wt) 1.0 0.5 0.0 i BCCP ACPs 18:0-Aer Desaturase Figure 31. Immunoblot analysis of BCCP, ACP and 18:0-ACP desaturase in leaf tissue of wild type and fatb-k0 Arabidopsis. A, total Arabidopsis protein was extracted from leaves of wild type and mutant and increasing amounts loaded on the gel (3.75, 7.5, 15 and 30 ug). BCCPs were detected with anti-biotin antibodies, ACPs and 18:0-ACP desaturases with specific antibodies against spinach ACP-I and avocado stearoyl-ACP desaturase, respectively. Relative molecular weight for BCCP-l is 35 kD, for ACP 10 kD and for 18:0-ACP desaturase 35 kD. When known, major leaf isoforms are noted; B, amount of each polypeptide was quantitated with ImageQuant software and is expressed as relative fold increase of mutant versus wild type. The fold values shown are the average of the fold values for each protein concentration and the bars denote the standard deviation of the average. Fold increase values correspond to BCCP-l, all ACPS detected and major 18:0-ACP desaturase band. The intensity of coomasie—blue stained bands was used to normalize western blot signals. 167 2003) and the resulting excess of C18 unsaturated fatty acids may surpass the capacity of cells to use them for membrane biosynthesis. This surplus leads then to increased fatty acid turnover (Figure 32). Importantly, although fatb-k0 leaves undergo substantial changes in fatty acid and glycerolipid metabolism, the cells maintain a constant composition of polar lipids. These observations demonstrate the importance of lipid homeostasis in plant cell membranes and how cells adapt glycerolipid metabolism to alternative fatty acid fluxes. Similar results were obtained with the Arabidopsis act] mutant in which a major disruption of the prokaryotic pathway for lipid biosynthesis does not affect significantly the lipid composition of the plant (Kunst et al., 1989). Induction of a futile cycle of fatty acid production and degradation is also observed when the california bay 12:0-ACP thioesterase (MCTE) is over-expressed in developing Brassica napus seeds (Eccleston and Ohlrogge, 1998). In this system, the re-direction of fatty acids towards 12:0 reprogram fatty acid metabolism in a way that both the catabolic and biosynthetic pathways are increased to remove the surplus of 12:0 and to maintain normal levels of C16 and C13 fatty acids, respectively. Thus, although by means of different mechanisms, the missexpression of acyl-ACP thioesterases either in mutant or transgenic plants has major impacts in lipid metabolism and triggers cellular responses to lessen those effects. These observations demonstrate the central role of acyl-ACP thioesterases in balancing and providing fatty acids for several different cellular processes. Increased rates of fatty acid biosynthesis and turnover in fatb-k0 168 11W! 3 Different labeled substrates have been used to assess the in vivo rate of fatty acid biosynthesis in plants, with [l-MC]-acetate, l4COz and 3H20 the most commonly used (Bao et al., 2000; Pollard and Ohlrogge, 1999; Browse et al., 1981; Jungas, 1968). The advantages of using labeled water over acetate or carbon dioxide to determine in vivo rates of product biosynthesis is that water crosses all membranes and rapidly equilibrates with all body water pools (Kelleher, 2001). Moreover, using labeled water eliminates the problem of the dilution of carbon tracers in inaccessible pools and also of different metabolite pool sizes in wild type versus mutant plants (Kelleher, 2001). For example, initial differences in the intracellular pools of acetate between wild type and mutant organisms could result in differential dilution of the tracers inside cells and therefore in different specific radioactivities in labeled fatty acids (Nunn et al., 1977; Cronan et al., 1975). A second important parameter when performing labeling experiments with plants is the use of intact plants versus tissue sections (e. g., leaf discs). A drawback of using the latter is that some artifacts may be introduced by induction of wounding responses that alter fatty acid metabolism (Nishiuchi et al., 1997; Conconi et al., 1996). Therefore, the observation that three different labeled substrates gave consistent increases in FAS rate of fatb-k0 compared to wild type, demonstrates that this mutant increases the rate of fatty acid production in leaves. Moreover, the use of cut and intact tissue shows that this effect is independent of differential wounding responses between the mutant and wild type. An important conclusion of these results is that fatty acid production appears not to limit the amount of total membrane lipid biosynthesis and consequently grth of fatb-k0 plants. Thus, the slower grth of the mutant is most likely the consequence of a 169 Figure 32. Comparative scheme of fatty acid metabolism in wild type and fatb-k0. FAS activation signals are triggered by mechanisms that sense low levels of saturated fatty acids in the cytosol. The same mechanisms may induce the up-regulation of FAS protein expression by transcriptional or post-transcriptional mechanisms. In addition, biochemical and metabolic activation of the FAS machinery are also considered. 170 553293 598.5 mam— $ 7 2:3 mtg.“ 68.255 .3 m_o>o_ >5..— . x 101 mU< Twfi AU< cum: ....... 9 a. \ 3% \ as: are b hw>¢=h5r~1 )11 011.11 (1 [11111111 33M 1' 1 3001110; 9 mam—13:3 j e (6 p Ave. .3: H b m9... 3: a \ c 4 3% \ "022“ base __ ho>¢=h=rF Figure 32 171 reduced supply of saturated fatty acids for cellular processes other than membrane lipid biosynthesis. However, changes in membrane properties as a result of changes in their fatty acids composition can not be ruled out at this point as a mechanism affectting growth of fatb-k0. Wild type and fatb-k0 leaves presented an average rate of fatty acid turnover of 9 % and 14 % per day respectively (Figure 26). Bao et al., (2000) previously reported a turnover rate of 4-5 % per day in wild type Arabidopsis plants. The higher rate obtained in our experiment may be the result of using different growth conditions and plant ages. In particular, the lightzdark period in our experiment was 18:6 hours whereas it was 13.52105 hours in the study by Bao et al., (2000). During light periods the rate of fatty acid turnover is accelerated and therefore longer light periods will increase decay of labeled fatty acids (Bao ct al., 2000). In addition, because fatb-k0 seedlings have slower growth rate than wild type (Bonaventure et al., 2003) and the decay of labeled fatty acids was followed for a period of 5 days, differences in the net accumulation of fatty acids per seedling were significant between wild type and mutant plants. Wild type seedlings increased by 2-fold their fatty acid content whereas fatb-k0 seedlings by 1.88-fold in this period of time (data not shown). Thus, the differences between the fatty acid turnover rates of wild type and fatb- ko (either 55 % when the average is considered or 70 % when the initial rate is considered) are slightly underestimated. 