Ema”: , 1.... h. . .m a? 4 we. 1,. 4. .. i... EBLT, (-er ul’a , W. «Wu MW . .2 in. 7. .Nfl. 1... .1... kw. awn . ..V k or 3 , ‘ can... ‘ «risk. 3...... u:- J—X" v , 1W . v a 1". aha-MN I! t." .. ... I!!! 3I$xvg ‘3 ‘2‘] a? find“... "HF 1,3510} ~ .i k. :s. .u I I LIBRARY Michip» ' State Uni ~+u VIVIUJ ~l..-' , " m’m':' This is to certify that the dissertation entitled ACQUISITION AND TRANSFER OF SULFUR BY THE MODEL ARBUSCULAR MYCORRHIZAL FUNGUS GLOMUS INTRARADICES presented by James William Baker Allen has been accepted towards fulfillment of the requirements for the Doctoral degree in Plant Biplgqy Y LAM Major Prof FEW 5 Signature £260 (1 Date MSU is an Affirmative Action/Equal Opportunity Employer —.—.-.-.-»-.---.-----.-.-.—.--—-.-—-— PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5/08 K‘/Prolecc&Pres/ClRC/DaleDuevindd ACQUISITION AND TRANSFER OF SULFUR BY THE MODEL ARBUSCULAR MYCORRHIZAL FUNGUS GLOM US IN T RARADICES By James William Baker Allen A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for a degree of DOCTOR OF PHILOSOPHY Plant Biology 2009 ABSTRACT ACQUISITION AND TRANSFER OF SULFUR BY THE MODEL ARBUSCULAR MYCORRHIZAL FUNGUS GLOM US IN T RARADICES By James William Baker Allen The amount of sulfur (S) in the environment is radically changing. Anthropogenic pollution has increased the annual worldwide terrestrial deposition of sulfur by more than 300 fold in the last 150 years. In recent decades, however, the legislated removal of S at the source has diminished S deposition by more than 6 fold in many areas of the United States, leading to increased reports of S deficiencies in crops. Technology in this area continues to simplify the process of source removal with the goal of achieving pre- industrial levels. Additionally, sulfate, the main form of S assimilated by plants, is highly mobile in soils and can leach into the water system at a rate equal to its addition to the soil surface depending on water addition. In response to these environmental changes, in recent years interest in plant S acquisition and assimilation has surged. The arbuscular mycorrhizal (AM) fungal symbiosis has co-evolved with plants since their transition to terrestrial environments, and AM fungi are now known to associate with an overwhelming majority of land plants. The primary role of the symbiosis is to enhance the acquisition of nutrients in support of plant growth and homeostasis. Although the transfer of P, and, to a lesser extent, N, have been thoroughly analyzed, S acquisition by endo-mycorrhizal symbioses is a neglected area of research with oficn contradictory results. The availability of a monoxenic culture system and development of a model organism for mycorrhizal research, Glomus intraradices, including a worldwide effort to sequence and annotate the genome, has opened the possibility to more definitively analyze S in relation to an AM fungal symbiosis. The goal of this thesis was to quantify and analyze the transfer of S in relation to plant demand, as well as pioneer research in the study of the regulation of S assimilation in AMF. This work complements the ultimate goal of mycorrhizal research: the use of AM fungi as a viable replacement for chemical fertilization in agriculture. The model AMF Glomus intraradices transferred physiologically significant amounts of oxidized and reduced S forms to host roots, suggesting that S acquisition is a likely role of this fungus in natural ecosystems. The assimilation of sulfate by the symbiotic fungus was less regulated by reduced S sources than was the transfer to the host roots. Additionally, the simultaneous assimilation of sulfate and met or cys by pre- syrnbiotic mycelium is novel among studied fungi, whose sulfate assimilatory pathways are often completely diminished in response to reduced S sources. The regulation of sulfate acquisition, reduction, and assimilation was found to be greatly effected by ammonia assimilation and intracellular GSH concentration, and, unlike other fungi, less effected by reduced S forms cys and met. During germination, the addition of met led to very similar gene expression patterns compared to the addition of GSH, which did not correlate with uptake data. In general, changes in gene expression were not reflective of changes in S uptake with the exception of the down-regulation of a putative high affinity sulfate permease by cys addition in both the symbiotic and pre-symbiotic tissue. Unlike other fungi studied, transcriptional regulation does not appear to be the primary point of control for S acquisition by germinating spores of G. intraradices. Copyfightby JAMES WILLIAM BAKER ALLEN 2009 ACKNOWLEDGEMENTS I would like to thank the members of my committee, John Ohlrogge, Frances Trail, and Raymond Hammerschmidt, for their role in making graduate school a pleasant learning experience. I would especially like to acknowledge the patience and support I received from my graduate advisor, Yair Shachar-Hill, who supported me after the failure of my first project and I would like to thank him for believing in my abilities even when I didn’t. He showed a great deal of compassion on a daily basis, not only to me, but to everyone who worked in his lab. He truly shaped my understanding and approach to science, and I couldn’t have hoped for a better advisor. I would also like to thank past and current members of Dr. Shachar-Hill’s lab who had an impact on my research including Heike Biicking, Doug Allen, Igor Libourel, and Eva Collakova for teaching, supporting, teasing, listening, and aggravating me. I’m a better scientist for having known them. I’ve also had the pleasure of working alongside my wife, Inga Krassovskaya, for several years. She has helped me succeed in many ways besides occasionally lending me her expertise. Without her patient ear, her love and support I certainly would not have succeeded in this endeavor. Finally I would like to thank my parents and my sisters for not bugging me about graduating and taking this whole thing seriously. My parents deserve special thanks for letting me move back in for a couple of years at the start of graduate school, and for giving us a place to vacation for free. TABLE OF CONTENTS ListofFigures.............,_.... ._ Chapter 1 ..,ix Introduction an Literature Review--- .. .. ,. .. . . ,. ,. .. Introduction.-- Life cycle Morphology and phylogeny Plant benefits from AM symbioses--. . .. .. Drought tolerance and disease reS1stance Phosphorus acquisition-.- Nitrogen and sulfur acquisition. - . . . .. Sulfate assimilation in plants. .. .. .. .. ._ . Sulfate uptake and reduction..-...-.-....... _. ._ _ . ., .. . .. Synthesis ofcysteine._,.....-.....-.-...._. ., ., ,, . _. . Transcriptional regulation--. _ .. .. _. .. .. .. . . _ Sulfur in agriculture.-. . . _ The role of microorgarusms in plant S nutntron Fungal sulfate a331m11atlon- ,. .. .. ., _. . . Saccharomyces cerevisiae. ,. .. .. .. . Aspergillus nidulans......-. .... .. . ,. . .. . . _ .....- Neurospora crassa..-....,..__...,..,..,.... .. ., .. .. - . .. Conclusions and rational.........- .. .. . . , . .. . . , References _. .. ._ .. __ .. ., ._ ._ .. .. _ Chapter 2 Sulfur Transfer through an Arbuscular Mycorrhiza Abstract... . Introduction ._ Experimental procedures Chemicals and reagents .. . ....-.55 Growth media . .. .. .. .. Growth of roots and mycorrhrzas Collection of roots and ERM... Extraction procedure............................_.._.. ., .... .. . . . ., .. .. .. . .56 _.. 57 ....58 35 . . Measurements of S in medta...-...-.... .. .. .. , ., ,. ., Non-mycorrhizal root measurements Addition of S metabolic regulators RNA extraction and putative gene fragment lSOlElthIl- .. .. . ., ,. .. ,. . 59 , 62 ,. ...62 Quantitative real time PCR measurements. Result. .................................................................................................................................................... .i “ “ Growth and sulfate uptake by 11011 mYcorrhizal roots . .. , . . vi O\OOO\JO\O\J>UJ-N'—' WWNNNr—‘r—Ip—Ar—dI—I NOOOQOOOOHONUI . .....48 53 53 55 59 Fungal uptake and transfer of sulfate is influenced by root access to sulfate ............................................................................................................................................................... . ........................... 65 Fungal transfer of S to the mycorrhizal roots versus direct uptake by roots ................................................................................................................................................................................................ 70 The effect of sulfur metabolites on sulfate uptake and transfer through the mycorrhizal symbiosis ............................................................................................................................... . ............. 78 Fungal uptake and transfer to host roots ofreduced S .. .. .. . .. ., .. ...81 Gene expression measurements-.........-..-.-..-...... ._ ,. .. ,. . 81 Drscussron ._ . . .. .. , .84 References-.--.-. _ .. .. .. _ . , 90 Chapter 3 Uptake and Assimilation of S during Presymbiotic Growth of an Arbuscular Mycorrhizal Fungus -- -m.--.-m-H.HH-m--.c-...----.-c-.--M---w H. ,,-..H- W.M96 Abstract . . ... ...... ..97 Introduction .. ., . .98 Expenmental procedure .. . .. .... .. ,, . ,. .. .. 102 Tissue culturing ., .. - .. .. ,, .. ...--102 Extraction and fracttonatron procedure .. .. .. .. .. .. . . . . ___...-_...103 Sample preparation for gene expression analysis. .. _. .. .. ,, .. ..---105 Real time quantitative PCR measurements .. ...-..-...107 Results ...................................................................................................................................................................... - . ......................... 109 Import and assimilation of sulfate by germinating spores and its regulation by S metabolites. - 109 Uptake and assimilation ofS from cysteme .. ., .. V 113 Simultaneous uptake of sulfate and reduced S spec1es . ... 113 The expression of putative S metabolism and transport genes . . , .. - .. 107 Discussion-.... --..-.....-._ ...-... .... .. .... .. .. , ,. ,. ...._...._.121 References.....---...-....., . .. .. . .. - .. 129 Chapter4 The Relationship between N and S Assimilation in Germinating Spores of Glomus intraradices- . .. . . . .. . - - ., . . , .. 136 Abstract .. _ . . .137 Introduction ,, . ,. .. .. ... .139 Experimental procedure . . . . . .. . . . , . _ . . .142 Amino acid quantlficatlon. .. ,. .. , . . . , .. -142 3:8- labeling experiments. - . _ .. .. .,. , . . , . . 143 15N-labeling experiments . . .. .. .. .. .-.--144 RNA extraction and gene expression measurements _ ,. .. . .. ., ...-......145 Results .......................................................................................................................................................... ................ . ....... 147 Free amino acid concentrations during germination .. .. _. .. .. ... 147 The effects of N sources on free amino acid concentrations . V . .,. . ...-154 The effects of N sources on sulfate uptake and assimilation..- ... .. ._. .. ...-.156 vii Interactions between sulfate and nitrogen assimilation pathways. ... .. ....160 Glutathione addition in relation to sulfate assimilation and free amino acid concentrations _ 161 Changes in gene expression of putative sulfate asmmrlatron genes in relation to nitrogen addition-......- .. . . . . . ...,....-169 Discussion..-.----... . .. .. .. _. . . - , . 171 References.........-. . .. . . .. 179 Chapter 5 Conclusions and Future Research-......--..--.. . , . .. . , _ -..-....183 Introduction. . ., .. ._ . . . ., . . . .. ., , . ......... 184 AMP and plant S nutrltron . ..... .. . _ - ,. , 185 Uptake and assimilation ofS by AMF . _ . - .. . . . 188 Future research _. . .. . . .. . . .. . . 191 References--....-.-... . .. , , ,_.. . . .. . 194 viii LIST OF FIGURES Fig. 1-1. Model ofsulfate assimilation in plant cells..- . - . . - . .. . . . 12 Fig. 1-2. Model ofsulfate assimilation in fungi. . . . , .. . .. . . . . . . .21 Fig. 2-1. The effect of sulfate concentration on the growth and sulfur incorporation of non-mycorrhizal transformed carrot roots .. _. . .. .. .. - .. . . .. ...--63 Fig. 2-2. Increased sulfate availability in the root compartments of split plates reduces transfer of S through the distal fungal ERM. 64 Fig. 2-3. Transfer of 358042- through the fungus versus direct root uptake. .. . . . . . ..67 Fig. 2-4. The effect of common sulfur metabolite repressors cysteine, methionine, and glutathione on the transfer and allocation of 358 in an arbuscular mycorrhiza .. .. . ...--72 Fig. 2-5. Transfer ofreduced S through a mycorrhizal symbiosis. . .. .. .. . ._ 80 Fig. 2-6. Fungal gene expression changes in response to the addition of 1 mM cysteine to ERM .................................................... ._ ........................................................................ . ..... _ ................... _ ................................ . ........... 83 Fig. 2-7. Model depicting S transfer through an endomycorrhizal symbiosis ...... . .,._ . 88 Fig. 3-1. Three hour uptake and incorporation of radio-labeled sulfate by germinating spores ........................................................ . ................................................................ . ................ . ............................................... . ..................... 111 Fig. 3-2. Radio-labeled sulfate uptake over time with the addition of sulfirr metabolites ........................................................................................................ 112 Fig 3-3. Combined uptake of sulfate and reduced S species cys and met . . . . . 115 Fig. 3-4. [3SS]Cysteine uptake and utilization in the presence of common sulfur metabohtes .. .. _.--....119 Fig. 3-5. Real-time PCR analysis of changes in sulfur assimilation pathway gene expression in response to sulfur metabolites- .. . .. ,_ . .. , .. . .. .. .. --.-120 Fig. 4-1. Time course of free amino acid concentration in germinating spores of Glomus intraradices--.-.- ., .. .... . . .148 Fig. 4-2. Free amino acid concentrations in the presence of exogenous N sources. .. . ...-150 Fig. 4-3. Influence of N sources on 355042- uptake, reduction, and allocation , . . 155 Fig. 4-4. Effect of SO42- (A) and S metabolites (B) on 15N-labeling of free amino acids .. ....... 157 Fig. 4-5. 35S-sulfate uptake, reduction, and allocation in the presence of ammonia and S metabol1tes .. -..._., .. .. .. ,. 163 Fig. 4- 6. Effects of ammonia and glutathione (GSH) on [3 5S]Cysteine uptake and utilization.- .. .. .. ._ ..- 164 Fig 4-7. Change in amino acid concentrations from glutathione addition. . . V 165 Fig. 4-8. Transcriptional changes of putative sulfur assimilation pathway genes afier + NH; addrtlon .. . 167 Fig 4-9. Working model of the regulation of sulfate assimilation in germinating spores of Glomus intraradices..-._---... ., . .. . _. . . , . ............ 177 Chapter 1 Introduction and Literature Review Sulfate Assimilation in Plants and Fungi Introduction Associations with symbiotic fungi, which facilitate the collection of nutrients easily accessible to aquatic but not soil borne roots, may have made possible the transition of plants from aquatic to terrestrial environments. Independent estimations place the time which plants first colonized land at approximately 500 or 800 million years ago (Wellman et al., 2003; Sanderson, 2003) while estimates of 600 to 1400 million years ago have been made for the divergence of the order Glomeromycota encompassing the arbuscular mycorrhizal fungi (AMF; Redecker et al., 2000; Heckman el al., 2001). Aside from vascular plants, the symbiosis also occurs between AMP and bryophytes, even cyanobacteria, suggesting that the symbiosis may have developed before plant roots (Read et al., 2000; James et al., 2006). In any case, the symbiosis is proven to have at least existed for the last 400 million years based on fossil evidence from the Rhynie chert in Scotland (Dotzler et al. 2009) and continues to be the dominant symbiosis on the planet (for a thorough review see Smith & Read, 2008). In the United States, the active removal of sulfur (S) from coal and oil has dramatically decreased the rate of anthropogenic S environmental emissions for the last three decades (Baumgardner et al., 2002; Lefohn et al., 1999). This change has sparked numerous studies on plant S nutrition, especially as crops begin to display signs of S deficiency. Although the AM symbiosis is now accepted as a positive, sometimes definitive, contributor to phosphorus (P) and nitrogen (N) nutrition in plants, their contribution to S nutrition is still not characterized, which is the focus of this thesis. Life Cycle Arbuscular mycorrhizal fungi (AMF) are obligate, asexual plant symbionts involved in plant nutrient acquisition which fimction and complete their life cycles using photosynthetic carbohydrate (for a review, see Smith & Read, 2008). Initially, spores germinate producing germ tubes which explore the soil for a symbiotic partner. If unsuccessfirl, growth of the germ tube is arrested by an unknown programmed mechanism, and the cytosol retreats into the spore in preparation for re-germination (Logi et al., 1998). Signals released by roots, specifically strigolactones, increase germ tube metabolism and grth (Akiyama et al., 2005). This is followed by the development of an appresorial structure and the initiation of symbiosis at the root surface (Besserer et al., 2008). Once the symbiosis is established, hyphae proliferate both intra- and extra- radically, forming two distinct metabolic environments (Shachar-Hill et al., 1995) specialized in either the acquisition and radical distribution of photosynthetically-derived carbon substrates, or in hyphal growth through the soil matrix, acquisition, and transfer of nutrients to the plant (Smith & Read, 2008). Intra-radical mycelium (IRM) and extra- radical mycelium (ERM) can be equal in mass; however the IRM only occupies approximately 2% of symbiotic roots (Olsson & Johansen, 2000). Mechanism(s) triggering sporulation are not characterized, but appear to be connected to root senescence. Sporulation involves a massive transfer of storage lipids, proteins, and a large number of nuclei, hundreds to several thousands (Hosny et al. , 1998), into developing spores. After approximately three months of growth, 90% of the ERM mass is in the form of these large multinucleate spores (Olsson & Johansen, 2000). The life cycle of AMF revolves around the conversion of plant photosynthate to triacylglycerols, which are the main nutritional form of carbon required for growth, sporulation, and germination. Glucose is imported from roots by IRM and converted to trehalose and glycogen, then to triacyl glycerides, the main storage form of C in AMF (ShacharHill et al., 1995; Pfeffer et al., 1999; Nagy et al., 1980). Lipid bodies transport C from the IRM to the ERM and are utilized for mycelial growth and homeostasis through interconversions by the glyoxylate and gluconeogenic pathways (Bago et al., 2002; Pfeffer et al., 1999; Lammers et al., 2001). ERM tissue is incapable of hexose uptake and is completely dependent on this transfer (Pfeffer et al., 1999). In G. intraradz'ces, the volume occupied by lipid bodies decreases from approximately 24% to 0.5% from host root to hyphal tip due to consumption along that length (Bago, 2000). Germinating spores and ERM can elongate and desaturate 16C fatty acids, but their synthesis is exclusive to IRM tissue (Trepanier et al., 2005; Pfeffer et al., 1999). The compartrnentalization of fatty acid synthesis is a possible reason for the obligate nature of the symbiosis. Spores of G. intarardices are 20% neutral lipid by mass, 80% of which are 16C triacyl-glycerides (Olsson & J ohansen, 2000), demonstrating the importance of this function for the completion of the fungal life cycle. Morphology and Phylogeny Gallaud (1905) was first to categorize the distinct structural classes of mycorrhizal fungi, which he termed the Arum, Paris, hepatic, and orchid types. Arum- and Paris-type endomycorrhiza predominate, and are characterized by the path of hyphal growth and morphology of absoption structures. Paris-type mycorrhizae develop entirely intracellularly with irregularly coiled hyphae which form non-terminal coiled arbusculate structures. The endomycorrhizal fungus Glomus intraradices is an example of an Arum- type mycorrhizal fungus, which proliferates through intercellular air spaces and form tree-like, terminal arbuscules. These arbuscules envaginate root cortical cells, forming a periplastic space between the root cell and arbuscular membranes that is generally accepted as the site of nutrient transfer between the fungus and plant (Bago, 2000; vanAarle et al., 2005). This is unlike Paris-type mycorrhizal interactions, where transfer of nutrients occurs at both the arbuscule and IRM (vanAarle et al., 2005). While still unclear, the availability of intercellular air spaces may influence the dominant