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he Relationship of Triacylglycerol and Starch Accumulation to Carbon and Energy Flows During Nutrient Deprivation in Chlamydomonas.1 ______________________ 1This Manuscript has been submitted for Publication to Plant Physiology under the same title. Authors: Matthew T. Juergens, Bradley Disbrow, and Yair Shachar-Hill. I generated and analyzed all data presented in this chapter as well as wrote this chapter. Bradley Disbrow assisted in sample processing and Yair Shachar-Hill assisted with writing and data interpretation. !!!!%KG! ABSTRACT Because of the potential importance of algae for green biotechnology, considerable effort has been invested in understanding their responses to nitrogen deprivation. The most frequently invoked reasons proposed for the accumulation of high cellular levels of triacylglycerol (TAG) and starch are variants of what may be termed the Overflow Hypothesis. According to this, growth inhibition results in the rate of photosynthetic energy and/or carbon input exceeding cellular needs; the excess input is directed into the accumulation of TAG and/or starch to prevent damage. The present study was aimed at providing a quantitative dataset and analysis of the main energy and carbon flows before and during nitrogen deprivation in a model system to assess alternative explanations. Cellular growth, biomass, starch and lipid levels as well as several measures of photosynthetic function were recorded for cells of Chlamydomonas reinhardtii cultured under 9 different autotrophic, mixotrophic and heterotrophic conditions during nutrient-replete growth and for the first four days of nitrogen deprivation. The results of a 13C labeling time course indicated that in mixotrophic culture, starch is predominantly made from CO2 and fatty acid synthesis is largely supplied by exogenous acetate, with considerable turnover of membrane lipids, so that total lipid rather than TAG is the appropriate measure of product accumulation. Heterotrophic cultures accumulated TAG and starch during N-deprivation, showing that these are not dependent on photosynthesis. We conclude that the overflow hypothesis is insufficient and suggest that storage may be a more universally important reason for carbon compound accumulation during nutrient deprivation. !!!!%KH! INTRODUCTION The current rate of utilization of fossil fuel products for energy and the chemical industry is unsustainable. Among the alternative renewable replacements, microalgal oil and biomass production have shown considerable promise (Atabani, et al. 2012, Chisti 2007, Ohlrogge, et al. 2009, Williams, et al. 2010) because many algae are capable of rapid photoautotrophic growth to high cell densities and can accumulate high dry weight percentages of triacylglycerol (TAG) (Hu, et al. 2008, Jones, et al. 2012, Wijffels, et al. 2010). TAG accumulation is induced in many microalgal taxa by a number of stresses including salt and nutrient deprivation (Murphy 2001). Beginning over 100 years ago (Beijerinck 1904) many researchers have reported that nitrogen (N) deprivation induces high levels of accumulation of TAG and starch in a range of microalgal species at the expense of decreased cell growth and slower metabolism (Granum, et al. 2002, Hu, et al. 2008, Liu, et al. 2013, Martin, et al. 1975a, Shifrin, et al. 1981, Spoehr, et al. 1949). This has led to interest in understanding the underlying physiological and adaptive functions of algal TAG production during N deprivation to help guide engineering of higher oil yields. In recent years, studies of the effects of nutrient deprivation have revealed many changes in the structure and function of networks across metabolism and other cellular functions, (Blaby, et al. 2013, Goodenough, et al. 2014, Juergens, et al. 2015, Miller, et al. 2010, Schmollinger, et al. 2014). Such studies and mutant screens have led to the identification of enzymes involved in TAG synthesis (Li, et al. 2012b, Merchant, et al. 2012b) and pointed to putative regulatory networks (Gargouri et al. 2015). However the interpretation of system-wide molecular changes and the choice of which among them to target for further investigation and practical purposes is !!!!%KI!uncertain and is strongly influenced by the prevailing views of the function of TAG accumulation. Several explanations have been proposed for the induction of TAG accumulation in algae under stress, ranging from storing reduced carbon as an energy source for survival and/or future recovery, to lipid reorganization during photosynthetic down-regulation and/or subsequent up-regulation, to photo-protection (Akita, et al. 2015, Grossman, et al. 2010a, Khozin-Goldberg, et al. 2005, Klok, et al. 2014, Kohlwein 2010, Murphy 2001). Roessler (Roessler 1990) appears to have been the first to postulate that algae accumulate TAG as a sink for excess photosynthetic energy and reductant to prevent photochemical damage. In this explanation, photosynthetic carbon and energy assimilation that can no longer be directed to growth when population increase is inhibited by nutrient deficiency or other stresses results in overflow products, particularly TAG. The concept of overflow metabolism has traditionally been associated with the export of metabolites by heterotrophic microbes (Neijssel, et al. 1975) and has more recently also been applied to photosynthetic metabolism in cyanobacteria (Courchesne, et al. 2009, Grundel, et al. 2012, Hays, et al. 2015) and higher plants (Weise, et al. 2011). In microalgal work, the idea of photosynthetic overflow (excess photosynthetic energy and carbon) as the driver of oil accumulation has become a widely accepted explanation (see for example (Hu, et al. 2008, Klok, et al. 2014, Li, et al. 2012b, Li, et al. 2013, Solovchenko 2012). This Overflow Hypothesis (OH) strongly influences research efforts and the interpretation of results in biological and engineering studies of microalgal metabolism and TAG accumulation and also has important implications for the ecophysiology of photosynthetic microbes. While frequently invoked in interpreting the results of nutrient deprivation studies, the OH has not been systematically assessed. !!!!%KJ! Such systematic assessments should include a quantitative analysis of the major carbon and energy flows during stress-induced oil accumulation, and should be carried out over a range of eco-physiologically and practically meaningful conditions in relevant model organisms (Smith et al. 2010). Despite a burgeoning literature on microalgal TAG accumulation under N deprivation, such studies are lacking. The green alga Chlamydomonas reinhardtii is the most widely studied species in research on microalgal oil production. Although it produces less TAG than some species and has a low tolerance of extreme conditions, its robust growth, ease of culture, ability to grow heterotrophically and the wealth of existing data and available genetic tools (including mutant collections, transformation tools, and a well annotated genome sequence), make it attractive for biofuel research (Merchant, et al. 2012b, Merchant, et al. 2007, Rochaix 2002). Carbon (C) and Energy (E) balances are established tools in ecology and metabolic engineering and are employed to assess growth and yields during commercial applications. Widely used for monitoring and improving bacterial and yeast fermentation processes, measuring cellular C and E influxes and outflows has also been used in several cyanobacterial and algal studies (Chapman, et al. 2015, Slade et al. 2013, Wagner et al. 2006, Watanabe et al. 1995). Figure 1 shows the C and E entering a Chlamydomonas cell (photosynthesis and/or acetate uptake) as well as the major C and E outputs (growth, maintenance, dissipation, and TAG and Starch production. Figure 1 also illustrates the overflow hypothesis, in which nutrient deprivation causes a redirection of output fluxes from reproduction to TAG and starch synthesis. Quantifying the relationships between C and E fluxes before and during stress conditions can point to processes such as dissipative metabolic or photochemical processes and can be used to !!!!%%K!make quantitative tests of predictions implied by the OH and alternative explanations for TAG accumulation. Figure 3.1. The main Carbon and Energy inputs and outputs of Chlamydomonas cells. CO2 provides C input, light provides energy input and acetate provides both. Incoming energy can be dissipated as heat, consumed for maintenance, or used in cell growth or biosynthesis of storage compounds (starch and oil). Incoming carbon is used in growth, stored in starch or oil, or respired as CO2. !!!!%%%!Here we present measurements of cellular growth, biomass, starch, TAG, and total fatty acid accumulation, changes in photosynthetic fluxes and acetate uptake rates across 5 light levels and 2 media compositions before and during the course of 96 h of N deprivation. Results from a 13C labeling time course experiment highlight the involvement of multiple lipid pools during N deprivation, so that (a) total lipid rather than TAG alone is the appropriate pool for C & E balances; and (b) lipid is preferentially formed from exogenous fixed carbon (when available) rather than from photosynthate, the latter contributing more to starch production. TAG and starch accumulated to significant levels in the dark after N depletion, demonstrating that TAG accumulation is not dependent upon photosynthetic activity. Using the photosynthetic, biomass and uptake rates, C and E influxes and outflows are reported and compared to predictions deduced from the overflow hypothesis. Several aspects of the findings do not support the OH in its straightforward form. We conclude that the overflow hypothesis is insufficient to explain carbon accumulation during nutrient deprivation and suggest that storage of biosynthetic precursors and/or chemical energy as adaptive survival traits may be more universal reasons. !!!!%%&! RESULTS Growth rates in N-replete conditions To characterize the effect of light and trophic conditions on Chlamydomonas reinhardtii growth rates, cells were grown in defined media either with acetate as a fixed carbon source (Tris Acetate Phosphate, TAP media) or without acetate (High Salt, HS media) at light levels of 0, 5, 15, 40, and 160 µmol photons m(2 s(1. Cell growth measured during exponential (steady state) growth is presented as specific growth rate in Figure 3.2. This range of light intensity spans heterotrophic growth in the dark in TAP media, light-limited autotrophic and mixotrophic conditions and saturating light levels. Illumination was kept below levels (at or above 200 µmol photons m(2 s(1) where cells showed chlorosis and reduced growth rates indicative of light stress and/or photo-inhibition (Bonente et al. 2012, Juergens, et al. 2015, Peers, et al. 2009). Increasing illumination from 40 to 160 µmol photons m(2 s(1 did not increase growth rate, suggesting that these cells are carbon limited (Ref on C limitation in Algae) (Goldman et al. 1974, Spalding 1989). This suggestion was supported by the observation that growth rates increased approximately twofold when humidified ambient air was circulated through the culture flasks to increase CO2 availability. !!!!%%'! Figure 3.2. Specific growth rates during exponential growth in nitrogen replete media. Growth with acetate (TAP media, squares) and or without (HS media, circles) under continuous illumination at light levels from 0 to 160 µ mol photons m-2 s-1. Specific growth rate (SG) is the rate constant for log growth and is inversely related to doubling time as SG = Ln (2)/ (doubling time). Error bars indicate +/- SD (n=3 biological replicates). 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0 50 100 150 Specific growth/Hr Light Intensity (µmol photons m-2 s-1 ) !!!!%%(! Biomass Biomass measured as ash free dry weight (AFDW) is shown in Figure 3.3A and Figure 3.3B. Cellular biomass in N replete cells is very similar across light levels, with mixotrophic and heterotrophic cultures (TAP medium) having approximately 25% more AFDW than autotrophic cultures. Cells cultured in TAP medium increased ~2 fold in cell biomass over 96 hrs whereas HS cells increased less in biomass, especially in the first 48h of deprivation. Biomass and C and N contents of cells were measured to quantify net carbon accumulation rates and cellular nitrogen contents as a proxy for nitrogenous biomass (protein plus nucleic acids). Cellular C and N contents measured by elemental analysis are shown in Figure 3.3C to Figure 3.3F and C:N ratio is shown in Figure 3.4. Carbon accounts for ~40-47% of cell dry weight during growth in N replete media (Figure 3.3C and Figure 3.3D), levels which changed little during N deprivation in TAP media - whereas C levels decreased in cells deprived of N in HS media. Nitrogen levels as a percentage of dry weight were higher in nitrogen replete cells growing autotrophically than in TAP media and fell significantly during N deprivation (except for heterotrophic cultures). The falling cellular %N during deprivation is due to the accumulation of biomass components lacking N, so that total N contents per cell were little changed. !!!!%%F! Figure 3.3. Cellular Biomass and % Carbon and Nitrogen before and during nitrogen deprivation. Biomass is given as Ash Free Dry Weight (AFDW) per million cells (A and B), %C (C and D), and %N (E and F) is reported for cultures grown in TAP (A C, and E) and HS (B, D and F) media and deprived of N at time 0. Empty diamonds, cells cultured at 160 µmol photons m-2 s-1 (µE); empty squares, 40 µE; empty triangles, 15 µE; filled diamonds, 5 µE; filled squares, 0 µE. Error bars indicate +/- SD (n=3). !!!!%%G! Figure 3.4. Carbon to Nitrogen (C:N) ratios during N deprivation. Data for TAP(A) and HS(B) cultures were taken before and during N deprivation. In all cases, 160 µmol m-1 s-1 is indicated by hollow diamonds, 40 µmol m-1 s-1 by hollowed squares, 15 µmol m-1 s-1 indicated by hollowed triangles , 5 µmol m-1 s-1 by solid diamonds, and 0 µmol m-1 s-1 by solid squares. Error bars indicate SD (n=3). The Accumulation of TAG and Starch The first compound known to accumulate to high levels during N deprivation is starch (Klein 1987) and the upregulation of starch synthesis gene expression has been reported during N deprivation (Ball, et al. 1990, Juergens, et al. 2015, Miller, et al. 2010). Increases in starch levels during N deprivation are shown in Figure 3.5; they account for the majority of the increases in cellular dry weight. Starch levels rose linearly for 48 hrs after N deprivation in both TAP and HS media, with higher rates at higher light levels and in TAP compared to HS media. After 48 hrs starch accumulation rates decreased, with cessation of starch accumulation by 96 hrs for all but the cells cultures at the highest light level in TAP media and those at the lowest light levels in HS media. !!!!%%H!During nitrogen-replete growth cultures contained close to 5*g of total fatty acid per million cells (Figure 3.5B and Figure 3.5C) and no TAG was detected (Figure 3.5E and 3.5F). During N deprivation in mixotrophic cultures, TAG levels were detected at 24 hours while little or no TAG was detected in HS cultures until 48h. Levels of starch and TAG were previously reported by Fan et al. (Fan, et al. 2012) and by Siaut et al. (Siaut, et al. 2011) for several C. reinhardtii strains cultured at 50 and 150 µmol photons m(2 s(1 respectively, after 2 or 3 days of N deprivation. Our results for these time points in cells supplied with 40 and 160 µmol photons m(2 s(1 are similar to those previously reported (except for mutant strains lacking starch). The observation that TAG accumulates during nitrogen deprivation in Chlamydomonas cells cultured heterotrophically in the dark separates the induction of TAG accumulation from photosynthesis. High levels of TAG accumulation have been reported in heterotrophically grown Chlorella species (Liu et al. 2011, Miao et al. 2004) and TAG accumulation was also observed in heterotrophically cultured Chlamydomonas after carbon starvation (Singh et al. 2014) and during N deprivation in the presence of high levels (60mM) of acetate (Fan, et al. 2012). !!!!%%I! Figure 3.5. Starch, FAME and TAG levels before and during nitrogen deprivation. Starch (A,B), FAME (C,D), and TAG (E,F) levels are reported in ug per million cells for cultures in TAP (A, C, E) and HS (B, D, F) media. Empty diamonds, cells cultured at 160 µmol photons m-2 s-1 (µE); empty squares, 40 µE; empty triangles, 15 µE; filled diamonds, 5 µE; filled squares, 0 µE. Error bars indicate +/- SD (n=3).!!!!!%%J!The increases in total cellular FAME levels we observed were lower and slower to start than increases in TAG. This is consistent with previous reports that membrane lipid levels fall during N deprivation and that some fatty acids used in TAG synthesis come from membrane lipids (Juergens, et al. 2015, Moellering et al. 2011, Yoon et al. 2012). However the extent to which fatty acids newly synthesized during N deprivation enter bulk membrane lipid pools is not known. The source(s) of carbon entering TAG and the fate of carbon assimilated during nutrient deprivation are relevant to the consideration of cellular metabolic input and output fluxes during TAG synthesis and were therefore probed in a labeling experiment. 13C labeling Results Chlamydomonas cells growing in nitrogen-replete TAP media were transferred to TAP media without N and containing 13C labeled acetate (100% 13C1,2). After 40 h cells were transferred to TAP media containing unlabeled acetate and lacking N for the remainder of the 96 h deprivation period. Cells cultured mixotrophically at 160 µmol photons m(2 s(1 were used as these conditions are similar to those used in many prior studies and showed the highest sustained starch and lipid accumulation rates. After 40 h of N deprivation in the presence of 13C-acetate, the proportion of starch-derived ions detected that were labeled with 13C was found to be low, with over 65% of the ions containing no 13C label (Figure 3.6A) and over 80% of the total carbon being 12C. Label levels increased modestly during the initial washout period between 40 and 72 h (~28% of starch carbon at 72 h originated from the 13C acetate taken up between 0 and 40 h). Label levels in starch did not significantly change between 72 and 96 h. Thus photosynthetic CO2 fixation and/or cellular biomass components made before deprivation is the dominant source of carbon for net starch synthesis during N deprivation under mixotrophic conditions. In !!!!