172 Turnover of Polar Lipids in wild type and fatb-k0 Analysis of the absolute turnover rates of C132C13 and C16:C18 molecular species of polar lipids reveal that after 5 days of chase period, approximately 70 % of the label lost in both wild type and fatb-k0 leaves was as C102C18 species of galactolipids and PG (Figures 4 and 5). This observation agrees with the fact that this pool of lipids contain more than 70 % of the fatty acids in leaves. Moreover, the similar percentage of label lost via galactolipids and PG in wild type and fatb-k0 firrther demonstrate the conservation of lipid homeostasis despite higher absolute rates of lipid turnover in the mutant. One intriguing question is why PC in fatb-k0 leaves has a bigger reduction in 16:0 content compared to the rest of extraplastidial glycerolipids (Bonaventure et al., 2003). In this study we demonstrate that C16 fatty acids are rapidly removed from PC afier incorporation in the mutant whereas they accumulate in wild type (Figure 28). One possibility is that the increased turnover of PC in the mutant is not selective for the type of fatty acid and therefore Clnglg as well as szClg species of this phospholipid are either degraded or interconverted to other components. Alternatively, more Clszlg species from PC may be used for plastid lipid biosynthesis in the mutant or C10 fatty acids selectively removed from PC to supply other cellular processes (e.g., sphingolipid synthesis or acylation reactions). The kinetics of fatty acid labeling also indicate that the ratio of flux of total fatty acids (C18 plus C16) through the prokaryotic and eukaryotic pathways is not substantially 173 altered in mutant leaves compared to wild type. Hence, the ratio of the distribution of initial masses of l4C-fatty acids into the different polar lipids between these two pathways was similar when wild type and fatb-k0 were compared (data not shown). Moreover, the data agree with previous results obtained from lipid biochemical analysis (Figure 25 in Chapter 3 and Figure 5 in Bonaventure et al., 2003). Up-regulation of F AS protein expression Disruption of the FATB gene in Arabidopsis has provided new insights into some of the regulatory mechanisms that affect expression of F AS enzymes. What are the mechanisms involved in up-regulation of BCCP, ACPS and 18:0-ACP desaturase protein levels in fatb-k0 leaves? Are there other fatty acid proteins up-regulated in fatb-k0 leaves? The use of large scale techniques for mRNA profiling will help to investigate whether up- regulation of FAS protein expression occurs at the mRNA level and also to identify additional FAS transcripts with altered expression in mutant versus wild type leaves. Moreover, this analysis could give additional information on changes in the expression of non-lipid genes that may help to understand changes in overall metabolism and perhaps to identify regulatory genes (Ruuska et al., 2002; Girke et al., 2000). Conclusions and Future Perspectives The rate of FAS in fatb-k0 leaves is increased by 40 % compared to wild type and this increment may be partially achieved by up-regulation of FAS protein expression. The 174 resulting increased supply of fatty acids above amounts needed for leaf growth leads to increased fatty acid turnover. Importantly, despite major changes in fatty acid and glycerolipid metabolisms in the mutant a constant composition of polar lipids is maintained in leaves. The increase in fatty acid synthesis and up-regulation of BCCP, ACPs and 18:0-ACP desaturase protein levels suggest that plant cells have mechanisms capable of sensing subnormal levels of saturated fatty acids and signaling the activation of the FAS machinery and protein expression in order to increase their production. The understanding of the mechanisms responsible for activation of fatty acid synthesis and lipid turnover as well as up-regulation of FAS protein expression in fatb-k0, may provide important clues for the understanding of the regulation of fatty acid and lipid synthesis in plants. MATERIALS AND METHODS Plant Material and Growth Conditions In all the experiments wild type Arabidopsis thaliana and fatb-k0 mutant plants (ecotype Wassilewska) (Bonaventure et al., 2003) were grown on a mixture of soilzvermiculite (1:1) for three weeks under white fluorescent light (100 nmol m"2 s") in a 18h:6h lightzdark photoperiod at 22°C in growth chambers. Sowed seeds were always stratified for four days at 4°C. Rate of Fatty Acid Biosynthesis by Arabidopsis Leaves 175 Rapidly expanding leaves from wild type and fatb-k0 plants (3 weeks old) were cut in strips (0.5 cm wide) and transferred to pre-weighed glass flasks containing 4.75 mL of incubation buffer (2.5 mM sodium MES (2-[N-Morpholino] ethanesulfonic acid) pH: 5.7, 0.0075 % (w/v) Tween-20 and 2.15 mg/mL of MS salts. Leaf strips from the same plant were randomly distributed in different flasks to ensure sample homogeneity between flasks. After sufficient samples had been prepared (approximately 0.2 g fresh weight per flask), the flasks were re-weighed to obtain gram fresh weight (gfw) values. The assay started by the addition of either 0.25 mL of3H20 (100 mCi/mL, 3.7 GBq/mL) (ARC Inc, American Radiolabeled Chemicals Inc; St. Louis, MO) or 0.025 mCi of [l-MC] sodium acetate (56 mCi/mmol) (ARC Inc, American Radiolabeled Chemicals Inc; St. Louis, M0) to each flask and incubating them for different times at 22°C in a temperature-controlled water bath with gentle agitation and continues illumination. All data points were performed in duplicate and the values presented in Table 13 represent initial rates of fatty acid synthesis (calculated using data from 0, 10, 20, 40, 60 min of continuous labeling). At the end of the assay period the incubation medium was removed and the tissue quickly washed twice with de-ionized water and quenched by heating in 10 mL of isopropanol for 10 min at 80°C. Lipids were extracted with hexane-isopropanol method (Hara and Radin, 1978). An aliquot of the lipid extract was suspended in acetone:water (4:1, v/v) to determine chlorophyll content (Amon, 1949). The methodology used to determine the rate of fatty acid biosynthesis using 14CO; was identical to the methodology used to determine the rate of fatty acid turnover (see below) with the only difference that plants were removed at different time points from the sealed bag using a double sealed air trap. 176 Rate of Fatty Acid Turnover by Arabiodpsis Leaves Wild type and fatb-k0 plants were grown for 3 weeks as indicated above. One day prior to the experiment, a total of 12 pots (6 pots with wild type and 6 withfatb-ko plants (15 plants per pot)) were randomly placed inside a transparent glove bag (40 litre gas space, 12R® Instruments for Research and Industry, Cheltenham, PA) with circulating air and same lighting and temperature conditions as indicated above. A 30 min pulse of 