arbuscular mycorrhizal type (Dickson et al., 2007). Surveys of plant/mycorrhizal interactions demonstrate that Arum-type fungi are mainly in symbiosis with fast-growing, high light plants, while Paris-type fungi are associated with slower growing low light or woody plants, possibly due to the presence of air spaces (Brundrett et al, 1990). A clear characterization of these types of AMF, however, is compromised by difficulties in discriminating between the structures, and therefore by possible misidentifications (Dickson, 2004). Phylogenetic analyses of 28S rRNA, the large subunit of RNA polymerase II, 18S rRNA, and B-tubulin gene sequences (DaSilva et al., 2006; Redecker & Raab, 2006; Shussler et al., 2001; Msiska & Morton, 2009) currently place arbuscular mycorrhizal fungi (AMF) into ten families in the phylum Glomeromycota, order Glomerales, Gigasporaceae, Glomaceae, Acaulosporaceae, Diversispora, Paraglomaceae, Geosiphoaceae, Ambisporaceae, Entrophosphosporaceae and Arcaesporaceae. Within the phylum, Glomus is the largest genus, with more than 70 species characterized by spores which bud from hyphal tips and germinate through the subtending hypha (Redecker & Raab, 2006). Lineages of Glomus are ubiquitous and often dominate AM fimgal communities (Opik et al., 2003), and several of them, chiefly Glomus intraradices, are commonly studied in laboratory settings (Smith & Read, 2008). This is the species whose genome is being sequenced (Martin et al., 2008) and is the subject of the work presented here. Plant benefits from AM symbioses The main benefit of endomycorrhizal associations with plant roots is considered to be the improved inorganic nutrient acquisition. This is a consequence of the fungal extra-radical hyphae increasing the range of the rhizosphere, the surface area available for absorption of nutrients, and the ability to access micro-environments in soil normally unavailable to roots (Allen & Allen, 1990). However, arbuscular mycorrhizal fungal (ANIF) symbioses also result in other ecologically significant benefits, including enhancement of drought tolerance and resistance to pathogenic fungal infection. Drought Tolerance and Pathogen Resistance As global climate change adjusts weather patterns, periods of drought are increasing in many areas of the world (Shein et al., 2005) and represent a major abiotic cause of growth, fecundity, and yield limitations (Boyer, 1996). Mycorrhizal colonization has been shown to alleviate water stress (For a review see Ruiz-Lozano, 2003), possibly because hyphae radiating into the soil bridge air gaps that normally impede bulk water flow (Allen, 2007). Colonization with Glomus intraradices or Gigaspora margarita increases stomatal conductance and drought tolerance in sorghum (Sorghum bicolor) and bean (Phaseolus vulgaris) plants independent of P nutrition (Cho et al., 2006; Auge et al., 2004a;2004b) Nutrient deficiency itself is an abiotic stress making plants more susceptible to pathogenic infections (Amtrnarm et al., 2008), and AMF colonization enhances nutrient uptake thus enhancing the plant defensive ability. However, the symbiosis is also known to reduce pathogenic fimgal frequency independent of nutritional effects (Yao et al. , 2002; Grandmaison et al., 1993). Colonization by AMF is linked to the accumulation of specific phenolic compounds in the cell walls of Allium cepa roots which induce hyphal branching and control AMF development in planta (Grandmaison et al., 1993). This effect, as well as the general resistance to fungal enzymatic cell wall degradation by phenolic compound deposition in root cell walls, provides a possible mechanism for the observed enhancement of the plant defense against invading fungi (Morandi et al., 1996; Yao et al., 2007). Phosphorus Aquisition Adsorption of phosphate is positively correlated with the concentration of A1, Mn, and Fe ions associated with clay particles (Lair et al., 2009), and is the least mobile plant macronutrient in the soil (Gahoonia & Nielsen, 1991; Hinsinger, 2001 ). The endomycorrhizal symbiosis is postulated to have developed in response to the challenge plants faced with P nutrition during their transition from aquatic to terrestrial life (Karandashov & Bucher, 2005) based on the immobility of phosphate in terrestrial environments as well as the clear benefits to plant P nutrition that the symbiosis confers (Hayrnan & Mosse,~1971; Bolan, 1991). The enhancement of P nutrition is therefore widely considered to be the major function of the endomycorrhizal symbiosis. Polyphosphates are synthesized and appear to be transported to the root depending on the availability of C for the fungus (Solaiman et al., 1999; Biicking & Shachar-Hill, 2005), and root phosphate transporters have been shown to be specifically induced in several plant species by the mycorrhizal symbiosis (J avot et al., 2007; Gomez et al., 2009, Grace et al., 2009; Nagy et al., 2009; Glassop et al., 2005). Recent evidence also exists which shows a connection between N and P nutrition, specifically that N fertilization may reduce colonization when P is non-limiting (Blanke et al., 2005). The relationship between P and other nutrient transfer has been little studied and may be an important aspect of the function of endomycorrhizal symbioses. Nitrogen and Sulfur Aquisition To date the majority of studies done on endomycorrhizal symbioses examine their role in plant P nutrition (Smith & Read, 2008). However, recent evidence suggests that the biologically significant transfer of other macro- and micro-nutrients by AMF also takes place under natural conditions. While inorganic N is the preferred source for plant uptake, the majority of N in soils is often in organic form and substantial quantities of N can be transferred by AMF. Leigh and coworkers (2009) demonstrated that Glomus intraradz'ces in association with Plantago lanceolata can transport enough N from labeled plant material to account for 20% of total plant N. The same fungus has been shown to increase the amount of N in mycorrhizal plants obtained from fertilizer by 38%, but only under low N conditions (Azcon et al., 2008). Compared to N, very little research has been done on S in relation to the endomycorrhizal symbiosis. Two studies showed little effect of AM fungi on the translocation of S (Monison, 1962; Cooper & Tinker, 1978), while two others found significant increases in S levels in the plant (Rhodes & Gerdeman, 1978; Banjeree et. al., 1999). Most dramatically, Rhodes & Gerdemann (1978) showed a 12 fold increase in the whole plant 35S content in onion plants due to mycorrhizal transfer. Further research into the role of AMF in plant S nutrition is needed to re-evaluate past research and accurately assess the importance of this transfer. Sulfate assimilation in plants Owing to its multiple oxidation states, S is linked in some way to most plant cellular processes, including protein synthesis, redox balancing, transcriptional/post- translational regulation, central metabolism through cofactors, and defense/ stress responses (Droux, 2004; Rausch & Wachter, 2005; Leustek & Saito, 1999). The cysteine component enables the efficient scavenging of oxidative metabolic byproducts by glutathione, which has been conserved since the initiation of aerobic life (N octor et al., 1998) and is necessary for the completion of the cell cycle in plants (May et al., 1998). In a larger sense, the synthesis of cysteine and methionine by plants supports animal S metabolism, making plant S aquisition and assimilation an indispensible aspect of planetary animal life. The enzymatic pathway of S uptake, reduction, and assimilation, as well as the allosteric/transcriptional regulation and the role of compartmentalization in the synthesis of the branch point of plant S metabolism, cysteine, are important aspects to the study of the role of mycorrhizal symbioses in plant S nutrition. Sulfate uptake and reduction Plant, fungal, and mammalian sulfate transporters share a common protein family and significant conservation on the amino acid level (Smith et al., 1995; Vidmar et al., 2000; Smith, 2003). S is imported by roots mostly as sulfate, moved around the plant with a variety of specialized transporters, and stored in vacuolar sap prior to reduction in the plastid (Ortiz-Lopez et al., 2000; Leustek & Saito, 1999). The conversion of sulfate to sulfide is an energy intensive process requiring the equivalent of ten ATP (Hell & Jost, 1997). The initial activation step (Figure 1-1a), catalyzed by sulfate adenosyltransferase (SAT), produces a high energy sulfur-phosphate acid anhydride bond through the hydrolysis of pyrophosphate to form adenosyl-S’-phosphosulfate (APS; Phartiyal et al. , 2006). Further phosphorylation, catalyzed by APS kinase (APSK), forms the molecule 3’-phosphoadenosyl-5’-phosphosulfate (PAPS). This molecule is utilized by plants solely for direct sulfation reactions catalyzed by sulfotransferases, most notably producing glucosinolates which give mustard its distinct taste (Vauclare et al., 2002). There is some debate over the mechanism by which activated sulfate is reduced to sulfite (Leustek & Saito, 1999). Based on sequence similarity with thioltransferases, it is likely that this reaction, catalyzed by APS sulfotransferase (APSS; Bick & Leustek, 1998), transfers sulfate to the cys moiety of glutathione (GSH) resulting in the spontaneous production of sulfite(F i gure 1-1a). Free sulfite is further reduced through the enzyme sulfite reductase (SIR) by oxidation of either light-generated ferredoxin (Takahashi et al, 1996) or 10 glutathione (Prior et al., 1999). Unlike other enzymes in the pathway which can have upwards of five isoforms, there is only one known gene encoding SIR in A. thaliana, and the localization of SIR to the plastid supports the exclusive reduction of sulfate in this compartment (Bork, 1998). 11 Figure 1-1 Model of sulfate assimilation in plant cells. Sulfate is imported into plant cell from the phloem and either sequestered in the vacuolar sap or transported into the plastid for reduction to sulfide. (A) In the plastid, sulfate is attached to adenosine through Sulfate Adenylyl Transferase (SAT) to form APS, which can be phosphorylated to PAPS by APS Kinase (APSK), or transfer the sulfate group to the cysteine moiety of glutathione through APS Sulfotransferase (APSS). PAPS production fuels direct sulfation reactions through a variety of SulfoTransferase enzymes. Sulfoglutathione reacts spontaneously with reduces glutathione to form oxidized glutathione and free sulfite, which is reduced by Sulfite Reductase (SIR) to sulfide. (B) O-acetylserine (OAS) is synthesized exclusively in the mitochondria by the cysteine synthase complex (CSC'). Sulfide and OAS are transported to the cytosol by a specific transporter (open circle), where free OAS Thiol-Lyase (OAS— T L) enzyme catalyzes their condensation to form acetate and cysteine. 12 Figure 1-1 ATP PM o- . f’ NH2 °=I=° SAT N/ N °' 0 o k > 11 \N N SULFATE ll °'-fi-0-l|’-O-CH2 o o- PLAS TID ADENYLYL- H H SULFATE (APS) OH OH O'-S- O- “I? O—CH2 r r :3 >/%)C O O' H H PHOSPH- _ _ _ ADENYLYL- 0 OH GS :SI 0 SULFATE (PAPS) O= P-o' SULFO- O I - GLUTATHIONE 0 cl)- GSH 82. SIR 0:? GSSH o- SULHDE SULFITE 6FD(RED) 6FD(OX) 13 Figure 1-1 (cont’d) 3/ i.’ \ CoA-O-C-CH3 ACETYL-CoA coA'SH o- 1:11-11,+ q (:30 O=C-(I3-l --C O- -C-CH3 H H2 0' NH2+ O-ACETYLSERINE o=c-c::- c- 0H (“3’ H H2 k MITOCHONDRIA SERINE CYTOSOL r 11 o=c—6.5, it is virtually non-existent due to the neutralization of pH dependent charges within the soil structure which otherwise bind anions, such as Al and Fe oxides (Scherer, 2001). The application of other anions, such as phosphate, lowers the adsorption and therefore increases the availability of sulfate to plants as well as to leaching effects. One effect of regular fertilization with phosphate is the decrease of sulfate in the soil by filling adsorption sites with phosphate anions, allowing sulfate ions to mobilize into unreachable areas for plant roots. Sulfate is highly soluble in water at normal pH ranges, resulting in efficient leaching from precipitation. The leaching of soil on an organic farm in Denmark was found to be 20 kg S ha'1 (Eriksen and Askegaard, 2000). This is likely the reason why approximately 98% of S found in soils is in an organically bound form (Scherer, 2001). Microorganisms in the soil convert available sulfate to organic esters or carbon- bonded S. In laboratory-grown firngal cultures, sulfate concentration is directly l8 proportional to the synthesis of a common organic form of S, choline sulfate (Saggar et al., 1981). Microbes re-mineralize organically bound sulfur when in need of carbon as an energy supply. The amount of sulfate present in soils is therefore governed by the relative rates of mineralization of organic S and the immobilization of free sulfate into organic compounds, which are both mainly the result of microbial activity. An average 6 to 39 kg S ha.1 of S is made available to plants in a variety of crop systems per year due to mineralization (Scherer, 2001). By contrast, Kiihn and Weller (1977) conducted lysimeter experiments on fields with or without S fertilization, and found that sulfate additions simply leached out of the soil. It is likely that fertilization with sulfate fails to increase the total amount of S in soil over time due to leaching, making mineralization of organic S by soil microbes the most constant source of available S for plant nutrition. Fungal sulfate assimilation Compared with plants, fungi have similarities and differences in their uptake, reduction, and assimilation of sulfate (Figure 1-2a), most notably in their use of 0- acetylhomoserine or O-acetylserine, or both, as final sulfide acceptors. In general, transcriptional regulation is coordinated among several assimilatory enzymes by a positive bZIP transcription factor. However, the regulation of this factor differs significantly between species, as does the allosteric regulation of the individual pathway enzymes. Most research in this area is done using one of three fungal model systems: Saccharomyces cerevisiae, Aspergillus nidulans, or Neurospora crassa. Aspects of S assimilation will be discussed in detail with respect to each of these fungi. 19 Saccharomyces cerevisiae The reduction of sulfate to sulfide in S. cerevisiae only differs in one aspect from reduction by plants, although the two systems differ significantly in sulfide assimilation. Sulfate is imported through single high or low affinity sulfate transporters (SUL1;SUL2) and activated by attachment to the terminal phosphate group of ATP by the enzyme S- adenosyl transferase (MET3) forming 5’-adenylylsulfate (APS). APS is further phosphorylated by APS kinase (MET14) to 3’-phospho-5’-adenylylsulfate (PAPS). PAPS is considered toxic to cells (Thomas & Surdin-Kerjan, 1997), which may explain why in plants APS is further reduced to sulfide and PAPS is primarily used for direct sulfation. In fungi, sulfate reduction takes place through PAPS (Figure 1-2a), and the toxicity is likely countermanded by the action of a kinase which both returns PAPS to APS and is necessary for the hydrolysis of PAP to AMP (Murguia et al. 1996). Also unlike plants, in S. cerevisiae sulfide is assimilated by reaction with O-acetylhomoserine (OAH) through the enzyme homocysteine synthase (MET25), not with O-acetylserine through cysteine synthase (Figure 1-2b). Cysteine is then synthesized from homocysteine through the intermediate molecule cystathionine by a pathway that is absent in plants (Ono et al. , 1999), the reverse-transsulfuration or cystathionine pathway (CT pathway). In the CT pathway, homocysteine reacts with serine to form cystathionine through the enzyme cystathionine B-synthase (STR4). Cystathionine is then catabolized to a-ketobutyric acid, ammonia, and cysteine by the enzyme cystathionine y-lyase (STRl). Homocysteine is also directly converted to methionine by homocysteine methyl transferase (MET6), making it the branch point of the pathway. 20 Figure 1-2 Model of sulfate assimilation in fungi. (A) Sulfate is imported by specialized high and low affinity sulfate perrneases (SUL). Reduction to sulfide requires activation through adenylation by sulfate adenyly transferase (SAT) and further phosphrylation of the adenosyl moiety by adenyly phosphosulfate kinase (APSK). The sulfate moiety is attached to a cys residue of the phosphoadenyly phosphosulfate reductase enzyme (PAPSR) and reduced by two electrons. Sulfite is further reduced by thioredoxin and six electrons to sulfide by sulfide reductase (SIR). (B) Sulfide can enter fungal metabolism at cysteine synthase (CS) or O-acetylhomoserine thiol-lyase (OAH-T L), the latter requiring an active reverse transsulfuration pathway for the conversion of homocysteine to cysteine through cystathionine beta synthase (CBS), forming the intermediate cystathionine from homocysteine and serine, and cystathionine gamma lyase (C GL), which splits cystathionine to cysteine and homoserine. Figure 1-2 A ATP PPi 9' ,2 NH2 0:?” SAT N/ N 0‘ o o t > SULFATE ll 11 N N o--i°l.-o—1:>-o-c1-12 O 0" ADENYLYL- H H SULFATE (APS) OH OH ATP APSK 2.. s ADP SULFIDE 1111-12 / N SIR 6TD(OX) |\ > 0 O \ 11 11 N N o--s-o-1=—o—CH2 6TD(RED) c") (I)_ o H H PHOSPH- ?' ADENYLYL- Cl) 0“ _ SULFATE (PAPS) _ - 0.3 #_ O=l|D O I "T o- o- SULFITE 22 mz_20.1._.<._.w>0 -32..- .._:_- OuoioIUIOImioioIouo N _ N_ N _ _ + :2 ... + ...z .0 szmm N: .... IOIOIOIOHO _ _ ..sz -o <00.._>._.m 04 h Ema nzolw101._.m0._.mo< ..NI: .0 ”.953 n_..eiwiolNoo -Nm 0 $60 .50 3.20.1255 Figure 1-2 (cont’d) -O N “I T...“ _.I_ _ Ou OIOIOIOI wIUIOIqu _ _ ..Nxz N: ..N_._z .o MIUIWIrso rum 0 _ _ +N:z -o 23 Sulfate reduction is a highly regulated process in S. cerevisiae involving feedback inhibition of enzymatic processes at multiple steps and strict control over the synthesis of new enzymes in the pathway (Marzluf, 1997). At the point of entry, the activity of both SULl and SUL2 is strongly inhibited by APS, and the synthesis of these two structural genes is completely suppressed by 0.8 mM methioine (met) and 0.2 mM 5’- adenosylmethionine (Adomet) (Breton and Surdin—Kerkan, 1977). SAT is a homohexamer in vivo (U llrich, 2001), and each monomer can be separated into four domains. Of these domains, the SAT catalytic domain is linked to a domain with structural homology to APS kinase which does not bind PAPS and has no kinase activity, and is possible evidence of an evolutionary relationship with a bi-functional enzyme (Lalor et al., 2003). The fungus Penicillium chrysogenum shares this APS kinase domain, which does bind to PAPS, decreasing the affinity of the catalytic domain for sulfate (MacRae et al., 2002). Allosteric inhibition of SAT by PAPS therefore may have been lost in this organism. However, in S. cerevisiae the catalytic activity of SAT is strongly inhibited by the product (APS) as well as by sulfide to a lesser extent (deVito & Dreyfuss, 1964). The synthesis of SAT is suppressed by met and, interestingly, de- repressed by exogenous cys. In contrast, synthesis of sulfite reductase is most strongly suppressed by cys and only partially suppressed by met (deVito & Dreyfuss, 1964). Sulfite reductase activity is also strongly regulated by NADP+, which completely represses the synthesis of sulfide (Yoshimato & Sato, 1967). Exposure to met has also been shown to increase the synthesis of the CT pathway enzymes, cystathionine [3- synthase and cystathionine y-lyase (Paszewski & Grabski, 1974). The toxicity of PAPS 24 and the high energetic cost of sulfate reduction are likely reasons for the multi-level control of this pathway in S. cerevisiae and other fungi. The transcription of genes involved in methionine biosynthesis is controlled by the intracellular concentration of Adomet sensed by a positive trans-acting basic leucine zipper transcriptional factor, Met4p (Thomas et al., 1992; Cherest et al., 1985), although methionyl-tRNA is also involved in regulation (Surdin-Kerjan et al. 1973). The regulatory function of Met4p depends on its association with a complex of proteins including a general transcription factor, centromere binding factor-1 (CBF-l), and a second S assimilation regulatory factor, Met28p, whose expression is itself controlled by Met4p (Kuras & Thomas, 1995). Under conditions of low met availability, the Met4p transcriptional activation complex associates with promoter regions of genes encoding S- adenosyl transferase (MET3), APS kinase (MET14), PAPS reductase (MET16), O-acetyl homoserine sulfydrolase (MET25), PAPS kinase (MET22), and the first enzyme for the synthesis of glutathione, glutamylcysteine synthetase (GSHl) (Kuras et al., 2002; Wheeler et al., 2002), activating transcription. The activity of the Met4p complex is controlled through the alteration or catabolism of Met4p by ubiquitylation (Kuras et al., 2002). Ubiquitylation of Met4p is conducted by ubiquitin ligase complex associated with an F -box protein, Met30p (Rouillon et al., 2000) whose activity is directly proportional to its concentration. Met30p transcription is positively