%&K!such cultures the total starch levels per cell doubled between 40 and 96 h (Figure 3.5A), during which time no 13C label was present in the medium, so we had expected a dilution of the 13C label in starch. The lack of a decrease in fractional labeling level between 40 and 96 h shows that 13C assimilated during nitrogen deprivation (0-40 h) contributed significantly to starch synthesis later. Labeling in fatty acids from the experiment described above is also shown in Figure 3.6. Figure 3.6B shows the distribution of mass isomers from a representative mass spectrum of palmitate (C16:0, as the methyl ester) from cells collected at 40 h. Labeling results for C16:0 from TAG and two membrane lipid fractions (galactolipids, which are major components of chloroplast membranes, and a polar lipid fraction containing lipids from both chloroplast and extra-plastidic membranes) are shown in Figure 3.6 C to Figure 3.6E. Analogous data for other abundant fatty acids are given in the Figures 3.7 and 3.8. After 40 h of labeling and nitrogen deprivation, approximately 70% of the fatty acid molecules from TAG and membrane lipids contained one or more 13C labeled carbons. In the majority of fatty acid molecules over half the carbon atoms were 13C and the most abundant labeled mass isomer was the fully labeled one (13C in all positions) and approximately half the total carbon in cellular fatty acids was 13C. The preponderance of fatty acid molecules in all pools were either unlabeled or highly labeled, and since the total fatty acid content of the cultures approximately doubled during the 40 h labeling period we infer that 75% or more of the carbon used to synthesize fatty acids during N deprivation is derived from acetate taken up during deprivation. !!!!%&%! Figure 3.6. 13C labeling of fatty acids and starch during N deprivation in cells cultured in TAP media at 160 µmol photons m-2 s-1. Cells grown to log phase in N replete media were moved at time 0 to TAP media lacking N with 100% uniformly labelled 13C acetate. After 40hrs the media was replaced with TAP media lacking N and containing unlabeled acetate for the remainder of the time course. The proportions of different mass isomers are plotted as a function of time. Panel A shows mass isomers of the m/z = 319 fragment ion containing carbons 2-6 of the glucose monomers of starch. Filled squares represent M+0, filled diamonds M+1, empty triangles m+2, XÕs M+3, and empty circles M+4. Panel B is a histogram of the mass isomer distributions for a representative 16:0 FAME mass spectrum showing the abundance of highly labeled molecules together with unlabeled molecules (M+0 and naturally-occurring M+1 masses). Panel C, D, and E represent mass isomer distributions for ions of intact C16:0 FAME molecules from TAG, MGDG, and Polar Lipid fractions respectively. Filled squares represent unlabeled molecules (M+0); filled diamonds, M+1 (one 13C atom); empty triangles, M+2; empty circles, the sum of M+3 through M+16 (fully labeled), Error bars indicate range (n=2). !!!!%&&!High fractional labeling in galactolipids and polar lipids (Figure 5C and Figure 5D) after 40 h of N deprivation shows that membrane lipids are synthesized at a significant rate during nitrogen deprivation. Since the cellular levels of membrane lipids fall by more than 50% during this time (Juergens, et al. 2015), it is clear that substantial rates of simultaneous synthesis and breakdown of membrane lipids are occurring, suggesting acyl chain trafficking through these lipid classes. Therefore we consider that total cellular lipid rather than TAG is the appropriate pool to be measured when evaluating net carbon and energy fluxes into accumulated intracellular compounds during nitrogen deprivation. After replacement of labeled with unlabeled acetate the proportion of fatty acid molecules that were highly 13C-labeled fell while the proportion of unlabeled molecules rose; this is consistent with dilution of highly labeled with newly synthesized unlabeled fatty acids. The similarity in C16:0 labeling patterns in different lipid pools is consistent with the continued flux of newly synthesized fatty acids into membrane lipids as well as TAG between 40 and 9 6h. In addition, during this ÒchaseÓ period there is an increase in the proportion of fatty acid molecules with low levels of labeling (M+1 and M+2) which we attribute to flow of C from non-lipid compounds synthesized during the initial stages of N deprivation into lipid synthesis at later stages. !!!!%&'! Figure 3.7. 13C labeling of C16:3 fatty acids during N deprivation in cells cultured in TAP media at 160 µmol m-1 s-1 . Cells grown to log phase in N replete media were moved at time 0 to TAP media lacking N with 100% uniformly labelled 13C acetate. After 40hrs the media was replaced with TAP media lacking N and containing unlabeled acetate for the remainder of the time course. The proportions of different mass isomers are plotted as a function of time. Panel A, B, and C represent mass isomer distributions for ions of intact C16:3 FAME molecules from TAG, MGDG, and Polar Lipid fractions respectively. Filled squares represent unlabeled molecules (M+0); filled diamonds, M+1 (one 13C atom); empty triangles, M+2; empty circles, the sum of M+3 through M+16 (fully labelled), Error bars indicate range (n=2). !!!!%&(! Figure 3.8. 13C labeling of C18 fatty acids during N deprivation in cells cultured in TAP media at 160 µmol m-1 s-1 . Cells grown to log phase in N replete media were moved at time 0 to TAP media lacking N with 100% uniformly labeled 13C acetate. After 40hrs the media was replaced with TAP media lacking N and containing unlabeled acetate for the remainder of the time course. The proportions of different mass isomers are plotted as a function of time. Panel A, B, and C represent mass isomer distributions for ions of intact C18:1 FAME molecules from TAG, MGDG, and Polar Lipid fractions respectively. Panel D, E, and F represent mass isomer distributions for ions of intact C18:2 FAME molecules from TAG, MGDG, and Polar Lipid fractions respectively. Panel G, H, and I represent mass isomer distributions for ions of intact C18:3 FAME molecules from TAG, MGDG, and Polar Lipid fractions respectively. Filled squares represent unlabeled molecules (M+0); filled diamonds, M+1 (one 13C atom); empty triangles, M+2; empty circles, the sum of M+3 through M+16 (fully labeled), Error bars indicate range (n=2). !!!!%&F!Light Absorption, Utilization Efficiency and Dissipation To gauge photosynthetic light capture rates, levels of chlorophyll (chl) and the efficiency of photosystem II were measured. Chl levels are shown in Figure 3.9A and Figure 3.9B for cells cultured in TAP and HS medium respectively. Nitrogen-replete cells growing under mixotrophic conditions contained more Chl than autotrophically growing cells, with the exception of mixotrophic cells at the lowest light levels. During nitrogen deprivation chl levels fell continuously in mixotrophically cultured cells, whereas under autotrophic conditions chl levels fell during the first two days and did not change significantly thereafter. Both theoretical quantum yield at photosystem II (FV/FM) and light driven yields ('II) were measured by fluorescence spectroscopy as measures of the efficiency with which absorbed light drives linear electron flow. In TAP media FV/FM and 'II (Figure 3.9C and Figure 3.9E respectively) declined through the nitrogen deprivation period, with the exception of cells under 5 µmol photons m-2 s-1. In autotrophic cells FV/FM and 'II (Figure 3.9D and Figure 3.9F) did not significantly change during nitrogen deprivation. !!!!%&G! Figure 3.9. Chlorophyll, Photosynthetic efficiency and Non Photochemical Quenching. !!!!%&H!Figure 3.9 (contÕd). Panel A and B represent Chlorophyll levels in Tap and HS conditions respectively. Panels C and D represent Maximum photosynthetic efficiency from dark adapted cells for TAP and HS conditions respectively. Panels E and F represent photosynthetic efficiency in the light for TAP and HS conditions respectively. Panels G and H represent photosynthetic efficiency in TAP and HS conditions respectively. In all cases, 160 µmol m-1 s-1 is indicated by hollow diamonds, 40 µmol m-1 s-1 by hollowed squares, 15 µmol m-1 s-1 indicated by hollowed triangles , 5 µmol m-1 s-1 by solid diamonds, and 0 µmol m-1 s-1 by solid squares. Error bars indicate SD (n=3). During photosynthetic stress conditions and at high light levels, cells dissipate excess absorbed light energy, by non- photochemical quenching (NPQ) (Muller, et al. 2001). We measured NPQ before and during N deprivation under mixotrophic (Figure 3.9G) and autotrophic (Figure 3.9H) conditions using the same light level for NPQ measurement as the level under which the cells had been cultured. NPQ levels recorded before and throughout nitrogen deprivation for all cultures were lower than values reported in previous studies in which much higher light levels were used for measurement than for growth (Niyogi, et al. 1997, Peers, et al. 2009, Terauchi et al. 2010). The absence of significant increases in NPQ after N-deprivation suggests that there is little or no light stress due to excess energy intake under these conditions. Photosynthetic Fluxes Photosynthetic fluxes were assessed by measuring oxygen evolution and electrochromic shift (ECS) absorbance spectroscopy as indicators of linear and cyclic electron flow respectively (Figure 3.10). Oxygen evolution rates at 160 µmol photons m(2 s(1 for both autotrophic and heterotrophic cultures decreased markedly during N deprivation (Figure 3.10A and Figure 3.10b). At 40 µmol photons m(2 s(1 and below, net oxygen fluxes were much less during nitrogen replete growth but remained largely unchanged during deprivation. Mixtrophically !!!!%&I!cultured cells at lower light levels and autotrophic cells were net oxygen consumers. ECS decay rates, which reflect the proton fluxes across thylakoid membranes that drive photosynthetic ATP synthesis, decreased several fold during nitrogen deprivation for cells under the highest light level for both mixotrophic and autotrophic cells (Figure 3.10C and Figure 3.10D). At lower light levels ECS values for N replete cells were significantly less than at 160 µmol photons m-2 s-1 and decreased moderately or were not significantly changed during nitrogen deprivation. Figure 3.10. Oxygen evolution and electrochromic shift measurements. Net oxygen evolution (A and B) and electrochromic shift decay rates (indicating proton fluxes across the thylakoid membrane, C and D) for cells grown in TAP (A and C) and HS (B and D) media. Results for cells cultured at 160 µmol m-2 s-1 are indicated by empty diamonds, 40 µmol m-2 s-1 empty squares, 15 µmol m-2 s-1 empty triangles, 5 µmol m-2 s-1 filled diamonds, and 0 µmol m-2 s-1 filled squares. Error bars indicate SD (n=3). !!!!%&J! Carbon Assimilation and Release Net CO2 uptake and efflux rates were measured before and during N deprivation (Figure 3.11A and Figure 3.11B). In TAP media, cultures under 160 and 40 µmol photons m-2 s-1 of illumination and in HS media at all light levels showed net CO2 assimilation when provided with nitrogen while cultures exposed to lower light levels or in the dark in TAP media were net CO2 producers. During nitrogen deprivation, net CO2 uptake and output rates decreased, with TAP-grown cells at 40 µmol photons m-2 s-1 becoming net CO2 producers after 2d of deprivation. Acetate uptake rates for mixotrophic and heterotrophic cultures (Figure 3.11C) were higher for cells at lower light intensities and in the dark and decreased during nitrogen deprivation. Thus both photosynthetic CO2 assimilation and consumption of external fixed carbon decreased strongly during N deprivation, with decreases being more marked at higher light levels. !!!!%'K! Figure 3.11. Carbon assimilation rates. !!!!%'%! ANALYSIS AND DISCUSSION The first observation of microalgal TAG accumulation under nitrogen deprivation was made over a century ago (Beijerinck 1904) and the potential to utilize this phenomenon for biofuels has been recognized for over 70 years (Harder et al. 1942a, Harder et al. 1942b). Among several explanations offered, the overflow hypothesis (OH), now at least 25 years old (Roessler 1990), currently holds sway over much of the interpretation of algal TAG accumulation data. The OH explains the synthesis of TAG and/or starch by algae during nutrient deprivation as a response to excess photosynthetic energy and/or carbon assimilation. The results of this study allow an assessment to be made of the OH in its straightforward form by making a series of inferences from it about the relationships between photosynthetic carbon and energy uptake rates and the rates of accumulation of TAG and starch and comparing these expectations to the data(Table 3.1). First, if accumulation is driven by surplus energy or carbon from photosynthesis, no accumulation is expected to occur in the dark. Starch accumulation rates in the dark were equal to or greater than rates for autotrophic cultures at all light levels and were exceeded only by mixotrophic cultures at the higher two light levels. We conclude either that starch accumulation during nitrogen deprivation is primarily driven by factors unrelated to photosynthetic overflow or that there is a separate additional explanation for its accumulation under heterotrophic conditions. TAG accumulation in the dark was also significant, although net Figure 3.11. (contÕd). Net CO2 uptake in TAP (A) and HS (B) media. Measurements of cells cultured at 160 µmol m-2 s-1 are indicated by empty diamonds, 40µmol m-2 s-1 by empty squares, 15µmol m-2 s-1 indicated by empty triangles , 5µmol m-2 s-1 by filled diamonds, and 0 µmol m-1 s-1 by filled squares. Acetate uptake in TAP cultures are in panel C for Nitrogen replete cells (White), 0-24hrs of N-deprivation (grey), and 24-96hrs of deprivation (black). Error bars indicate SD (n=3). !!!!%'&!fatty acid accumulation was modest compared to cultures in the light. Indeed heterotrophic cells from taxa as diverse as fungi bacteria and mammals have been shown to accumulate TAG under nutrient deprivation (Alvarez et al. 2002, Frenk et al. 1958, Morin et al. 2011, Murphy 2001, Packter et al. 1995). !!!!%''! Table 3.1. Overflow Hypothesis predictions and verification. This table summarizes the predictions associated with the Overflow Hypothesis of TAG accumulation and whether they were supported or not by this study. !!!!%'(!Second, the OH links growth before nutrient deprivation to TAG production afterwards; initially we examine carbon balances. Figure 3.12 shows a comparison of carbon accumulation rates before and during the first 24 h after N deprivation. The first 24 h after deprivation were selected for comparison with N-replete growth since at this time photosynthetic function is most similar to pre-deprivation rates. In this work, as elsewhere (Fan, et al. 2012, Siaut, et al. 2011) starch was found to be the dominant carbon sink, with total FAME accounting for less than 15% of net carbon accumulation during the first 24 h. Total FAME accumulation rates only match and begin to exceed those for starch after 48 h, by which time photosynthetic fluxes are substantially lower. For autotrophic cells across all light levels the total carbon accumulation rates after deprivation are close to 90% of those during N replete growth, which is consistent with the OH as applied to total C. For cells cultured under mixotrophic or heterotrophic conditions, total carbon accumulation rates after deprivation are not well explained by the rates during N replete growth. Although there is a trend with increasing light levels in TAP media towards greater total C accumulation rates before and after deprivation, the slope corresponds to only 11% and the trend does not extrapolate meaningfully to low light and lower growth rates. Concerning the origins of carbon from which TAG and starch are synthesized, 13C acetate labeled both these pools in mixotrophic cultures that were grown at the highest light level to maximize photosynthetic inputs. Photosynthesis is therefore not the dominant metabolic source of acetyl-CoA used in TAG synthesis in mixotrophy. Fan et al. (Fan, et al. 2012) have suggested that TAG accumulation is dependent upon total carbon precursor availability rather than only CO2 from photosynthesis. !!!!%'F! Figure 3.12. Cellular carbon accumulation rates before and during the first 24hrs after nitrogen deprivation. Accumulation during nitrogen replete growth was derived from growth rates and total cellular carbon contents; carbon accumulation after deprivation is due to starch and lipid production. Filled symbols represent autotrophic conditions (HS media), while empty ones represent mixotrophic and heterotrophic conditions (TAP media). Squares indicate values for total Carbon accumulation (starch plus total FAME) while diamonds and circles represent starch and total FAME respectively. Within each series, increasing light levels correspond to symbols from left to right Error bars depict +/- SD (n=3). The dashed line corresponds to a slope of 1 and the solid line is the least-squares best fit to the autotrophic total carbon accumulation results. !!!!