14C02 was given to the plants by mixing 2 mCi of 14C—NaHCO3 (56 mCi/mmol) (ARC Inc; St. Louis, MO) with concentrated sulfuric acid inside the sealed bag and air circulated by using a small battery-driven fan. For the chase period, the radioactive atmosphere was rapidly vented and the plants were placed in a normal (non-radioactive) atmosphere for different times in the same growth conditions as described above. At each time point , 15 wild type and 15 fatb-k0 plants were randomly removed from different pots and separated in two individual samples (7-8 plants/sample for wild type and mutant). Leaf tissue was immediately weighed, frozen in liquid nitrogen and stored at -80°C for subsequent lipid extraction with hexane—isopropanol method (Hara and Radin, 1978) and for chlorophyll content determination (Amon, 1949). Fatty Acid Analysis of Lipid Extracts from Leaf Tissue Transmethylation of fatty acids from total lipid extracts was performed similarly for all the experiments. An aliquot of lipid extracts was heated at 90°C for 1 hour in 1 mL of 177 10% (v/v) boron-trichloride/methanol (Sigma). After acidification with aqueous acetic acid, fatty acid methyl esters (FAMES) were extracted two times with hexane and radioactivity in the sample (either 3 H or 14C) analyzed by scintillation counting (Beckman Instruments Inc, Fullerton, CA). A second aliquot of total lipid extract from each sample was transmethylated and F AMES were separated by thin layer chromatography (TLC) on K6 silica plates (Whatman Inc, Clifton, PA) using 90/10 (v/v) hexane/diethyl-ether. Radioactive bands corresponding to FAMES were localized by scanning in an Instant Imager system (Packard, Meriden, CT). The bands corresponding to FAMES were recovered from the plates and radioactivity measured by scintillation counting (Beckman Instruments Inc, Fullerton, CA). Individual Glycerolipid Analysis. Glycerolipid classes from total lipid extracts were separated by thin layer chromatography (TLC) on K6 silica plates (Whatman Inc, Clifton, PA) impregnated with 0.15 M ammonium sulphate and activated for 3 hours at 1100 C (Kahn and Williams, 1977). The TLC plates were developed three times with 91/30/8 (v/v/v) acetone:toluene:water and scanned for radioactivity using an Instant Imager (Packard, Meriden, CT), both to quantitate radioactivity and to locate the appropiate bands for recovery. Lipids were also detected after spraying with 0.2 % (w/v) 2’-7’- dichlorofluorescein/methanol and viewing under UV light. Standards were used to identify the different glycerolipid classes. Lipids were eluted from the silica with chloroform/methanol/water (5/5/1) and fatty acid methyl esters prepared as described 178 above. FAMES from the different lipid classes were hydrogenated using hydrogen at slightly greater than atmospheric pressure with a platinum (IV) oxide catalyst in methanol. The reaction was performed for at least 2 hours giving complete reduction of unsaturated to saturated FAMES. Analysis and isolation of 18 and 16 carbon FAMES were by KC18 reverse phase TLC (Whatman, Clifton, PA), developed half and then fully with acetonitrile/methanol/water (130/70/1, v/v/v). Plates were scanned for radioactivity using an Instant Imager (Packard, Meriden, CT) to quantitate radioactivity and to locate the appropiate bands for recovery. Scintillation counting was also used to measure radioactivity after scrapping-off the bands corresponding to 16 and 18 carbon fatty acids from the plates. SDS-PAGE and Immunoblotting Extraction of proteins from leaves was performed as follows. Up to 0.1 g of leaf tissue was harvested and placed into a 1.5 mL plastic microcentrifuge tube. Plant material was pulverized with a microcentrifuge pestle in the presence of liquid nitrogen. Powder was immediately reconstituted in 0.2 mL of extraction buffer (2 % (v/v) 2-mercaptoethanol, 50 mM HEPES (pH: 7.8), 100 mM NaCl, 0.05 % (w/v) SDS) by vortexing. Insoluble debris was collected by centrifugation for 15 min at 12,000g. Supernatant was removed and placed into a fresh microcentrifuge tube. Protein concentration was determined by dye-binding protein assay using bovine serum albumin as the standard (Bradford, 1976). 179 SDS-PAGE and protein transfer to nitrocellulose was performed using standard conditions. Nitrocellulose membranes were blocked for at least 1 h with 10 mM Tris- HCl, pH: 8.0, 0.15 M sodium chloride, 0.3 % (v/v) Tween 20 (TBS-T), and 2 % (w/v) nonfat dry milk. Anti-biotin antibodies conjugated to alkaline phosphatase (Kirkegaard and Perry Laboratories, Gaithersburg, MD) were directly added at a 1:5,000 dilution to detect BCCP. Antisera raised against avocado stearoyl-ACP desaturase (Shanklin and Somerville, 1991) and spinach acyl carrier protein-I (ACP-I) (Post-Beittenmiller et al., 1991) were added at a 1:1,000 dilution. Probing proceeded for 16 h at 4° C followed by 1 h at 25° C after which membranes were briefly rinsed with TBS-T. Antibody bound proteins were detected by incubating blots for l h at 250 C with alkaline-phosphatase- conjugated anti-rabbit secondary antibodies (Kirkegaard and Perry Laboratories, Gaithersburg, MD). The membranes were then washed 6 times for 5 min each with TBS- T. Blots were then washed 5 min with developing solution (0.1 M Tris-HCl, pH: 9.5, 0.1 M sodium chloride and 5 mM magnesium chloride) before colorimetric detection in developing solution containing 0.33 mg/mL p-nitro blue tetrazolium chloride and 0.17 mg/mL 5-bromo-4-chloro-3-indoyl phosphate p-toluidine salt. Immunoblot signals were quantitated using ImageQuant software (Molecular Dynamics, Sunnyvale, CA), which accounted for band area plus intensity. Blot signals were corrected for differences in protein loading (as determined by Coomasie Brilliant Blue R 250 staining) and normalized against values of the control plants. Increasing amounts of total protein were resolved by SDS-PAGE for immunoblot analyses to ensure that both major and minor bands were within the linear sensitivity range of the detection system. 