regulated by Met4p when Adomet is abundant, so Met4p regulates its own breakdown (Smothers et al., 2000). However, ubiquitylated Met4p is only degraded when S. cerevisiae is grown in minimal medium amended with met. In rich media, ubiquitylated Met4p is not only stable, but selectively activates genes encoding S-adenosyl methioine synthetic enzymes to supply one carbon 25 metabolism (Kuras et al., 2002), thus serving as two distinct transcription factors. The complex between ubiquitin ligases and Met30p is stabilized by met, linking S metabolism to the regulation of S assimilation (Mountain et al., 1993). Interestingly, the transcription of MET4 is not regulated by S source (Mountain et al., 1993), but is regulated by another bZIP transcription factor, Gcn4p, which also regulates amino acid synthetic genes. Synthesis of met and cys is therefore coordinately regulated with general amino acid synthesis as well as the intracellular Adomet concentration. Aspergillus nidulans Sulfate reduction in A. nidulans is analogous to the S. cerevisiae pathway with few exceptions. Sulfate perrnease, while competitively inhibited by sulfate, accumulates intracellular sulfate up to a concentration of 0.04 M in a knock-out mutant of S-adenosyl transferase (sC), and a down-stream product is responsible for its allosteric inhibition (Bradfield et al., 1970). S-adenosyl transferase is competitively inhibited by its product, APS, but unlike S. cerevisiae, it is allosterically inhibited by PAPS (Borgeswalmsley et al., 1995; Renosto et al., 1990). Growth in a high concentration of met (5 mM) completely abolishes sulfate assimilation measured by 35S labeling (Paszewski et al., 1984), however transcription of APS kinase (sD) and cysteine synthase (cysB) does not appear to be affected (Clarke et al. 1997; Topezewski et al, 1997). In S. cerevisiae, exogenously supplied met leads to the accumulation of met and Adomet, while in A. nidulans it results in cystathionine accumulation (Pieniazck et al., 1973), which may explain the difference in transcriptional regulation by met between Saccharomyces and Aspergillus. The likely point of transcriptional regulation of sulfate 26 reduction is at the level of entry. Sulfate perrnease transcripts are reduced by at least four fold in high met (Grynberg et al., 2001). An early study of pathway regulation indicates that sulfate perrnease, S-adenosyl transferase, and sulfite reductase protein levels are significantly reduced in cys, homocysteine, and met amended media (Paszewski & Grabski, 1974), which is antithetical to current gene expression data and has not since been repeated. These data led to the commonly held belief that the pathway is regulated by cys (Marzluf, 1997). Few studies to date have analyzed transcriptional regulation of sulfate reduction enzymes by cys, making this hypothesis difficult to evaluate. Unlike S. cerevisiae, A. nidulans possesses fimctional OAS and CT pathways. The CT pathway is now known to be the primary source of cys only when the OAS pathway is blocked (Paszewski & Grabski 1974; Pieniazck et al., 1974), and the only biologically relevant pathway of cys synthesis under normal growth conditions in A. nidulans is through cysteine synthase (Brzywczy et al., 2007). Akin to S. cerevisiae, a positive acting bZIP regulatory protein was discovered for A. nidulans, MetRp, now known to be responsible for the transcriptional activation of homocysteine synthase (cysD), S-adenosyl transferase (sC), and cysteine synthase (cysB) (N atorff et al., 2003). A knock-out mutation of MetR reduces incorporation of S from exogenous sulfate by 87% (Natorff et al., 2003). Neither MetR nor Met4 are transcriptionally regulated by S sources themselves (N atorff et al., 2003; Mountain et al., 1993), but are regulated on the protein level apparently by similar mechanisms. Additionally, four negative regulatory proteins have also been discovered through mutagenesis studies, sulfur controller proteins (soon) A, B, C, and D (Natorff et al. 1993). The genetic sequence of sconB is homologous to met30 from S. cerevisiae and 27 contains an F-box motif for associating with ubiquitin ligase complexes as well (N atorff et al, 1998). Based on these similarities, it is likely, although not currently proven, that S. cerevisiae and A. nidulans contain similar ubiquitylation-based control systems for the regulation of sulfate assimilation. Neurospora crassa Although N. crassa was the model fungal system in which much of the research on global regulation of the sulfate assimilation pathway was first conducted, considerably less is known about regulation on the enzymatic level compared to S. cerevisiae or A. nidulans. Two sulfate perrneases have been characterized, specific to either mycelial growth (cys-14) or germination (cys-13), and are highly expressed under S-limiting conditions (Ketter et al., 1991). The synthesis of either isozyme is repressed completely by high external concentrations of met and feedback inhibited by sulfate up to a maximum of 60% (Marzluf, 1972). Interestingly, there is a time period in which there is no effect of met addition on the synthesis of cys-13 which lasts for approximately the first six hours of spore germination (Marzluf, 1972). This appears to be the only instance when gerrnination-specific regulation of sulfate reduction was measured in fimgi, and suggests that during this time transcriptional control of the pathway may differ from normal growth. Like A. nidulans, P. chrysogenum, and P. duponti, N. crassa sulfate adenyly] transferase enzyme is allosterically regulated by PAPS (Renosto et al., 1990). Also similar to A. nidulans, N. crassa possessed functional CT and OAS pathways (Flavin & Slaughter, 1964; Marzluf, 1997), and cystathionine y-lyase is strongly feedback regulated in by cysteine both fungi, as much as 92% in N. crassa (Paszewski & Grabski 28 1974; F lavin & Segal, 1964). The relative contribution of the two pathways has not been measured in N. crassa as it has in A. nidulans, but the overall similarity between the two fungal pathways suggests that the OAS pathway may be dominant in N. crassa as well. The sulfate assimilatory pathway in N. crassa is transcriptionally regulated by a positive and negative acting control system, Cys3/Scon2, analogous to the Met3/Met30 system in S. cerevisiae and the MetR/SconB system in A. nidulans (Fu & Marzluf, 1990; Fu et al., 1989; Paietta, 1990; Sizemore et al., 2002; for a review see Marzluf, 1997). The F-box motif, now recognized as an important component of cell cycle regulation, was first discovered during the analysis of the negative-acting sulfur controller gene scon2 by Kumar and Paietta (1998), who also demonstrated that scon2 remains functional when its F-box motif is replaced with the analogous sequence from Met30. Scon2 is transcriptionally activated by the positive-acting transcriptional regulatory protein Cys3, analogous to the activation of Met30 by Met4 in S. cerevisiae (Paietta, 1990; Smothers et al. 2000). Another negative acting regulatory protein which associates with scon2 in the formation of a ubiquitin ligase complex, scon3, is homologous to sconC from A. nidulans, and like sconC and MetR, is not transcriptionally regulated by Cys3 (Sizemore et al., 2002). CysB is necessary for the activation of genes encoding enzymes of the sulfate assimilatory pathway including cysl4 and cysl 3 (J arai & Marzluf, 1991; Marzluf & Metzenberg, 1968). 29 Conclusion and Rationale The rise and fall of sulfur emissions is among the greatest anthropogenic environmental changes in history. According to ice core analysis, the amount of sulfate deposited from sea water per year has risen by 75% since 1850 (Mayaweski et al, 1990), resulting in the addition of approximately 4.5 billion metric tons globally (Lefohn et al, 1999). Although these emissions continue to increase on a global scale, the Clean Air Act has led to a striking decrease in S pollution in the US over the last three decades. In rural areas of the eastern part of the US, sulfate deposition fell by 29% between 1990 and 2003, a figure which correlates closely with the 33% emissions drop from power plants in the area over the same time period (Baumgardner et al., 2002; Lefohn et al., 1999). Based on current legislation, emissions are projected to decrease a further 11% by the year 2025 with nothing likely to prevent their future decline to near pre-Industrial Revolution levels. The simplest solution to the impending necessity to replace the S now available to crop plants from atmospheric deposition is the fertilization of crops with sulfate. This is an inexpensive and readily available solution as the legislation requiring removal of S from coal and oil has led to a large excess of available sulfate. However, the soil mobility of sulfate can result in roughly equivalent rates of environmental input from soil leaching (Kiihn & Weller, 1977). Therefore the use of sulfate to amend S deficiencies in crops will necessarily lead to the reintroduction of a pollutant into the water system, at least partially negating one of very few environmental success stories of the twentieth century. This thesis provides a foray into an alternative means of supplementing plant sulfate by 30 examining the transfer of S through a mycorrhizal symbiosis already known to enhance P and N nutrition in plants. The role of AMF in plant S acquisition has been a largely neglected scientific problem. Previous studies produced inconclusive, often conflicting results (Morrison, 1962; Cooper & Tinker, 1978; Rhodes & Gerdeman, 1978a; 1978b; Banjeree et. al., 1999). Recently, greater attention has been given to the role and regulation of S metabolism in plants (Kopriva, 2006; Rausch & Wachter, 2005; Bloem et al., 2005) which, with environmental reasons already discussed, make investigations into S transfer through an AMF important. In response, three related studies were conducted of the quantification and regulation of S transfer using a monoxenic culture system (chapter 2), the regulation of S assimilation in the pre-symbiotic stage (chapter 3), and of connections between the regulation of N and S assimilation (chapter 4). The results of these studies demonstrate the ability of an AMF to import and assimilate both organic and inorganic S in the pre-symbiotic and symbiotic stages of development, as well as transfer physiologically relevant quantities of S to host roots. 31 References Akiyama K, Matsuzaki K, Hayashi H. 2005. Plant sesquiterpenes induce hyphal branching in arbuscular mycorrhizal fimgi. Nature 435(7043): 824-827. Allen MF, Allen EB. 1990. Carbon source of va mycorrhizal fungi associated with chenopodiaceae from a semiarid shrub-steppe. Ecology 71(5): 2019-2021. Allen MF. 2007. Mycorrhizal fungi: Highways for water and nutrients in arid soils. Vadose Zone Journal 6(2): 291-297. Amtmann A, Troufflard S, Armengaud P. 2008. The effect of potassium nutrition on pest and disease resistance in plants. Physiologia Plantarum 133(4): 682-691. Auge RM, Moore JL, Sylvia DM, Cho KH. 2004a. Mycorrhizal promotion of host stomatal conductance in relation to irradiance and temperature. Mycorrhiza 14(2): 85-92. Auge RM, Sylvia DM, Park S, Buttery BR, Saxton AM, Moore JL, Cho KH. 2004b. Partitioning mycorrhizal influence on water relations of Phaseolus vulgaris into soil and plant components. Canadian Journal of Botany-Revue Canadienne De Botanique 82(4): 503-5 14. Azcon R, Rodriguez R, Amora-Lazcano E, Ambrosano E. 2008. Uptake and metabolism of nitrate in mycorrhizal plants as affected by water availability and N concentration in soil. European Journal of Soil Science 59(2): 131-138. Bago B. 2000. Putative sites for nutrient uptake in arbuscular mycorrhizal fungi. Plant and Soil 226(2): 263-274. Bago B, Pfeffer PE, Zipfel W, Lammers P, Shachar-Hill Y. 2002. Tracking metabolism and imaging transport in arbuscular mycorrhizal fimgi. Metabolism and transport in AM fungi. Plant and Soil 244(1-2): 189-197. Banerjee MR, Chapman SJ, Killham K. 1999. Uptake of fertilizer sulfur by maize from soils of low sulfur status as affected by vesicular-arbuscular mycorrhizae. Canadian Journal of Soil Science 79(4): 557-559. Baumgardner RE, Lavery TF, Rogers CM, Isil SS. 2002. Estimates of the atmospheric deposition of sulfur and nitrogen species: Clean Air Status and Trends Network, 1990- 2000. Environmental Science & Technology 36(12): 2614-2629. Berkowitz O, Wirtz M, Wolf A, Kuhlmann J, Hell R. 2002. Use of biomolecular interaction analysis to elucidate the regulatory mechanism of the cysteine synthase complex from Arabidopsis thaliana. Journal of Biological Chemistry 277(34): 30629- 30634. 32 Besserer A, Becard G, J auneau A, Roux C, Sejalon-Delmas N. 2008. GR24, a synthetic analog of strigolactones, stimulates the mitosis and growth of the arbuscular mycorrhizal fimgus Gigaspora rosea by boosting its energy metabolism. Plant Physiology 148(1): 402-413. Bick JA, Leustek T. 1998. Plant sulfur metabolism - the reduction of sulfate to sulfite. Current Opinion in Plant Biology 1(3): 240-244. Blanke V, Renker C, Wagner M, Fullner K, Held M, Kuhn AJ, Buscot F. 2005. Nitrogen supply affects arbuscular mycorrhizal colonization of Artemisia vulgaris in a phosphate-polluted field site. New Phytologist 166(3): 981-992. Bloem E, Haneklaus S, Schnug E. 2005. Significance of sulfur compounds in the protection of plants against pests and diseases. Journal of Plant Nutrition 28(5): 763-784. Bolan NS. 1991. A critical-review on the role of mycorrhizal fungi in the uptake of phosphorus by plants. Plant and Soil 134(2): 189-207. Borgeswalmsley MI, Turner G, Bailey AM, Brown J, Lehmbeck J, Clausen 1G. 1995. Isolation and characterization of genes for sulfate activation and reduction in aspergillus-nidulans - implications for evolution of an allosteric control region by gene duplication. Molecular & General Genetics 247(4): 423-429. Bork C, Schwenn JD, Hell R. 1998. Isolation and characterization of a gene for assimilatory sulfite reductase from Arabidopsis thaliana. Gene 212(1): 147-153. Boyer J S. 1996. Advances in drought tolerance in plants. Advances in Agronomy, Vol 56, 187-218. Bradfield G, Somerfield P, Meyn T, Holby M, Babcock D, Bradley D, Segel IH. 1970. Regulation of sulfate transport in filamentous fungi. Plant Physiology 46(5): 720- 727. Breton A, Surdinkerjan Y. 1977. Sulfate uptake in saccharomyces-cerevisiae - biochemical and genetic study. Journal of Bacteriology 132(1): 224-232. Brundrett MC. 2002. Coevolution of roots and mycorrhizas of land plants. New Phytologist 154(2): 275-304. Brzywczy J, N atorff R, Sienko M, Paszewski A. 2007. Multiple fimgal enzymes possess cysteine synthase activity in vitro. Research in Microbiology 158(5): 428-436. Biicking H, Shachar-Hill Y. 2005. Phosphate uptake, transport and transfer by the arbuscular mycorrhizal fungus Glomus intraradices is stimulated by increased carbohydrate availability. New Phytologist 165(3): 899-912. 33 Ceccotti SP 1996. Plant nutrient sulphur - A review of nutrient balance, environmental impact and fertilizers. Fertilizer Research 43:117-125. Cherest H, Thao NN, Surdinkerjan Y. 1985. Transcriptional regulation of the met3- gene of saccharomyces-cerevisiae. Gene 34(2-3): 269-281. Cho KH, Toler H, Lee J, Ownley B, Stutz JC, Moore JL, Auge RM. 2006. Mycorrhizal symbiosis and response of sorghum plants to combined drought and salinity stresses. Journal of Plant Physiology 163(5): 517-528. Clarke DL, Newbert RW, Turner G. 1997. Cloning and characterisation of the adenosyl phosphosulphate kinase gene from Aspergillus nidulans. Current Genetics 32(6): 408-412. Cooper KM, Tinker PB. 1978. Translocation and transfer of nutrients in vesicular- arbuscular mycorrhizas .2. Uptake and translocation of phosphorus, zinc and sulfur. New Phytologist 81(1): 43-47. DaSilva DKA, Freitas ND, Cuenca G, Maia LC, Oehl F. 2008. Scutellospora pemambucana, a new fungal species in the Glomeromycetes with a diagnostic germination orb. Mycotaxon 106: 361-370. Devito PC, Dreyfuss J. 1964. Metabolic regulation of adenosine triphosphate sulfirrylase in yeast. Journal of Bacteriology 88(5): 1341-1345. Dickson S, Smith FA, Smith SE. 2007. Structural differences in arbuscular mycorrhizal symbioses: more than 100 years after Gallaud, where next? Mycorrhiza 17(5): 375-393. Dickson S. 2004. The Arum-Paris continuum of mycorrhizal symbioses. New Phytologist 163(1): 187-200. Dotzler N, Walker C, Krings M, Hass H, Kerp H, Taylor TN, Agerer R. 2009. Acaulosporoid glomeromycotan spores with a germination shield from the 400-million- year-old Rhynie chert. Mycological Progress 8(1): 9-18. Droux M. 2004. Sulfur assimilation and the role of sulfur in plant metabolism: a survey. Photosynthesis Research 79(3): 331-348. Eriksen J, Askegaard M. 2000. Sulphate leaching in an organic crop rotation on sandy soil in Denmark. Agriculture Ecosystems & Environment 78(2): 107-114. Feldman-Salit A, Wirtz M, Hell R, Wade RC. 2009. A Mechanistic Model of the Cysteine Synthase Complex. Journal of Molecular Biology 386(1): 37-59. Flavin M, Segal A. 1964. Purification and properties of cystathionine cleavage enzyme of 1. Journal of Biological Chemistry 239(7): 2220-2227. 34 Flavin M, Slaughter C. 1964. Cystathionine cleavage enzymes of neurospora. Journal of Biological Chemistry 239(7): 2212-2224. Fu YH, Marzluf GA. 1990. Nit-2, the major positive-acting nitrogen regulatory gene of neurospora-crassa, encodes a sequence-specific dna-binding protein. Proceedings of the National Academy of Sciences of the United States of America 87(14): 5331-5335. Fu YH, Paietta JV, Mannix DG, Marzluf GA. 1989. Cys-3, the positive-acting sulfur regulatory gene of neurospora-crassa, encodes a protein with a putative leucine zipper dna-binding element. Molecular and Cellular Biology 9(3): 1120-1127. Gahoonia TS, Nielsen NE. 1991. A method to study rhizosphere processes in thin soil layers of different proximity to roots. Plant and Soil 135(1): 143-146. Gallaud J. 1905. Etude sur les mycorrhizes endotrophes. Revue General Botany 17(5): 5-48. Glassop D, Smith SE, Smith FW. 2005. Cereal phosphate transporters associated with the mycorrhizal pathway of phosphate uptake into roots. Planta 222(4): 688-698. Gomez SK, J avot H, Deewatthanawong P, Torres-Jerez I, Tang YH, Blancaflor EB, Udvardi MK, Harrison MJ. 2009. Medicago tnmcatula and Glomus intraradices gene expression in cortical cells harboring arbuscules in the arbuscular mycorrhizal symbiosis. Bmc Plant Biology 9:10. Grace EJ, Cotsaftis O, Tester M, Smith FA, Smith SE. 2009. Arbuscular mycorrhizal inhibition of growth in barley cannot be attributed to extent of colonization, fungal phosphorus uptake or effects on expression of plant phosphate transporter genes. New Phytologist 181(4): 938-949. Grandmaison J, Olah GM, Vancalsteren MR, Furlan V. 1993. Characterization and localization of plant phenolics likely involved in the pathogen resistance expressed by endomycorrhizal roots. Mycorrhiza 3(4): 155-164. Grynberg M, Piotrowska M, Pizzinini E, Turner G, Paszewski A. 2001. The Aspergillus nidulans metE gene is regulated by a second system independent from sulphur metabolite repression. Biochimica Et Biophysica Acta-Gene Structure and Expression 1519(1-2): 78-84. Haas FH, Heeg C, Queiroz R, Bauer A, Wirtz M, Hell R. 2008. Mitochondrial serine acetyltransferase functions as a pacemaker of cysteine synthesis in plant cells. Plant Physiology 148(2): 1055-1067. Hayman DS, Mosse B. 1971. Plant grth responses to vesicular-arbuscular mycorrhiza .1. Growth of endogone-inoculated plants in phosphate-deficient soils. New Phytologist 70(1): 19-25. 35 Heckman DS, Geiser DM, Eidell BR, Stauffer RL, Kardos NL, Hedges SB. 2001. Molecular evidence for the early colonization of land by fimgi and plants. Science 293(5532): 1129-1133. Hell R, Jost R. 1997. Differential molecular and metabolic responses of Arabidopsis thaliana to short-tenn sulfate deficiency. Plant Physiology 114(3): 748-748. Hinsinger P. 2001. Bioavailability of soil inorganic P in the rhizosphere as affected by root-induced chemical changes: a review. Plant and Soil 237(2): 173-195. Hosny M, Gianinazzi-Pearson V, Dulieu H. 1998. Nuclear DNA content of 11 fungal species in Glomales. Genome 41(3): 422-428. Howarth JR, Dominguez-Solis JR, Gutierrez-Alcala G, Wray JL, Romero LC, Gotor C. 2003. The serine acetyltransferase gene family in Arabidopsis thaliana and the regulation of its expression by cadmium. Plant Molecular Biology 51(4): 589-598. James TY, Kauff F, Schoch CL, Matheny PB, Hofstetter V, Cox CJ, Celio G, Gueidan C, Fraker E, Miadlikowska J, et al. 2006. Reconstructing the early evolution of Fungi using a six-gene phylogeny. Nature 443(7113): 818-822. Jarai G, Marzluf GA. 1991. Sulfate transport in neurospora-crassa - regulation, turnover, and cellular-localization of the cys-l4 protein. Biochemistry 30(19): 4768-4773. Javot H, Penmetsa RV, Terzaghi N, Cook DR, Harrison MJ. 2007. A Medicago truncatula phosphate transporter indispensable for the arbuscular mycorrhizal symbiosis. Proceedings of the National Academy of Sciences of the United States of America 104(5): 1 720- 1 725. Karandashov V, Bucher M. 2005. Symbiotic phosphate transport in arbuscular mycorrhizas. Trends in Plant Science 10(1): 22-29. Ketter JS, Jarai G, Fu YH, Marzluf GA. 1991. Nucleotide-sequence, messenger-ma stability, and dna recognition elements of cys-14, the structural gene for sulfate perrnease-ii in neurospora-crassa. Biochemistry 30(7): 1780-1787. Kopriva S, Buchert T, Fritz G, Suter M, Benda RD, Schunemann V, Koprivova A, Schurmann P, Trautwein AX, Kroneck PMH, Brunold C. 2002. The presence of an iron-sulfur cluster in adenosine 5 '-phosphosulfate reductase separates organisms utilizing adenosine 5 '-phosphosulfate and phosphoadenosine 5 '-phosphosulfate for sulfate assimilation. Journal of Biological Chemistry 277(24): 21786-21791. Kopriva S. 2006. Regulation of sulfate assimilation in Arabidopsis and beyond. Annals of Botany 97(4): 479-495. 