%'G! Third, regarding redox energy balance the OH posits that the function of TAG synthesis during nutrient deprivation is to consume excess photosynthetically produced reductant and thereby prevent photodamage due to over reduction of electron transport chain components. We would expect from this that cellular fatty acid levels should rise most rapidly when net oxygen production is highest (early on in nutrient deprivation), which is not the case (compare Figure 4C and 4D with Figure 7A and 7B). Since cells at the highest light level were CO2 limited during N-replete growth, these might be expected to show either high rates of NPQ or elevated fatty acid accumulation rates compared to starch early during deprivation, neither of which was observed. Related to the potential role of TAG synthesis in mitigating over reduction in the electron transport chain is a fourth prediction, one that has received apparent support from past studies. In this version of the OH, following N deprivation cells experience stress from excess light energy uptake leading to the activation of energy dissipative mechanisms such as non-photochemical quenching and/or the Mehler reaction in addition to TAG synthesis. Increases in NPQ under nutrient deprivation have been previously observed, reaching levels that account for a substantial proportion of light energy reaching the photosystems (Allen et al. 2008, Antal et al. 2006, White et al. 2011, Wilson et al. 2007). However, those measurements of NPQ were made under much higher light levels than those under which cultures were maintained. Measurements of NPQ under light levels under which cells had been cultured (Figure 3.9) show stable or decreasing NPQ in TAP media while in HS media cells exhibit decreasing NPQ following N deprivation. Slow decreases in photosynthetic efficiency, Chl levels and gas exchange rates in parallel with low and decreasing levels of NPQ indicate that a coordinated downregulation of photosynthetic structure and function that is used to match energy supply and demand under a wide range of !!!!%'H!energy input rates, making the idea of overflow from a mismatch of supply and demand less appealing. An alternative perspective on energy balances is to assess the relationship between photosynthetic energy input and the accumulation of TAG and starch. Figure 3.13A and Figure 3.13B show the values of an estimate of the rates of cellular light energy intake: the product of illumination level, chlorophyll content and 'II. This parameter accounts for light absorbed and the quantum efficiency with which light drives photosynthetic electron flow (Genty, et al. 1989). While only part of the energy capture estimated by this parameter becomes available to cellular functions in the form of ATP and NADPH, the dominant losses involved, such as that associated with light of wavelength < 680nm and the stoichiometries and/or irreversibilities of electron transport and H+/ATPÕase fluxes are such that the proportion lost is likely to be similar across the conditions used here. The photosynthetic energy capture rates span a range of approximately 50 fold under nitrogen-replete conditions, a range which narrows during N deprivation since the capture rates fall more strongly in cells cultured with lower light levels. The extent to which light capture rate is reflected in lipid, starch and total C accumulation rate is shown in Figure 3.13C to Figure 3.13H as the ratio of carbon accumulated to the light capture parameter. This ratio varies widely among culture conditions at any given time, with cells exposed to higher light levels accumulating less starch or FAME as a proportion of apparent light energy captured. The OH would lead one to expect higher rather than lower proportions of light energy to be used for carbon accumulation at higher light levels. Alternatively or additionally a threshold of light capture rates might be expected to be needed to result in photosynthetic energy overflow that triggers C accumulation; a comparison of Figures 3.13A and 3.13B with Figure 3.5 shows that no such threshold is evident. Starch and TAG rates vs. acetate is presented in Figure 3.14. !!!!%'I! Figure 3.13. Accumulated Carbon V.S Potential absorbed light !!!!%'J!Figure 3.13 (ContÕd). A metric of potential usable light by photosystem II was calculated from 'II times Chl levels* light intensity per million cells. TAP cultures are in panel A and HS culture data is in panel B. Panel C and D show the rate of starch accumulation vs. absorbed light for TAP and HS respectively. Panel E and F show rates of FAME accumulation vs. absorbed light for TAP and HS respectively. G and H display rates of total carbon accumulation vs. absorbed light. For TAP and HS repectively In all cases, 160 µmol photons m-2 s-1 (µE) is indicated by hollow diamonds, 40 µE by hollowed squares, 15 µE indicated by hollowed triangles , 5 µE by solid diamonds. Error bars indicate SD (n=3). The OH can also be considered in terms of the relationship between captured light energy and the energy deposited in accumulated compounds; for the latter we used heats of combustion for starch and glycerolipid (and for nitrogen replete growth also of protein/nucleic acids). This comparison is made in Figure 3.15A and Figure 3.15B for mixotrophic and autotrophic cultures respectively. For autotrophic cells at lower light levels and during N deprivation but not before, there is a correlation between photosynthetic energy input (X axis of figure 3.15B) and the accumulation of chemical potential energy in starch plus TAG (Y axis). For N replete, mixotrophic cultures and higher light levels, this correlation is not evident. The ratio of light energy input to chemical energy accumulated is shown as a function of deprivation time in Figure 3.15C and 3.15D. This ratio ranges by 20 fold across light levels with cells having higher light input rates accumulating less of it in starch and TAG, which is not expected if C compound accumulation is a large contributor to coping with light energy input. !!!!%(K! Figure 3.14. µg Starch and FAME per uptaken acetate. Panels A and B represent Starch and FAME accumulation rates respectively for TAP conditions. In all cases, 160 µmol m-1 s-1 is indicated by hollow diamonds, 40 µmol m-1 s-1 by hollowed squares, 15 µmol m-1 s-1 indicated by hollowed triangles , 5 µmol m-1 s-1 by solid diamonds, and 0 µmol m-1 s-1 by solid squares. Error bars indicate SD (n=3). !!!!%(%! Figure 3.15 Light energy intake and energy stored in accumulated biomass (as heat of combustion) in TAP (panels A and C) and HS (panels B and D). The relationship between this energy output into biomolecules and the energy input from photosynthetic light capture is shown in panels A and B. The ratio of chemical potential energy accumulated in each 24 h period to the light energy intake rate is shown in panels C and D. For A and B, Time 0 hrs is indicated by empty diamonds, 24hrs N- is indicated by empty squares, 48 hrs N- is indicated by empty triangles, 72hrs N- is indicated by empty circles, and 96hrs is indicated by XÕs. For C and D, 160 µmol photons m-2s-1 (µE) is indicated by empty diamonds, 40 µE by empty squares, 15 µE indicated by empty triangles, 5 µE by filled diamonds. Error bars indicate +/- SD (n=3). !!!!%(&!CONCLUSION This study was aimed at obtaining sufficient data to assess of the OH. Some of the results, such as the trend towards higher rates of TAG accumulation at higher light levels, and the correlation between total net carbon assimilation before and after nutrient deprivation in autotrophic cells are consistent with the OH, although that they are also compatible with alternative, storage based explanations. Other observations, including starch and TAG accumulation in heterotrophic cultures, the timing of TAG accumulation relative to maximal photosynthetic rates, and the apparent absence of correlation between light stress or excess reductant and TAG or starch accumulation show that the OH is insufficient as an explanation for the C accumulating response of algae during nutrient deprivation. The results of a number of studies have been interpreted in terms of the OH hypothesis. One such study concerns the Chlamydomonas PGD1 galactolipase mutant. This mutant accumulates half as much TAG as wild type cells and is found to undergo severe chlorosis and cell death during N deprivation unless it is given DCMU (an inhibitor of PSII electron transfer). If excess photosynthetic energy is normally used for fatty acid synthesis when growth is halted by nutrient deprivation, reduction of TAG synthesis in this mutant could cause cell damage due to energy overflow. In light of the fact that most carbon and energy during N deprivation enter starch rather than TAG it would be important to measure starch levels to determine C and E balances. Also the OH explanation for the PDG1 phenotype assumes that the mutation has no significant effect on the chloroplast membrane and thus the environmnent around the photocenters, which may induce stress on the organism during N deprivation that would be alleviated by inhibiting PSII. It would also be of interest to measure such dissipative mechanisms as NPQ and the Mehler reactions in the PDG1 mutant under N deprived culture conditions, to !!!!%('!verify whether there is evidence for increased light stress. Finally, while PGD1 cells accumulate ROS, this was only found to accumulate 7 days after N deprivation - well past the point where photosynthetic fluxes are substantial. This suggests (as the authors note) that ROS is more likely linked to autophagy than to energy overflow. We suggest that alternative explanations for starch and TAG accumulation by microalgae under nutrient deprivation should also be assessed experimentally. A plausible alternative role for TAG accumulation during N is that TAG serves as a storage site for, and a subsequent source of, acyl chains to aid in membrane degradation and resynthesis. One study following 14C labeled arachidonic acid found that this fatty acid was mobilized from TAG to chloroplast membrane during recovery {Khozin-Goldberg, 2005 #921}. The correlation between nutrient deprivation and the accumulation of TAG in photosynthetic algae has been known for over a century, its significance for green biotechnology development as well as for the ecophysiology of nutrient cycles makes understanding the adaptive causal links involved important and should be subjected to further experimental assessment. Future experiments should include 14C labeling during N deprivation and following the label during N recovery in total biomass, starch, TAG, amino acids and membrane lipids. !!!!%((! METHODS Culturing Chlamydomonas reinhardtii strain cc400 cw-15 mt++ was obtained from the Chlamydomonas Research Center and grown at 23 ¡C in liquid Tris-Acetate-Phosphate (TAP) media (Gorman, et al. 1965) and Sueoka High Salt (HS) media (Sueoka 1960) in 1L flasks shaken at 125 rpm under continuous illumination at 160, 40, 15, 5 and 0 µmol photons m(2 s(1 and ambient CO2 concentrations. Cell growth was determined by optical density measurements at 750 nm using a DU 800 spectrophotometer (Beckman-Coulter). Cultures were grown to cell densities of between 0.15 and 0.3 O.D. to minimize self-shading which becomes significant in denser cultures. Cells were counted using a Z series Coulter Counter cell and particle counter (Beckman-Coulter). For N deprivation, cells were centrifuged and resuspended in TAP media lacking ammonium chloride (nitrogen source). Chlorophyll Concentration Cells were collected by centrifugation of 1 ml culture and Chl was extracted in 1 ml of 80 % acetone for 20 minutes from pelleted samples after the supernatant culture medium was removed. After extraction, samples were pelleted by centrifugation and the supernatant was used for analysis. Chl was quantified spectroscopically as described in (Ritchie 2006) using a DU 800 spectrophotometer (Beckman Coulter). !!!!%(F!Ash Free dry weight 50 ml culture volume was harvested at each time point and cells were filtered onto Millipore Glass Fibre Prefilters (Millipore) under reduced pressure. Filtered cells were dried in an oven at 100¡C for 4 h and weighed on a Sartorius CP225D analytical balance. Weighed samples were then heated 540¡C to incinerate the cell biomass and the filters were reweighed (Zhu et al. 1997). Ash free dry weight (AFDW) was taken as the difference between sample weights before and after incineration. Carbon and Nitrogen Contents Aliquots of 50 ml of culture were harvested by centrifugation at each time point. Cells were then pelleted by centrifugation and frozen at -78¡C then lyophilized for 12h. Dried samples were weighed on a Sartorius CP225D analytical balance in tin capsules, which were then submitted to Duke Environmental Stable Isotope Laboratory (Duke University) for elemental analysis. Lipid analyses Cells harvested by centrifugation at 0oC and total lipids were extracted from samples containing the equivalent of ~10 mg of dry cell mass by the method of Folch (Folch, et al. 1957). Briefly, cells were extracted with 1ml of CHCl3:MeOH (1:2, vol/vol), vortexing samples for 2 min. The sample was then centrifuged and the supernatant was collected; the extraction was repeated two more times and the supernatant extracts were pooled. Extracts were dried under flowing N2 gas at room temperature and stored at -20 ¡C until analysis. Total TAG was quantified using Thin Layer Chromatography (TLC) separation and densitometry. Briefly total lipid extracts were loaded onto Analtech uniplate silica gel HL plates !!!!%(G!(Analtech) and separated with hexane: diethyl ether: acetic acid (70:30:1 vol/vol). After iodine staining, samples were scanned with a Cannon image class mf4690 scanner. ImageJ software (National Institute of Health) was used to quantitate TAG levels from the images. Conversion of integrated density into weight of TAG was performed by comparison with a biological standard sample (96hN-deprived Chlamydomonas TAG extract quantified by Gas Chromatography Flame Ionization Detection (GCFID) of its Fatty Acid Methyl Esters (FAMEs). FAMEs were quantified by GCFID after derivatization of lipid samples. C15 TAG in toluene was first added to total cell lipid extracts as internal standard. Total lipid extracts were treated with 150ul 2M methanolic KOH in 1ml hexane and vortexed for 5 min at room temperature completion. 6N HCl was added to neutralize the pH, samples were vortexed for 1 min and the hexane phase was transferred to another glass tube and dried under flowing N2 gas at room temperature and stored at -20¡C. Remaining free fatty acids were derivatized with silylation agent N-tert-butyldimethylsilyl- N-methyltrifluoroacetamide (MTBSTFA) (Sigma, Saint Louis, MO) in pyridine. Samples were dried, resuspended in 1 ml hexane, and quantified using an Agilent GCFID with a DB-23 column (Agilent) as previously described(Pollard et al. 2015). Oxygen Production and Consumption Rates Changes in dissolved oxygen in cell suspensions were measured with a NEOFOX analyzer FOXY-R probe with a FOXY-AF-MG coating (Ocean Optics) as previously described (Juergens, et al. 2015).The O2 sensor was immersed in 2 ml of culture in a capped 3ml cuvette with stirring. Net oxygen evolution was measured for 5 minutes at the same illumination level as !!!!%(H!the culture had been grown in and O2 consumption was measured for 1 minute in the dark immediately after the light period. In vivo fluorescence spectroscopy All spectroscopic measurements were performed with biological triplicates at each time point as previously described (Juergens, et al. 2015). Light-induced absorbance and chlorophyll fluorescence yield were measured using a kinetic spectrophotometer/fluorometer (Hall C, et al. 2011, Livingston, et al. 2010, Sacksteder, et al. 2001) modified for liquid samples by replacing the leaf holder with a temperature-controlled, stirring enabled cuvette holder (standard 1 cm pathlength). Cells were maintained under far red light LED (730nm) 20 min to oxidize the plastoquinone pool for accurate F0 measurements. After dark/ far red adaptation, the first saturating pulse for Chl fluorescence measurements was applied with a pulsed measuring beam (505 nm peak emission LED) filtered through a BG18 (Edmund Optics) glass filter. The sample was then illuminated with the its respective photosynthetic photon flux density (PPFD) using a pair of light emitting diodes (LEDs) (Luxeon III LXHL-PD09, Philips) with maximal emission at 620 nm, directed towards opposite sides of the cuvette, perpendicular to the measuring beam. Fluorescence yields from saturating pulses were measured under actinic light and averaged over 6 measurements, separated by 120s intervals. Both absorption and fluorescence measuring pulses were 20-35 *s in duration and attenuated to produce less than 0.1% increase in chlorophyll fluorescence yield in dark-adapted samples. The first dark interval relaxation kinetics trace measuring the electrochromic shift (ECS) kinetics (one trace per biological replicate) was recorded after three minutes of actinic illumination, followed by one !!!!%(I!minute of dark. Actinic LEDs were calibrated using a Licor LI190 PAR quantum sensor. CO2 Production and Consumption Rates CO2 exchange measurements were carried out with a LICOR XT 6400 (LICOR) infrared gas analyzer. Air was continuously circulated through 250ml flasks flask containing 50ml cultures, which were maintained under culture incubation conditions, and CO2 levels were recorded for the air entering and leaving the flask. Input CO2 levels were adjusted so that returning air contained 400 ppm. Starch Analysis Total glucose contained in starch was measured after amyloglucosidase and amylase digestion with the Megazyme total starch analysis kit (Megazyme, Ireland), similar to Work et al. (Work, et al. 2010). Briefly, pellets remaining after extraction of lipids from cells with 2:1 methanol:chloroform were autoclaved for 1 hour in 0.1 M Acetate buffer pH 4.8 then treated with $-amylase and amyloglucosidase for 1 hour at 55oC. Free glucose was quantitated with a colorimetric assay at 510nm using a starch assay kit (Megazyme, Ireland) according to the manufacturerÕs instructions. Acetate Analysis Total acetate in culture media was measured with an acetate analysis Kit (K-ACETRM, Megazyme, Ireland). TAP media samples were diluted fivefold to be within the linear response range and 1.5 ml sub-samples were used for NADH consumption measurement by absorbance changes at 340nm. Briefly, acetic acid and ATP are converted to acetyl phosphate and ADP by !!!!