180 LITERATURE CITED Arnon DJ. (1949) Copper enzymes in isolated chloroplasts: polyphenol oxidase in Beta vulgaris. Plant Physiol, 24: 1-15. Bao X., Focke M., Pollard M., Ohlrogge J. (2000) Understanding in vivo carbon precursor supply for fatty acid synthesis in leaf tissue. Plant J, 22: 39-50. Beisson R, Koo A.J., Ruuska S., Schwender J., Pollard M., Thelen J.J., Paddock T., Salas J.J., Savage L., Milcamps A., Mhaske V.B., Cho Y., Ohlrogge J.B. (2003) 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 Physiol, 132: 681-697. Bonaventure G., Salas J.J., Pollard M.R., Ohlrogge J.B. (2003) Disruption of the FATB gene in Arabidopsis demonstrates an essential role of saturated fatty acids in plant growth. Plant Cell, 15: 1020-1033. Bradford M.M. (1976) A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem, 72: 248-254. Branen J.K., Shintani D.K., Engeseth NJ. (2003) Expression of antisense acyl carrier protein-4 reduces lipid content in Arabidopsis leaf tissue. Plant Physiol, 132: 748-56. Browse J., Kunst L., Anderson S., Hugly S., Somerville C. (1989) A mutant of Arabidopsis deficient in the chloroplast 16:1/18:1 desaturase. Plant Phys, 90: 522-529. Browse J., Roughan G., Slack R. (1981) Light control of fatty acid synthesis and diurnal fluctuations of fatty acid composition in leaves. Biochem J, 196: 347-354. Browse J., Somerville C. (1991) Glycerolipid synthesis - biochemistry and regulation. Annu Rev Plant Physiol Plant Mol Biol, 42: 467-506. Browse J., Warwick N., Somerville C.R., Slack CR. (1986) Fluxes through the prokaryotic and eukaryotic pathways of lipid synthesis in the '16:3' plant Arabidopsis thaliana. Biochem J, 235: 25-31. Conconi A, Miquel M, Browse JA, Ryan CA (1996) Intracellular Levels of Free Linolenic and Linoleic Acids Increase in Tomato Leaves in Response to Wounding. Plant Physiol, 111: 797-803. Cronan J.E., Rock C.O. (1996) Biosynthesis of membrane lipids. In Escherichia colli and Salmonella: Cellular and Molecular Biology, F.C. Neidhart, R. Curtiss III, J.L. 181 turm- 1 h. Ingraham, E.C.C. Lin, K.B. Low, B. Magasanik, W.S. Reznikoff, M. Riley, M. Schaechter, and HE. Umbarger, eds. (Washington, DC: ASM Press), pp. 612-636. Cronan J.E., Weisberg L.J., Allen R.G. (1975) Regulation of membrane lipid synthesis in Escherichia coli: accumulation of free fatty acids of abnormal chain length during inhibition of phospholipid synthesis. J Biol Chem, 250: 5835-5840. Doermann P, Voelker T, Ohlrogge (1995) Cloning and expression in Escherichia coli of a novel thioesterase from Arabidopsis thaliana specific for long chain acyl-acyl carrier proteins. Arch Biochem Biophys, 316: 612-618. Eccleston V.S., Ohlrogge J.B. (1998) 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, 10: 613- 22. Hara A., Radin N.S., (1978) Lipid extraction of tissues with a low toxicity solvent. Anal Biochem, 90: 420-426. Hlousek-Radojcic A, Post-Beittenmiller D, Ohlrogge JB (1992) Expression of constitutive and tissue specific acyl carrier protein isoforms in Arabidopsis. Plant Physiol 98: 206-214. Jungas R.L. (1968) Fatty acid synthesis in adipose tissue incubated in tritiated water. Biochemistry, 7: 3708-3717. Kahn M.U., Williams J.P. (1977) Improved thin layer chromatographic method for the separation of major phospholipids and glycolipids from plant lipid extracts and phosphatidylglycerol and bis(monoacylglycerol) phosphate from animal lipid extracts. J Chromatogr, 140: 179-185. Kelleher JK (2001) Flux estimation using isotopic tracers: common ground for metabolic physiology and metabolic engineering. Metab Eng, 32100-10. Kolattukudy PE (2003) Polyesters in higher plants. Adv Biochem Eng Biotechnol, 71: 1-49. Kunst L., Browse J., Somerville C. (1989) A mutant of Arabidopsis deficient in desaturation of palmitic acid in leaf lipids. Plant Physiol, 90: 943-947. Lightner J., Wu J.R., Browse J. ( 1994) A mutant of Arabidopsis with increased levels of stearic-acid. Plant Physiol, 106: 1443-1451. Lynch D.V. (1993) Sphingolipids. In Lipid Metabolism in Plants, T.S. Moore, ed. (Boca Raton, FL: CRC Press), pp. 285-303. 182 11"“ mi. —I Mekhedov S., de Ilarduya O.M., Ohlrogge J. (2000) Toward a functional catalog of the plant genome. A survey of genes for lipid biosynthesis. Plant Physiol, 122: 389-402. Mou 2., He Y., Dai Y., Liu X., Li J. (2000) Deficiency in fatty acid synthase leads to premature cell death and dramatic alterations in plant morphology. Plant Cell, 12: 405- 418. Nandi A, Krothapalli K, Buseman CM, Li M, Welti R, Enyedi A, Shah J (2003) Arabidopsis sfd Mutants Affect Plastidic Lipid Composition and Suppress Dwarfing, Cell Death, and the Enhanced Disease Resistance Phenotypes Resulting from the Deficiency of a Fatty Acid Desaturase. Plant Cell, 15: 2383-98. Nishiuchi T, Hamada T, Kodama H, Iba K (1997) Wounding changes the spatial expression pattern of the arabidopsis plastid omega-3 fatty acid desaturase gene (FAD7) through different signal transduction pathways. Plant Cell, 9: 1701-12. Nunn WD, Kelly DL, Stumfall MY (1977) Regulation of fatty acid synthesis during the cessation of phospholipids biosynthesis in Escherichia coli. J Bacteriol, 132: 526-31. Ohlrogge J.B., Jaworski J.G. (1997) Regulation of fatty acid synthesis. Annu Rev Plant Physiol Plant Mol Biol, 48: 109-136. Ohlrogge JB, Kuhn DN, Stumpf PK (1979) Subcellular localization of acyl carrier protein in leaf protoplasts of Spinacia oleracea. Proc Natl Acad Sci U S A, 76: 1194-8. Pollard M.R., Ohlrogge J.B. (1999) Testing models of fatty acid transfer and lipid synthesis in spinach leaf using in viva oxygen-18 labeling. Plant Physiol, 121: 1217- 1226. Post-Beitenmiller D., Jaworski J.G., Ohlrogge J.B. (1991) In vivo pools of free and acylated acyl carrier proteins in spinach: Evidence for sites of regulation of fatty acid biosynthesis. J Biol Chem, 266: 1858-1865. Post-Beittenmiller D. (1996) Biochemistry and molecular biology of wax production in plants. Annu Rev Plant Physiol Plant Mol Biol, 47 : 405-430. Roughan P.G. (1970) Turnover of the glycerolipids of pumpkin leaves, the importance of phosphatidylcholine. Biochem J, 117: 1-8. Savage L.J., Ohlrogge J.B. (1999) Phosphorylation of pea chloroplast acetyl-COA carboxylase. Plant J, 18: 521-527. Schuller H.J., Hahn A., Troster F., Schutz A., Schweizer E. (1992) Coordinate genetic control of yeast fatty acid synthase genes FASl and FAS2 by an upstream activation site common to genes involved in membrane lipid biosynthesis. EMBO J, 11: 107-114. 183 “'1" 1m Shanklin J.S., Somerville C. (1991) Stearoyl-acyl—carrier-protein desaturase from higher plants is structurally unrelated to the animal and fungal homologs. Proc Natl Acad Sci USA, 88: 2510-2514. Shintani D.K., Ohlrogge J.B. (1995) Feedback inhibition of fatty acid synthesis in tobacco suspension cells. Plant J, 7: 577-587. Shintani D.K., Roesler K., Shorrosh B., Savage L., Ohlrogge J.B. (1997) Antisense expression and overexpression of biotin carboxylase in tobacco leaves. Plant Physiol, 114: 881-886. Somerville C., Browse J., Jaworski J.G., Ohlrogge J. (2000) Lipids. In B Buchanan, W Gruissem, R Jones, eds, American Society of Plant Physiologists, Rockville, Maryland, 456—527. Thelen J.J., Mekhedov S., Ohlrogge J.B. (2001) Brassicaceae express multiple isoforms of biotin carboxyl carrier protein in a tissue specific manner. Plant Phys, 125: 2016-2028. Voelker T.A., Jones A., Cranmer A.M., Davies H.M., Knutzon D.S. (1997) Broad- range and binary-range acyl-acyl-carrier protein thioesterases suggest an alternative mechanism for medium-chain production in seeds. Plant Physiol, 114: 669-77. Wallis J.G., Browse J. (2002) Mutants of Arabidopsis reveal many roles for membrane lipids. Prog Lipid Res, 41: 254-78. Wells G.B., Lester R.L. (1983) The isolation and characterization of a mutant strain of Saccharomyces cerevisiae that requires a long chain base for grth and for synthesis of phosphosphingolipids. J Biol Chem, 258: 10200-10203. Wu J.R., James D.W., Dooner H.K., Browse J. (1994) A mutant of Arabidopsis deficient in the elongation of palmitic acid. Plant Physiol, 106: 143-150. Yalovsky S, Rodriguez-Concepcion M, Gruissem W (1999) Lipid modification of proteins slipping in and out of membranes. Trends Plant Sci 4: 439-445. 