36 Krueger S, Niehl A, Martin MC, Steinhauser D, Donath A, Hildebrandt T, Romero LC, Hoefgen R, Gotor C, Hesse H. 2009. Analysis of cytosolic and plastidic serine acetyltransferase mutants and subcellular metabolite distributions suggests interplay of the cellular compartments for cysteine biosynthesis in Arabidopsis. Plant Cell and Environment 32(4): 349-367. Kilhn H, Weller H. 1977. 6 jahige untersuchung fiber schwefelzufuhr durch niederschlage und schwefelverluste durch auswaschung (lysimeter). Z. Pflnzenernaehr. Bodenk. 140: 431-440. Kumar A, Paietta JV. 1998. An additional role for the F-box motif: Gene regulation within the Neurospora crassa sulfur control network. Proceedings of the National Academy of Sciences of the United States of America 95(5): 2417-2422. Kuras L, Rouillon A, Lee T, Barbey R, Tyers M, Thomas D. 2002. Dual regulation of the Met4 transcription factor by ubiquitin-dependent degradation and inhibition of promoter recruitment. Molecular Cell 10(1): 69-80. Kuras L, Thomas D. 1995. Functional-analysis of met4, a yeast transcriptional activator responsive to s-adenosylmethionine. Molecular and Cellular Biology 15(1): 208-216. Lair GJ, Zehetner F, Khan ZH, Gerzabek MH. 2009. Phosphorus sorption-desorption in alluvial soils of a young weathering sequence at the Danube River. Geoderma 149(1- 2): 39-44. Lalor DJ, Schnyder T, Saridakis V, Pilloff DE, Dong A, Tang H, Leyh TS, Pai EF. 2003. Structural and functional analysis of a truncated form of Saccharomyces cerevisiae ATP sulfurylase: C-terrninal domain essential for oligomer formation but not for activity. Protein Engineering 16(12): 1071 -1 079. Lammers PJ, Jun J, Abuhaker J, Arreola R, Gopalan A, Bago B, Hernandez- Sebastia C, Allen JW, Douds DD, Pfeffer PE, Shachar—Hill Y. 2001. The glyoxylate cycle in an arbuscular mycorrhizal fungus. Carbon flux and gene expression. Plant Physiology 127(3): 1287-1298. Lee SM. 2005. Cysteine does not repress adenosine-5 '-phosphosulfate reductase through its conversion to either sulfate or glutathione. Journal of Plant Biology 48(1): 57-63. Lefohn AS, Husar JD, Husar RB. 1999. Estimating historical anthropogenic global sulfur emission patterns for the period 1850-1990. Atmospheric Environment 33(21): 3435-3444. Leigh J, Hodge A, Fitter AH. 2009. Arbuscular mycorrhizal fungi can transfer substantial amounts of nitrogen to their host plant from organic material. New Phytologist 181(1): 199-207. 37 Leustek T, Martin MN, Bick JA, Davies JP. 2000. Pathways and regulation of sulfur metabolism revealed through molecular and genetic studies. Annual Review of Plant Physiology and Plant Molecular Biology 51: 141 - l 65. Leustek T, Saito K. 1999. Sulfate transport and assimilation in plants. Plant Physiology 120(3): 637-643. Logi C, Shrana C, Giovannetti M. 1998. Cellular events involved in survival of individual arbuscular mycorrhizal symbionts growing in the absence of the host. Applied and Environmental Microbiology 64(9): 3473-3479. MacRae IJ, Segel 1H, Fisher AJ. 2002. Allosteric inhibition via R-state destabilization in ATP sulfurylase from Penicillium chrysogenum. Nature Structural Biology 9(12): 945- 949. Martin F, Gianinazzi-Pearson V, Hijri M, Lammers P, Requena N, Sanders IR, Shachar-Hill Y, Shapiro H, Tuskan GA, Young JPW. 2008. The long hard road to a completed Glomus intraradices genome. New Phytologist 180(4): 747-750. Maruyama-Nakashita A, Inoue E, Watanabe-Takahashi A, Yarnaya T, Takahashi H. 2003. Transcriptome profiling of sulfur-responsive genes in Arabidopsis reveals global effects of sulfirr nutrition on multiple metabolic pathways. Plant Physiology 132(2): 597-605. Marzluf GA, Metzenberg RI. 1968. Positive control by cys-3 locus in regulation of sulfur metabolism in neurospora. Journal of Molecular Biology 33(2): 423-427. Marzluf GA. 1972. Control of synthesis activity and turnover of enzymes of sulfur metabolism in neurospora-crassa. Archives of Biochemistry and Biophysics 150(2): 714- 721. Marzluf GA. 1997. Molecular genetics of sulfur assimilation in filamentous fungi and yeast. Annual Review of Microbiology 51: 73-96. May MJ, Vernoux T, Leaver C, Van Montagu M, Inze D. 1998. Glutathione homeostasis in plants: implications for environmental sensing and plant development. Journal of Experimental Botany 49(321): 649-667. Mayewski PA, Lyons WB, Spencer MJ, Twickler MS, Buck CF, Whitlow S. 1990. An ice-core record of atmospheric response to anthropogenic sulfate and nitrate. Nature 346(6284): 554-556. Morandi D 1996. Occurrence of phytoalexins and phenolic compounds in endomycorrhizal interactions, and their potential role in biological control. Plant and Soil 185(2): 241-251. 38 Morrison TM 1962. Uptake of sulphur by mycorrhizal plants. New Phytologist 62: 21- 27. Mountain HA, Bystrom AS, Korch C. 1993. The general arnino-acid control regulates met4, which encodes a methionine-pathway-specific transcriptional activator of saccharomyces-cerevisiae. Molecular Microbiology 7(2): 215-228. Msiska Z, Morton JB. 2009. Phylogenetic analysis of the Glomeromycota by partial beta-tubulin gene sequences. Mycorrhiza 19(4): 247-254. Murguia JR, Belles JM, Serrano R. 1996. The yeast HAL2 nucleotidase is an in vivo target of salt toxicity. Journal of Biological Chemistry 271(46): 29029-29033. Nagy R, Drissner D, Amrhein N, J akobsen I, Bucher M. 2009. Mycorrhizal phosphate uptake pathway in tomato is phosphorus-repressible and transcriptionally regulated. New Phytologist 181(4): 950-959. Nagy S, Nordby HE, Nemec S. 1980. Composition of lipids in roots of 6 citrus cultivars infected with the vesicular-arbuscular mycorrhizal fimgus, glomus-mosseae. New Phytologist 85(3): 377-384. Natorff R, Balinska M, Paszewski A. 1993. At least 4 regulatory genes control sulfirr metabolite repression in aspergillus-nidulans. Molecular & General Genetics 238(1-2): 185-192. Natorff R, Sienko M, Brzywczy J, Paszewski A. 2003. The Aspergillus nidulans metR gene encodes a bZIP protein which activates transcription of sulphur metabolism genes. Molecular Microbiology 49(4): 1081-1094. Nicholls P. 1975. Effect of sulfide on cytochrome-aa3 isosteric and allosteric shifts of reduced alpha-peak. Biochimica et Biophysica Acta 396(1): 24-35. Noctor G, Arisi ACM, Jouanin L, Kunert KJ, Rennenberg H, Foyer CH. 1998. Glutathione: biosynthesis, metabolism and relationship to stress tolerance explored in transformed plants. Journal of Experimental Botany 49(321): 623-647. Noji M, Inoue K, Kimura N, Gouda A, Saito K. 1998. Isofonn-dependent differences in feedback regulation and subcellular localization of serine acetyltransferase involved in cysteine biosynthesis from Arabidopsis thaliana. Journal of Biological Chemistry 273(49): 32739-32745. Olsson PA, Johansen A. 2000. Lipid and fatty acid composition of hyphae and spores of arbuscular mycorrhizal fungi at different grth stages. Mycological Research 104: 429- 434. 39 Ono B, Hazu T, Yoshida S, Kawato T, Shinoda S, Brszczy J, Paszewski A. 1999. Cysteine biosynthesis in Saccharomyces cerevisiae: a new outlook on pathway and regulation. Yeast 15(13): 1365-1375. Opik M, Moora M, Liira J, Koljalg U, Zobel M, Sen R. 2003. Divergent arbuscular mycorrhizal fungal communities colonize roots of Pulsatilla spp. in boreal Scots pine forest and grassland soils. New Phytologist 160(3): 581-593. Ortiz-Ceballos AI, Pena-Cabriales JJ, Fragoso C, Brown GG. 2007. Mycorrhizal colonization and nitrogen uptake by maize: combined effect of tropical earthworms and velvetbean mulch. Biology and Fertility of Soils 44(1): 181-186. Paietta JV. 1990. Molecular-cloning and analysis of the scon-2 negative regulatory gene of neurospora—crassa. Molecular and Cellular Biology 10(10): 5207-5214. Paszewsk.A, Grabski J. 1974. Regulation of s-amino acids biosynthesis in aspergillus- nidulans - role of cysteine and-or homocysteine as regulatory effectors. Molecular & General Genetics 132(4): 307-320. Paszewski A, Prazmo W, Nadolska J, Regulski M. 1984. Mutations affecting the sulfur assimilation pathway in aspergillus-nidulans - their effect on sulfirr amino-acid- metabolism. Journal of General Microbiology 130(MAY): 1 113-1 121. Pfeffer PE, Douds DD, Becard G, Shachar-Hill Y. 1999. Carbon uptake and the metabolism and transport of lipids in an arbuscular mycorrhiza. Plant Physiologz 120(2): 587-598. Phartiyal P, Kim WS, Cahoon RE, Jez JM, Krishnan HB. 2006. Soybean ATP sulfurylase, a homodirneric enzyme involved in sulfur assimilation, is abundantly expressed in roots and induced by cold treatment. Archives of Biochemistry and Biophysics 450(1): 20-29. Pieniazck NJ, Stepien PP, Paszewski A. 1973. Aspergillus-nidulans mutant lacking cystathionine beta-synthase - identity of l-serine sulfl1ydrylase with cystathionine beta- synthase and its distinctness from o-acetyl-l-serine sulfliydrylase. Biochimica Et Biophysica Acta 297(1): 37-47. Porcel R, Aroca R, Cano C, Bago A, Ruiz-Lozano JM. 2006. Identification of a gene from the arbuscular mycorrhizal fungus Glomus intraradices encoding for a 14-3-3 protein that is up-regulated by drought stress during the AM symbiosis. Microbial Ecology 52(3): 575-582. Porcel R, Aroca R, Cano C, Bago A, Ruiz-Lozano JM. 2007. A gene from the arbuscular mycorrhizal fungus Glomus intraradices encoding a binding protein is up- regulated by drought stress in some mycorrhizal plants. Environmental and Experimental Botany 60(2): 251-256. 40 Prior A, Uhrig J F, Heins L, Wiesmann A, Lillig CH, Stoltze C, 8011 J, Schwenn JD. 1999. Structural and kinetic properties of adenylyl sulfate reductase from Catharanthus roseus cell cultures. Biochimica et Biophysica Acta-Protein Structure and Molecular Enzymology 1430(1): 25-38. Querejeta J1, Allen MF, Alguacil MM, Roldan A. 2007. Plant isotopic composition provides insight into mechanisms underlying growth stimulation by AM fungi in a semiarid environment. Functional Plant Biology 34(8): 683-691. Quintero FJ, Garciadeblas B, RodriguezNavarro A. 1996. The SAL] gene of Arabidopsis, encoding an enzyme with 3'(2'),5'-bisphosphate nucleotidase and inositol polyphosphate l-phosphatase activities, increases salt tolerance in yeast. Plant Cell 8(3): 529-537. Rausch T, Wachter A. 2005. Sulfur metabolism: a versatile platform for launching defence operations. Trends in Plant Science 10(10): 503-509. Read DJ, Duckett JG, Francis R, Ligrone R, Russell A. 2000. Symbiotic fungal associations in 'lower' land plants. Philosophical Transactions of the Royal Society of London Series B-Biological Sciences 355(1398): 815-830. Redecker D, Kodner R, Graham LE. 2000. Glomalean fungi from the Ordovician. Science 289(5486): 1920-1921. Redecker D, Raab P. 2006. Phylogeny of the Glomeromycota (arbuscular mycorrhizal fungi): recent developments and new gene markers. Mycologia 98(6): 885-895. Renosto F, Martin RL, Wailes LM, Daley LA, Segel IH. 1990. Regulation of inorganic sulfate activation in filamentous fiingi - allosteric inhibition of atp sulfurylase by 3'- phosphoadenosine-S'-phosphosulfate. Journal of Biological Chemistry 265(18): 10300- 10308. Rhodes LH, Gerdemann JW. 1975. Phosphate uptake zones of mycorrhizal and nonmycorrhizal onions. New Phytologist 75(3): 555-&. Rhodes LH, Gerdemann JW. 1978a. Hyphal translocation and uptake of sulfur by vesicular-arbuscular mycorrhizae of onion. Soil Biology & Biochemistry 10(5): 355-360. Rhodes LH, Gerdemann JW. 1978b. Influence of phosphorus nutrition on sulfur uptake by vesicular-arbuscular mycorrhizae of onion. Soil Biology & Biochemistry 10(5): 361- 364. Riley NG, Zhao FJ, McGrath SP. 2000. Availability of different forms of sulphur fertilisers to wheat and oilseed rape. Plant and Soil 222(1-2): 139-147. 41 Rouillon A, Barbey R, Patton EE, Tyers M, Thomas D. 2000. Feedback-regulated degradation of the transcriptional activator Met4 is triggered by the SCFMet3O complex. Embo Journal 19(2): 282-294. Ruiz-Lozano JM. 2003. Arbuscular mycorrhizal symbiosis and alleviation of osmotic stress. New perspectives for molecular studies. Mycorrhiza 13(6): 309-317. Saggar S, Bettany JR, Stewart JWB. 1981. Measurement of microbial sulfur in soil. Soil Biology & Biochemistry 13(6): 493-498. Sanderson MJ. 2003. Molecular data from 27 proteins do not support a Precambrian origin of land plants. American Journal of Botany 90(6): 954-956. Scherer HW. 2001. Sulphur in crop production - invited paper. European Journal of Agronomy 14(2): 81-111. Schussler A, Schwarzott D, Walker C. 2001. A new fungal phylum, the Glomeromycota: phylogeny and evolution. Mycological Research 105: 1413-1421. Shacharhill Y, Pfeffer PE, Douds D, Osman SF, Doner LW, Ratcliffe RG. 1995. Partitioning of carbohydrate-metabolism in vesicular-arbuscular mycorrhizal leek revealed by NMR-spectroscopy. Plant Physiology 108(2): 22-22. Shein KA, Waple AM, Menne MJ, Christy JC, Levinson DH, Lawrimore J H, Wuertz DB, Xie P, J anowiak JE, Robinson DA, et al. 2006. State of the climate in 2005. Bulletin of the American Meteorological Society 87(6): S6-SlOZ. Sizemore ST, Paietta JV. 2002. Cloning and characterization of scon-3(+), a new member of the Neurospora crassa sulfur regulatory system. Eukaryotic Cell 1(6): 875- 883. Smith SE, Read DJ. 2008. Mycorrhizal symbiosis, 3rd edn. London, UK: Academic Smith SE, Smith FA, J akobsen I. 2003. Mycorrhizal fungi can dominate phosphate supply to plants irrespective of growth responses. Plant Physiology 133(1): 16-20. Smothers DB, Kozubowski L, Dixon C, Goebl MG, Mathias N. 2000. The abundance of met30p limits SCFMet30p complex activity and is regulated by methionine availability. Molecular and Cellular Biology 20(21): 7845-7852. Solaiman MZ, Ezawa T, Kojima T, Saito M. 1999. Polyphosphates in intraradical and extraradical hyphae of an arbuscular mycorrhizal fungus, Gigaspora margarita. Applied and Environmental Microbiology 65(12): 5604-5606. 42 Surdinkerjan Y, Cherest H, Derobichonszulmajster H. 1976. Regulation of methionine synthesis in saccharomyces cerevisiae operates through independent signals - methionyl-transfer-ma (met) and s-adenosylmethionine. Acta Microbiologica Academiae Scientiarum Hungaricae 23(2): 109-120. Takahashi S, Yoshida Y, Tamura G. 1996. Purification and characterization of ferredoxin-sulfite reductases from leek (Allium tuberosum) leaves. Journal of Plant Research 109(1093): 45-52. Thomas D, Jacquemin I, Surdinkerjan Y. 1992. Met4, a leucine zipper protein, and centromere-binding factor-i are both required for transcriptional activation of sulfur metabolism in saccharomyces-cerevisiae. Molecular and Cellular Biology 12(4): 1719- 1727. Thomas D, SurdinKerjan Y. 1997. Metabolism of sulfur amino acids in Saccharomyces cerevisiae. Microbiology and Molecular Biology Reviews 61(4): 503-515. Topezewski J, Sienko M, Paszewski A. 1997. Cloning and characterization of the Aspergillus nidulans cysB gene encoding cysteine synthase. Current Genetics 31(4): 348- 356. Trepanier M, Becard G, Moutoglis P, Willemot C, Gagne S, Avis TJ, Rioux JA. 2005. Dependence of arbuscular-mycorrhizal fungi on their plant host for palmitic acid synthesis. Applied and Environmental Microbiology 71(9): 5341-5347. Truong DH, Eghbal MA, Hindmarsh W, Roth SH, O'Brien PJ. 2006. Molecular mechanisms of hydrogen sulfide toxicity.Drug Metabolism Reviews 38(4): 733-744. Ullrich TC, Huber R. 2001. The complex structures of ATP sulfurylase with thiosulfate, ADP and chlorate reveal new insights in inhibitory effects and the catalytic cycle. Journal of Molecular Biology 313(5): 1 1 17-1 125. VanAarle 1M, Cavagnaro TR, Smith SE, Smith FA, Dickson S. 2005. Metabolic activity of Glomus intraradices in Arum- and Paris-type arbuscular mycorrhizal colonization. New Phytologist 166(2): 61 1-618. Vauclare P, Kopriva S, Fell D, Suter M, Sticher L, von Ballmoos P, Krahenbuhl U, den Camp RO, Brunold C. 2002. Flux control of sulphate assimilation in Arabidopsis thaliana: adenosine 5 '-phosphosulphate reductase is more susceptible than ATP sulphurylase to negative control by thiols. Plant Journal 31(6): 729-740. Vidmar JJ, Tagmount A, Cathala N, Touraine B, Davidian JCE. 2000. Cloning and characterization of a root specific hi gh-affinity sulfate transporter from Arabidopsis thaliana. Febs Letters 475(1): 65-69. 43 Wang YY, Vestberg M, Walker C, Hurme T, Zhang XP, Lindstrom K. 2008. Diversity and infectivity of arbuscular mycorrhizal fungi in agricultural soils of the Sichuan Province of mainland China. Mycorrhiza 18(2): 59-68. Watanabe M, Mochida K, Kato T, Tabata S, Yoshimoto N, Noji M, Saito K. 2008. Comparative Genomics and Reverse Genetics Analysis Reveal Indispensable Functions of the Serine Acetyltransferase Gene Family in Arabidopsis. Plant Cell 20(9): 2484- 2496. Wellman CH, Richardson JB. 1993. Terrestrial plant microfossils from silurian inliers of the midland valley of scotland. Palaeontology 36: 155-193. Wheeler GL, Quinn KA, Perrone G, Dawes IW, Grant CM. 2002. Glutathione regulates the expression of gamma-glutamylcysteine synthetase via the Met4 transcription factor. Molecular Microbiology 46(2): 545-556. Wirtz M, Berkowitz O, Droux M, Hell R. 2001. The cysteine synthase complex from plants - Mitochondrial serine acetyltransferase from Arabidopsis thaliana carries a bifunctional domain for catalysis and protein-protein interaction. European Journal of Biochemistry 268(3): 686-693. Wirtz M, Droux M, Hell R 2004. O-acetylserine (thiol) lyase: an enigmatic enzyme of plant cysteine biosynthesis revisited in Arabidopsis thaliana. Journal of Experimental Botany 55(404): 1785-1798. Wirtz M, Hell R. 2006. Functional analysis of the cysteine synthase protein complex from plants: Structural, biochemical and regulatory properties. Journal of Plant Physiology 163(3): 273-286. Yao MK, Tweddell RJ, Desilets H. 2002. Effect of two vesicular-arbuscular mycorrhizal fungi on the growth of micropropagated potato plantlets and on the extent of disease caused by Rhizoctonia solani. Mycorrhiza 12(5): 235-242. Yao Q, Zhu HH, Zeng RS. 2007. Role of phenolic compounds in plant defence: Induced by arbuscular mycorrhizal fungi. Allelopathy Journal 20(1): 1-13. Yoshimato A, Sato R. 1968. Studies on yeast sulfite reductase .i. Purification and characterization. Biochimica et Biophysica Acta 153(3): 555-563. Zhao FJ, Wood AP, McGrath SP. 1999. Effects of sulphur nutrition on growth and nitrogen fixation of pea (Pisum sativum L.). Plant and Soil 212(2): 209-219. 44 Chapter 2 Sulfur Transfer through an Arbuscular Mycorrhiza The contents of this chapter are modified from an article published in Plant Physiology. Allen JW, Shachar-Hill Y. 2009. Sulfur transfer through an arbuscular mycorrhiza. Plant Physiology 149(1): 549-560. 45 Abstract Despite the importance of sulfur for plant nutrition, the role of the arbuscular mycorrhizal symbiosis in sulfur uptake has received little attention. To address this issue, 35 . . . S labeling experiments were performed on mycorrhizas of transformed carrot roots and Glomus intraradices grown monoxenically on bi-compartmental Petri dishes. The uptake and transfer of 358042- by the fungus and resulting 358 partitioning into different metabolic pools in the host roots was analyzed when i) altering the sulfate concentration available to roots and ii) supplying the firngal compartment with cysteine, methionine or glutathione. Additonally, the uptake, transfer and partitioning of 35S from the reduced S sources [35S]cys and [35$]met was determined. Sulfate was taken up by the firngus and transferred to mycorrhizal roots, increasing root S contents by 25% in a moderate (not grth limiting) concentration of sulfate. High sulfate levels in the mycorrhizal root compartment halved the uptake of ”$042- from the fngal compartment. The addition of 1 mM methionine, cysteine, or glutathione to the fungal compartment reduced the transfer of sulfate by 26%, 45% and 80% respectively over one month. Similar quantities of 358 were transferred to mycorrhizal roots whether 358042-, [3SS]cys, or [3SS]met was supplied in the fungal compartment. Fungal transcripts for putative sulfur assimilatory genes were identified, indicating the presence of the trans-sulfuration pathway. The suppression of fungal sulfate transfer in the presence of cysteine coincided with a reduction in putative sulfate perrnease and not sulfate adenylyltransferase transcripts, 46 suggesting a role for fungal transcriptional regulation in S transfer to the host. A testable model is proposed describing root S acquisition through the AM symbiosis. 