%(J!acetate kinase. The forward reaction is maintained by phosphotransacetlyase reaction of acetyl phosphate and coenzyme A to form acetyl-CoA and inorganic phosphate. Added phosphoenolpyruvate and ADP are converted to pyruvate and ATP by pyruvate kinase to maintain ATP levels and to produce pyruvate in proportion to the original levels of acetate. D-lactate dehydrogenase then converts the pyruvate and NADH into D-lactic acid and NAD+. Calculation of Carbon and Energy Balances Acetate and oxygen uptake rates were converted into µg C/ 106 cells/ hr. Net oxygen fluxes were assumed to represent net CO2 fluxes at a 1:1 ratio with 12g C/mol O2. Acetate consumption was multiplied by 2 moles C/mole acetate. Carbon accumulation rates in biomass were obtained from the percent carbon of the biomass and the biomass accumulated. FAME and Starch mass accumulation rates were converted to µg C/ 106 cells/ h by multiplying the by the mass fraction of carbon for those molecules (72 g C per 162g polymerized starch and 204gC /267g FAME). 13C labeling Algal cells were fed 100% uniformly 13C labeled Acetate in TAP media without N for the first 40h after N deprivation. Cells were then centrifuged and resuspended in unlabeled TAP media without N for a further 56 h (96 h of deprivation total). Samples were collected at 40, 72, and 96 h following N deprivation. Starch and Lipids were extracted and treated as above, glucose from starch samples was derivatized with methoxyamine and TMS as described in (Roessner et al. 2001). Labeling in FAME samples was analyzed by GC-MS as described previously (Allen et al. 2007) while glucose labeling was analyzed on GC-MS using the same chromatography and MS parameters previously used for amino acids by (Chen et al. 2011b). !!!!%FK! ACKNOWLEDGEMENTS We thank Dr.Õs David Kramer and Thomas Sharkey for generously making available to us equipment for making photosynthesis-related measurements and Dr. Sean Weise for advice on CO2 gas exchange measurements. !!!!%F%! REFERENCES !!!!%F&!REFERENCES Akita, T. and Kamo, M. (2015) Theoretical lessons for increasing algal biofuel: Evolution of oil accumulation to avert carbon starvation in microalgae. Journal of theoretical biology, 380, 183-191. Allen, A.E., Laroche, J., Maheswari, U., Lommer, M., Schauer, N., Lopez, P.J., Finazzi, G., Fernie, A.R. and Bowler, C. (2008) Whole-cell response of the pennate diatom Phaeodactylum tricornutum to iron starvation. Proc Natl Acad Sci U S A, 105, 10438-10443. Allen, D.K., Shachar-Hill, Y. and Ohlrogge, J.B. (2007) Compartment-specific labeling information in 13C metabolic flux analysis of plants. Phytochemistry, 68, 2197-2210. Alvarez, H.M. and Steinbuchel, A. (2002) Triacylglycerols in prokaryotic microorganisms. Appl. Microbiol. Biotechnol., 60, 367-376. Antal, T.K., Volgusheva, A.A., Kukarskikh, G.P., Krendeleva, T.E., Tusov, V.B. and Rubin, A.B. (2006) [Examination of chlorophyll fluorescence in sulfur-deprived cells of Chlamydomonas reinhardtii]. Biofizika, 51, 292-298. Atabani, A.E., Silitonga, A.S., Badruddin, I.A., Mahlia, T.M.I., Masjuki, H.H. and Mekhilef, S. (2012) A comprehensive review on biodiesel as an alternative energy resource and its characteristics. Renew. Sust. Energ. Rev., 16, 2070-2093. Ball, S.G., Dirick, L., Decq, A., Martiat, J.C. and Matagne, R.F. (1990) Physiology of Starch Storage in the Monocellular Alga Chlamydomonas-Reinhardtii. Plant Sci, 66, 1-9. Beijerinck, M. (1904) Das Assimilationsprodukt der Kohlens−ure in den Chromatophoren der Diatomeen. Recueil Travaux Botaniques N”erlandais, 1, 28-32. Blaby, I.K., Glaesener, A.G., Mettler, T., Fitz-Gibbon, S.T., Gallaher, S.D., Liu, B.S., Boyle, N.R., Kropat, J., Stitt, M., Johnson, S., Benning, C., Pellegrini, M., Casero, D. and Merchant, S.S. (2013) Systems-Level Analysis of Nitrogen Starvation-Induced Modifications of Carbon Metabolism in a Chlamydomonas reinhardtii Starchless Mutant. The Plant cell, 25, 4305-4323. Bonente, G., Pippa, S., Castellano, S., Bassi, R. and Ballottari, M. (2012) Acclimation of Chlamydomonas reinhardtii to different growth irradiances. J Biol Chem, 287, 5833-5847. Chapman, S.P., Paget, C.M., Johnson, G.N. and Schwartz, J.M. (2015) Flux balance analysis reveals acetate metabolism modulates cyclic electron flow and alternative glycolytic pathways in Chlamydomonas reinhardtii. Frontiers in plant science, 6, 474. !!!!%F'!Chen, X.W., Alonso, A.P., Allen, D.K., Reed, J.L. and Shachar-Hill, Y. (2011) Synergy between C-13-metabolic flux analysis and flux balance analysis for understanding metabolic adaption to anaerobiosis in E. coli. Metab Eng, 13, 38-48. Chisti, Y. (2007) Biodiesel from microalgae. Biotechnology Advances, 25, 294-306. Courchesne, N.M., Parisien, A., Wang, B. and Lan, C.Q. (2009) Enhancement of lipid production using biochemical, genetic and transcription factor engineering approaches. Journal of biotechnology, 141, 31-41. Fan, J., Yan, C., Andre, C., Shanklin, J., Schwender, J. and Xu, C. (2012) Oil accumulation is controlled by carbon precursor supply for fatty acid synthesis in Chlamydomonas reinhardtii. Plant & cell physiology, 53, 1380-1390. Folch, J., Lees, M. and Sloane Stanley, G.H. (1957) A simple method for the isolation and purification of total lipides from animal tissues. J Biol Chem, 226, 497-509. Frenk, S., Gomez, F., Ramos-Galvan, R. and Cravioto, J. (1958) Fatty liver in children; kwashiorkor. The American journal of clinical nutrition, 6, 298-309. Gargouri, M., Park, J.J., Holguin, F.O., Kim, M.J., Wang, H., Deshpande, R.R., Shachar-Hill, Y., Hicks, L.M. and Gang, D.R. (2015) Identification of regulatory network hubs that control lipid metabolism in Chlamydomonas reinhardtii. J Exp Bot, 66, 4551-4566. Genty, B., Briantais, J.M. and Baker, N.R. (1989) The Relationship between the Quantum Yield of Photosynthetic Electron-Transport and Quenching of Chlorophyll Fluorescence. Biochimica et biophysica acta, 990, 87-92. Goldman, J.C., Oswald, W.J. and Jenkins, D. (1974) Kinetics of Inorganic Carbon Limited Algal Growth. J Water Pollut Con F, 46, 554-574. Goodenough, U., Blaby, I., Casero, D., Gallaher, S.D., Goodson, C., Johnson, S., Lee, J.H., Merchant, S.S., Pellegrini, M., Roth, R., Rusch, J., Singh, M., Umen, J.G., Weiss, T.L. and Wulan, T. (2014) The path to triacylglyceride obesity in the sta6 strain of Chlamydomonas reinhardtii. Eukaryot Cell, 13, 591-613. Gorman, D.S. and Levine, R.P. (1965) CYTOCHROME F AND PLASTOCYANIN - THEIR SEQUENCE IN PHOTOSYNTHETIC ELECTRON TRANSPORT CHAIN OF CHLAMYDOMONAS REINHARDTII. P Natl Acad Sci USA, 54, 1665-&. Granum, E., Kirkvold, S. and Myklestad, S.M. (2002) Cellular and extracellular production of carbohydrates and amino acids by the marine diatom Skeletonema costatum: diel variations and effects of N depletion. Mar Ecol Prog Ser, 242, 83-94. Grossman, A.R., Gonzalez-Ballester, D., Shibagaki, N., Pootakham, W., Moseley, J. and Pootakham, W. (2010) Responses to Macronutrient Deprivation. !!!!%F(!Grundel, M., Scheunemann, R., Lockau, W. and Zilliges, Y. (2012) Impaired glycogen synthesis causes metabolic overflow reactions and affects stress responses in the cyanobacterium Synechocystis sp. PCC 6803. Microbiology, 158, 3032-3043. Hall C, Cruz J, Wood M, Zegara R, De Mars D, Carpenter J, Kanazawa A and D, K. (2011) Photosynthesis: research for food, fuel and future-15th international conference on photosynthesis. Zhejiang University Press, Bejing, 184-188. Harder, R. and Witsch, H.v. (1942a) Berichtr Versuche zur Fettsynthese mittels autotropher Microorganismen. . Forschungsdienst Sonderheft, 270-275. Harder, R. and Witsch, H.v. (1942b) Die Massenkultur von Diatomeen. Ber Deutsch Bot Ges, 146-152. Hays, S.G. and Ducat, D.C. (2015) Engineering cyanobacteria as photosynthetic feedstock factories. Photosynth. Res., 123, 285-295. Hu, Q., Sommerfeld, M., Jarvis, E., Ghirardi, M., Posewitz, M., Seibert, M. and Darzins, A. (2008) Microalgal triacylglycerols as feedstocks for biofuel production: perspectives and advances. Plant J, 54, 621-639. Jones, C.S. and Mayfield, S.P. (2012) Algae biofuels: versatility for the future of bioenergy. Curr Opin Biotechnol, 23, 346-351. Juergens, M.T., Deshpande, R.R., Lucker, B.F., Park, J.J., Wang, H., Gargouri, M., Holguin, F.O., Disbrow, B., Schaub, T., Skepper, J.N., Kramer, D.M., Gang, D.R., Hicks, L.M. and Shachar-Hill, Y. (2015) The Regulation of Photosynthetic Structure and Function during Nitrogen Deprivation in Chlamydomonas reinhardtii. Plant Physiol, 167, 558-573. Khozin-Goldberg, I., Shrestha, P. and Cohen, Z. (2005) Mobilization of arachidonyl moieties from triacylglycerols into chloroplastic lipids following recovery from nitrogen starvation of the microalga Parietochloris incisa. Bba-Mol Cell Biol L, 1738, 63-71. Klein, U. (1987) Intracellular Carbon Partitioning in Chlamydomonas reinhardtii. Plant Physiol, 85, 892-897. Klok, A.J., Lamers, P.P., Martens, D.E., Draaisma, R.B. and Wijffels, R.H. (2014) Edible oils from microalgae: insights in TAG accumulation. Trends in biotechnology, 32, 521-528. Kohlwein, S.D. (2010) Triacylglycerol homeostasis: insights from yeast. J Biol Chem, 285, 15663-15667. Li, X., Moellering, E.R., Liu, B., Johnny, C., Fedewa, M., Sears, B.B., Kuo, M.H. and Benning, C. (2012) A galactoglycerolipid lipase is required for triacylglycerol accumulation and survival following nitrogen deprivation in Chlamydomonas reinhardtii. The Plant cell, 24, 4670-4686. !!!!%FF!Li, Y., Han, D., Yoon, K., Zhu, S., Sommerfeld, M. and Hu, Q. (2013) Molecular and Cellular Mechanisms for Lipid Synthesis and Accumulation in Microalgae: Biotechnological Implications. In Handbook of Microalgal Culture. Liu, B.S. and Benning, C. (2013) Lipid metabolism in microalgae distinguishes itself. Current Opinion in Biotechnology, 24, 300-309. Liu, J., Huang, J., Sun, Z., Zhong, Y., Jiang, Y. and Chen, F. (2011) Differential lipid and fatty acid profiles of photoautotrophic and heterotrophic Chlorella zofingiensis: assessment of algal oils for biodiesel production. Bioresour Technol, 102, 106-110. Livingston, A.K., Cruz, J.A., Kohzuma, K., Dhingra, A. and Kramer, D.M. (2010) An Arabidopsis mutant with high cyclic electron flow around photosystem I (hcef) involving the NADPH dehydrogenase complex. The Plant cell, 22, 221-233. Martin, N.C. and Goodenough, U.W. (1975) Gametic differentiation in Chlamydomonas reinhardtii. I. Production of gametes and their fine structure. The Journal of cell biology, 67, 587-605. Merchant, S.S., Kropat, J., Liu, B., Shaw, J. and Warakanont, J. (2012) TAG, You're it! Chlamydomonas as a reference organism for understanding algal triacylglycerol accumulation. Current Opinion in Biotechnology, 23, 352-363. Merchant, S.S., Prochnik, S.E., Vallon, O., Harris, E.H., Karpowicz, S.J., Witman, G.B., Terry, A., Salamov, A., Fritz-Laylin, L.K., Marechal-Drouard, L., Marshall, W.F., Qu, L.H., Nelson, D.R., Sanderfoot, A.A., Spalding, M.H., Kapitonov, V.V., Ren, Q., Ferris, P., Lindquist, E., Shapiro, H., Lucas, S.M., Grimwood, J., Schmutz, J., Cardol, P., Cerutti, H., Chanfreau, G., Chen, C.L., Cognat, V., Croft, M.T., Dent, R., Dutcher, S., Fernandez, E., Fukuzawa, H., Gonzalez-Ballester, D., Gonzalez-Halphen, D., Hallmann, A., Hanikenne, M., Hippler, M., Inwood, W., Jabbari, K., Kalanon, M., Kuras, R., Lefebvre, P.A., Lemaire, S.D., Lobanov, A.V., Lohr, M., Manuell, A., Meier, I., Mets, L., Mittag, M., Mittelmeier, T., Moroney, J.V., Moseley, J., Napoli, C., Nedelcu, A.M., Niyogi, K., Novoselov, S.V., Paulsen, I.T., Pazour, G., Purton, S., Ral, J.P., Riano-Pachon, D.M., Riekhof, W., Rymarquis, L., Schroda, M., Stern, D., Umen, J., Willows, R., Wilson, N., Zimmer, S.L., Allmer, J., Balk, J., Bisova, K., Chen, C.J., Elias, M., Gendler, K., Hauser, C., Lamb, M.R., Ledford, H., Long, J.C., Minagawa, J., Page, M.D., Pan, J., Pootakham, W., Roje, S., Rose, A., Stahlberg, E., Terauchi, A.M., Yang, P., Ball, S., Bowler, C., Dieckmann, C.L., Gladyshev, V.N., Green, P., Jorgensen, R., Mayfield, S., Mueller-Roeber, B., Rajamani, S., Sayre, R.T., Brokstein, P., Dubchak, I., Goodstein, D., Hornick, L., Huang, Y.W., Jhaveri, J., Luo, Y., Martinez, D., Ngau, W.C., Otillar, B., Poliakov, A., Porter, A., Szajkowski, L., Werner, G., Zhou, K., Grigoriev, I.V., Rokhsar, D.S. and Grossman, A.R. (2007) The Chlamydomonas genome reveals the evolution of key animal and plant functions. Science, 318, 245-250. Miao, X.L. and Wu, Q.Y. (2004) High yield bio-oil production from fast pyrolysis by metabolic controlling of Chlorella protothecoides. Journal of biotechnology, 110, 85-93. !!!!%FG!Miller, R., Wu, G., Deshpande, R.R., Vieler, A., Gartner, K., Li, X., Moellering, E.R., Zauner, S., Cornish, A.J., Liu, B., Bullard, B., Sears, B.B., Kuo, M.H., Hegg, E.L., Shachar-Hill, Y., Shiu, S.H. and Benning, C. (2010) Changes in transcript abundance in Chlamydomonas reinhardtii following nitrogen deprivation predict diversion of metabolism. Plant Physiol, 154, 1737-1752. Moellering, E.R. and Banning, C. (2011) Galactoglycerolipid metabolism under stress: a time for remodeling. Trends in plant science, 16, 98-107. Morin, N., Cescut, J., Beopoulos, A., Lelandais, G., Le Berre, V., Uribelarrea, J.L., Molina-Jouve, C. and Nicaud, J.M. (2011) Transcriptomic Analyses during the Transition from Biomass Production to Lipid Accumulation in the Oleaginous Yeast Yarrowia lipolytica. PloS one, 6. Muller, P., Li, X.P. and Niyogi, K.K. (2001) Non-photochemical quenching. A response to excess light energy. Plant Physiol, 125, 1558-1566. Murphy, D.J. (2001) The biogenesis and functions of lipid bodies in animals, plants and microorganisms. Prog Lipid Res, 40, 325-438. Neijssel, O.M. and Tempest, D.W. (1975) The regulation of carbohydrate metabolism in Klebsiella aerogenes NCTC 418 organisms, growing in chemostat culture. Archives of microbiology, 106, 251-258. Niyogi, K.K., Bjorkman, O. and Grossman, A.R. (1997) Chlamydomonas xanthophyll cycle mutants identified by video imaging of chlorophyll fluorescence quenching. The Plant cell, 9, 1369-1380. Ohlrogge, J., Allen, D., Berguson, B., DellaPenna, D., Shachar-Hill, Y. and Stymne, S. (2009) Driving on Biomass. Science, 324, 1019-1020. Packter, N.M. and Olukoshi, E.R. (1995) Ultrastructural studies of neutral lipid localisation in Streptomyces. Archives of microbiology, 164, 420-427. Peers, G., Truong, T.B., Ostendorf, E., Busch, A., Elrad, D., Grossman, A.R., Hippler, M. and Niyogi, K.K. (2009) An ancient light-harvesting protein is critical for the regulation of algal photosynthesis. Nature, 462, 518-U215. Pollard, M., Martin, T.M. and Shachar-Hill, Y. (2015) Lipid analysis of developing Camelina sativa seeds and cultured embryos. Phytochemistry, 118, 23-32. Ritchie, R.J. (2006) Consistent sets of spectrophotometric chlorophyll equations for acetone, methanol and ethanol solvents. Photosynth Res, 89, 27-41. Rochaix, J.D. (2002) Chlamydomonas, a model system for studying the assembly and dynamics of photosynthetic complexes. FEBS Lett, 529, 34-38. !!!!%FH!Roessler, P.G. (1990) Environmental-Control of Glycerolipid Metabolism in Microalgae - Commercial Implications and Future-Research Directions. Journal of Phycology, 26, 393-399. Roessner, U., Luedemann, A., Brust, D., Fiehn, O., Linke, T., Willmitzer, L. and Fernie, A. (2001) Metabolic profiling allows comprehensive phenotyping of genetically or environmentally modified plant systems. The Plant cell, 13, 11-29. Sacksteder, C.A., Jacoby, M.E. and Kramer, D.M. (2001) A portable, non-focusing optics spectrophotometer (NoFOSpec) for measurements of steady-state absorbance changes in intact plants. Photosynth Res, 70, 231-240. Schmollinger, S., Muhlhaus, T., Boyle, N.R., Blaby, I.K., Casero, D., Mettler, T., Moseley, J.L., Kropat, J., Sommer, F., Strenkert, D., Hemme, D., Pellegrini, M., Grossman, A.R., Stitt, M., Schroda, M. and Merchant, S.S. (2014) Nitrogen-Sparing Mechanisms in Chlamydomonas Affect the Transcriptome, the Proteome, and Photosynthetic Metabolism. The Plant cell. Shifrin, N.S. and Chisholm, S.W. (1981) Phytoplankton Lipids - Interspecific Differences and Effects of Nitrate, Silicate and Light-Dark Cycles. Journal of Phycology, 17, 374-384. Siaut, M., Cuine, S., Cagnon, C., Fessler, B., Nguyen, M., Carrier, P., Beyly, A., Beisson, F., Triantaphylides, C., Li-Beisson, Y.H. and Peltier, G. (2011) Oil accumulation in the model green alga Chlamydomonas reinhardtii: characterization, variability between common laboratory strains and relationship with starch reserves. Bmc Biotechnol, 11. Singh, H., Shukla, M.R., Chary, K.V. and Rao, B.J. (2014) Acetate and bicarbonate assimilation and metabolite formation in Chlamydomonas reinhardtii: a 13C-NMR study. PloS one, 9, e106457. Slade, R. and Bauen, A. (2013) Micro-algae cultivation for biofuels: Cost, energy balance, environmental impacts and future prospects. Biomass Bioenerg, 53, 29-38. Smith, V.H., Sturm, B.S., Denoyelles, F.J. and Billings, S.A. (2010) The ecology of algal biodiesel production. Trends Ecol Evol, 25, 301-309. Solovchenko, A.E. (2012) Physiological Role of Neutral Lipid Accumulation in Eukaryotic Microalgae under Stresses. Russ. J. Plant Physiol., 59, 167-176. Spalding, M.H. (1989) Photosynthesis and Photorespiration in Fresh-Water Green-Algae. Aquat Bot, 34, 181-209. Spoehr, H.A. and Milner, H.W. (1949) The Chemical Composition of Chlorella - Effect of Environmental Conditions. Plant Physiol, 24, 120-149. Sueoka, N. (1960) Mitotic Replication of Deoxyribonucleic Acid in Chlamydomonas Reinhardi. Proc Natl Acad Sci U S A, 46, 83-91. !!!!