184 Chapter 5 FINAL CONCLUSIONS, PERSPECTIVES AND FUTURE DIRECTIONS The work presented herein used several approaches to increase our understanding of the signals and regulatory mechanisms that influence fatty acid biosynthesis (FAS) in plants. These approaches range from gene expression analysis to detailed biochemical analysis of lipid metabolism in a mutant and the results have provided new insights into regulation of this primary metabolic pathway. In addition, work has focused on the partition of fatty acids from their source inside plastids to an array of biosynthetic pathways outside this organelle. One goal of plant biotechnology is to use these organisms as large scale chemical factories for production of edible and industrial products and as an alternative to petroleum-derived products (Thelen and Ohlrogge, 2002). The understanding of the regulatory mechanisms of fatty acid synthesis and partitioning should contribute to the development of better strategies to manipulate fatty acid production in plants. Conclusions and future directions on F AS gene expression in Arabidopsis Major conclusions of this thesis research related to FAS gene expression are: 1. Different environmental and metabolic signals regulate ACP expression in Arabidopsis depending on the ACP isoforms and their tissue specific expression. 185 2. Up-regulation of a subset of fatty acid and lipid genes and the steady expression of others appear to be sufficient to match new requirements for fatty acids in plant tissues. 3. Mechanisms that sense low levels of saturated fatty acids may signal the up- regulation of FAS proteins in Arabidopsis leaves. In Chapter 2, the genes encoding for acyl carrier proteins (ACPS) were selected for the initial experiments and results presented in that chapter demonstrate that regulation of these genes depends on the protein isoforms and their tissue specific expression. For example, ACP4 is the most abundant isoform in Arabidopsis leaves and it is regulated by light and carbon, similarly to photosynthetic genes. By contrast, the less abundant leaf isoforms ACP2 and ACP3 are not light-regulated but instead respond to growth, suggesting a more critical role for them in rapidly dividing tissue (e.g., meristems). One obvious question that arises from the analysis of ACP expression is whether ACP4 is the only fatty acid gene regulated by light in Arabidopsis. A recent report based on GeneChip analysis of phytochrome mutants and different lighting conditions indicates that at least 4 additional genes involved in fatty acid and lipid synthesis respond also to light (Tepperrnan et al., 2001). These genes include ACP4, chloroplast linoleate desaturase (FAD7), omega-3-desaturase (FAD3), chloroplast omega-6 desaturase (FAD6) and stearoyl-ACP desaturase. Since the microarrays used in that study covered approximately one third of the predicted genes in the Arabidopsis genome (Affymetrix 8K GeneChip), additional fatty acid and lipid genes may also be light regulated. 186 Are ACP2 and ACP3 the only fatty acid genes that are growth/cell cycle regulated in Arabidopsis? A recent report describes a GeneChip analysis of synchronized Arabidopsis cell suspension cultures used to identify cell-cycle controlled genes (Menges et al., 2002). The results indicate that only ACP2 and 3, two putative acyl-CoA synthetases and chloroplast omega-3-desaturase (F AD7-8) are under cell-cycle regulation in Arabidopsis. Again, only one third of the Arabidopsis genome was covered by the GeneChip used in that study and therefore the list of fatty acid and lipid genes regulated by cell-cycle may extend in the future. When an Arabidopsis cDNA microarray representing 3,500 genes was used to examine changes in gene expression during Arabidopsis seed development and in particular during the phase of oil accumulation (8-12 days after flowering), it was found that only a subset of genes involved in fatty acid and lipid metabolism showed significant changes in expression (Ruuska et al., 2002). For example, ACCase subunits, KAS I and II, ACP1, and F AD2 and 3 show peaks of expression between 8-12 days after flowering. In summary, it appears from these observations that rather than up-regulating all or most genes for fatty acid and lipid synthesis, plant cells up-regulate a limited number of them according to the tissue and its specific needs for fatty acids and lipids. Thus, instead of “global” regulatory mechanisms for all or most FAS genes, up-regulation of a subset of them and the steady expression of others seem to be sufficient to match new requirements for fatty acids in different tissue or under changing environmental conditions. Although 187 we focused our research on ACP genes, a logical extension can be applied to other fatty acid genes and we speculate that most likely all of them respond differentially to diverse signals and can be up-regulated according to tissue requirements (Menges et al. 2002, Ruuska et al., 2003; Tepperrnan et al., 2001). In Arabidopsis, most enzymes involved in fatty acid and lipid synthesis are encoded by multiple genes, and one intriguing question has been why Arabidopsis requires so many different FAS enzyme isoforms (Beisson et al., 2002). Their differential tissue expression and response to signals under variable conditions sheds some light on this question. Analysis of gene expression in fatb-k0 leaves, indicates that the protein levels of BCCP, ACPS and 18:0-ACP desaturase are increased between 1.5 to 2-fold in young leaves. Thus, these results suggest that mechanisms that sense low levels of saturated fatty acids can increase the expression of at least some F AS proteins. Therefore a new signal can be added to the repertoire of signals that, by still unknown mechanisms, activate fatty acid protein expression. In addition, it remains to be tested whether additional fatty acid gene products also respond to this regulatory mechanism. Thus, is there a global up-regulation of fatty acid proteins in fatb-k0 leaves or, are only a subset of critical genes up-regulated? An answer to this question will provide interesting information to correlate induction of specific FAS genes with increased rates of fatty acid production in leaves. Future directions Although most of the structural genes for fatty acid and lipid biosynthesis are known or have strong putative candidates, information about regulatory factors for this metabolic 188 pathway is still poor (Beisson et al., 2003; Mekhedov et al., 2000). Thus, one breakthrough step in plant fatty acid and lipid research will be to isolate regulatory