47 Introduction Plants and arbuscular mycorrhizal (AM) firngi have been co-evolving since the Devonian period (Simon et al., 1993; Brundrett, 2002). Today the AM symbiosis involves an estimated 80% of land plant species and 92% of plant families (Wang & Qiu, 2006). This symbiosis augments nutrient uptake by roots, which deplete the rhizosphere at a rate dependent on water availability, as well as the concentration and soil mobility of the nutrient (Bhat et al., 1976; Li et al., 1991; Gahoonia et al., 1994; Wang et al., 2004). Phosphorus is particularly immobile in soils (Gahoonia & Nielsen, 1991) and the AM symbiosis can increase plant growth in P deficient soils 19 fold (Hayman & Mosse, 1971). Although P acquisition is a major benefit of the symbiosis, a growing body of research points towards a more complex role for arbuscular mycorrhizas in nutrient uptake (Lambers et al., 2008; Liu et al., 2002). Nitrogen nutrition, for instance, is improved in a number of crop species (Pemer et al., 2008; Ortiz-Ceballos et al., 2007) and a monoxenic culture system consisting of transformed roots in symbiosis with the mycorrhizal fimgus Glomus intraradices has been used to demonstrate the capacity of this symbiosis to transfer N (Govindaraj alu et al., 2005; J in et al. 2005). Plants take up S primarily as the sulfate anion (Leustek, 1996), which is often found in low concentrations in the soil. The uptake and utilization of S by plants is reviewed by Leustek (2000) and Rennenberg et al. (2007). Sulfate is mobile and commonly lost through soil leaching (Eriksen & Askegaard, 2000). Because of this, typically 95% of the soil 8 is in organically bound forms such as sulfate esters (Scherer, 2001; Tabatabai, 1984) synthesized by soil microorganisms (Fitzgerald, 1976; Spencer & 48 Harada, 1960). Such forms of S are not thought to be significant sources for plants (Leustek, 1996). In the past, the effects of leaching have likely been masked by the high input of S fiom atmospheric pollution. However, the drastic reductions in S deposition in North America and Europe over the last decade have led to an equally dramatic rise in reported cases of S deficiency in crop species (Lefohn et al., 1999; Baumgardner et al., 2002; Riley et al., 2000). Under conditions of low S availability, an arbuscular mycorrhizal symbiosis can increase the percentage of S in pot grown onions (Guo et al., 2007) and maize (Banjeree et al., 1999). The ability of mycorrhizal fungi to transfer N and P from organic compounds (Joner & Jakobsen, 1994; Hawkins et al., 2000, Govindarajulu et al. 2005) also suggests the possibility that mycorrhizal plants might obtain S from organic sources. The small number of reports on the effects of AM colonization on the uptake of sulfur, which is present in amount equal to P in plants, have presented differing results (Gray & Gerdemann, 1973; Rhodes & Gerdemann, 1978a; Cooper & Tinker, 1978). The pioneering study by Gray & Gerdemann (1973) showing that mycorrhizal colonization in . . 35 , clover and maize increased S uptake compared to non-mycorrhizal plants was repeated by Rhodes & Gerdemann (1978a) in onion, but shown to be heavily influenced by P nutritional benefits. Rhodes & Gerdemann (1978b) were also the first to examine the transfer of 358042- through a mycorrhizal fungus exposed to an isolated source unavailable to the host plant. In a study published the same year, mycorrhizal clover grown in sterile bi-compartmental soil/agar plates failed to show a significant amount of transfer of isolated S through the fimgus (Cooper & Tinker, 1978). More recent research on ectomycorrhizas is more definitive on the subject (For a review see Rennenberg, 49 1999). In the association between Beech seedlings and the ectomycorrhizal fungi Laccaria laccata or Paxillus involutus, only L. laccata was shown to increase S uptake in S deficient conditions, and only when N was supplied (Kreuzwieser & Rennenberg, 1998). Interestingly, an analysis of free-living Laccaria bicolor showed that in contrast with the results from other filamentous firngi and yeast, sulfate uptake was unaffected by the presence of sulfur metabolites (Mansouri-Bauly et al., 2006). This may be an adaptation enabling the fungus to continue supplying sulfate to the host after its own needs are met (Mansouri-Bauly et al., 2006). Though little is known about the mechanism of S assimilation and its regulation in mycorrhizal fungi, these pathways have been extensively characterized in other fungal species (For a review see Marzluf, 1997). Cysteine (cys) is the major entry point of reduced S into metabolism, and is formed by the reaction of sulfide (made by the successive reduction of sulfate to sulfite and of sulfite to sulfide) with O-acetylserine catalyzed by cysteine synthase. However, through the reverse trans-sulfirration pathway, both Aspergillus and Neurospora species are able to efficiently convert methionine (met) to cys (Marzluf, 1993), making met an efficient sulfur source for these fungi. This pathway is crucial in Saccharomyces cerevisiae, where sulfide is assimilated not though cysteine synthase, but through the sulihydrolation of O-acetylhomoserine to form homocysteine from which met is made (Ono et al., 1999; Thomas & Surdin-Kerj an, 1997). By contrast, Schizosaccharomyces pombe has been shown to be deficient in reverse trans-sulfuration pathway enzymes making it dependent on cysteine synthase for S assimilation (Brzywczy etal., 2002). This results in a reduced regulatory role for met in S. pombe (Brzywczy et al., 2002). There is also evidence that the S assimilation enzymes 50 are transcriptionally regulated in S. cerevisiae specifically by levels of cys (Hansen & J ohannesen, 2000). Both positive and negative global regulatory proteins have been discovered in N. crassa, A. nidulans, and S. cerevisiae that share homology at the protein level (Marzluf, 1997). In N. crassa and S. cerevisiae, the expression of the S assimilatory genes is almost completely absent in the presence cys or met, and is greatly reduced by glutathione (gsh) addition (Ono et al., 1999; Ono et al., 1991; Marzluf, 1997). These same enzymes in Aspergillus nidulans, on the other hand, are not down regulated in response to sulfur metabolites (Natorff et al., 2003; Sienko & Paszewski, 1999; Clark et al., 1997). The transcription of sulfate perrneases is suppressed in all species studied by low concentrations of S metabolites (Grynberg et al., 2001; Pilsyk et al., 2007; Bradfield et al., 1970; Cherest et al., 1997; Ketter & Marzluf, 1988; Ketter et al., 1991; Marzluf, 1993; 1997). A putative cysteine synthase transcript has been identified from the AM fimgus Glomus intraradices (http://darwin.nmsu.edu/~plammers/glomus/), which appears to be the only report on the identity or regulation of AM fungal genes related to S metabolism, transport, or regulation. This study was aimed at determining the forms of S taken up, metabolized, and transferred to its host roots by the fungal partner in the AM symbiosis. A second goal was to identify effectors of S transfer to host roots, and to begin to identify AM fungal genes of S handling and their regulation. Using an AM symbiosis of monoxenic cultured roots the transfer of S was shown to be nutritionally significant and to be regulated by S availability to the host and by S metabolite availability to the extraradical mycelia. The uptake and transfer of reduced S through the symbiosis was also demonstrated. Putative homologs of sulfur uptake and metabolism genes were also identified and their 51 transcriptional regulation examined, indicating that sulfate uptake is transcriptionally regulated at the level of sulfate perrnease. These findings allow the development of a testable model of S uptake and transfer in the AM symbiosis. 52 Experimental Procedures Chemicals and reagents Gel-gro gellan (MP Biomedical, Solon, OH) was used for the solidification of growth media. Radioactively labeled sulfate was obtained as Na23SSO4 from MP Biomedicals (Solon, OH). Labeled cys and met were separated from a crude mixture of 35S-labeled compounds (TRAN35S-LABELTM; MP Biomedicals, Solon, OH) containing equal to or less than 70% [3SS]-met and 15% [3SS]-cys. The compounds were separated using an amino acid analyzer (Hitachi L8800). Retention times for the amino acids were determined using post-column ninhydrin bonding and UV absorption. The radioactive compounds were then collected in a ninhydrin-free environment by manual collection of the appropriate fractions. Growth media All experimental procedures used M medium (Fortin et al., 2002) with modifications to reduce the sulfate content. This was autoclaved and, for culture plates, solidified with 3.5 gL-1 Gel Gro gellan (MP Biomedicals, Solon, OH). The 3 mM MgSO4 normally present in M medium was replaced with 0.1 mM MgC12. An additional 2 mM of calcium was added as CaClz to replace the divalent cation concentration needed for the solidification of Gel-gro. Sulfate was supplied as NaZSO4 at different concentrations as indicated. Normal microelement concentrations for M media containing 53 0.02 mM sulfate were used for each experiment. These were included in all experiments with the exception of the measurement of non-mycorrhizal root growth (Figure 1a.). The medium in the root- but not fungal-compartments contained 10 gL-1 (29.2 mM) sucrose. NaZSO4 was added to the fungal compartments at the time of labeling as 0.2 mL of concentrated solution. Growth of roots and mycorrhizas Uncolonized Ri T-DNA transformed roots of Daucus carrota (DC2; Diop et al., 1992) were grown at 25°C with 0.02, 0.12, 0.5, or 3 mM NaZSO4 in modified liquid M media. Roots were inoculated at plate inception as previously described (Pfeffer et al., 1999). The roots and fungus were allowed to proliferate on both sides of bi- compartmented Petri plates at 25°C until the fungal extraradical mycelium was well developed (~six weeks). The colonized roots and media in each compartment were transferred to empty compartments of new plates in which the other compartment contained new medium. At the time of transfer, the media in which the mycorrhizal roots were growing was supplemented with one quarter of the original nutrient contents including sucrose and phosphorus, but excluding CaClz, added as ~0.5-1mL of a sterile solution. Root growth over the barrier after transplantation was prevented by pruning, and the fungal ERM typically grew over the barrier within one week of the transfer, colonizing the empty compartment. Fungal compartments were labeled when ERM had growth into at least half of the media, or within two weeks of crossing the barrier, whichever was first. 54 Collection of roots and ERM Root material was collected with forceps and rinsed for 5 min in deionized water . 35 . . to remove extemal- and reduce apoplastrc- S. The collected roots were rinsed again with deionized water in a separate container, blotted dry, frozen in liquid nitrogen, and lyophilized. ERM was collected by blending the solidified medium at high speed in 10 mM sodium citrate buffer (pH 6.0) at an approximate gel: buffer ratio of 1 :2.5 for 2 min, which dissolves the gellan (Pfeffer et al., 1999). Tissue was collected on a sieve which was rinsed with four 30 mL aliquots of cold 10 mM NaZSO4, followed by an equal quantity of cold deionized water. The final flow through was checked for radioactivity, which was not detectable by scintillation counting in a 0.5 mL aliquot. To collect ERM from root compartments, the roots were first removed from the media under a dissecting microscope. Fungal mycelium collected from the sieve was blotted briefly, and frozen in liquid nitrogen. Tissue was weighed after overnight lyophilization. Extraction procedure For biochemical fractionation, lyophilized fungal mycelium was pulverized with two 3 mm stainless steel beads using a bead mill (Retsch MM301). Due to the high lipid content, 0.1 mL of cold methanolzwater (70:30) was added to aid in disruption. The samples were shaken at 30 Hz for 4 min, and 2 uL samples were analyzed by dissecting microscope to ensure that hyphae and any spores had been broken. After disruption, 0.9 mL of cold methanolzwater (70:30) was added and the sample was vortexed for 5 min. Samples were then centrifuged and supematants collected. The cold aqueous methanol 55 extraction was repeated twice more using 1 mL each time, and the supematants pooled. 0.5 mL of supernatant solution was scintillation counted after adding to 5 mL of BioSafe 11 (MP Biomedicals) scintillation cocktail. In order to determine the amount of sulfate in the samples, the sulfate was precipitated from an aliquot of the aqueous alcohol extract by adding 0.1 mL of 100 mM NaZSO4 and 0.3 mL of 10 mM HCl solution to 0.5 mL of the sample followed by vortexing and the addition of 0.1 mL of 100 mM BaClz. Samples were then incubated for 30 min at 100°C. The resulting barium sulfate precipitate was removed by centrifugation, and the sulfate-free supernatant was scintillation counted. The amount of sulfate in each sample was determined by the subtraction of these counts from those obtained from an aliquot of the original aqueous alcohol extract. Tests with standards showed that 99.8 +/- 0.1 % of sulfate was removed by this procedure (three trials with three samples each) and that 96.4 +/- 1.3 % of 35S-Cys and 95.6 +/- 1.6 % of 35 . . . S-Met remained in solutron. Residues after aqueous ethanol extraction were extracted using lmL of protein extraction buffer containing 9 M urea, 1 % SDS, 25 mM Tris-HCl pH 6.8, 1 mM EDTA, and 0.7 M 2-mercaptoethanol as described by Osherov and May (1998). After vortexing and centrifugation at 10,000 rpm for 5 minutes, 0.5 mL of the supernatant was added to 5 ml of scintillation cocktail and counted. The remaining cellular debris was transferred to sealed glass vials containing 0.5 mL TS-2 tissue solubilizer (MP Biomedicals) and incubated at 70°C for 1-2 weeks until solubilized. The solubilized solution was mixed with 5 mL of scintillation fluid and titrated to pH 7 with HCl. This solution was diluted ten fold to diminish color quenching and analyzed by scintillation counting. 56 Dried roots were pulverized in 15 mL centrifirge tubes with ten 5 mm metal beads for 30 min using a paint shaker. Five mL of cold methanolzwater (70:30) was added to the powder and the tubes were vortexed for five minutes. While keeping particulates suspended, a 1 mL aliquot of the solution was transferred to a microcentrifuge tube and centrifuged. Subsequent steps were as for firngal samples. Measurements of 353 in media Three 1 cm diameter wells were excavated in the gel at the time of labeling with a sterilized cork borer as far from one another as possible. 0.1-0.2 mL of media of the same composition as that in the compartment being analyzed was added to each well, and this volume was maintained through weekly refilling as needed. At the start of the experiment, label was added to one of the wells, and 35S in each of the three wells were measured after one week to ensure efficient diffusion of the radioactivity through the plate. 358 contents were measured by adding 50 pl of liquid media to 5 mL of scintillation fluid, followed by scintillation counting. Non-mycorrhizal root measurements Roots were grown on solidified M media containing 0.12 mM S as NaZSO4 for four weeks and ~5cm root segments were aseptically removed, weighed, and added to 20mL of liquid M media without S (ZnSO4* 7H20 and CuSO4* 5H20 replaced by 1.2 mg of ZnCl and 0.5 mg of CuClz) for measurements of sulfur-mediated growth 57 limitation. For each sulfate concentration (0, 0.005, 0.01, 0.02, 0.05, 0.1, 0.5, and 3 mM), roots were weighed then placed on 20 mL of solidified media per side of three bi- compartrnental petri plates and allowed to grow for three months. The roots were collected, dried at 70°C overnight, and weighed. To measure S uptake, roots were grown at 0.12 mM sulfate as above and transferred into 20mL of S- media solidified with gellan containing 0.02, 0.12, or 3mM sulfate labeled with 11.5 uCi of Naz[3SSO4]. After four weeks, the roots were removed fi'om the media, rinsed several times in deionized water to remove external radioactivity, lyophilized, and weighed. The extraction procedure for the radio-labeled tissue was simplified to include only the aqueous alcohol and solubilization steps. Addition of S metabolic regulators Cys, met, and gsh were added to a final concentration of 1 mM (as 0.2-0.5 ml of sterile aqueous solutions) to the fungal compartments of split plates with ten week old roots growing with 0.52 mM NaZSO4. The uptake of 37.4 uCi of Naz[3SSO4], which was added to fimgal compartments simultaneously with metabolic regulators and 0.12 mM Na2804, was monitored by analyzing aliquots of media from a liquid filled well as described above. The plates were collected at 2, 4, and 6 weeks. The fungal mycelium was collected by dissolving the gellan in a blender with 10 mM sodium citrate (pH 6), then sieving the blended solution. The roots were extracted from the media using forceps and rinsed in deionized water for five minutes before freezing. The fungal and root tissues were extracted as described above. 58 Using cultures grown as for the above experiment, 1 mM of cys labeled with 25.7 uCi of [3SS]cys or 1 mM met labeled with 54.2 uCi of [3SS]met was applied to the fungal sides of 15 plates as the sole S source. The plates were incubated at 25°C for one month prior to collecting and extracting the biochemical fractions as described above. RNA extraction and putative gene fragment isolation Sequences of putative sulfur gene fragments were identified fiom an EST database (Jun et al., 2002) and from additional high throughput sequencing data (PJ Lammers, Shachar-Hill and coworkers unpublished results) from cDNA made from extraradical mycelium. To confirm their identity, total RNA was extracted using the RNeasy Plant Mini Kit (Qiagen, Valencia, CA) from frozen G. intraradices germinating spore tissue, disrupted as already described using a bead mill, followed by DNA removal using RNase-fiee DNase (Turbo DNA-free; Ambion, Austin, TX). Samples were split into two aliquots following quantification by absorbance for the synthesis of cDNA and a control without reverse transcriptase. cDNA was synthesized using SuperScript II reverse transcriptase(1nvitrogen corp., Carlsbad, CA). Primer sets were developed from EST and high throughput sequencing data as follows: Putative sulfate perrnease forward primer 5’- TAGCAATAATTACAAGAATACCAG-3’, and reverse primer 5’- GGGATTCTTATCTTGGAA-3’; putative cystathionine B-synthase forward primer 5’- CTAGCCACTCCTGTAATTGTTCCTC -3’, and reverse primer 5- GAGAAAGCAGGAATACTTAAACCAGG -3’; putative cystathionine y-lyase forward primer 5’- GGTGGAATGTTAAGTTTTAGGATTAAGG -3 ’, and reverse primer 5’- CATCAACATCTTCGATACCGATG-3’. PCR products were separated by and isolated 59 from agarose gel using the QIAquick Gel Extraction Kit (QIAGEN GmbH, Germany) and sequenced by an ABI PRISM® 3100 Genetic Analyzer. (Applied Biosystems, CA, USA). Resulting sequences showed single open reading frames. The sequences studied were submitted to GenBank and given the following accession numbers: Putative high affinity sulfate perrnease (FJ 161947); putative sulfate adenylyltransferase (FJ 161948); putative cystathionine y-lyase (FJ 161949); putative cystathionine [El-synthase (FJ 161950). Quantitative real time PCR measurements Mycorrhizal split plates were grown until the fungal compartment was approximately half colonized. To the fungal compartment 1 mL of a filter sterilized solution containing either sodium sulfate or sodium sulfate and cys was applied to give a final concentration of 0.12 mM sulfate and 1 mM cys. Plates were incubated for 24 hrs before tissue from five to seven plates was collected as described above and immediately frozen in liquid nitrogen. RNA was extracted and converted to cDNA as described above. The initial qRT-PCR reaction mixture containing primers at a concentration of 300 nM and 1 ng of cDNA template was made to 0.015 mL in a 96 well plate, 0.01 mL of which was transferred to a 384 well plate and the PCR reactions were monitored using an ABI Prism 7900 HT Sequence Detection System (Applied Biosystems, Austin, TX) with the following cycling program: 50°C for 2 min, 95°C for 10 min followed by 40 cycles of 95°C for 15 sec and 60°C for l min. Power SYBR Green 2-Step Master Mix (Applied Biosystems, Austin, TX) was used for all real time PCR assays. The AACT, and comparative CT, methods were utilized for the determination of relative gene expression (Livak and Schmittgen, 2001). The expression of an S4 ribosomal protein was used to 60 normalize relative gene expression data as described by Govindaraj alu et al. (2005). Other primer sequences (IDT, Syntegen, Skokie, IL) for QRT-PCR were designed using Primer Express software from Applied Biosystems (Austin, TX) as follows: Putative high affinity sulfate perrnease forward 5’-TTGGATCATTCTTTCATGCGTATC-3’, reverse 5’- GACACCGGCTAATGGAGTACGT-3’; putative S-adenosyl transferase forward 5’- CCGGAGTTGATGATCCTTACG-3’, reverse 5’- ACTGACACACTGTTTACTAACATCAACAA-3’; putative cystathionine B-synthase forward 5’-TGCTTCAGTTGGTGTACGAACAA-3’, reverse 5’- AAATGAGCCAGGAAAAGGTTGA-3’; putative cystathionine y-synthase forward 5’- GACCAGCGTGAGTCATTTTAGAAG-3 ’, reverse 5 ’- AGCTGAATCTTTGGGTGGTGTT-3’. Experimental samples were measured using six technical replicates for each of three biological replicates per condition, and errors were expressed as standard error of the mean. 61 Results Growth and sulfate uptake by non-mycorrhizal roots To establish conditions for studying S transfer between host roots and AM fungus, the relationship between sulfate availability, root growth, and sulfate utilization was investigated in uncolonized roots (Figure 2-1). In the absence of available sulfate there was a three fold decrease in root growth compared to growth under saturating sulfate levels (3 mM; Figure 2-1A). Final dry weights were not significantly different when compared singly at sulfate concentrations above 0.005 mM, however there was a consistent trend of increasing dry weight with increased S availability. Root growth was detected even in roots exposed to medium containing no S, probably due to internal S stores within the initial root segments. A much greater difference was seen in the uptake and incorporation of sulfate when roots were labeled with 35804-2 at high (3 mM), low (0.014 mM), or intermediate (0.114 mM) sulfate levels (Figure 1B). There was a close to four fold increase in both aqueous alcohol-soluble and —insoluble S content from low to intermediate S conditions, and an eleven fold increase between low and high S. A reduction in the root compartment sulfate concentration from 3 mM to 0.12 mM resulted in a reduction in uptake of 64% (Figure 2-1B), with little or no reduction (27% difference in mean values, not statistically significant at the 95% confidence level) in growth (Figure 2-1A). Lowering the sulfate available to mycorrhizal roots to less than 0.05 mM had a negative impact on the symbiosis, reducing the number of plates where the firngal ERM grew over the barrier from greater than 80% to less than 20% (data not shown). There was no such reduction when 0.12 mM sulfate was utilized. 