%FI!Terauchi, A.M., Peers, G., Kobayashi, M.C., Niyogi, K.K. and Merchant, S.S. (2010) Trophic status of Chlamydomonas reinhardtii influences the impact of iron deficiency on photosynthesis. Photosynth Res, 105, 39-49. Wagner, H., Jakob, T. and Wilhelm, C. (2006) Balancing the energy flow from captured light to biomass under fluctuating light conditions. New Phytol, 169, 95-108. Watanabe, Y., Delanoue, J. and Hall, D.O. (1995) Photosynthetic Performance of a Helical Tubular Photobioreactor Incorporating the Cyanobacterium Spirulina-Platensis. Biotechnology and bioengineering, 47, 261-269. Weise, S.E., van Wijk, K.J. and Sharkey, T.D. (2011) The role of transitory starch in C(3), CAM, and C(4) metabolism and opportunities for engineering leaf starch accumulation. J Exp Bot, 62, 3109-3118. White, S., Anandraj, A. and Bux, F. (2011) PAM fluorometry as a tool to assess microalgal nutrient stress and monitor cellular neutral lipids. Bioresour Technol, 102, 1675-1682. Wijffels, R.H. and Barbosa, M.J. (2010) An Outlook on Microalgal Biofuels. Science, 329, 796-799. Williams, P.J.L. and Laurens, L.M.L. (2010) Microalgae as biodiesel & biomass feedstocks: Review & analysis of the biochemistry, energetics & economics. Energ Environ Sci, 3, 554-590. Wilson, A., Boulay, C., Wilde, A., Kerfeld, C.A. and Kirilovsky, D. (2007) Light-induced energy dissipation in iron-starved cyanobacteria: roles of OCP and IsiA proteins. The Plant cell, 19, 656-672. Work, V.H., Radakovits, R., Jinkerson, R.E., Meuser, J.E., Elliott, L.G., Vinyard, D.J., Laurens, L.M., Dismukes, G.C. and Posewitz, M.C. (2010) Increased lipid accumulation in the Chlamydomonas reinhardtii sta7-10 starchless isoamylase mutant and increased carbohydrate synthesis in complemented strains. Eukaryot Cell, 9, 1251-1261. Yoon, K., Han, D., Li, Y., Sommerfeld, M. and Hu, Q. (2012) Phospholipid:Diacylglycerol Acyltransferase Is a Multifunctional Enzyme Involved in Membrane Lipid Turnover and Degradation While Synthesizing Triacylglycerol in the Unicellular Green Microalga Chlamydomonas reinhardtii. The Plant cell, 24, 3708-3724. Zhu, C.J. and Lee, Y.K. (1997) Determination of biomass dry weight of marine microalgae. J. Appl. Phycol., 9, 189-194. !!!!%FJ! CHAPTER-4 Flux Balance Analysis of Chlamydomonas during Nitrogen Deprivation !!!!%GK!ABSTRACT Algae have been suggested as sources for biofuel production due to their ability to grow quickly and accumulate carbon in the form of starch and triacylglycerols. Despite much research on these accumulations, there is little consensus on the best approaches to increase production as well as the reasons behind the accumulations. In Chapter 3 we measured various carbon accumulations, photosynthetic properties, acetate uptake, and gas exchange rates during nitrogen deprivation in Chlamydomonas across 5 light levels and two media conditions. In this chapter we use that data as constraints for flux balance analysis using a genome scale model of Chlamydomonas, iCre1355. Flux results demonstrate that the model is robust across multiple growth conditions and representative of current knowledge about metabolism under each condition. Nitrogen deprivation experiments show a single pathway for starch accumulation across all light and media conditions but different metabolic routes for TAG accumulation depending upon the growth condition used. We additionally correct the iCre1355 model to have proper photosynthetic reaction stoichiometry as well as add several reactions left out in the original model. This work provides a foundation for a deeper metabolic understanding of carbon accumulation during nutrient deprivation. !!!!%G%!INTRODUCTION The current utilization of petroleum resources as fuel and chemical substrate is unsustainable (Solomon 2010). Further, variable fluctuations in the prices of these foreign derived commodities has dramatic effects on our economy (Brown et al. 2013). Additional costs of petroleum usage are found in deleterious effects on our environment, global warming, and public health. Developing dependable, clean, and sustainable renewable carbon and energy sources to replace petroleum is at the forefront of current bioenergy research. Renewable resources being developed rely on wind, solar, and the burning of photosynthetically derived biomass for energy. Biomass sources can be burned directly for electricity or processed through natural or artificial means to produce liquid and gas fuels as well as syngas and target molecules for the chemical industry. Crop plants such as maize and sugar cane were initially studied as feedstocks for biofuel (bioethanol); however, the focus has significantly shifted to developing fuels from single celled photosynthetic microalgae (Chisti 2007, Martin 2010, Ohlrogge, et al. 2009, Solomon 2010). Many algae have been studied for their capacity as industrial biofuel strains including Nannochloropsis (Radakovits, et al. 2012), Chlorella (Liu, et al. 2011), and Galdiaria (Selvaratnam, et al. 2014). Much of the research on biofuel production in microalgae however has been done on the model green alga Chlamydomonas reinhardtii (Fan, et al. 2012, Liu, et al. 2013, Siaut, et al. 2011). This alga has been studied extensively over the last fifty years (Goodenough 2015, Harris 2009b, Merchant, et al. 2012b) and has a plentiful list of resources such as a fully sequenced genome (Blaby, et al. 2014, Merchant, et al. 2007), multiple transcriptomic and proteomic analyses (Wang, et al. 2014), as well as various metabolic libraries (Bolling, et al. 2005, Lee do, et al. 2008, May, et al. 2008). Further, an extensive library of !!!!%G&!characterized mutants have been generated for Chlamydomonas enabling more rapid and extensive studies on gene-phenotype relationships (Li, et al. 2016). In recent years, several large scale studies have been done on Chlamydomonas during biofuel accumulation conditions, especially various nutrient deprivation regimes. These studies provide guidance for rational engineering and strain development for biofuel and engineering purposes. Aside from describing the mechanisms behind a given phenotype, one of the ultimate goals of systems biology is to be able to accurately direct genetic engineering efforts through predictive modeling. Omics data can be used to determine the set of enzymes, metabolic reactions and metabolites present in an organism and allow the formation of a metabolic model (Borodina et al. 2005). Metabolic reactions are then described in the form of linear reaction stoichiometries, accounting for all energy, reductant, carbon, and other components necessary for biology. Measured physiological data such as biomass, metabolite uptake rates, and known reaction rates can be used as constraints for Flux Balance Analysis (FBA), a systems biology techniqueused to study the generated metabolic model (Baroukh et al. 2015, Shachar-Hill 2013). FBA uses linear computational programming to balance all of the reaction stoichiometries in the model based upon the given constraints, with flux values for each reaction as output. Generated fluxes are used to describe the flow of carbon/ metabolites/ energy in an organism and, through perturbation, aid in the predictions of hypothesis driven genetic engineering efforts. FBA has been used in the prediction of several successful engineering efforts, including those in Corynebacterium glutamicum (Vallino et al. 1993), Sacchromyces cerevisiae (Famili et al. 2003), and Staphylococcus aureus (Heinemann et al. 2005). Despite the importance of understanding the relationship between phenotypes and metabolic flux, FBA has only recently been applied to photosynthetic organisms. Early phototrophic !!!!%G'!metabolic models of cyanobacteria Synechocystis and Spirulina were simplistic in nature, modeling only central metabolism and simplistic photosynthetic expressions (Cogne et al. 2003, Yang et al. 2000). Shastri and Morgan generated another model of Synechocystis, this time taking into consideration the reactions validated by the sequenced genome (Shastri et al. 2005). At the same time, efforts to characterize higher plant metabolism involved single network C13 metabolic flux analysis, a steady state analysis requiring expensive C13 label and intensive analytical chemistry analysis (Allen et al. 2009, Kruger et al. 2009, Libourel et al. 2008). This process only allowed for the description of a network, unable to predict the effects of gene knockouts or other genetic engineering perturbations. Boyle and Morgan built upon the Synechocystis model, generating one to describe the eukaryotic, green alga Chlamydomonas reinhardtii and metabolic compartmentation (Boyle, et al. 2009). This provided groundwork for the production of larger scale models in algae and higher plants. FBA genome scale models have thus been generated for Arabidopsis thaliana (de Oliveira Dal'Molin et al. 2010, Poolman et al. 2009) and Zea maize (Saha et al. 2011) with smaller, literature based models generated for barley (Grafahrend-Belau et al. 2009) and Brassica napus seeds (Hay et al. 2011, Pilalis et al. 2011). These models have helped determined where gaps are in our genome annotations, what reactions are in the same network, as well as described some interactions between metabolites in different compartments such as the cytosol, plastid, and mitochondria. Due to the vast amount of omic data generated and genome sequencing (Merchant, et al. 2007) in recent years, FBA has been applied to the green alga Chlamydomonas. Morgan and Boyle generated the first metabolic reconstruction of central metabolism in Chlamydomonas reinhardtii. This model included three compartments (plastid, cytosol, and mitochondria), 484 metabolic reactions, and led to enzyme annotation (Boyle, et al. 2009). This model was then !!!!%G(!followed by a transcript verified model (Manichaikul et al. 2009). Significant improvements were made in three larger, genome scale models (GEMs), AlgaGEM (derived from an Arabidopsis model (Dal'Molin, et al. 2011), iRC1080 (generated from a bottom up approach) (Chang et al. 2011b), and iCre1355 (a recently updated iRC1080) (Imam, et al. 2015). These models have been used to study the effect of CO2 and acetate upon photosynthesis (Chapman, et al. 2015, Melo et al. 2014), transcript regulation of nitrogen deprivation (Imam, et al. 2015, Lopez Garcia de Lomana, et al. 2015), and network reduction (Rugen et al. 2012). While these models are highly informed from genome, transcript, and protein systems analysies, they lack rigorous constraints based upon functional physiological measurements. In this work we used flux balance analysis to study the metabolic carbon and energy fluxes during normal growth and nutrient deprivation in Chlamydomonas. The iCre1355 metabolic model was modified to correct photosynthetic reaction stoichiometries and physiological measurements taken in chapter 3 were used to constrain the feasible flux space solutions. Carbon flux solutions generated gave results in congruence with current dogma on heterotrophic, autotrophic, and mixotrophic growth demonstrating the robustness of the model across several conditions. Photosynthetic fluxes and cellular photon demand results also provided support to conclusions of chapter 3. Nutrient deprivation experiments demonstrated a single source for starch synthesis in the plastid across all conditions while triacylglycerol carbon can be derived from several metabolic pathways depending upon the growth conditions of the cell. This work provides a foundation to explore ways to increase biofuel production in algae as well as study how metabolism responds to different growth environments. !!!!%GF!METHODS Model Curation In the work we chose to use the genome scale model iCre1355 updated by Imam et al. (Imam, et al. 2015) from the iCR1080 model by Chang (Chang, et al. 2011b). This model includes updated gene annotation and gene protein relations as well as deletions of unsupported gene reactions. Further, while other models for Chlamydomonas such as AlgaGem were generated from other organisms including Arabidopsis (Dal'Molin, et al. 2011, de Oliveira Dal'Molin, et al. 2010), iCre1355/ iCR1080 were built from the ground up and separate network reactions such as photosynthesis were separated into multiple equations instead of single stoichiometries. One example of this is linear electron flow where in AlgaGEM there is a single reaction while in iCre1355/iCR1080 there are separate reactions for PSII, b6f complex, PSI, ferredoxin NADPH reductase, and so on, allowing the model to more realistically demonstrate biology. !!!!"##! Table 4.1. Changes made to model related to photosynthesis. Reactions were corrected to match current stoichiometric understandings as well as remove free energy generation from sodium fluxes. !!!!"#$!Despite other works having used iCR1080 and iCre1355 and ÒupdatedÓ them, we found several cases of incorrect photosynthetic stoichiometries and free energy generation through sodium transport reactions (Chapman, et al. 2015, Cogne et al. 2011, Gomez et al. 2016, Lopez Garcia de Lomana, et al. 2015, Rugen, et al. 2012). Presented in Table 4.1, we show corrections in the stoichiometries of photosystem I and II, cyclic electron flow (CEF), and Ferredoxin NADPH reductase, and b6f complex. Chapman et al. 2015 claimed to correct CEF stoichiometries and send flux through plastoquinone instead of plastocyanin pools; however, these are not found in their published model. Additionally, four reactions were added for photosynthesis; the Mehler reaction which generates oxygen radicals at PSI(Mehler 1951a, Mehler 1951b, Mehler et al. 1952), the Plastid Terminal Oxidase (PTOX2, Cre03.g172500) which dissipates excess photosynthetic electrons by dumping them onto water (Houille-Vernes, et al. 2011), the second CEF pathway through NADPH dehydrogenase (NDH), using NADPH to reduce the quinone pool (Peng et al. 2009, Saroussi et al. 2016), and finally the Q cycle used by the b6f complex. The radical oxygen generated from the Mehler reaction Òo2RÓ is now reduced to water through the ascorbate peroxidase pathway, whereas previously this pathway was non- functional. Additional photosynthetic changes included removing the Òdummy O2Ó demand function and adding a useful photon count reaction to aid photon constraints. PSII generated oxygen now enters the thylakoid O2 pool instead of simply dissipating away. Finally the Q cycle is split into two reactions, the first takes one electron from plastoquinol and sends it to the plastocyanin while the second reaction takes the second electron and uses it to reduce another plastoquinone. Figure 4.1 shows the current photosynthetic pathways and areas of possible perturbation in the model. !!!!"#%! Other general changes include the deletion of sodium demand functions from organelles which generated free energy for the cell through sodium flows. The model was additionally altered to remove constraints on oxygen evolution, RuBisCO activity, CO2 and other components not strictly controlled in the media. Next, from our experiments, the ATP maintenance costs used in Imam et al. were too high (model was infeasible with our constraints) so we used the lower ATP value in Chang et al. Further, the ATP maintenance cost flux was moved from the cytosol Figure 4.1 Photosynthetic fluxes allowed in model. Fluxes at each of these sites can be constrained or knocked out depending on the needs of the modeler. ATPase: ATP synthase; B6F: B6F complex with quinone cycle; P680: 680nm photons; P673: 673nm photons; PQ: Plastoquinone; PGH2: Plastoquinol; PSI: Photosystem I; PSII: Photosystem II; FDXOX: oxidized ferrodoxin; FDXRD; reduced ferrodoxin; PTOX: Plastid Terminal Oxidase; NDH: NADPH hydrogenase; PGR5: Proton Gradient Regulation 5 !!!!!"#&!to the plastid to remove a H2O2 and proton gradient generating reaction that helped moved ATP from the plastid to the cytosol. The biomass equations presented in the model were adjusted placing half of the ATP demand in the cytosol and half in the plastid. Finally, a NADPH burning reaction was added to the plastid as a feature to allow the study of the effects of increased reductant upon the system. Flux Balance Analysis FBA was performed using a curated form of the pre-existing model of C. reinhardtii metabolism, iCre1355, as mentioned above. Measured CO2, biomass, TAG, starch, and acetate uptake rates were constrained in the models. Minimizing light use for autotrophic and CO2 emission for mixo- and heterotrophic cultures was used as an objective function. Flux predictions were generated for the various light and nutrient environments measured in Chapter 3. These constraints included growth rates, acetate uptake, light intensity, Ash free dry weights, carbon uptake, starch and TAG accumulation rates, and photosynthetic efficiencies. The Growth LED designed by Chang et al. (year) was chosen as the light source for the model to simplify the photon fluxes to either PSI or PSII. Flux Balance simulations were carried out using the COBRA Toolbox in the MATLAB programming environment (MathWorks, Inc. http://www.mathworks.com/ products/matlab). Gurobi Optimizer was used as the linear solver (www.Gurobi.com). Flux variability analysis was used to determine the max and min flux for each of the reactions during N+ and N- conditions (Mahadevan et al. 2003). !!!!"$'!Cell Cultivation, growth, and constraint measurements See Chapter 3 for growth and uptake measurement details. These include measurements of biomass % carbon, starch and TAG levels, as well as acetate, O2, and CO2 fluxes at 0, 5, 15, 40, and 160 µmol photons/m2/s light intensities in Tris-acetate- phosphate (mixotrophic) and high salt (autotrophic) conditions. !!!!"