factors for the expression of FAS proteins. One important consequence of the identification of these factors would be to manipulate lipid metabolism in genetically engineered plants. The advantage of expressing regulatory factors rather than single enzymes is that an entire metabolic pathway could be up-regulated and for example increase the yield of desired products (Gantet and Memelink, 2002; Memelink et al., 2001; Grotewold et al., 1998). Nevertheless, one difficulty presented by primary metabolism is that it often is not very amenable to metabolic engineering, in part because it operates in parallel with multiple pathways and therefore has complex interconnections with them (Carman and Henry, 1999). For example, carbon and nitrogen metabolism are closely interconnected and genes involved in nitrogen metabolism are regulated by carbon (Oliveira and Coruzzi, 1999). In addition, alteration of primary metabolism can cause unwanted secondary effects that have to be controlled before engineered plants can be used for commercial purposes (e.g., deformed morphology and size of seeds with reduced saturated fatty acids, Chapter 3-Figure 20). Thus, in order to successfully manipulate fatty acid metabolism in plants, it is helpful to know and consider the complex interconnections between metabolic pathways and their regulation. The studies performed in Chapter 2 open the possibility to identify transcription factors involved in regulation of fatty acid gene expression. One procedure to achieve this goal is to develop a system in which signal—specific (e.g., growth) changes in gene expression can be followed to perform functional promoter analysis (Sun et al., 2003; Chen et al., 189 2002; Fujiki et al., 2000). Thus, in Chapter 2 (Figure 15) a system to identify. elements involved in expression and regulation of the Arabidopsis ACP2 promoter by growth was developed. Further characterization (deletions and site-specific mutations) of this promoter in tobacco BY-2 cells will allow the identification of short DNA elements important for growth-regulation of this gene. Once these elements have been identified, the isolation of transcription factors can be approached by techniques such as yeast one- hybrid system, oligo-affinity chromatography, and electromobility shift assays (EMSA) (Carles et al., 2002; Uno et al., 2000; Jarrett, 1993). Keeping in mind our longer-term interest in increasing oil production in seeds, one evident question is whether the isolation of factors involved in regulation of ACP2 by growth will also be important for expression of ACP2 or fatty acid genes in seeds. On the one hand, the same transcription factor could be expressed in different tissues and have similar functions perhaps by association with alternative partners. On the other hand, completely different regulatory factors may be used in different tissues. In this case however, the ectopic expression in seeds of, for example, a meristem specific factor, could still have effects on fatty acid metabolism in seeds. Similarly, a functional promoter analysis of ACP4 can allow the identification of transcription factors involved in light-regulation of this gene. For this purpose an ACP4 promoter deletion series fused to luciferase can be created and analyzed in leaf tissue of transgenic plants. In parallel, the same plants can be used for the analysis of mechanisms involved in sugar-repression of the ACP4 gene. 190 um . 1‘ I What are the mechanisms involved in up-regulation of BCCP, ACPS and 18:0-ACP desaturase protein levels in fatb-k0 leaves? Is this up-regulation at the level of transcription or at downstream processes such as protein turnover? Are there other fatty acid genes up-regulated in fatb-k0 leaves? One experiment that may provide important information about increases in mRNA abundance of BCCP, ACPs and 18:0-ACP desaturase together with additional fatty acid genes is to compare transcript profiles between fatb-k0 and wild type leaves by microarray analysis. Moreover, this analysis could give additional information on changes in the expression of non-lipid genes that may help to understand changes in overall metabolism and perhaps to identify regulatory genes. One limitation of microarrays is that this technique only searches for differences in transcript abundance and information on translational control or protein degradation are not considered. In addition, transcript abundance for regulatory proteins is usually low and difficult to detect by this method (Vainrub and Pettitt, 2003; Watson and Akil, 2002). Thus, additional experiments will be required to obtain a more complete picture of changes in fatty acid gene expression including translational control (e.g., polysome association) and protein turnover. What is the role of the C T C CGCC box and polypyrimide tracts in the 5 ’ leader of A CP mRNAs? The results in Chapter 2 (Figure 9) indicate that the 5’UTR of ACPS have an active role on gene expression. One possible scenario is that specific factors bind to the leader and facilitate the interaction between ribosomes and ACP transcripts. Alternatively, a tertiary 191 “T‘mr‘ I -4!!! structure in the RNA may form to facilitate ribosome association and presumably translation initiation (Bolle et al., 1996; Staub and Maliga, 1994; Danon and Mayfield, 1991). Several approaches can be followed to answer these questions, for example the conserved elements in the ACP leaders can be replaced with random sequences to test how the association with polysomes and expression of a reporter gene are affected in transgenic plants. Second, the binding capacity of factors present in cytosolic fractions to ACP 5’UTRs can be investigated by electromobility-shift assays (EMSA) (Danon and Mayfield, 1991). Transcript stability can be analyzed by techniques such as mRNA turnover after inhibiting RNA Pol II activity with a-amanitin in seedlings grown in liquid culture (Hua et al., 2001). Conclusions and future directions on fatty acid metabolism and its regulation in fatb-k0. Major conclusions of this thesis research related to the role of saturated fatty acid are: 1. FATB is a major regulator of saturated fatty acid production in all tissues. 2. Saturated fatty acids have a critical role in plant growth and development. 3. Low levels of saturated fatty acids in the cytosol triggers signals to increase the rate of fatty acid biosynthesis in leaf plastids. Chapter 3 investigates the production and partition of saturated fatty acids from plastids to other cellular pathways. Analysis of the fatb-k0 mutant revealed that FATB is a major enzyme responsible for export of saturated fatty acids from plastids in all tissues. Moreover, reduction of cytosolic saturated fatty acid pools brought about by the lack of 192 if“..