62 30- 25‘ :: -Jr’irJl—i Root Growth (mg DW) 5)“ SS— 0 0.025 0.05 0.075 0.1 0.5 3. O 8042' Concentration in Media (mM) to S Incorporation (ug mg'1 DW) 0.02 0.12 3 8042' Concentration (mM) Figure 2-1 The effect of sulfate concentration on the growth and sulfur incorporation of non-mycorrhizal transformed carrot roots. (A) Root growth measured as dry weight increase at different concentrations of NaZSO4 in 20mL of modified liquid M media (St.Arnaud et.al., 1996). (B) The incorporation of S from solid media containing 0.02, 0.12, or 3 mM NaZSO4 labeled with 11.5 uCi of Na2[35SO4] was measured in roots after four weeks of labeled growth. Labeled root compounds were extracted with cold MeOHzHZO (70:30) (white bars) before the cellular residue was solubilized in tissue solubilizer (gray bars). The fractions were analyzed by liquid scintillation counting. Error is expressed as standard error of the mean. 63 S Removed from Media (pg) 0) O o 10 20 30 4o 50 60 70 3 Days Labeled NNWQ§§ §$8$88 1501 “Iii S Incorporation (ng mg"I DW) 01 O O Fungal S- Fungal S+ Root 8- Root S+ Figure 2-2 Increased sulfate availability in the root compartments of split plates reduces transfer of S through the distal fungal ERM. (A) The uptake of S by ERM from plates labeled with 30 uCi of Na2[3 5 504] in 0.12 mM Na2SO4 in the fungal compartments was measured as the removal of radioactivity from the media when the root compartments were supplied with moderate S (0.12 mM NaZSO4; solid line) or high S (3mM NaZSO4; broken line). (B) Measurements of S incorporated into the roots (Root S+/-) and distal fungal (Fungal S+/-) ERM from the above experiment after 66 days of labeling; S+ refers to high sulfate levels (3mM) in root compartments and S- to moderate levels (0.12mM). Radioactivity was measured by scintillation counting after a crude separation by cold MeOHzHZO (70:30) extraction (white bars) followed by solubilization of the remaining tissue (gray bars). The aqueous alcohol extraction dissolves sulfate, sulfite, S amino acids, and potentially also unidentified sulfated esters while the remainder includes proteins, sulfolipids, and other unknown products. Values are presented as mean +/- s.e.m (n25). 64 Fungal uptake and transfer of sulfate is influenced by root access to sulfate The extraradical mycelium of Glomus intraradices takes up sulfate at significant rates (Figure 2-2A) and transfers most of it to the mycorrhizal roots (Figure 2-2B). The 35 . . . total amount of S in ERM tissue was less than 5% of the amount found in the root tissue, owing to the greater than twenty-fold increase in biomass between the two (tissue mass data not shown). The availability of S in the root compartment strongly influences fungal uptake in the distal ERM compartment. More than twice as much sulfate was removed from the fungal compartment when the corresponding root compartment was exposed to 0.12 mM versus 3 mM S (Figure 2-2A). When roots were grown in 0.12 mM sulfate, uptake by the fungus was linear up to day 40 (R2 = 0.98). The fungus removed 0.98 pg of S from the media per day during this time. Fungal uptake of S when 3 mM sulfate was supplied to roots was linear (0.83 pg of S per day, R2 =0.92) only for the first 20 days before the rate of removal slowed. The total amount of S transferred from the fungal compartment to roots grown at high sulfate levels was about one quarter of that in roots exposed to moderate sulfate levels (Figure 2-2B). In contrast, there was no significant effect of sulfate levels in the root compartment on the incorporation of S by the distal ERM. Additionally, in roots growing at moderate S levels the fraction containing sulfate and the sulfur amino acids . . . 35 . . . contained approxrmately twrce as much S as did the same fraction in the ERM. 35 . . . The amount of S in the media of the colonlzed root compartments was monitored throughout the experiment (see methods) and found to be low. The movement into the root compartment media from the fungal compartment was linear (R2 = 0.98) at 65 63 ng S per day when the root compartment media contained 0.12 mM sulfate. This represents 6.4% of the amount transfer to the roots. The movement into the high S root media was also linear (R2 = 0.99) with 55 ng S per day being moved across the barrier, representing 6.6% of the amount transferred to the roots. 66 Figure 2-3 Transfer of 358042- through the fungus versus direct root uptake. (A) The uptake of S over 104 days from root and fungal compartment media. The root (proximal) compartments of split plates were labeled with 17.3 pCi of Naz[35SO4] in 0.2 mM NaZSO4 (171 pg S). Half of the fungal (distal) compartments were labeled with 17.3 pCi of Na2[35SO4] in 0.2 mM Na2SO4 (171 pg S, open diamonds, solid line), while the others contained 0.02 mM sulfate without S (closed squares, broken line). Distal uptake was monitored from labeled plates (closed triangles, solid line). A straight broken line denotes the total initial S content of the proximal compartment media. (B) Incorporation of 358 by roots when the fungal side was supplied with low unlabeled (0.02 mM) or moderate 35S-labeled (0.2 mM) sulfate was measured after 110 days of labeling by chemical fractionation into sulfate- (white bars), amino acid- (light gray bars), and protein-containing (medium gray bars) pools, and solubilized remaining tissue (dark gray bars) pools measured by liquid scintillation counting. (C) Incorporation of S into different pools of ERM collected from the root or fungal side compartments of split plates when the filngal compartments contained 0.02mM unlabeled sulfate (S-) or 0.2mM of 358 labeled (S+); as above the root side compartments of all plates contained 0.2mM 358 labeled sulfate . Fractions are depicted as in panel A. Values represent means +/- s.e.m. (n24). 67 Figure 2-3 00.. our 563 9:; ow b n. in \IIM .....M NH. ON ow cm oo oo v ON _. ov _. om _. cm 9 (6n) elpew won peaouleu s 68 Figure 2-3 (cont’d) 1.6- 1.4‘ 1.2: le—i o.a - -. 0.6 - 1 0.4 - 0.2 l S Incorporation (pg mg'1 DW) 0.02 mM 0.12 mM $042- in Fungal Side Media (mM) 600 - 500 l 400 - 300 . 200 ' S Incorporation (ng mg-1 DW) Fungal Side ERM Root Side ERM 69 Fungal transfer of s to the mycorrhizal roots versus direct uptake by roots The relative contribution of S obtained from the fungal ERM to total root S was determined at intermediate S levels (0.22 mM) by comparing filngal derived 358 in roots with 358 directly absorbed by the roots themselves. Plates were labeled with 358042- in the root compartment to determine the amount of S incorporated by direct root uptake. Half of these plates were also labeled in the fungal compartment with 358042- at 0.22 mM sulfate while no S was added to the fungal compartments of the remaining plates. The roots took up most of the labeled sulfate initially provided within 8 weeks (Figure 2- 3A). This uptake was not affected by the presence of sulfate in the fungal compartments. The uptake of S from the fimgal compartment media began when the fungus crossed the barrier (~30 days after inoculation) and was comparable to direct root uptake rates in the next few weeks (Figure 2-3A). S incorporation by the root and fungal tissue after ~10 weeks of grth is shown in figures 33 and 3C. When the fungal compartment contained 358042: the total incorporation of S by the roots increased by ~25% (Figure 2-3B). The largest change was in the sulfate fraction, which increased by 60%. There were no significant differences in root grth observed when sulfate was added to the fungal compartments (root weight data not shown), which is consistent with the lack of significant grth increase between moderate and high sulfate growth conditions (Figure 2-1). When the fungal side was labeled, total incorporation of 35$ by the fungus was similar in ERM collected from the fungal and root compartments when the fungal side was labeled (Figure 2-3C), which is consistent with the absence of any known physiological differences between the ERM in the two compartments, and showing that 70 the availability of S to the distal ERM did not affect incorporation by the ERM in the root compartment. 358 accumulated in the fungal ERM from the unlabeled fungal compartment showing that in a low S environment (0.02 mM) S can be moved by the fungus from the root compartment to the distal ERM (Figure 2-3C). 71 > 140- A E . .I. o 120 To: A AB E 100- .. or B 3 5 so g ‘ c ‘3' .2 B 4.: S 60‘ B o O. '5 o 40~ .E ID .. 20- O O C m ‘ ‘ A 0-1 8042' cvs MET GSH Figure 2-4 The effect of common sulfur metabolite repressors cysteine, methionine, and glutathione on the transfer and allocation of 35$ in an arbuscular mycorrhiza. The incorporation of 358 into roots and fungal ERM was analyzed afier labeling the distal compartments with 37.4 pCi of Na2[35SO4] in 0.12 mM Na2s04 (91 pg S) alone (8042-), or with lmM cysteine (CYS), lmM methionine (MET), or lmM reduced glutathione (GSH). Root (panels A-C) and fungal (panels D-F) was analyzed after two (A,D), four (B,E), and six (C,F) weeks. Fungal ERM was collected from the distal compartments of same plates as those from which root tissue was collected from the proximal compartments. Tissue was analyzed by chemical fractionation into pools containing sulfate (white bars), amino acids (light gray bars), proteins (medium gray bars), and the tissue solubilized biomass remaining afier extraction (dark gray bars). Fractions were analyzed by liquid scintillation counting. Letters within the bars represent ANOVA single factor analyses between bars of that type. Letters above the complete bar graphs are the analyses of total S uptake. Values are reported as mean +/- s.e.m. (n25). 72 Figure 2-4 (cont’d) c c _a CB B A B A B B B A l A A 1 fl - .t — — [—I 0 O 0 0 0 0 5 0 5 0 W 0 3 2 2 1 1 cso F.mE w 9.: :0320385 m «com CYS MET GSH $042- 73 Figure 2-4 (cont’d) AB 300 - 250 - 0 .25 :9: m o 5 0 5 1 m5 :ozfloahoofi w «com GSH MET CYS 74 Figure 2-4 (cont’d) D 250 - 200 r 150 ' 0 5 :50 F.mE w 9: 53209.85 w 395.... GSH MET CYS 75 Figure 2—4 (cont’d) 400 ~ 0 0 0 o 5 0 5 2 1 1 :50 F.9: w 55 cozfionuoofi w _umcam 350 r o . 250 ~ GSH MET CYS 76 Figure 2-4 (cont’d) F 300 ‘ 250 ‘ 200 * 150 ‘ 100 ‘ 50 r :50 F.mE w as cos—20985 m 395.... GSH MET CYS 77 The effect of sulfur metabolites on sulfate uptake and transfer through the mycorrhizal symbiosis Radio-labeled sulfate was added to the fungal compartments of split plates together with methionine (met), cysteine (cys), and glutathione (gsh) in amounts previously shown to suppress sulfate assimilation in other filngi (Kuras & Thomas, 1995; Ono et al., 1999; Ketter & Marzluf, 1988; Ketter et al., 1991; Grynberg et al., 2001). The . . . 35 . . . distributlon of S in root and fungal tissues was then measured after two, four, and srx weeks. Substantial quantities of S were transferred to the roots over the course of the experiment despite the presence of 0.5 mM sulfate in the root compartments (Figure 2-4A to C). The presence of cys decreased S transfer to the mycorrhizal roots by an average of 45% throughout the experiment, with the sulfate- and amino acid- containing pools of S in the roots showing larger fractional decreases. Methionine had no significant effect on . . . 35 total S transfer although there were srgnlficant decreases in the S transferred to the low molecular weight soluble metabolite pools in mycorrhizal roots. Glutathione dramatically 35 . . . . . . reduced S accumulation 1n the roots at all time pornts, With an average reduction of 80%. There were no significant effects of the presence of cys, met or gsh on total S incorporation from labeled sulfate by the distal ERM and little or no effect on the partitioning of S within this tissue (Figures 2-4D to F). Glutathione greatly reduced the growth of fungal ERM in the distal compartment though not the root grth (not shown) so it is unclear whether the reduction in the transfer of S by gsh treated ERM is due to the regulation of sulfate uptake or simply to reduced fungal growth. Neither cys nor met significantly affected the grth of fungal 78 ERM or roots, with ERM biomass approximately doubling between weeks two and four (data not shown). Thus the effect of cys on S transfer to mycorrhizal roots does not appear to be due to growth inhibition. 79 350 ] 300 i 250 - 200 - 150 - 100 - 50* CYS ROOT CYS ERM MET ROOT MET ERM Figure 2-5 Transfer of reduced S through a mycorrhizal symbiosis. Mycorrhizal roots were grown in 0.5 mM Na2804 and the filngal partner allowed to grow into distal compartments containing either lmM cysteine (CYS) labeled with 25.7 pCi [ S]-Cys or lmM methionine (MET) labeled with 54.2 pCi of [3SS]-Met as the sole S source. Roots and fungal distal ERM were collected afier four weeks of labeling and the radioactive counts split into different biochemical fractions. Fractions included SO4 _ (white bars), amino acids (light gray bars), proteins (medium gray bars), and the remaining solubilized cellular material (dark gray bars) and were obtained through scintillation counting. Values are reported as mean +/- s.e.m. (n=9). 8O Fungal uptake and transfer to host roots of reduced 3 Fungal ERM imports and transfers substantial amounts of reduced S supplied as cys or met (Figure 2-5). The amount of S transferred when 35S-labeled cys is the sole source of S in the fungal compartment is comparable to the S transfer when plates are labeled with sulfate (Figures 2-2B, 2-3B, and 2-4A to C). Roughly half of this amount was transferred when met was provided (Figure 2-5). The incorporation of S by the fungal mycelium per mg dry weight when labeled cys is supplied to the fungal compartment is less than the level found in the roots by about 40%. Utilizing met as an S source led to a 46% increase in the amino acid fraction in the ERM (Figure 2-5). There were striking differences between the relative amount of S in the solubilized fractions in both roots and fungal mycelia when reduced S was supplied compared to sulfate. Labeling with cys and met resulted in a much higher percentage of labeling in the solubilized root tissue pool (78% and 88%, respectively) than when sulfate was supplied as the primary S source (7-21%; Figures 2-4A to D). The opposite relationship was found in fungal mycelia, where applying cys or met as primary sources of S led to solubilized tissue fractions that were at most 8% of total S incorporated (Figure 2-5). When labeled sulfate was the primary S source, 5-22% of the total S incorporation is found in the tissue solubilized fraction (Figures 2-4D to F). Gene expression measurements The expression of fungal transcripts encoding putative genes involved in sulfate assimilation was analyzed in relation to the application of cys to the distal ERM. Partial sequences of transcripts representing putative sulfur assimilatory genes were identified by high throughput 454 sequencing of cDNA from fungal ERM grown in split plates in M 81 medium (3mM sulfate). These included a 234 bp sequence with 67% identity at the amino acid level to SULl from S. cerevisiae, and a 247 bp sequence with 67 % and 70% identities at the amino acid level to sulfate adenylyltransferase from A. nidulans and A. terreus, respectively. Sequences with homology to enzymes defining the reverse trans- sulfuration pathway were also identified. These included a 403 bp sequence with 79% and 71% identities at the amino acid level to cystathionine B-synthase from A. fumigatus and CYS4 from S. cerevisiae. Additionally, a 220 bp sequence was identified with 58% and 69% amino acid identities to CYS3 from S. cerevisiae and cystathionine y-lyase from A. fumigatus. The presence of transcripts for these putative sulfur genes is are consistent with sulfate uptake and assimilation via the usual assimilation pathway, and also strongly suggests the presence of the reverse trans-sulfuration pathway in AM fungi. The expression of the sulfate perrnease and reverse trans-sulfuration enzyme candidates were analyzed in ERM after exposure to 0.12 mM sulfate, with or without the presence of 1 mM cys. The conditions were chosen to enable the comparison between gene expression changes and the uptake and transfer of radio-labeled sulfate with cys addition (Figure 2- 4A to C). There was a reduction in the expression of the candidate high affinity sulfate perrnease sequence by a factor of 3.15 +/- 1.15. The expression of sequences with homology to sulfate adenylyltransferase and cystathionine y-lyase was not significantly different between treatments. The expression with cys addition of the putative cystathionine y-lyase sequence studied was significantly different. However, the difference was small (less than two fold). 82 Fold Change SUL SAT cpsyn Cv s Figure 2-6 Fungal gene expression changes in response to the addition of 1 mM cysteine to ERM. Expression relative to control plates of sequences with homology to high affinity sulfate perrnease (SUL), S-adenosyl transferase (SAT), cystathionine B-synthase (CBsyn), and cystathionine y-lyase (Cylse) was analyzed by quantitative real-time PCR using a ribosomal protein sequence as a control as previously reported (Govindaraj ulu et al., 2005; see methods). RNA was extracted from ERM tissue collected after 24 hr incubation with 0.12 mM S042; and with or without 1 mM cys added to the distal compartments of five split plates per sample. Values are reported as mean +/- s.e.m. (n=3 biological replicates each consisting of tissue from 5 to 7 plates). 