$"! RESULTS Measurements of biomass composition, oxygen evolution, carbon dioxide fluxes, and acetate uptake from chapter 3 were used as constraints for flux balance analysis in Chlamydomonas reinhardtii. Conditions tested included autotrophic, mixotrophic, and heterotrophic growth regimes during N replete growth, and 24 and 96 hrs after nitrogen depletion. N-depleted samples were constrained additionally with measured starch and triacylglycerol (TAG) accumulations with their respective demand functions. For cultures grown in the light (mixotrophic and autotrophic), the objective function was the minimization of light use (Pcount flux) as otherwise the model could use as much light as possible and lead to wasteful energy usage. Heterotrophic cultures objective function was set to minimize CO2 emissions, forcing a conservative use of carbon. Flux data generated from the balance analysis is presented in figure 4.2, simplified to primary carbon metabolism. Numbers presented indicate mmol Carbon/ gDW/Hr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itrogen Replete Conditions The flux maps presented under each of the three conditions demonstrates the robustness of the model to study various media conditions. Known metabolic activities of cells under each condition are supported - i.e., autotrophic growth has dominant fluxes through the Calvin cycle, mixotrophic uses both acetate and Calvin Benson cycle for carbon and energy, and the heterotrophic condition uses acetate and has an inactive Calvin Benson cycle. Both light using autotrophic and mixotrophic conditions also show a linear relationship between light consumed and growth rates demonstrating the models ability to represent different light conditions (Figure 4.3). Also, mixotrophic conditions receive more biomass per photon as acetate is present to aid biosynthesis and growth. y = 709.85x + 0.1457 R! = 1 y = 1331.6x - 53.317 R! = 0.99898 0 5 10 15 20 25 30 35 40 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 Photon demand (mmol/gDW/hr) specific growth rate(Hr-1) Figure 4.3. Photon demand vs. growth rate at 5, 40, 15, and 160 µmol photon/m2/s. Hollow circle represent autotrophic conditions and hollow squares represent mixotrophic conditions. !!!!"#$!These models additionally support the idea that cells at even moderate light intensities are exposed to more light than than needed. Figure 4.4 shows the calculated photon demands at 5, 40, 15, and 160 µmol photon/m2/s. vs. the relative light absorbed by the cell that was calculated in chapter 3. Here we find that although cells have the capacity to absorb significant amounts of light with increasing light intensity, only a small amount is actually needed. From chapter 3 we also found that our cells are carbon limited which will account for decreased photon demands. Autotrophic data shows this the most with almost no increase in light demand from 40 to 160 µmol photon/m2/s even though the potential absorbed light increases 3 fold. Aside from interesting photosynthetic outputs, these flux balances present insight into general metabolism. First, none of the maps showed flux through plastidic fructose bisphospate aldolase, instead using oxidative pentose phosphate pathway (OPP) reactions to generate fructose bisphosphate for starch production. This is interesting as this is normally a major gluconeogenic pathway. Second, we find 3-phosphoglycerate to be the dominant triose phosphate exported from the plastid from photosynthesis while glyceraldehyde-3-phosphate is generally depicted of as the main trios exported. Next, the maps generally depict little to no flux through traditional citric acid cycle, opting for flux through the glyoxolate cycle in the mitochondria(activity naturally found in the peroxisome). Finally, for all the maps there is a general flux from the plastid to the mitochondria in the form of aspartate and alpha ketoglutarate to oxaloacetate, moving ammonia and carbon. !!!!"#%! Figure 4.4. Photon demand vs. relative absorbed light at 5, 40, 15, and 160 µmol photon/m2/s. Hollow circle represent autotrophic conditions and hollow squares represent mixotrophic conditions. Relative absorbed light calculated from !ii * µg Chlorophyll/million cells* light intensity from data in chapter 3 for each light intensity. 0 5 10 15 20 25 30 35 40 0 20 40 60 80 100 Photon Demand (mmol/gDW/ Hr) Relative absorbed light !!!!"#&!Photosynthesis During N Deprivation In chapter three, I measured decreases in photosynthesis during nitrogen deprivation in both mixorophic and autotrophic conditions. Here it was found that mixotrophic cultures had higher photosynthetic activities during nitrogen replete conditions than autotrophic conditions and fell to values significantly lower than autotrophic during later timepoints of N deprivation. We linked this to mixotrophic cells being able to rely on fed acetate while autotrophic cells were reliant upon photosynthesis for energy and carbon even under N deprivation. We find the same trends in each growth condition with the flux modeling, presented in Figure 4.5. Figure 4.5. Light Demand predictions during N deprivation. A represents mixotrophic cultures while B represents autotrophic cultures. Black squares are 160, black diamonds are 40, hollow triangles are 15, and hollow circles are 5 µmol photons/ m2/s light intensities. Mixotrophic cells demonstrate a higher flux during N replete conditions and a lower photosynthetic flux during nutrient depletion. !!!!"#'!To assess how efficient cells are using light at different light intensities, we compared flux balance predictions to measured maximum available photons. Predicted photon demands were converted from mmol photon/gdw/hr to µmol photon/million cells/hr. Total available photons were calculated from measured light intensity, chlorophyll(capacity to abosorb light) and "II(efficiency of light use around PSII), making the assumption that light was hitting both photosynthetic complexes equally and that each complex uses light with the same efficiency. Absorbed light was calculated using the extinction coefficient of chlorophyll, chlorophyll abundance, and BeerÕs law, Absorbance = #(extinction coefficienct)*chlorophyll concentration*L (path length of light). FBA predicted photon use divided by the calculated maximum available light to the cell is presented in Figure 4.6. There are a number of conclusions one can derive from this figure. First, as light intensity increases, there is a decreasing efficiency of use from 80% at low light intensities down to 13% for higher light intensities. Second, both autotrophic and mixotrophic have similar light use efficiencies at higher light intensities but less so at low light intensities. Finally, it shows light constraints are relieved with increasing light intensity. The decreases in photosynthesis found under mixotrophic conditions extends beyond just photon demand. Table 4.2 shows how each photosynthetic flux changes during N deprivation. !!!!"#(! !!)!)*"!)*$!)*%!)*&!)*'!)*(!)*+!)*#!)*,!)!')!"))!"')!$))!!"#$%&'(')$*+,-.$%'++/01,$%&'(')+$2/3&($/)(,)+/(4$5*6'1$%&'(').67.+8$-./0!1230!Figure 4.6. FBA predicted light use vs. calculated light available for photosynthesis. Predicted photons from FBA were divided by the calculated photons available to photosynthesis under nutrient replete conditions across multiple light intensities. Filled squares indicate mixotrophic conditions(TAP media) while hollow squares indicate autotrophic conditions(HS media). !!!!"#+! Table 4.2 Photosynthetic Fluxes. Flux values generated through flux balance analysis at each light intensity and media condition at N+, 24 hrs after N- and 96 hrs after N-. All values are in units of mmol C/gDW/Hr. !!!!"##!Nitrogen Deprivation Metabolism Chlamydomonas was studied using Flux balance analysis at 24 and 96 hrs after deprivation. Under these conditions, nitrate and ammonium sources were constrained to zero, eliminating all biomass production. During these conditions the only drivers of synthesis are that of starch and TAG, removing all metabolism related to normal biomass production. During the first 24 hrs, we measured some cell growth as reported in chapter 3 but were unable to factor that in for this experiment. Under all growth conditions, starch precursors are produced from the fructose bisphosphate to glucose 1 phosphate and so on using the OPP pathway reactions. TAG carbon is derived from multiple sources depending upon the growth condition with primary chains coming from the plastid and elongation reactions derived from the mitochondria AcCoA pools. Mixotrophic TAG is derived predominantly from acetate (in congruence with chapter 3 conclusions) while autotrophic TAG is derived from plastidic pyruvate to amino acid to AcCoa metabolism. Meanwhile heterotrophic TAG is predicted to come from the mitochondrial AcCoA derived from acetate and transported to the plastid for fatty acid synthesis. With much interest in biofuel production this data suggests that the route to increased biofuel production depends upon the growth conditions one uses. !!!!"#,!DISCUSSION Here, the stoichiometric model iCre1355 was used to explore Chlamydomonas metabolism during both nitrogen replete and deplete conditions and found to produce accurate representations of known metabolism under three different growth conditions. Further, imposing measured constraints such as biomass and substrate uptakes upon the system gave photosynthetic results similar to those suggested in chapter three. In congruence with that data, actual light needed for starch and TAG biosynthesis is many times lower than that available to the cell. This is in congruence with evidence against the overflow hypothesis provided in chapter three. Finally this work showed that while nitrogen deprivation starch production comes from the plastid through fructose bisphosphate, TAG accumulations can come from various sources and pathways dependent upon growth conditions. The iCre1355 model (Imam, et al. 2015, Lopez Garcia de Lomana, et al. 2015), although newly updated with current gene lists, has been found to be missing correct reaction stoichiometries and known gene reactions for photosynthesis. This model did present itself as functioning and somewhat predictive but requires more intensive biochemistry annotation. Updates that I have added to the model allow for a number of activities. First, photosynthetic reactions have been corrected allowing for more realistic modeling results. Second, a counter of photons as well as an NADPH burning reaction have been added to control photon usage and system reduction respectively. Finally, The next step is to push the model to its limits, determining how much starch and TAG the system can make with current and potential carbon and energy inputs and how those more active metabolisms work and move carbon and energy. Metabolic pathways !!!!",)!with heightened fluxes may give rise to insight for engineering efforts. Additionally, the actual movement of fatty acids through membranes can be more fully characterized to give insight into lipid trafficking pathways and possible metabolic reasons for using one or another for TAG filling during nutrient deprivation. One way to do this might include adding an endoplasmic reticulum compartment and link the plastid lipids to it instead of the mitochondria as in this model. !!!!","!MODEL CHALLENGES Before the model demonstrated fluxes representing known activities, several problems had to become overcome. Initial modeling efforts in both autotrophic and mixotrophic conditions demonstrated large fluxes through small molecule, radical oxygen, and fermentative pathways. Small molecule (carbon monoxide and nitrous oxide), and hydrogen peroxide production pathways were constrained (set to .001 mmol C/gDW/Hr), forcing flux through the fermentative pathways to produce ethanol, acetate, formate, and lactate. Attempts to constrain (limit) these pathways led to protons being dumped from NADPH onto O2, forming water though the added photosynthetic PTOX reaction. Limiting PTOX caused PSII fluxes to remain unchanged. We believed that PSII activity was too high in relation to the needs of the cell, causing an overly reduced environment. Attempts to minimize PSII gave the same flux as before suggesting PSII or downstream NADPH production and dissipation was necessary for the modeling environment. Interestingly, if we constrained PSII to a low value while keeping minimization of light use as an objective function, cycle electron flow activity increased. There are a number of transporters requiring a proton motive force to move molecules across membranes including those for ATP. Moving the ATP maintenance cost from the cytosol to the plastid reduced the small molecule/ peroxide/ fermentative fluxes several fold demonstrating protons requirements in transport were partly responsible for the NADPH consumption. Attempts to further pin point the exact cause of the NADPH consumption and the hidden roles of proton generation in the model led to several dead ends and infeasible models. To address the problem, we installed a NADPH consumption reaction in the plastid that simply dissipates NADPH into NADP+ and H+. !!!!",$!Solving the model with this flux removed most reactions that appeared to be wastefully burning NADPH or producing excessive amounts of small molecules. All flux maps presented in this work were carried out with the NADPH burning reaction active. Intensive probing and revaluation of reaction stoichiometries of the system will be required to determine the true cause of the NADPH burn requirement. As mentioned before, another challenge laid in making sure the photosynthetic stoichiometries were correct. iCre1355 and its predecessor iCre1080 both demonstrated incorrect photosynthetic proton transport, incomplete cyclic electron flow reactions, and lack of Q cycle, Mehler reaction, PTOX, or NADPH Plastoquinone reductase. Previous studies demonstrated no flux through cyclic electron flow during autotrophic conditions however after correcting the reactions and constraining the system, we find significant flux though this reaction under nitrogen repletion. Additionally, the original model allowed for free energy generation through free sodium transport. Sodium was generated outside the cell and destroyed inside the cell generating a free flow of sodium that allowed free transmembrane transport of various metabolites. Further, the original model used a Òdummy O2Ó reaction that removed PSII generated oxygen instead of it entering the thylakoid lumen oxygen pool. !!!!",%!FUTURE DIRECTIONS The work presented in this chapter is ongoing. While the fluxes calculated through our modeling efforts are representative of actual trends measured in the lab, they may not be completely true in respect to the actual metabolism active in the organism. The genome scale model used in this study is extensive and has been generated from all gene reactions currently supported by transcript and proteomic analysis or required for basic metabolism activity however only roughly half of the Chlamydomonas genome has been functionally annotated at this time (Blaby, et al. 2014, Grossman et al. 2010b). This leaves large room for improvement to the total reactome (summation of all reactions) of the model, which may drastically alter the final fluxes. Further, the model used is static, stoichiometric one meaning without extensive user given constraints, it does not have the ability to reflect changes in substrate concentrations, negative or positive feedback regulation, changes in proton and other gradients, or other dynamic situations. Additionally, without extensive constraints on metabolisms known to be nonfunctional under tested conditions, they may be active. In our case we saw extensive nitric oxide, carbon monoxide, fermentation, and reactive oxygen species generation under normal N+ growth conditions before constraining to low fluxes. Groups such as Imam et al. (Imam, et al. 2015) used transcript data during N deprivation to constrain their flux space but again only half of the known genes have been annotated and may leave much room for improvement upon further annotation. Similarly, we chose light or CO2 use as the objective functions in this study however this may not be the actual. During normal growth it is not known what the primary objective function of an algae is and simply saying it is to grow as much as !!!!",&!possible or use as little light as possible may be very misleading. Further, using these constraints during nutrient deprivation may be equally far off the actual function of the cell. Cells are already flush with more light than they need under these conditions, what is stopping them from using as much light as they want wastefully, would it not provide an advantage in nature to absorb as much light as possible and prevent competitors from growing? Using different objective functions can give wildly different results and should be explored in future studies. To more properly constrain flux maps in the future and have a more accurate representation of the metabolism in Chlamydomonas, 13C metabolic flux analysis should be used(Allen, et al. 2009, Chen, et al. 2011b, Kim et al. 2008, Libourel, et al. 2008, O'Grady et al. 2012, Shachar-Hill 2013, Young, et al. 2011, Zamboni et al. 2009). This methodology uses 13C labeled substrate such as acetate or bicarbonate to monitor metabolism. Chlamydomonas grown in labeled substrate containing media will reach steady state incorporation and percent 13C labeling and patterning which can be analyzed through analytical methods such as mass spectrometry or nuclear magnetic resonance. Only a certain range of possible flux solutions exist to give rise to the labeling pattern seen in the organism and that can be calculated computationally. As this is a measure of actual metabolism in the organism, it can be used to check the validity of and guide constraints in flux balance analysis. Nitrogen deprivation is a continuously changing condition and will require more intensive dynamic labeling experiments instead of the steady state ones used for nitrogen replete conditions. Models can be further validated by using 13C substrates labeled at different positions as different carbons on a molecule may be directed to different metabolic end points. !!!!",'! In conclusion, while the model used in this study may be extensive it is by no means complete and further, more intensive work needs to be done to constrain the models so they are more representative of current metabolism. Once validated, the flux balance modeling technique can be used to predict the outcomes of gene knockouts, over expressions, and other genetic engineering efforts. !!!!",(!ACKNOWLEDGEMENTS We would like the thank Saheed Imam, Stephen Chapman, and Jean-Marc Schwartz for their help in understanding their models and flux balance assistance. !!!!",+! REFERENCES !!!!",#!REFERENCES Allen, D.K., Libourel, I.G.L. and Shachar-Hill, Y. (2009) Metabolic flux analysis in plants: coping with complexity. Plant Cell Environ, 32, 1241-1257. Baroukh, C., Munoz-Tamayo, R., Steyer, J.P. and Bernard, O. (2015) A state of the art of metabolic networks of unicellular microalgae and cyanobacteria for biofuel production. Metab Eng, 30, 49-60. Blaby, I.K., Blaby-Haas, C.E., Tourasse, N., Hom, E.F., Lopez, D., Aksoy, M., Grossman, A., Umen, J., Dutcher, S., Porter, M., King, S., Witman, G.B., Stanke, M., Harris, E.H., Goodstein, D., Grimwood, J., Schmutz, J., Vallon, O., Merchant, S.S. and Prochnik, S. (2014) The Chlamydomonas genome project: a decade on. Trends in plant science, 19, 672-680. Bolling, C. and Fiehn, O. (2005) Metabolite profiling of Chlamydomonas reinhardtii under nutrient deprivation. Plant Physiol, 139, 1995-2005. Borodina, I. and Nielsen, J. (2005) From genomes to in silico cells via metabolic networks. Curr Opin Biotechnol, 16, 350-355. Boyle, N.R. and Morgan, J.A. (2009) Flux balance analysis of primary metabolism in Chlamydomonas reinhardtii. BMC systems biology, 3, 4. Brown, S.P.A. and Kennelly, R. (2013) Consequences of U.S. Dependence on Foreign Oil. University of Nevada, Las Vegas: National Energy Policy Institute. Chang, R.L., Ghamsari, L., Manichaikul, A., Hom, E.F.Y., Balaji, S., Fu, W., Shen, Y., Hao, T., Palsson, B.