- FATB caused changes in the distribution of saturated fatty acids between different pathways that reveal a hierarchy in the metabolism of these molecules. For example, although there is more than a 50 % reduction in exported saturated fatty acids in the mutant, sphingoid base amounts in leaves were similar to wild type. In contrast, changes in wax amounts could be detected between these two plant classes. These observations are interpreted to indicate that maintaining sphingolipid levels is more essential than maintaining wax amounts for the plant. Some of the changes in saturated fatty acid- derived pathways have clearly negative effects in plant performance and seed development at standard growth conditions. Thus, saturated fatty acids have a critical role in plant growth and development and the understanding of the cellular steps affected will be important to manipulate fatty acid composition in plants without affecting their performance (see below). Analysis of fatty acid metabolism in fatb-k0 also gave new insights into the mechanisms and signals that increase the rate fatty acid biosynthesis in plant cells (Chapter 4). Low levels of saturated fatty acids in the cytosol not only triggers signals to up-regulate fatty acid protein expression but also to increase the rate of fatty acid biosynthesis. These results demonstrate that plants can uncouple rate of growth from rate of fatty acid production and therefore they have the inherent capacity to increase fatty acid production although growing at lower rates than normal. This is an important observation because it suggests that cells can be programmed to produce more fatty acids independent of a developmental program or rate of growth that sets the amount of fatty acids to be made. 193 Future directions What signal(s) activate fatty acid synthesis in plants? How is the rate of fatty acid synthesis increased? These are crucial questions that if answered will give important insights in the regulation of fatty acid biosynthesis in plants. This thesis research suggests that fatb-k0 provides a new system to tackle these questions. The rate of fatty acid biosynthesis in fatb—k0 leaves is increased by approximately 40 % compared to wild type and unraveling the underlying activation mechanisms could provide a breakthrough in understanding regulation of fatty acid biosynthesis. One critical issue to begin to answer these questions is to understand where in the cell the signal(s) to activate FAS are being generated. In Chapter 4, several mechanisms have been proposed to explain possible signal sources such as sensing of extraplastidial acyl-COA pools, rate of sphingoid base synthesis, rate of protein acylation, changes in membrane properties that affect transport or vesiculation. One possible approach to identify the actual mechanisms is to isolate suppressors of fatb-k0 (see below). What is the target of these activation mechanisms in fatb-k0 leaves? Is AC Case activity altered in fatb—k0 leaves? Several lines of evidence suggest that acetyl-COA carboxylase (ACCase) is a major regulatory enzyme in fatty acid production in plants and therefore it could be one target of the activation mechanisms (Ohlrogge and Jaworski, 1997). The B-CT subunit of ACCase is phosphorylated and this event is correlated with activation of its activity (Savage and Ohlrogge, 1999). Thus, analysis of the phosphorylation status of B—CT in 194 fatb-Ito and wild type leaves could provide information on ACCase as a target of the activation mechanism induced in the mutant. In addition, evidence for a negative feedback regulatory mechanism acting on ACCase comes from experiments where tobacco cells were incubated in the presence of exogenous fatty acids (Shintani and Ohlrogge, 1995). Could ACCase in fatb-k0 leaves be released from some sort of feedback control? It has been proposed that in bacteria ACCase is inhibited by acyl-COA products, and this inhibition can be released by overexpression of acyl-CoA thioesterases (Jiang and Cronan, 1994). However, plant ACCase appears to be non responsive to long acyl-ACP products in vitro (Roesler et al., 1996). Other potential feedback inhibitors of FAS enzymes could be acyl-CoA, free fatty acids and glycerolipids. Because FAS occurs inside the plastids but the major utilization of the products of fatty acid synthesis is at the ER membranes, it is possible that feedback regulation requires communication across the plastid envelop. Could cytosolic saturated fatty acids be feedback controllers of FAS synthesis? What is limiting growth of the fatb-k0 mutant? Different hypothesis can be proposed to explain the reduced growth of fatb—k0 mutant as outlined in Chapter 3. However, pleiotropic effects brought about by the genetic lession could be difficult to interpret and one approach to study this complex system is to identify suppresors of the grth phenotype of fatb-k0. The mutant grows slower than the wild type plants and identification of seedlings that performed better than the mutant could be identified in mutageneized fatb-k0 populations. Mutagenesis of fatb-k0 can be 195 performed chemically by using EMS or by T-DNA activation tagging. One advantage of using EMS is that it creates point mutations that could lead to amino acid substitutions and change protein properties without killing its activity (J ander et al., 2003; Me Callum et al., 2000). On the other hand, activation tagging finds suppressors by overexpression of targeted genes and therefore a different group of genes from those targeted by EMS is selected (Memelink, 2003; van der Fits et al., 2001). Suppresors of fatb-k0 may fall in different categories, for example genes that increase saturated fatty acids levels, genes involved in biochemical pathways different from FAS, genes involved in regulation of FAS, genes that participate in membrane transport or vesiculation, and changes in protein acylation or properties of acylated proteins. Thus, as mentioned above the suppressor approach could help to unravel both: mechanisms that limit fatb-k0 growth as well as regulatory mechanisms of fatty acid synthesis. For instance, if a particular cellular process is affected and limits growth, a mutation that bypass this limitation could reside in a regulatory protein of an interdependent process. To illustrate this idea, suppressor mutants of yeast deficient in Golgi vesicle formation were mapped in genes corresponding to transcription factors (HACl) and kinases (IREl) that are involved in the unfolded protein response (UPR) signaling pathway (Higashio and Kohno, 2002). Thus, activation of UPR indirectly stimulates ER to Golgi vesiculation and reconstitutes growth of yeast affected in this process. In a different experiment, yeast suppressors of sphingolipid biosynthesis where isolated and named SLC (sphingolipid compensation) (Lester et al., 1993). These strains are viable without synthesizing sphingoid bases and therefore lack ceramides. SLC mutants cells synthesize novel PI (phosphatidylinositol) substituted with long (26 carbons) fatty acids in the sn-2 position of glycerol. Thus, the 196 SLC suppressors