83 Discussion The uptake of 35$ by non-mycorrhizal roots was dependent on the external sulfate concentration (Figure 2-1B), as has been reported for tomato seedlings and carrot storage root sections (Lopez et al., 2002; Cram, 1983). In carrot, Cram (1983) showed a proportional increase up to an external concentration of 50 mM, leading the authors to conclude that feedback inhibition of sulfate uptake is absent. Vegetative root growth was less affected by alterations in available sulfate concentrations than was uptake or reduction (Figure 2-1 A). Subsequent measurements of mycorrhizal uptake activity were conducted using sulfate concentrations that increased potential differences in root uptake and contents and minimized the limitation of growth by S. Bi-compartmental Petri plate cultures are a well established model mycorrhizal system (St. Amaud et al. 1996) used extensively in studies of mycorrhizal metabolism and nutrient transfer (e. g. Govindarajulu et al., 2005). This system made possible the use of defined media and the isolation of the majority of extra-radical mycelium (ERM) from the rhizosphere, allowing S movement through the ERM to be quantified. With moderate S levels available to the roots (0.12 mM), uptake and transfer by the fungus increased the amount of S in the roots by 25% (Figure 2-3B). The movement of 358 into roots was also observed in all the experiments regardless of the S-source utilized in quantities much higher than those found in fungal ERM. Fungal uptake and transfer of S was influenced by the concentration of sulfate available to the host roots. Increasing the sulfate concentration in the root compartment from 0.12 mM to 3 mM resulted in a halving of fungal uptake (Figure 2-2A), while approximately one fifth of the amount of transferred S 84 was found in roots grown in 3 mM sulfate compared to the amount found in roots grown in 0.12 mM sulfate (Figure 2-23). These reductions may be due to host signaling and/or to changes in the concentration gradient of S at the rootzfungal interface. After supplying 35S-sulfate to the fungal compartment under a range of conditions, labeled sulfate was invariably found in roots in amounts that dwarfed the total quantities in the fungal mycelium. For example, after six weeks of labeling the fungal . 35 2- . . compartment With 804 , total sulfate in the fungus was 5.1% of the amount in the roots (Figures 2-4C and F). Roots in that experiment were grown in 0.5 mM sulfate and the distal ERM was supplied with 0.12 mM, and the same was true (5.3%) when the roots were grown in 0.22 mM sulfate and the fungus was also labeled with 0.22 mM sulfate (Figure 2-3B and C). Based on this evidence one may conclude that the sulfate anion is transferred by G. intraradices to host roots. However, as shown in Figure 5, the fungus is clearly also capable of the uptake and transfer of reduced forms of S at rates comparable to sulfate. Thus the ERM can supply the host plant with organic forms of S from the soil. Since an estimated 95% of the S in soils is in organic forms, the ability of mycorrhizal plants to access S from this source is potentially important for plant nutrition. The first steps towards understanding the regulation of sulfate uptake by a mycorrhizal fungus in the symbiotic state were taken by analyzing the effect of common sulfur metabolite repressors cys, met, and gsh on the uptake and transfer of radio-labeled sulfate through the fungus. The addition of 1 mM met was shown to suppress the expression of sulfur assimilation pathway genes in yeast (Kuras & Thomas, 1995) and the complete suppression of uptake has been demonstrated in the presence of ten times lower levels (Breton & Surdin Kerjan, 1977). This finding was later refined by mutant analyses 85 and the suppression shown to be due to the synthesis of cys through the reverse trans- sulfuration pathway (Ono et al., 1999). The sulfate perrneases $3 in A. nidulans and Cysl4 in N. crassa are also strongly suppressed by met (Ketter & Marzluf, 1988; Ketter et al., 1991; Grynberg et al., 2001). In contrast to these organisms, the addition of 1 mM met to the fungal compartments of mycorrhizal split plates had no significant effect on total sulfate uptake and transfer (Figure 2-4). By contrast the application of 1 mM cys to the ERM resulted in S transfer from fungus to mycorrhizal roots being approximately halved, (Figures 2-4A to C). The amount by which sulfate transfer is reduced in the presence of cys (Figure 2-4) is similar to the level of S that is taken up as cys and transferred to the roots (Figure 2-5). Labeled gsh has been shown to be imported by ectomycorrhizal oak trees at rates comparable to sulfate (Seegmiiller & Rennenberg, 2002). However, the addition of 1 mM gsh to endomycorrhizal distal ERM resulted in a suppression of the growth of this tissue. While the transfer of sulfate continued and the assimilation by the ERM was largely unaffected on a per mg basis, S transfer was greatly diminished. Since gsh also suppressed the expression of the putative sulfate perrnease (data not shown), it is unclear how much of the effect of gsh was due to regulation of transport and how much to repression of ERM growth. Thus the regulation of S uptake and metabolism in G. intraradices is different from that in other filamentous fungi studied to date. The changes in expression of genes putatively involved in sulfate transport and assimilation indicates a role for transcriptional regulation of sulfate perrnease in the reduction in sulfate transport in the presence of cys. The reduction in S transfer of 45% with the application of 1 mM cys to distal ERM (Figure 2-4C) correlated with a 3.8 +/- 86 1.2 fold reduction in the expression of mRNA encoding a putative sulfate perrnease. The lack of a significant change in the expression of a cDNA sequence with high homology to S-adenosyl transferases suggests that the reduction in S transfer is a function of reduced uptake rather than sulfate reduction. However, the analysis of more S assimilation pathway genes and compartment-specific sulfate pennease isozymes is needed before concluding that the transcriptional regulation of this sulfate perrnease gene explains the reduction in S transfer with the addition of cys. Sequences encoding putative genes with homology to cystathionine B-synthase and cystathionine y-lyase were also identified. Their expression is indicative of a functional reverse trans-sulfirration pathway in Glomus intraradices, akin to A. nidulans and S. saccharomyces, but unlike S. pombe (Ono et al., 1992a, 1992b; Paszewski and Grabski, 1975; Brzywczy et al., 2002). The putative cystathionine B-synthase sequence was slightly down-regulated (1.49 +/- 0.41 fold) in the presence of cys while the sequence with homology to cystathionine y-lyase showed no significant change. 87 S042'(our) ROOT 2- —» 5 REDUCED $04 (ROOT) ( ) “ Protein Protein IRM I ——..’. S(REDUCED) ......— _____ 5 4| 4| _ L.._—_ 2- _ SC’4 (our) lip \ rise—--- I l/ S042'(our) Figure 2-7 Model depicting S transfer through an endomycorrhizal symbiosis. Roots and fungal mycelium both import SO42- from external sources, and fungal uptake can supply isolated mycelium. The transfer of S04} through the mycorrhizal symbiosis is inversely proportional to root uptake. (23ys and met are imported by the fungus, resulting in a reduction of fungal uptake of S04 _ (dashed line) and the transfer of a reduced form of S to the root (S(REDUCED)). The reduction of 8042- and uptake of reduced S leads to incorporation in the protein pools in both roots and fungus. Putative steps in the assimilation pathway based on sequence data are depicted as gray arrows. Steps involving putative sulfur assimilation genes are labeled in gray letters as follows: (a) high affinity sulfate perrnease; (b) sulfate adenylyltransferase; (c) y-cystathionine lyase; (d) B- cystathionine synthetase. IRM, intraradical mycelium; ERM, extraradical mycelium. 88 Figure 2-7 is a working model consistent with the findings presented. We have demonstrated uptake, assimilation, and transfer of sulfate (Figures 2-3B and 2-4A to C) and of reduced S (Figure 2-5) by the fungal partner to the host roots in physiologically significant quantities, along with bidirectional transport of sulfate within the fungal ERM (Figure 2-3C). The inverse relationship between the rhizospheric sulfate concentration and the amount of sulfate uptake from the fungal compartment (Figures 2-2A to B) suggests that an increase in root intracellular or apoplastic sulfate can suppress both the transfer of sulfate from the intraradical hyphae to the host and the uptake by the extraradical hyphae. This relationship is depicted in the model as the negative regulation of sulfate transfer across the rootzhyphal interface by root sulfate levels, which subsequently results in an increase in the fungal intracellular sulfate concentration and suppression of uptake in the distal ERM. The relationship between the transfer of a reduced form of S when cys or met are imported by the fungus and the uptake of sulfate in the presence of cys and met suggest a simultaneous transfer of sulfate and reduced. S and/or a regulatory response. Relative real-time PCR measurements of expression for a putative high affinity sulfate perrnease gene revealed a ~73% reduction in transcript levels when the ERM was exposed to 1 mM cys, indicating a regulatory effect on sulfate uptake (Figures 2-6 and 2-7). The working model proposed is a testable interpretation of the findings presented that invites future experiments to verify gene identities, localize expression within the host roots and analyze regulation in mycorrhizas of whole plants. 89 References Banerjee R, Evande R, Kabil O, tha S, Taoka S. 2003. Reaction mechanism and regulation of cystathionine beta-synthase. Biochimica Et Biophysica Acta-Proteins and Proteomics 1647(1-2): 30-3 5. Baumgardner RE, Lavery TF, Rogers CM, Isil SS. 2002. Estimates of the atmospheric deposition of sulfur and nitrogen species: Clean Air Status and Trends Network, 1990- 2000. Environmental Science & Technology 36(12): 2614-2629. Bhat KKS, Nye PH, Baldwin JP. 1976. Diffusion of phosphate to plant roots in soil .4. Concentration distance profile in rhizosphere of roots with root hairs in a low-P soil. Plant and Soil 44(1): 63-72. Bradfield G, Somerfield P, Meyn T, Holby M, Babcock D, Bradley D, Segel IH. 1970. Regulation of sulfate transport in filamentous fungi. Plant Physiology 46(5): 720- 732. Breton A, Surdinkerjan Y. 1977. Sulfate uptake in“saccharomyces-cerevisiae - biochemical and genetic study. Journal of Bacteriology 132(1): 224—232. Brundrett MC. 2002. Coevolution of roots and mycorrhizas of land plants. New Phytologist 154(2): 275-304. Brzywczy J, Sienko M, Kucharska A, Paszewski A. 2002. Sulphur amino acid synthesis in Schizosaccharomyces pombe represents a specific variant of sulphur metabolism in fungi. Yeast 19(1): 29-35. Cherest H, Davidian JC, Thomas D, Benes V, Ansorge W, SurdinKerjan Y. 1997. Molecular characterization of two high affinity sulfate transporters in Saccharomyces cerevisiae. Genetics 145(3): 627-635. Clarke DL, Newbert RW, Turner G. 1997. Cloning and characterisation of the adenosyl phosphosulphate kinase gene from Aspergillus nidulans. Current Genetics 32(6): 408-412. Cooper KM, Tinker PB. 1978. Translocation and transfer of nutrients in vesicular- arbuscular mycorrhizas .2. Uptake and translocation of phosphorus, zinc and sulfur. New Phytologist 81(1): 43-55. Cram J. 1983. Characteristics of sulfate transport across plasmalemma and tonoplast of carrot root-cells. Plant Physiology 72(1): 204-211. 90 Diop T, Becard G, Pichey Y. 1992. Long-term invitro culture of an endomycorrhizal filngus, Gigaspora margarita, on Ri T-dna transformed roots of carrot. Symbiosis 12: 249-259. Eriksen J, Askegaard M. 2000. Sulphate leaching in an organic crOp rotation on sandy soil in Denmark. Agriculture Ecosystems & Environment 78(2): 107-114. Fitzgerald JW. 1976. Sulfate ester formation and hydrolysis - potentially important yet often ignored aspect of sulfur cycle of aerobic soils. Bacteriological Reviews 40(3): 698- 72 1 . Fortin JA, Becard G, Declerck S, Dalpe Y, St-Arnaud M, Coughlan AP, Piche Y. 2002. Arbuscular mycorrhiza on root-organ cultures. Canadian Journal of Botany-Revue Canadienne De Botanique 80(1): 1-20. Gahoonia TS, Nielsen NE. 1991. A method to study rhizosphere processes in thin soil layers of different proximity to roots. Plant and Soil 135(1): 143-146. Gahoonia TS, Raza S, Nielsen NE. 1994. Phosphorus depletion in the rhizosphere as influenced by soil-moisture. Plant and Soil 159(2): 213-218. Govindarajulu M, Pfeffer PE, Jin HR, Abuhaker J, Douds DD, Allen JW, Bucking H, Lammers PJ, Shachar-Hill Y. 2005. Nitrogen transfer in the arbuscular mycorrhizal symbiosis. Nature 435(7043): 819-823. Gray LE, Gerdemann JW. 1973. Uptake of sulphur-35 by vesicular-arbuscular mycorrhizae. Plant and Soil 39: 687-689 Grynberg M, Piotrowska M, Pizzinini E, Turner G, Paszewski A. 2001. The Aspergillus nidulans metE gene is regulated by a second system independent from sulphur metabolite repression. Biochimica et Biophysica Acta-Gene Structure and Expression 1519(1-2): 78-84. Guo T, Zhang JL, Christie P, Li XL. 2007. Pungency of spring onion as affected by inoculation with arbuscular mycorrhizal firngi and sulfur supply. Journal of Plant Nutrition 30(7-9): 1023-1034. Hansen J, Johannesen PF. 2000. Cysteine is essential for transcriptional regulation of the sulfur assimilation genes in Saccharomyces cerevisiae. Molecular and General Genetics 263(3): 535-542. Hawkins HJ, Johansen A, George E. 2000. Uptake and transport of organic and inorganic nitrogen by arbuscular mycorrhizal fungi. Plant and Soil 226(2): 275-285. 91 Hayman DS, Mosse B. 1971. Plant growth responses to vesicular-arbuscular mycorrhiza .1. Growth of endogone-inoculated plants in phosphate-deficient soils. New Phytologist 70(1): 19-27. Jin H, Pfeffer PE, Douds DD, Piotrowski E, Lammers PJ, Shachar—Hill Y. 2005. The uptake, metabolism, transport and transfer of nitrogen in an arbuscular mycorrhizal symbiosis. New Phytologist 168: 687-696. Joner EJ, J akobsen I. 1994. Contribution by 2 arbuscular mycorrhizal fungi to p-uptake by cucumber (cucumis-sativus 1) from p-32 labeled organic-matter during mineralization in soil. Plant and Soil 163(2): 203-209. Jun J, J A, Rehrer C, Pfeffer PE, Shachar—Hill Y, Lammers PJ. 2002. Expression in an arbuscular mycorrhizal fungus of genes putatively involved in metabolism, transport, the cytoskeleton and the cell cycle. Plant and Soil 244: 141-148. Ketter JS, J arai G, Fu YH, Marzluf GA. 1991. Nucleotide-sequence, messenger-ma stability, and dna recognition elements of cys-l4, the structural gene for sulfate permease-ii in neurospora-crassa. Biochemistry 30(7): 1780-1787. Ketter JS, Marzluf GA. 1988. Molecular-cloning and analysis of the regulation of cys- 14+, a structural gene of the sulfur regulatory circuit of neurospora-crassa. Molecular and Cellular Biology 8(4): 1504-1508. Kreuzwieser J, Rennenberg H. 1998. Sulphate uptake and xylem loading of mycorrhizal beech roots. New Phytologist 140(2): 319-329. Kuras L, Thomas D. 1995. Functional-analysis of met4, a yeast transcriptional activator responsive to s-adenosylmethionine. Molecular and Cellular Biology 15(1): 208-216. Lambers H, Raven JA, Shaver GR, Smith SE. 2008. Plant nutrient-acquisition strategies change with soil age. Trends in Ecology & Evolution 23(2): 95-103. Lefohn AS, Husar JD, Husar RB. 1999. Estimating historical anthropogenic global sulfur emission patterns for the period 1850-1990. Atmospheric Environment 33(21): 3435-3444. Leustek T. 1996. Molecular genetics of sulfate assimilation in plants. Physiologia Plantarum 97(2): 41 1-419. Leustek T, Martin MN, Bick JA, Davies JP. 2000. Pathways and regulation of sulfur metabolism revealed through molecular and genetic studies. Annual Review of Plant Physiology and Plant Molecular Biology 51: 141-165. Li XL, George E, Marschner H. 1991. Extension of the phosphorus depletion zone in va-mycorrhizal white clover in a calcareous soil. Plant and Soil 136(1): 41-48. 92 Liu A, Hamel C, Elmi A, Costa C, Ma B, Smith DL. 2002. Concentrations of K, Ca and Mg in maize colonized by arbuscular mycorrhizal fungi under field conditions. Canadian Journal of Soil Science 82(3): 271 -278. Livak KJ, Schmittgen TD. 2001. Analysis of relative gene expression data using real- time quantitative PCR and the 2(-Delta Delta C(T)). Methods 25: 402-408 Lopez J, Bell CI, Tremblay N, Dorais M, Gosselin A. 2002. Uptake and translocation of sulphate in tomato seedlings in relation to sulphate supply. Journal of Plant Nutrition 25(7): 1471-1485. Mansouri-Bauly H, Kruse J, Sykorova Z, Scheerer U, Kopriva S. 2006. Sulfur uptake in the ectomycorrhizal fungus Laccaria bicolor S23 8N. Mycorrhiza 16(6): 421-427. Marzluf GA. 1993. Regulation of sulfur and nitrogen-metabolism in filamentous fungi. Annual Review of Microbiology 47: 31-55. Marzluf GA. 1997. Molecular genetics of sulfur assimilation in filamentous fungi and yeast. Annual Review of Microbiology 51: 73-96. N atorff R, Sienko M, Brzywczy J, Paszewski A. 2003. The Aspergillus nidulans metR gene encodes a bZIP protein which activates transcription of sulphur metabolism genes. Molecular Microbiology 49(4): 1081-1094. Ono B, Hazu T, Yoshida S, Kawato T, Shinoda S, Brszczy J, Paszewski A. 1999. Cysteine biosynthesis in Saccharomyces cerevisiae: a new outlook on pathway and regulation. Yeast 15(13): 1365-1375. Ono B, Heike C, Yano Y, Inoue T, Naito K, Nakagami S, Yamane A. 1992a. Cloning and mapping of CYS4 gene of Saccharomyces cerevisiae. Current Genetics 21: 285-289. Ono B, Tanaka K, Naito K, Heike C, Shinoda S, Yamamoto S, Ohmori S, Oshima T, Toh-e A. 1992b. Cloning and characterization of CYS3(CYIl) gene of Saccharomyces cerevisiae. Journal of Bacteriology 174: 3393-3 347. Ono BI, Naito K, Shirahige YI, Yamamoto M. 1991. Regulation of cystathionine gamma-lyase in Saccharomyces-cerevisiae. Yeast 7(8): 843-848. Ortiz-Ceballos AI, Pena-Cabriales JJ, Fragoso C, Brown GG. 2007. Mycorrhizal colonization and nitrogen uptake by maize: combined effect of tropical earthworms and velvetbean mulch. Biology and Fertility of Soils 44(1): 181-186. Osherov N, May GS. 2001. The molecular mechanisms of conidial germination. F ems Microbiology Letters 199(2): 153-160. 93 Paszewski A, Grabski J. 1975. Enzymatic lesions in methionine mutants of Aspergillus nidulans: role and regulation of an alternative pathway of cysteine and methionine synthesis. Journal of Bacteriology 124: 893-904 Perrier H, Rohn S, Driemel G, Batt N, Schwarz D, Kroh LW, George E. 2008. Effect of nitrogen species supply and mycorrhizal colonization on organosulfur and phenolic compounds in onions. Journal of Agricultural and Food Chemistry 56(10): 3538-3 545. Pfeffer PE, Douds DD, Becard G, Shaehar-Hill Y. 1999. Carbon uptake and the metabolism and transport of lipids in an arbuscular mycorrhiza. Plant Physiology 120(2): 587-598. Pilsyk S, Natorff R, Sienko M, Paszewski A. 2007. Sulfate transport in Aspergillus nidulans: A novel gene encoding alternative sulfate transporter. Fungal Genetics and Biology 44(8): 715-725. Rennenberg H. 1999. The significance of ectomycorrhizal flmgi for sulfur nutrition of trees. Plant and Soil 215(2): 115-122. Rennenberg H, Herschbach C, Haberer K, Kopriva S. 2007. Sulfur metabolism in plants: Are trees different? Plant Biology 9: 620-637. Rhodes LH, Gerdemann JW. 1978a. Hyphal translocation and uptake of sulfur by vesicular-arbuscular mycorrhizae of onion. Soil Biology & Biochemistry 10(5): 355-360. Rhodes LH, Gerdemann JW. 1978b. Influence of phosphorus nutrition on sulfur uptake by vesicular-arbuscular mycorrhizae of onion. Soil Biology & Biochemistry 10(5): 361- 364. Riley NG, Zhao FJ, McGrath SP. 2000. Availability of different forms of sulphur fertilisers to wheat and oilseed rape. Plant and Soil 222(1-2): 139-147. Scherer HW. 2001. Sulphur in crop production - invited paper. European Journal of Agronomy 14(2): 81-111. Seegmuller S, Rennenberg H. 2002. Transport of organic sulfur and nitrogen in the roots of young mycorrhizal pedunculate oak trees (Quercus robur L.). Plant and Soil 242: 291-297. Sienko M, Paszewski A. 1999. The metG gene of Aspergillus nidulans encoding cystathionine beta-lyase: cloning and analysis. Current Genetics 35(6): 638-646. Simon L, Bousquet J, Levesque RC, Lalonde M. 1993. Origin and diversification of endomycorrhizal fungi and coincidence with vascular land plants. Nature 363: 67-69. 94 Spencer B, Harada T. 1960. Role of choline sulphate in the sulphur metabolism of fungi. Biochemical Journal 77: 305-315. StArnaud M, Hamel C, Vimard B, Caron M, Fortin JA. 1996. Enhanced hyphal grth and spore production of the arbuscular mycorrhizal fungus Glomus intraradices in an in vitro system in the absence of host roots. Mycological Research 100: 328-332. Tabatabai MA. 1986. Sulfur in Agriculture. American Society of Agronomy, Madison, WI. Thomas D, SurdinKerjan Y. 1997. Metabolism of sulfur amino acids in Saccharomyces cerevisiae. Microbiology and Molecular Biology Reviews 61(4): 503-515. Wang B, Qiu YL. 2006. Phylogenetic distribution and evolution of mycorrhizas in land plants. Mycorrhiza 16(5): 299-363. Wang ZY, Kelly JM, Kovar JL. 2004. In situ dynamics of phosphorus in the rhizosphere solution of five species. Journal of Environmental Quality 33(4): 1387-1392. 95 Chapter 3 Uptake and Assimilation of S during Presymbiotic Growth of an Arbuscular Mycorrhizal Fungus 96 Abstract Arbuscular mycorrhizal fungal spores germinate in pure water yet readily import C and N. The biological significance of this uptake is unknown and the use of other macronutrients including S is unexplored. The uptake of 35$042., [35S]cys, and [35$]met, and the incorporation of 35S into different metabolic pools during the presymbiotic growth of Glomus intraradices were measured by scintillation counting. Sulfate assimilation was analyzed in relation to sulfate concentration and exposure to unlabeled and labeled cys and met, glutathione (GSH) and buthionine sulfoximine (BSO). Corresponding changes in relative gene expression of putative sulfate assimilation pathway genes were measured by quantitative real-time PCR. Glomus intraradices showed little preference for reduced S sources over sulfate. Exposure to cys or GSH initially reduced sulfate uptake by 35% and 63% respectively. Met addition did not decrease sulfate uptake while gene expression measurements suggested the presence of a functional cystathionine pathway. Gene expression measurements revealed a two fold decrease in S-related genes after cys addition while GSH or met more than doubled the expression of all genes analyzed. S acquisition by germinating spores of G. intraradices appears to be regulated post-transcriptionally and to a lesser extent than in other fungi studied. 