¯., Salehi-Ashtiani, K. and Papin, J.A. (2011) Metabolic network reconstruction of Chlamydomonas offers insight into light-driven algal metabolism. Molecular systems biology, 7, 518-518. Chapman, S.P., Paget, C.M., Johnson, G.N. and Schwartz, J.M. (2015) Flux balance analysis reveals acetate metabolism modulates cyclic electron flow and alternative glycolytic pathways in Chlamydomonas reinhardtii. Frontiers in plant science, 6, 474. Chen, X.W., Alonso, A.P., Allen, D.K., Reed, J.L. and Shachar-Hill, Y. (2011) Synergy between C-13-metabolic flux analysis and flux balance analysis for understanding metabolic adaption to anaerobiosis in E. coli. Metab Eng, 13, 38-48. Chisti, Y. (2007) Biodiesel from microalgae. Biotechnology Advances, 25, 294-306. !!!!",,!Cogne, G., Gros, J.B. and Dussap, C.G. (2003) Identification of a metabolic network structure representative of Arthrospira (spirulina) platensis metabolism. Biotechnology and bioengineering, 84, 667-676. Cogne, G., Rgen, M., Bockmayr, A., Titica, M., Dussap, C.-G., Cornet, J.-F. and Legrand, J. (2011) A model-based method for investigating bioenergetic processes in autotrophically growing eukaryotic microalgae: Application to the green algae Chlamydomonas reinhardtii. Biotechnology progress, 27, 631-640. Dal'Molin, C.G., Quek, L.E., Palfreyman, R.W. and Nielsen, L.K. (2011) AlgaGEM--a genome-scale metabolic reconstruction of algae based on the Chlamydomonas reinhardtii genome. Bmc Genomics, 12 Suppl 4, S5. de Oliveira Dal'Molin, C.G., Quek, L.E., Palfreyman, R.W., Brumbley, S.M. and Nielsen, L.K. (2010) AraGEM, a genome-scale reconstruction of the primary metabolic network in Arabidopsis. Plant Physiol, 152, 579-589. Famili, I., Fırster, J., Nielsen, J. and Palsson, B.O. (2003) Saccharomyces cerevisiae phenotypes can be predicted by using constraint-based analysis of a genome-scale reconstructed metabolic network. P Natl Acad Sci USA, 100, 13134-13139. Fan, J., Yan, C., Andre, C., Shanklin, J., Schwender, J. and Xu, C. (2012) Oil accumulation is controlled by carbon precursor supply for fatty acid synthesis in Chlamydomonas reinhardtii. Plant & cell physiology, 53, 1380-1390. Gomez, J.A., Hoffner, K. and Barton, P.I. (2016) From sugars to biodiesel using microalgae and yeast. Green Chemistry, 18, 461-475. Goodenough, U. (2015) Historical perspective on Chlamydomonas as a model for basic research: 1950Ð1970. The Plant Journal, 82, 365-369. Grafahrend-Belau, E., Schreiber, F., Koschutzki, D. and Junker, B.H. (2009) Flux Balance Analysis of Barley Seeds: A Computational Approach to Study Systemic Properties of Central Metabolism. Plant Physiol, 149, 585-598. Grossman, A.R., Karpowicz, S.J., Heinnickel, M., Dewez, D., Hamel, B., Dent, R., Niyogi, K.K., Johnson, X., Alric, J., Wollman, F.A., Li, H. and Merchant, S.S. (2010) Phylogenomic analysis of the Chlamydomonas genome unmasks proteins potentially involved in photosynthetic function and regulation. Photosynth Res, 106, 3-17. Harris, E.H. (2009) Preface to Volume 1. In The Chlamydomonas Sourcebook (Second Edition) (Harris, E.H., Stern, D.B. and Witman, G.B. eds). London: Academic Press, pp. ix. Hay, J. and Schwender, J. (2011) Metabolic network reconstruction and flux variability analysis of storage synthesis in developing oilseed rape (Brassica napus L.) embryos. Plant J, 67, 526-541. !!!!$))!Heinemann, M., Kummel, A., Ruinatscha, R. and Panke, S. (2005) In silico genome-scale reconstruction and validation of the Staphylococcus aureus metabolic network. Biotechnology and bioengineering, 92, 850-864. Houille-Vernes, L., Rappaport, F., Wollman, F.A., Alric, J. and Johnson, X. (2011) Plastid terminal oxidase 2 (PTOX2) is the major oxidase involved in chlororespiration in Chlamydomonas. Proc Natl Acad Sci U S A, 108, 20820-20825. Imam, S., Schauble, S., Valenzuela, J., Lopez Garcia de Lomana, A., Carter, W., Price, N.D. and Baliga, N.S. (2015) A refined genome-scale reconstruction of Chlamydomonas metabolism provides a platform for systems-level analyses. Plant J, 84, 1239-1256. Kim, H.U., Kim, T.Y. and Lee, S.Y. (2008) Metabolic flux analysis and metabolic engineering of microorganisms. Mol. Biosyst., 4, 113-120. Kruger, N.J. and Ratcliffe, R.G. (2009) Insights into plant metabolic networks from steady-state metabolic flux analysis. Biochimie, 91, 697-702. Lee do, Y. and Fiehn, O. (2008) High quality metabolomic data for Chlamydomonas reinhardtii. Plant methods, 4, 7. Li, X., Zhang, R., Patena, W., Gang, S.S., Blum, S.R., Ivanova, N., Yue, R., Robertson, J.M., Lefebvre, P., Fitz-Gibbon, S.T., Grossman, A.R. and Jonikas, M.C. (2016) An indexed, mapped mutant library enables reverse genetics studies of biological processes in Chlamydomonas reinhardtii. The Plant cell. Libourel, I.G.L. and Shachar-Hill, Y. (2008) Metabolic flux analysis in plants: From intelligent design to rational engineering. Annu Rev Plant Biol, 59, 625-650. Liu, B.S. and Benning, C. (2013) Lipid metabolism in microalgae distinguishes itself. Current Opinion in Biotechnology, 24, 300-309. Liu, J., Huang, J., Sun, Z., Zhong, Y., Jiang, Y. and Chen, F. (2011) Differential lipid and fatty acid profiles of photoautotrophic and heterotrophic Chlorella zofingiensis: assessment of algal oils for biodiesel production. Bioresour Technol, 102, 106-110. Lopez Garcia de Lomana, A., Schauble, S., Valenzuela, J., Imam, S., Carter, W., Bilgin, D.D., Yohn, C.B., Turkarslan, S., Reiss, D.J., Orellana, M.V., Price, N.D. and Baliga, N.S. (2015) Transcriptional program for nitrogen starvation-induced lipid accumulation in Chlamydomonas reinhardtii. Biotechnology for biofuels, 8, 207. Mahadevan, R. and Schilling, C.H. (2003) The effects of alternate optimal solutions in constraint-based genome-scale metabolic models. Metab Eng, 5, 264-276. !!!!$)"!Manichaikul, A., Ghamsari, L., Hom, E.F., Lin, C., Murray, R.R., Chang, R.L., Balaji, S., Hao, T., Shen, Y., Chavali, A.K., Thiele, I., Yang, X., Fan, C., Mello, E., Hill, D.E., Vidal, M., Salehi-Ashtiani, K. and Papin, J.A. (2009) Metabolic network analysis integrated with transcript verification for sequenced genomes. Nature methods, 6, 589-592. Martin, M.A. (2010) First generation biofuels compete. New biotechnology, 27, 596-608. May, P., Wienkoop, S., Kempa, S., Usadel, B., Christian, N., Rupprecht, J., Weiss, J., Recuenco-Munoz, L., Ebenhoh, O., Weckwerth, W. and Walther, D. (2008) Metabolomics- and proteomics-assisted genome annotation and analysis of the draft metabolic network of Chlamydomonas reinhardtii. Genetics, 179, 157-166. Mehler, A.H. (1951a) Studies on reactions of illuminated chloroplasts. I. Mechanism of the reduction of oxygen and other Hill reagents. Archives of biochemistry and biophysics, 33, 65-77. Mehler, A.H. (1951b) Studies on reactions of illuminated chloroplasts. II. Stimulation and inhibition of the reaction with molecular oxygen. Archives of biochemistry and biophysics, 34, 339-351. Mehler, A.H. and Brown, A.H. (1952) Studies on reactions of illuminated chloroplasts. III. Simultaneous photoproduction and consumption of oxygen studied with oxygen isotopes. Archives of biochemistry and biophysics, 38, 365-370. Melo, D.O.P., Moncada, R.J.-P., Winck, F.V. and Barrios, A.F.G. (2014) In Silico Analysis for Biomass Synthesis under Different CO2 Levels for Chlamydomonas reinhardtii Utilizing a Flux Balance Analysis Approach. In Advances in Computational Biology: Proceedings of the 2nd Colombian Congress on Computational Biology and Bioinformatics (CCBCOL) (Castillo, F.L., Cristancho, M., Isaza, G., PinzŠn, A. and Rodr™guez, C.J.M. eds). Cham: Springer International Publishing, pp. 279-285. Merchant, S.S., Kropat, J., Liu, B., Shaw, J. and Warakanont, J. (2012) TAG, You're it! Chlamydomonas as a reference organism for understanding algal triacylglycerol accumulation. Current Opinion in Biotechnology, 23, 352-363. Merchant, S.S., Prochnik, S.E., Vallon, O., Harris, E.H., Karpowicz, S.J., Witman, G.B., Terry, A., Salamov, A., Fritz-Laylin, L.K., Marechal-Drouard, L., Marshall, W.F., Qu, L.H., Nelson, D.R., Sanderfoot, A.A., Spalding, M.H., Kapitonov, V.V., Ren, Q., Ferris, P., Lindquist, E., Shapiro, H., Lucas, S.M., Grimwood, J., Schmutz, J., Cardol, P., Cerutti, H., Chanfreau, G., Chen, C.L., Cognat, V., Croft, M.T., Dent, R., Dutcher, S., Fernandez, E., Fukuzawa, H., Gonzalez-Ballester, D., Gonzalez-Halphen, D., Hallmann, A., Hanikenne, M., Hippler, M., Inwood, W., Jabbari, K., Kalanon, M., Kuras, R., Lefebvre, P.A., Lemaire, S.D., Lobanov, A.V., Lohr, M., Manuell, A., !!!!$)$!Meier, I., Mets, L., Mittag, M., Mittelmeier, T., Moroney, J.V., Moseley, J., Napoli, C., Nedelcu, A.M., Niyogi, K., Novoselov, S.V., Paulsen, I.T., Pazour, G., Purton, S., Ral, J.P., Riano-Pachon, D.M., Riekhof, W., Rymarquis, L., Schroda, M., Stern, D., Umen, J., Willows, R., Wilson, N., Zimmer, S.L., Allmer, J., Balk, J., Bisova, K., Chen, C.J., Elias, M., Gendler, K., Hauser, C., Lamb, M.R., Ledford, H., Long, J.C., Minagawa, J., Page, M.D., Pan, J., Pootakham, W., Roje, S., Rose, A., Stahlberg, E., Terauchi, A.M., Yang, P., Ball, S., Bowler, C., Dieckmann, C.L., Gladyshev, V.N., Green, P., Jorgensen, R., Mayfield, S., Mueller-Roeber, B., Rajamani, S., Sayre, R.T., Brokstein, P., Dubchak, I., Goodstein, D., Hornick, L., Huang, Y.W., Jhaveri, J., Luo, Y., Martinez, D., Ngau, W.C., Otillar, B., Poliakov, A., Porter, A., Szajkowski, L., Werner, G., Zhou, K., Grigoriev, I.V., Rokhsar, D.S. and Grossman, A.R. (2007) The Chlamydomonas genome reveals the evolution of key animal and plant functions. Science, 318, 245-250. O'Grady, J., Schwender, J., Shachar-Hill, Y. and Morgan, J.A. (2012) Metabolic cartography: experimental quantification of metabolic fluxes from isotopic labelling studies. J Exp Bot, 63, 2293-2308. Ohlrogge, J., Allen, D., Berguson, B., DellaPenna, D., Shachar-Hill, Y. and Stymne, S. (2009) Driving on Biomass. Science, 324, 1019-1020. Peng, L., Fukao, Y., Fujiwara, M., Takami, T. and Shikanai, T. (2009) Efficient operation of NAD(P)H dehydrogenase requires supercomplex formation with photosystem I via minor LHCI in Arabidopsis. The Plant cell, 21, 3623-3640. Pilalis, E., Chatziioannou, A., Thomasset, B. and Kolisis, F. (2011) An In Silico Compartmentalized Metabolic Model of Brassica Napus Enables the Systemic Study of Regulatory Aspects of Plant Central Metabolism. Biotechnology and bioengineering, 108, 1673-1682. Poolman, M.G., Miguet, L., Sweetlove, L.J. and Fell, D.A. (2009) A Genome-Scale Metabolic Model of Arabidopsis and Some of Its Properties. Plant Physiol, 151, 1570-1581. Radakovits, R., Jinkerson, R.E., Fuerstenberg, S.I., Tae, H., Settlage, R.E., Boore, J.L. and Posewitz, M.C. (2012) Draft genome sequence and genetic transformation of the oleaginous alga Nannochloropsis gaditana. Nature Communications, 3. Rugen, M., Bockmayr, A., Legrand, J. and Cogne, G. (2012) Network reduction in metabolic pathway analysis: elucidation of the key pathways involved in the photoautotrophic growth of the green alga Chlamydomonas reinhardtii. Metab Eng, 14, 458-467. Saha, R., Suthers, P.F. and Maranas, C.D. (2011) Zea mays iRS1563: A Comprehensive Genome-Scale Metabolic Reconstruction of Maize Metabolism. PloS one, 6. !!!!$)%!Saroussi, S.I., Wittkopp, T.M. and Grossman, A.R. (2016) The type II NADPH dehydrogenase facilitates cyclic electron flow, energy dependent quenching and chlororespiratory metabolism during acclimation of Chlamydomonas reinhardtii to nitrogen deprivation. Plant Physiol. Selvaratnam, T., Pegallapati, A.K., Montelya, F., Rodriguez, G., Nirmalakhandan, N., Van Voorhies, W. and Lammers, P.J. (2014) Evaluation of a thermo-tolerant acidophilic alga, Galdieria sulphuraria, for nutrient removal from urban wastewaters. Bioresour Technol, 156, 395-399. Shachar-Hill, Y. (2013) Metabolic network flux analysis for engineering plant systems. Current Opinion in Biotechnology, 24, 247-255. Shastri, A.A. and Morgan, J.A. (2005) Flux balance analysis of photoautotrophic metabolism. Biotechnology progress, 21, 1617-1626. Siaut, M., Cuine, S., Cagnon, C., Fessler, B., Nguyen, M., Carrier, P., Beyly, A., Beisson, F., Triantaphylides, C., Li-Beisson, Y.H. and Peltier, G. (2011) Oil accumulation in the model green alga Chlamydomonas reinhardtii: characterization, variability between common laboratory strains and relationship with starch reserves. Bmc Biotechnol, 11. Solomon, B.D. (2010) Biofuels and sustainability. Annals of the New York Academy of Sciences, 1185, 119-134. Vallino, J.J. and Stephanopoulos, G. (1993) Metabolic flux distributions in Corynebacterium glutamicum during growth and lysine overproduction. Biotechnology and bioengineering, 41, 633-646. Wang, H., Gau, B., Slade, W.O., Juergens, M., Li, P. and Hicks, L.M. (2014) The global phosphoproteome of Chlamydomonas reinhardtii reveals complex organellar phosphorylation in the flagella and thylakoid membrane. Molecular & cellular proteomics : MCP, 13, 2337-2353. Yang, C., Hua, Q. and Shimizu, K. (2000) Energetics and carbon metabolism during growth of microalgal cells under photoautotrophic, mixotrophic and cyclic light-autotrophic/dark-heterotrophic conditions. Biochemical Engineering Journal, 6, 87-102. Young, J.D., Shastri, A.A., Stephanopoulos, G. and Morgan, J.A. (2011) Mapping photoautotrophic metabolism with isotopically nonstationary C-13 flux analysis. Metab Eng, 13, 656-665. Zamboni, N., Fendt, S.M., Ruhl, M. and Sauer, U. (2009) (13)C-based metabolic flux analysis. Nature protocols, 4, 878-892. !!!!$)&! CHAPTER-5 Conclusions and Future Directions !!!!$)'!CARBON ACCUMULATIONS DURING NUTRIENT LIMITATIONS In nature, one of the most common problems life faces is nutrient limitation, especially that of nitrogen. Across biology, organisms respond with slowing metabolisms and accumulations of carbon, be it glycogen (Hasunuma et al. 2013), starch (Ball, et al. 1990), triacylglycerols (TAGs) (Merchant, et al. 2012b, Siaut, et al. 2011), or other energy dense molecules. In the field of biofuel technology, researchers seek to use these accumulated molecules as sources of energy as fuels or for carbon substrate for the chemical industry. Algae are photosynthetic organisms that can accumulate these compounds under nutrient stress and have been the target of extensive research over the last several decades, especially for the production of TAG. Despite this work, the reasons cells accumulate one carbon compound or another as well as the role these stored carbon compounds play in the nutrient deplete/ replete life cycle of organisms remain undefined. Several explanations have been proposed for the induction of TAG accumulation in algae under stress, ranging from storing reduced carbon as an energy source for survival and/or future recovery, to lipid reorganization during photosynthetic down-regulation and/or subsequent up-regulation, to photo-protection from excess photosynthetic energy and carbon (Akita, et al. 2015, Grossman, et al. 2010a, Hu, et al. 2008, Khozin-Goldberg, et al. 2005, Klok, et al. 2014, Kohlwein 2010, Murphy 2001, Roessler 1990). These roles can be divided into four classifications: (i) sink for excess fixed carbon, (ii) sink for excess photosynthetic energy, (iii) carbon storage to aid nutrient recovery, and (iv) energy storage to aid nutrient recovery. The work in this thesis has sought to deepen the understanding of the relationship between photosynthesis and !!!!$)(!TAG accumulation in microalgae and either support or counter the above motivations through the analysis of carbon and energy fluxes. PHOTOSYNTHESIS AND NUTRIENT DEPRIVATION In Chapter 2, I present the effects of nitrogen deprivation on photosynthesis in Chlamydomonas reinhardtii. This work demonstrates drastic decreases in the photosynthetic capacity and efficiency during the first 48 hrs of N deprivation. Using transcript and protein profiles along with functional photosynthetic measurements we came to the conclusion that cells decrease photosynthesis in a controlled and coordinated manner. Further when non-photochemical quenching (a measure of photosynthetic stress) was calculated from chlorophyll fluorescence measurements, it was found that this parameter decreased overtime suggesting the cells were not under photosynthetic stress and were able to dissipate any ÒexcessÓ energy. Evidence for posttranslational control of photosynthetic activity demonstrates cells are able to decrease photosynthetic parameters well before there are changes in protein abundances. Another manuscript from the same study found transcript and protein increases for genes related to nitrogen assimilation demonstrating that the cells focus on priming themselves for the return of nitrogen (Park, et al. 2014). This suggests algae have evolved an encoded systemic response to nutrient deprivation that provides them some advantage to survive the stress. !!!!