overcome the essential function of sphingolipids by producing novel PI that structurally mimics sphingolipids. In conclusion, the identification of regulatory factors and the unraveling of the regulatory mechanisms responsible for FAS activation in fatb-k0 together with the origin and nature of the inducing signal(s) may provide essential clues for the understanding of the regulation of fatty acid and lipid synthesis in plants. LITERATURE CITED Baerson SC, Vander Heiden MG, Lamppa GK (1998) Overlapping yet distinct promoter activities of two Arabidopsis genes coding for nearly identical isoforms of the acyl carrier protein. Int J Plant Sci 159: 533-538 Baerson SR, Lamppa GK (1993) Developmental regulation of an acyl carrier protein gene promoter in vegetative and reproductive tissues. Plant Mol Biol 22: 255-67 Battey JF and Ohlrogge JB (1990) Evolutionary and tissue specific control of expression of multiple acyl carrier protein isoforms in plants and bacteria. Planta 180: 352-360 Beisson F, Koo AJ, Ruuska S, Schwender J, Pollard M, Thelen JJ, Paddock T, Salas JJ, Savage L, Milcamps A, Mhaske VB, Cho Y, Ohlrogge JB (2003) 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 Physiol, 132: 681-97 Bolle C, Herrmann RG, Oelmuller R (1996) Different sequences for 5'-untranslated leaders of nuclear genes for plastid proteins affect the expression of the beta- glucuronidase gene. Plant Mol Biol 32: 861-8 Carles C, Bies-Etheve N, Aspart L, Leon-Kloosterziel KM, Koornneef M, Echeverria M, Delseny M (2002) Regulation of Arabidopsis thaliana Em genes: role of ABIS. Plant J, 30: 373-83 Carman GM, Henry SA (1999) Phospholipid biosynthesis in the yeast Saccharomyces cerevisiae and interrelationship with other metabolic processes. Prog Lipid Res, 38: 361- 399 197 Chen PW, Lu CA, Yu TS, Tseng TH, Wang CS, Yu SM (2002) Rice alpha-amylase transcriptional enhancers direct multiple mode regulation of promoters in transgenic rice. J Biol Chem, 277: 13641-9 Danon A, Mayfield SP (1991) Light regulated translational activators: identification of chloroplast gene specific mRNA binding proteins. EMBO J, 10: 3993-4001 Gantet P, Memelink J (2002) Transcription factors: tools to engineer the production of pharrnacologically active plant metabolites. Trends Pharmacol Sci, 2: 563-9 Grotewold E, Chamberlin M, Snook M, Siame B, Butler L, Swenson J, Maddock S, Clair GS, Bowen B (1998) Engineering secondar y metabolism in maize cells by ectopic expression of transcription factors. Plant Cell, 10: 721-40 Hannapel DJ, Ohlrogge JB (1988) Regulation of acyl carrier protein messenger RNA levels during seed and leaf development. Plant Physiol, 86: 1174-1178 Higashio H, Kohno K (2002) A genetic link between the unfolded protein response and vesicle formation from the endoplasmic reticulum. Biochem Biophys Res Commun, 296: 568-74. Hlousek-Radojcic A, Post-Beittenmiller D, Ohlrogge JB (1992) Expression of constitutive and tissue specific acyl carrier protein isoforms in Arabidopsis. Plant Physiol, 98: 206-214 Hua XJ, Van de Cotte B, Van Montagu M, Verbruggen N (2001) The 5' untranslated region of the At-PSR gene is involved in both transcriptional and post-transcriptional regulation. Plant J, 26: 157-69 Jander G, Baerson SR, Hudak JA, Gonzalez KA, Gruys KJ, Last RL (2003) Ethylmethanesulfonate saturation mutagenesis in Arabidopsis to determine frequency of herbicide resistance. Plant Physiol, 131: 139-46 Jarrett HW (1993) Affinity chromatography with nucleic acid polymers. J Chromatogr, 618: 315-39 Jiang P, Cronan JE ( 1994) Inhibition of fatty acid synthesis in Eschericia coli in the absence of phospholipid synthesis and release of inhibition by thioesterase action. J Bacteriol, 176: 2814-2821 Lester RL, Wells GB, Oxford G, Dickson RC (1993) Mutant strains of Saccharomyces cerevisiae lacking sphingolipids synthesize novel inositol glycerolipids that mimic sphingolipid structures. J Biol Chem, 268: 845-856. 198 L _I .1: pin—1.1:. km Li G, Chandrasekharan MB, Wolffe AP, Hall TC (2001) Chromatin structure and phaseolin gene regulation. Plant Mol Biol, 46: 121-9 McCallum CM, Comai L, Greene EA, Henikoff S (2000) Targeted screening for induced mutations. Nat Biotechnol, 18: 455-7 Mekhedov S, Martinez de Ilarduya O, Ohlrogge J (2000) Toward a functional catalog of the plant genome. A survey of genes for lipid biosynthesis. Plant Physiol, 122: 389- 401 Memelink J (2003) T-DNA Activation Tagging. Methods Mol Biol, 236: 345-62 Memelink J, Kijne JW, van der Heijden R, Verpoorte R (2001) Genetic modification of plant secondary metabolite pathways using transcriptional regulators. Adv Biochem Eng Biotechnol, 72: 103-25 Menges M, Hennig L, Gruissem W, Murray JA (2002) Cell cycle-regulated gene expression in Arabidopsis. J Biol Chem, 277: 41987-2002. Ohlrogge JB, Jaworski JG (1997) Regulation of fatty acid synthesis. Annu Rev Plant Physiol Plant Mol Biol, 48: 109-136 Oliveira IC, Coruzzi GM (1999) Carbon and amino acids reciprocally modulate the expression of glutamine synthetase in Arabidopsis. Plant Physiol, 121: 301-10 Ruuska S, Girke T, Benning C, Ohlrogge J (2002) Contrapuntal networks of gene expression during Arabidopsis seed filling. Plant Cell, 14: 1191-1206 Savage LJ, Ohlrogge JB (1999) Phosphorylation of pea chloroplast acetyl-COA carboxylase. Plant J, 18: 521-527 Shintani DK, Ohlrogge JB (1995) Feedback inhibition of fatty acid synthesis in tobacco suspension cells. Plant J, 7: 577-587 Staub JM, Maliga P (1994) Translation of psbA mRNA is regulated by light via the 5'- untranslated region in tobacco plastids. Plant J, 4: 547-53 Sun C, Palmqvist S, Olsson H, Boren M, Ahlandsberg S, Jansson C (2003) A novel WRKY transcription factor, SUSIBAZ, participates in sugar signaling in barley by binding to the sugar-responsive elements of the isol promoter. Plant Cell, 15: 2076-92. Tepperrnan J M, Zhu T, Chang HS, Wang X, Quail PH (2001) Multiple transcription- factor genes are early targets of phytochrome A signaling. Proc Natl Acad Sci U S A, 98: 943 7-42 199 Thelen JJ, Ohlrogge JB (2002) Metabolic engineering of fatty acid biosynthesis in plants. Metab Eng, 4: 12-21 Uno Y, Furihata T, Abe H, Yoshida R, Shinozaki K, Yamaguchi—Shinozaki K (2000) Arabidopsis basic leucine zipper transcription factors involved in an abscisic acid- dependent signal transduction pathway under drought and high-salinity conditions. Proc Natl Acad Sci U S A, 97: 11632-7 Vainrub A, Pettitt BM (2003) Sensitive quantitative nucleic acid detection using oligonucleotide microarrays. J Am Chem Soc, 125: 7798-9 van der Fits L, Hilliou F, Memelink J (2001) T-DNA activation tagging as a tool to isolate regulators of a metabolic pathway from a genetically non-tractable plant species. Transgenic Res, 10: 513-21. Watson SJ, Akil H (2002) Evaluation of Affymetrix Gene Chip sensitivity in rat hippocampal tissue using SAGE analysis. Serial Analysis of Gene Expression. Eur J Neurosci, 16: 409-13. 200 11111111111111111111111