$)+!OVERFLOW HYPOTHESIS CONCLUSIONS Several works have suggested the motivation for starch and TAG accumulation during nutrient deprivation is related to photosynthetic carbon or energy overflow(Hu, et al. 2008, Li, et al. 2012b, Roessler 1990). The most prominent study in support of the overflow hypothesis involves analyses of the Chlamydomonas PGD1 mutant which accumulates only half as much TAG as its parent line. My arguments against the mutants support of the overflow hypothesis are given in chapter three in more detail. My work in chapter 2 did not support this hypothesis. To expand upon this and determine if there was evidence for these overflows I studied the relationship between photosynthesis and accumulated starch and TAG molecules during nutrient deprivation across 2 media conditions and 5 light intensities (presented in chapter 3). I chose to work at light intensities representative of normal physiological growth where the organism would not be expected to experience photosynthetic stress, allowing N deprivation to be the sole inducer if any existed. Further using both autotrophic and mixotrophic conditions would allow me to see the affects increased total carbon available on the system would be. If photosynthetic overflows are the main drivers of carbon accumulation I would expect a number of things: (i) No carbon would accumulate in the dark during N deprivation, (ii) there would be no carbon accumulation at low light intensities where light would be expected to be growth limiting, (iii) All carbon accumulations would come from photosynthetic ally fixed carbon, even under mixotrophic conditions, (iv) Cells would demonstrate significant photosynthetic stress in the form of NPQ, (v) there would be a proportionate increase in carbon accumulation per !!!!$)#!increases in light intensity, and finally (vi) carbon accumulations would be highest when photosynthetic rates were highest, decreasing over time. In this work I found the contrary to these points. Algae were able to accumulate significant amounts of starch and TAG at low light intensities and in the dark demonstrating carbon accumulations do not depend upon photosynthetically fixed carbon or energy. Second, under our conditions, algae demonstrated a decrease in total NPQ over the timecourse suggesting they are able to safely control their energy intake and are not under significant photosynthetic stress. Third, there was not a proportionate increase in starch or TAG accumulations with increasing light intensity for both mixotrophic and autotrophic conditions pointing to other regulatory mechanisms controlling the accumulations. Under mixotrophic conditions, through 13C acetate labeling during n deprivation, we found almost all TAG carbon coming from acetate with a minority of it supporting starch accumulations suggesting TAG and starch accumulations are under separate regulatory and carbon networks. Additionally, if there indeed is a photosynthetic overflow sink, it would be starch due to accumulation of photosynthetic carbon and it being the first pool to accumulate. Finally, I found TAG accumulation rates to be at their highest during the later phases of nutrient deprivation, well after photosynthetic capacity and efficiency have decreased excessively, a result counter-intuitive to the overflow hypothesis. While this work has not described what the fundamental motivations of algal oil accumulation are, this contributes a strong step forward in saying what are not the drivers. My data strongly suggests TAG accumulations are not induced by photosynthetic stresses. This points to a regulated storage role for these molecules that !!!!$),!aids recovery from nutrient deprivation through either/ both carbon and energy stores( iii and iv above). Further, chapter 3 data points to a relationship between initial N+ growth rates and carbon accumulation rates with faster growing cells consuming and utilizing environmental N at a faster rate than slower growing cells, causing them to enter the N deprived carbon accumulating stage faster. Future research efforts should then be driven to look at the regulatory network, especially that around growth, for hints on controls of carbon accumulation. MODELING OF NUTRIENT DEPRIVED METABOLISM In chapter 4, I used a stoichiometric model of Chlamydomonas to study N+ and N- carbon metabolism to determine the most important metabolisms for growth, starch, and TAG accumulations. Here, physiological measurements from chapter 3 were used as constraints for Flux Balance Analysis to probe the genome scale metabolic model. Nitrogen deprivation experiments showed that across all light and media conditions, starch was being produced from pentose phosphate pathway fructose generation while TAG carbon came from multiple different locations depending upon the media source and light condition. Reduced carbon such as acetate is demonstrated as a preferred substrate for oil accumulation over photosynthetically derived carbon. Further corrections made to the photosynthetic reaction stoichiometries allows for more realistic flux studies on photosynthesis and its interaction with metabolism. !!!!$")!FUTURE EXPERIMENTS AND DIRECTION The preliminary data presented in Appendix 1 demonstrates interesting relationships between starch and TAG breakdown and cell growth during recovery from nutrient deprivation. First, it is demonstrated that starch degradation precedes TAG degradation suggesting unique roles for each storage pool. Second, photosynthesis was found to increase in efficiency and capacity during the first 24 hrs of N recovery suggesting starch and TAG degradation may help in the recovery process. This data leaves many questions remaining. First, what are the individual roles of starch and TAG in nutrient recovery? Does starch carbon aid nitrogen fixation or is it burned for energy? Does TAG provide acyl chains for membrane synthesis, especially in the thylakoid where large amounts of membrane are required for photosynthesis? Finally how does the photosynthetic environment change over the course of nutrient recovery and how does that change with or without acetate? To approach the roles of starch and TAG during nutrient recovery, I propose a 14C labeling experiment where cells are labeled with 14C bicarbonate during nitrogen deprivation and monitored during recovery with no label. Total counts of 14C in biomass per culture volume during nutrient recovery would show whether a significant amount of labeled carbon is retained during recovery or removed. If 14C is retained in total biomass, it would demonstrate that starch and TAG play carbon storage roles, providing carbon for new biomass, and if there is a small amount retained it would suggest these compounds play energy storage roles as they would simply be burned after N recovery. Next counting the change in 14C label in TAG and other glycerolipid pools after TLC separation might show flows of 14C label from one pool to another. I believe that there will be a significant amount of label going from TAG to MGDG !!!!$""!and DGDG pools as TAG would help restore photosynthetic membranes. Expanding upon this, if cells were labeled with 14C during 12-40 hrs(starch labeling) and from 40 to 96 hrs (TAG labeling) (See Chapter 3) during N deprivation, one could get label dominantly into either starch or TAG pools and be able to follow specific labeling from each stored compound. Counting 14C in other biomass such as protein pools may point to the starch and TAG utility for the assimilation of nitrogen. With this data trends, one may be able to tease out of starch, TAG, or both pools are in line with photosynthetic measurement increases, contributing to its recovery. One earlier study used labeled arachidionic acid to suggest TAG is used to help MGDG production during nutrient flow however this was in a different alga strain and only one Fatty acid species (Khozin-Goldberg, et al. 2005). Understanding the roles of starch and TAG during nutrient recovery may shed light on ways to increase their production during nutrient deprivation for bioenergy purposes. !!!!$"$!FINAL Determining the final destination of starch and TAG carbon helps determine reasons for the accumulation of these compounds. Further, systems approaches identifying gene/ proteins responsible for storage compound degradation may provide direction for engineering efforts to have algae that accumulate large amounts of carbon during normal growth. Although an inverse relationship between cell growth and carbon storage has been continually demonstrated there may still be a way to unhinge these relations. This may lie in deregulating photosynthetic controls decreasing potential carbon and energy fixation, or in regulation controlling N+ TAG and starch pool sizes. In this dissertation I have shown TAG accumulations are not caused by photosynthetic stress/ overflow during nutrient deprivation under mixotrophy at normal physiological conditions. Additionally, I have shown that starch increases before TAG during deprivation and is used before TAG during nutrient recovery(see Appendix). This points to a regulatory control switch between starch and TAG synthesis suggesting one look at the regulatory environment of the cell when it switches from starch to TAG synthesis. Tuning those signals may allow for increased TAG and decreased starch synthesis. Next as autotrophic cells do not decrease their photosynthetic components as much as mixotrophic cells during N dep, it may be possible to compare the two conditions and pull out what regulatory controls are at play to help engineer an organism that has heightened energy and carbon capture during N- to aid carbon accumulation. !!!!$"%! In homage to all the general biology courses I have ever taken: --THE MITOCHONDRIA IS THE POWERHOUSE OF THE CELL-- !!!!$"&! APPENDIX !!!!!$"'! To address the functional roles of starch and TAG accumulations, these compounds need to be studied during nutrient recovery. Very little literature exists on nutrient recovery in algae, especially on carbon metabolism and the ultimate destination of the starch and TAG carbon and energy. Here, I have gathered preliminary data on the cell physiology, photosynthetic parameters, and kinetics of nutrient recovery to aid future research endeavors in this area. ! After 96 hrs of N deprivation in TAP media, nitrogen is added back to the media and the Chlamydomonas cells were followed. Figure 5.1 (Panel A) shows cells do not begin to divide until after 24 hrs. During this 24 hr lag period, cells synthesized new chlorophyll to reach near N+ level (Panel B). Further, cells decreased their Ash Free Dry Weights (AFDW) suggesting starch and TAG and being degraded within this period. Total FAMES were also found to decrease in parallel. Both TAG and starch levels were back to N+ amounts by 36 hrs (Figure 5.2). This data demonstrates starch mobilization hrs before TAG suggesting starch is a first energy or carbon source for the cell which would enable the rapid uptake of nitrogen into amino acids. As TAG levels fall before total FAME, this demonstrates fatty acid distributions in the cell is changing and may represent movement of TAG fatty acids into membrane lipids. !!!!$"(! Figure 5.1. Changes in cell counts, biomass and Chl after nitrogen readdition. (A) Cell counts, (B) Chlorophyll abundance, and (C) Ash Free Dry Weights(AFDW) were measured at each timepoint after N readdition. Error bars indicate standard deviation, N=3. !0 2 4 6 8 10 12 14 16 0 20 40 60 80 million cells/ml Time(HRS) 0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 ug AFDW/ million cells Time (HRS) A C B 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 20 40 60 80 ugChl/milion cells Time(HRS) !!!!$"+! To learn about the relationship between starch and TAG and cell recovery, we measured the disappearance of these molecules (Figure 5.2). In Panel A of 5.2, starch rates were found to decrease almost immediately after the return of nitrogen. TAG levels were found to decrease starting after 6 hrs of recovery and Figure 5.2. TAG and Starch Degradation during N recovery. Starch levels (A) begin to fall before TAG (hollow square) and FAME (hollow triangle) Levels (B) during N recovery. Error bars indicate standard deviation, N=3. !!!!!$"#! !-0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0 10 20 30 40 nmol O2/min/ million cells Time (hrs) 0 0.0005 0.001 0.0015 0.002 0.0025 0.003 0.0035 0 10 20 30 40 !h+/s/million cells Time (Hrs) 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 20 40 60 80 Fv/FM, !ii, NPQ Time (Hrs) 0 0.2 0.4 0.6 0.8 1 1.2 0 20 40 60 80 Qp Time (Hrs) A C B D Figure 5.3. Photosynthetic measures during N recovery. Chlorophyll fluorescence measurements (A) show FV/FM(hollow square), $ii (hollow square), and NPQ (hollow diamond) values reach N replete levels within 24 Hrs. Qp(B) measures demonstrate PSII quinone pools are open to N+ levels within 12 hrs of recovery. Thylakoid proton efflux rates (C) increase to N+ levels between 24- 36hrs. Net oxygen evolution(D) (hollow squares) and consumption ( hollow triangles) reach N+ levels by 36 hours. Error bars indicate standard deviations, N=3. !!!!!$",! The Recovery of Photosynthesis was studied through chlorophyll fluorescence, absorbance spectroscopy, and gas measurements (Figure 5.3). PSII photosynthetic efficiency was found to return to N+ levels in the first 24 hrs of recovery and photosynthetic stress (NPQ) was found to return within the same period. The PSII quinone pool returned to N+ levels within the first 12 hrs demonstrating quick repair of PSII components. Proton efflux rates, a measure of ATP synthase, showed continued increase in rates through 36hrs suggesting cells continue to restore photosynthetic components through this time point. Oxygen evolution values show a similar trend as the total proton efflux supporting continued restoration. These collective data suggest Figure 5.4. Acetate uptake during nutrient recovery. Acetate levels do not appear to decrease rapidly until after 48hrs(A). Acetate uptake levels do no increase until after 48 hrs of N recovery(B). Error bars equal ranges of measurements, N=2. !!!!!$$)!cells rapidly restore the intrinsic efficiency of their photosynthetic components but continue to increase their capacity for photosynthesis even after 36 hrs of recovery. It is important to note that starch and TAG decrease when photosynthetic capacity is most rapidly restored suggesting these molecules play a part in the restoration. As a measure of metabolic demand during recovery, media acetate levels were measured. Starting off slow, acetate began to be uptake at a faster rate from 12 to 48 hs with dramatic increases in uptake from 48 to 72 hrs. Acetate uptake per cell values are presented in panel B of figure 5.4. This suggests that cells start with a slowed metabolism during n deprivation which increases to N+ levels by 48 hrs. This coincides with increases in photosynthetic activity and cell growth. !!!!$$"! REFERENCES !!!!$$$!REFERENCES Akita, T. and Kamo, M. (2015) Theoretical lessons for increasing algal biofuel: Evolution of oil accumulation to avert carbon starvation in microalgae. Journal of theoretical biology, 380, 183-191. Ball, S.G., Dirick, L., Decq, A., Martiat, J.C. and Matagne, R.F. (1990) Physiology of Starch Storage in the Monocellular Alga Chlamydomonas-Reinhardtii. Plant Sci, 66, 1-9. Grossman, A.R., Gonzalez-Ballester, D., Shibagaki, N., Pootakham, W., Moseley, J. and Pootakham, W. (2010) Responses to Macronutrient Deprivation. Hasunuma, T., Kikuyama, F., Matsuda, M., Aikawa, S., Izumi, Y. and Kondo, A. (2013) Dynamic metabolic profiling of cyanobacterial glycogen biosynthesis under conditions of nitrate depletion. J Exp Bot, 64, 2943-2954. Hu, Q., Sommerfeld, M., Jarvis, E., Ghirardi, M., Posewitz, M., Seibert, M. and Darzins, A. (2008) Microalgal triacylglycerols as feedstocks for biofuel production: perspectives and advances. Plant J, 54, 621-639. Khozin-Goldberg, I., Shrestha, P. and Cohen, Z. (2005) Mobilization of arachidonyl moieties from triacylglycerols into chloroplastic lipids following recovery from nitrogen starvation of the microalga Parietochloris incisa. Bba-Mol Cell Biol L, 1738, 63-71. Klok, A.J., Lamers, P.P., Martens, D.E., Draaisma, R.B. and Wijffels, R.H. (2014) Edible oils from microalgae: insights in TAG accumulation. Trends in biotechnology, 32, 521-528. Kohlwein, S.D. (2010) Triacylglycerol homeostasis: insights from yeast. J Biol Chem, 285, 15663-15667. Li, X., Moellering, E.R., Liu, B., Johnny, C., Fedewa, M., Sears, B.B., Kuo, M.H. and Benning, C. (2012) A galactoglycerolipid lipase is required for triacylglycerol accumulation and survival following nitrogen deprivation in Chlamydomonas reinhardtii. The Plant cell, 24, 4670-4686. Merchant, S.S., Kropat, J., Liu, B., Shaw, J. and Warakanont, J. (2012) TAG, You're it! Chlamydomonas as a reference organism for understanding algal triacylglycerol accumulation. Current Opinion in Biotechnology, 23, 352-363. Murphy, D.J. (2001) The biogenesis and functions of lipid bodies in animals, plants and microorganisms. Prog Lipid Res, 40, 325-438. !!!!$$%!Park, J., Wang, H., Gargouri, M., Deshpande, R., Skepper, J., Holguin, F.O., Juergens, M., Shachar-Hill, Y., Hicks, L. and Gang, D.R. (2014) The response of Chlamydomonas reinhardtii to nitrogen deprivation: A systems biology analysis. Submitted. Roessler, P.G. (1990) Environmental-Control of Glycerolipid Metabolism in Microalgae - Commercial Implications and Future-Research Directions. Journal of Phycology, 26, 393-399. Siaut, M., Cuine, S., Cagnon, C., Fessler, B., Nguyen, M., Carrier, P., Beyly, A., Beisson, F., Triantaphylides, C., Li-Beisson, Y.H. and Peltier, G. (2011) Oil accumulation in the model green alga Chlamydomonas reinhardtii: characterization, variability between common laboratory strains and relationship with starch reserves. Bmc Biotechnol, 11.