97 Introduction The arbuscular mycorrhizal (AM) fiingal symbiosis evolved more than 400 million years ago (Dotzler et al., 2006) and remains the most widespread underground symbiosis. AM fungi can be found in association with more than 90% of plant families in all biomes filled by land plants excluding arctic tundra (Wang & Qiu, 2006). Although the AM symbiosis helps plants take up mineral nutrients, especially phosphorus, from the soil (Cooper & Tinker, 197 8; Li et al., 1991; Gahoonia et al., 1994; Wang et al., 2004; Hepper & Warner, 1983), the application of AM fungi to agriculture has been limited by their obligate biotrophic nature and resulting difficulties in efficient large scale propagation of inoculum (Douds et al., 2006). It is therefore a goal of mycorrhizal research to increase the practical use of AM fungi by determining a culturing method that enables their propagation in the absence of host roots. The transition fi'om pre-symbiotic to symbiotic growth was an early focus of efforts to identify a method of root-less propagation of AM fungi (Hepper, 1984; Carr et al., 1985). Fungal spores germinate without external mineral or carbon sources although they are imported when available (Bucking et al., 2008; Smith & Read, 2008; Gachomo et al., 2009), and germ tubes severed from the spore rely on them to extend normally (Hepper, 1983). Germinating spores are limited in the uptake and use of exogenous carbon sources and lack some primary metabolic pathways including de novo fatty acid synthesis (Bago et al., 1999). Recently, chemicals in root exudates including an isolated signaling molecule, strigolactone, have been shown to alter the metabolism and gene expression of AM ftmgal germinating spores in the absence of a host (Besserer et al., 98 2008). The addition of root exudates alters gene expression in Gigaspora rosea, inducing changes in physiological state in preparation for root penetration (Tamasloukht et al. , 2003). A better understanding of the regulation of metabolic functions and nutrient allocation during pre-symbiotic growth may thus reveal signaling pathways that govern the developmental stages of the firngus. This study examined the regulation of sulfiir assimilation in germinating spores of Glomus intraradices as part of the ongoing effort to achieve axenic propagation of AM fungi. Direct reduction of sulfate with NADPH is a highly endergonic reaction, and is bypassed in fungi through adenylation and firrther phosphorylation of the adenosyl group (Lowe, 1991; Marzluf, 1997). These steps are carried out by S-adenylyl-transferase (SAT) and adenylyl-sulfate kinase (APSK), respectively, and are followed by the two step reduction of the activated sulfate moiety to sulfide. The initial reduction of the sulfate moiety by thioredoxin oxidation is mediated by the enzyme 3’-phosphoadenosine- 5’-phosphosulfate reductase (PAPSR); sulfite reductase is responsible for further reduction to sulfide. Sulfide is incorporated into either O-acetylserine (OAS) or 0- acetylhomoserine (OAH), forming cysteine or homocysteine, respectively. Fungi and archaebacteria are the only organisms known to possess both assimilation pathways (Marzluf, 1997; Zhou and White, 1991). The dominant path of sulfide incorporation varies among fungi studied (Marzluf, 1997). The assimilation of sulfide through OAH requires the conversion of homocysteine to cysteine by a specific two-step enzymatic pathway, the Cystathionine (CT) pathway, which involves the conversion of homocysteine and serine to cystathionine by the enzyme cystathionine B-synthase (CBS) followed by the cleavage of cystathionine to a- 99 ketobutyric acid, ammonia, and cysteine by cystathionine y-lyase (CGL). In Saccharomyces cerevisiae, cysteine synthesis occurs solely through the CT pathway (Cherest & Surdin-Kerjan, 1992; Ono et al., 1999), while the pathway is absent in the closely related yeast Candida valida (Piotrowska, 1993). This is also true for Schizosaccharomyces pombe (Brzywczy et al., 2002), although the possession of both pathways is the norm. Aspergillus nidulans and Neurospora crassa contain the enzymes to assimilate sulfide into either cysteine or homocysteine, as do the yeast Yarrowia lipolytica, Cephalosporium acremonium, and T richosporon cutaneum (Brzywczy & Paszewski, 1993; Paszewski & Grabski, 1974; Morzycka & Paszewski, 1979; Jackobsen & Metzenberg, 1977; Lewandowska & Paszewski, 1987). although in A. nidulans the synthesis of cysteine is dominant (Brzywczy et al.. 2007). Other fungi may have similar biases. Regulation of S assimilation in fungi is mediated by the transcription and allosteric control of S assimilation enzymes (Marzluf, 1997; Foster et al., 1994; Renosto et al., 1990; Bradfield et al., 1970). A homologous positive regulator controlling a host of S assimilation genes has been identified for Neurospora crassa (CYS3; Marzluf and Metzenberg, 1968), Aspergillus nidulans (METR; Natorff et al., 2003), and Saccharomyces cerevisiae (MET4; Thomas et al., 1992). Repression of sulfate assimilation by the addition of a reduced S source such as met is mediated by controlling the activity of the positive regulatory protein through transcriptional regulation (Marzluf, 1997) or ubiquitination (Kuras et al., 2002). Additionally, SAT is strongly inhibited by the product of the second enzymatic step, 3’-phosphoadenosine-5’-phosphosulfate (PAPS), in A. nidulans, Penicillium chrysogenum, Penicillium duponti, and N. crassa 100 (Renosto et al., 1990), and sulfate uptake is completely suppressed by an order of magnitude lower concentration of met in S. cerevisiae (Breton & Surdin-Kelj an, 1977). The results presented here show that germinating spores of Glomus intraradices can assimilate S from sulfate, cysteine (cys) and methionine (met). The addition of cys reduced sulfate assimilation, but less than it does in other fungi, and with diminishing potency after its addition. Met addition had little effect on the utilization of S from sulfate 'or cys, but increased the expression of putative genes of the sulfate assimilatory pathway. In contrast to labeling data, putative genes encoding CT pathway enzymes were up- regulated by more than four fold in response to met, suggesting the induction of the pathway from met to cys. In contrast to other fungi, assimilation of reduced and oxidized forms of S was simultaneous and non-preferential. Exposure to reduced glutathione (GSH) resulted in a reduction in sulfate uptake similar to cys addition and changes in gene expression similar to met addition, whereas buthionine sulfoximine, a specific inhibitor of GSH synthesis, increased sulfate utilization and had no effect on gene expression. The results demonstrate that presymbiotically growing G. intraradices assimilates S and possesses a complex regulatory system for S metabolism with distinctive features compared to all other fungi studied to date. 101 Experimental Procedures Tissue culturing Standardized germination cultures were prepared under sterile conditions by blending spores (Shenk & Smith; DAOM 181602) with solidified Gel-gro agar (MP Biomedicals, Solon, OH) and transferring 2 mL into each well of a sterile 12-well polystyrene plate. Glomus intraradices spores purchased from Premier Tech Biotechnologies (Reviere-du-Loup, Québec, Canada) were used throughout. Prior to labeling, spore cultures contained only Gel-gro, spores, 2 mM CaClz, and 2 mM MES buffer (pH 6.0). This process resulted in consistent sample weights and a uniform spore density of approximately 2500 spores mL-l. MES contains S, however its addition did not alter radio-labeled sulfate uptake (not shown). Germination plates were stored at 4°C for a minimum of one week before use. All experiments were conducted using germination plates unless, as in the 3 hr time point tested (Fig. 3-1) and the germination of tissue for gene expression analysis (Fig. 3-5), the timing of tissue collection was critical for the success of the experiment. For experiments where germination plates were impractical, spores were germinated in liquid media. Radio-labeled sulfate (N a2[3SSO4]; MP Biomedicals) was supplied to spores with 0.1 mM Na2SO4 with two exceptions. The examination of 35S-labeling from amino acids and sulfate combined (Fig. 3-3) was conducted with 0.1 mM total S, each source comprising 0.05 mM S. Also, the measurement of uptake at three hours from various sulfate concentrations (Fig. 3-1) utilized 0.01, 0.014, 0.022, 0.04, 0.1 or 0.2 mM NaZSO4. 102 Labeled cys and met (L-[3SS]cysteine and L-[35S]methionine; MP Biomedicals) were supplied with either 0.1 mM (Fig. 3-3) or 0.05 mM (Fig. 3-4) of the corresponding non- labeled amino acid. In all experiments conducted filter-sterilized stock solutions were added as a 1 mL solution to each well and spores germinated in a 30°C incubator. Samples in blended Gel- gro were collected by first transferring the contents of a well to a plastic bottle with 100 mL of 10 mM sodium citrate buffer (pH 6.0) to disintegrate the gellan. Surface radioactivity was removed by a sieving procedure described previously (Allen & Shachar-Hill, 2009). Briefly, 30 mL of cold 10 mM Na2804 was allowed to flow through the sample. This was repeated three times followed by 30 mL of cold deionized water. Samples were immediately frozen in liquid nitrogen, lyophilized, and weighed. In the case of 35S-labeled samples not germinated in Gel-gro (Fig. 3-1), the sodium citrate buffer was replaced by 100mL of cold MiliQ water. Spores germinated for the measurement of gene expression were only washed with approximately 10 mL of MiliQ water over a sieve followed by immediate collection and freezing in liquid nitrogen. The collection of 3"SS-labeled spores and mycelia lasted approximately five minutes while the collection of material for gene expression measurements took less than one minute. Samples for gene expression analysis were stored in a -80°C freezer until extracted. Extraction and fractionation procedure The relative allocation of 353 was determined by the crude separation of the cellular components according to chemical solubility (Allen & Shachar-Hill, 2009). In 103 this way, fractions containing, but not limited to, sulfate, the free amino acids cysteine and methionine, proteins, and sulfonated lipids were isolated from one another. Two 3 mm stainless steel beads were added to each dried sample in 1.5mL screw-top centrifuge vials along with 0.1 mL of methanol:water solution (70:30). Fungal spore and mycelial structures were disintegrated using a bead mill (Retsch MM301) set at 30Hz for four minutes at 4°C. A representative sample was taken from each milling and examined under a dissecting microscope to ensure complete breakage. Another 0.9 mL of methanol:water solution (70:30) was added and the samples vortexed for five minutes to isolate ionic and small organic compounds including sulfate, sulfite, free amino acids, and glutathione. The samples were centrifuged and supematants collected. A portion of the alcohol extract (0.5 mL) was added to a solution containing Na2S04 and slightly acidified with 0.05 mL of 1 N HCl, followed by the addition of BaClz solution. The final concentrations of sulfate and barium were 10 mM and 11 mM, respectively. Samples were placed in a 100°C heat block for 30min resulting in the precipitation of BaSO4. In control experiments (not shown), this procedure was determined to be more than 99% effective in removing a known amount of radio-labeled sulfate from solution. The comparison between radioactivity levels in the pre- and post-barium precipitation . . 35 . . . . solutions determined the amount of S in the sulfate and ammo ac1d pools. Protelns were extracted from the remaining pellets using Urea/SDS buffer consisting of 9 M urea, 1% SDS, 0.7 M B-mercaptoethanol, 25 mM Tris-HCl (pH 6.8), and 1 mM EDTA (Osherov and May, 1998). To completely solubilize the proteins, samples were heated to 100°C on a heat block for 5 min, vortexed for l min, and heated again for 1 min. After 104 centrifugation at 14,000 g for 5 min, 0.5 mL of the supernatant was analyzed by scintillation counting. The effect on counting efficiency of adding 1 M urea to scintillation fluid (BioSafe 11, MP Biomedicals) was analyzed and found to be negligable (not shown). Pellets were transferred to glass ampules containing 0.5 mL TS-2 tissue solubilizer (MP Biomedicals), sealed over a gas flame, and incubated at 70°C for one week. Scintillation fluid (Bio-Safe II, MP biomedicals) was added directly to the tissue solubilizing solution followed by scintillation counter analysis, and was titrated with l N HCl solution until neutral before scintillation counting. Sample preparation for gene expression analysis Samples were pulverized using a bead mill (Retsch MM301) set to 30 Hz with 100 pL of RNA extraction buffer and two 3 mM stainless steel beads. Care was taken not to allow drawing by milling the samples for 2 min intervals followed by re-freezing in liquid nitrogen for three to four cycles. A small aliquot of two samples were checked under a dissecting microscope for tissue disruption. Total RNA was then extracted using the RNeasy Plant Mini Kit (Qiagen, Valencia, CA) immediately followed by DNA removal using RNase-fiee DNase (Turbo DNA-free; Ambion, Austin, TX). Following quantification of RNA by spectroscopy, an equal concentration of cDNA for each sample was synthesized with SuperScript II reverse transcriptase (Invitrogen corp., Carlsbad, CA) from each sample. Simultaneously, a set of reverse transcriptase free control samples were created. Sequences of gene fragments with high homology to known sulfur assimilatory genes from filamentous filngi and yeast were obtained through a combination of high 105 throughput sequencing data (PJ Lammers, Y Shachar-Hill and coworkers unpublished results) and a published EST database (Jun et al., 2002). These sequence conti gs were verified by standard PCR using cDNA (obtained as already described) derived from spores germinated in water for one week followed by sequencing using an ABI PRISM® 3100 Genetic Analyzer (Applied Biosystems, CA, USA). A QIAquick Gel Extraction Kit (QIAGEN GmbH, Germany) was used to isolate the DNA fragments after separation by electrophoresis. The primer sequences (IDT, Syntegen, Skokie, IL) used and fragment sizes are listed here: putative high affinity sulfate perrnease (482 bp), forward 5’- CGGGATTCCAGATGCACTT-3 ’ , reverse 5 ’ -GGGAAGTTCATTCCATGGTC-3 ’; putative sulfate adenylyl-transferase (902 bp), forward 5’- ATATCGACCATTATACAAGAGTAAGAGTC-3 ’, reverse 5 ’- CGATATATCCGTCCTTTTCGA-3’; putative phosphoadenylyl sulfate reductase (764 bp), forward 5’-ACCCGCGAATCAGTATATCG-3’, reverse 5’- CCCACCTTCCATCTCTTTCA-3’; putative cysteine synthase (1404 bp), forward 5’- AAACCAACAAAGGGACAACG-3 ’, reverse 5 ’-TTTATCCAGTTCCGGCAGTC-3 ’; putative cystathionine B-synthase (605 bp), forward 5’- ATGGAGTCTAATGACAAAGAAAATAATTG-3 ’, reverse 5 ’- GAGAGATTTTGCAACTTGAATTGCTGCC-3’; putative cystathionine y-lyase (922 bp), forward 5’-CCAAATCGAAGTGCATTTGAAGAGTC-3’, reverse 5’- CCTAGCAAATCATCAACATCTTCGATACC-3 ’ . 106 Real time quantitative PCR measurements Gene expression was analyzed using an ABI Prism 7900 HT Sequence Detection System (Applied Biosystems, Austin, TX) using the following cycling program: 50°C for 2 min, 95°C for 10 min followed by 40 cycles of 95°C for 15 sec and 60°C for 1 min. Power SYBR Green 2-Step Master Mix (Applied Biosystems, Austin, TX) was used to produce 0.015 mL samples containing 1 ng of cDNA and 300 nM primers per sample. From these, 0.01 mL was transferred to a 384 well plate for analysis. Gene expression of putative S assimilation genes was measured relative to the expression of an S4 ribosomal protein as described by Govindarajalu et al. (2005) using the AAC1~ (Livak & Schmittgen, 2001) and comparative CT methods. Primer Express software (Applied Biosystems, Austin, TX) was used to create qRT-PCR specific primer sets for all sequences studied. Primer sequences for qRT-PCR are listed here: putative high affinity sulfate perrnease forward 5’-CAATTCTGGGATTTTATGGTCTTTTT-3’, reverse 5’-GATTCCATATTCGACGGTAGTGAA-3’; putative sulfate adenylyl- transferase forward 5’-AGCTTTTGTAGCAGAGCTAAC-3’, reverse 5’- GGCAAACGGAGCAATTGG-3’; putative phosphoadenylyl sulfate reductase forward 5’-AAAATCTTGGAATGGGCTCTT-3’ and reverse 5’- GTGAGGCCGAAAGCAGTTG-3’; putative cystathionine B-synthase forward 5’- CAATTATTGAGCCAACTTCTGGAA-3’, reverse 5’- TCCTTTGATAGCAGCCGCTAA-3’; putative cystathionine y-lyase , forward 5’- AGCAAAGAAACAGGCAAGAGGAT-3 ’, reverse 5 ’- TGCTTCCTTAAAACCACCATTAAT-3’; S4 ribosomal protein forward 5’- 107 AACAGGTGGTAGAAATATGGGAAGA-3’ and reverse 5’- AAGCCGCCTACGTGTCGTT-3 ’. 108 Results Import and assimilation of sulfate by germinating spores and its regulation by S metabolites The uptake and assimilation of exogenous sulfate during presymbiotic growth was measured by supplying radio-labeled sulfate to spores germinated 7 days prior. The rate of import increased linearly (R2 = 0.983) when sulfate was supplied at concentrations ranging from 10-100 pM sulfate (Fig. 3-1). In this range there was also a linear relationship between external sulfate concentration and incorporation into different biochemical fractions: sulfate (R2 = 0.978), amino acids (R2 = 0.964), extractable proteins (R2 = 0.989), and tissue solubilized cell material (R2 = 0.901). In proportion to the total S uptake, incorporation into the amino acid fraction decreased by 5 fold, while intracellular sulfate doubled between 10 and 100 pM external sulfate. There was no firrther increase in uptake or assimilation of S into any fraction with concentrations exceeding 100 pM (Fig. 3-1; data not shown). Import, reduction, and incorporation of 35S from sulfate was continuous for at least 7 days and was affected by some, but not all, sulfur metabolites tested (Fig. 3-2). . . 35 . . 2 . The total incorporation of S over 2, 4, and 7 days increases linearly (R = 0.977) wrth analogous increases in pools containing proteins and aqueous-alcohol soluble S- containing molecules. The uptake of sulfate was reduced by the addition of cysteine (cys) or reduced glutathione (GSH), but not methionine (met). Cysteine initially reduced total 109 35S uptake by 35 i 4% by day 2, however, the effect of cys diminished to a 21 :l: 4% reduction by day 4 and finally a 13 i 4% reduction by day 7. Less than 10% of the cys available in the media was consumed during the experiment. GSH addition and the inhibition of GSH synthesis had the most dramatic effects on sulfate uptake by the spores (Fig. 3-2). Uptake was reduced by 63 i 1% and 64 fl: 3% at days 2 and 4, respectively, when 1 mM GSH was added. At day 7, however, this effect was diminished to the point that a difference in uptake was not statistically significant between the control and GSH samples, similar to the diminished effect of cys addition. Buthionine sulfoximine (BSO) irreversibly inhibits GSH synthesis at the point of glutamate-cysteine ligase (Griffith & Meister, 1979). The effect of adding 10 mM B80 was to increase sulfate uptake, compared to control samples by 30 +/- 2% (day 2), 47 +/- 9% (day 4), and 82 +/- 10% (day 7). 35$ incorporation into the sulfate and reduced sulfur-containing small molecule pools increased most (Fig 3-2). Taken together the effects of GSH and BSO strongly suggest that intracellular GSH concentrations regulate sulfate assimilation. The uptake of sulfate in germinating spores exposed to exogenous met was not statistically significantly different fi'om the uptake by spores germinated in sulfate alone at any time point. 110 1.2 . 1 . E E § l 9 0.8 . 'c» E U) E 0.6 . C .9 '5 0 e- o.4 . O O E D a) c f if" 0.2 . ""‘ }. " * L B A et- Fe 0 - "7“, ‘1’ 1o 14 22 40 100 200 Sulfate Concentration (pM) Figure 3-1 Three hour uptake and incorporation of radio-labeled sulfate by germinating spores. Spores were germinated in an S-free media for five days prior to the addition of 10, 14, 22, 40, 100, 120, 150, 200, or 300 pM sulfate labeled with 353042' at a specific activity of 7.2 pCi pg-1 S. Tissue was collected and counts fractionated into pools containing but not limited to sulfate (white bars), amino acids (light gray bars), proteins (gray bars), and tissue solubilized cell debris (dark gray bars). Error bars depict standard error of the mean (11 = 5). Statistical analyses done by ANOVA single factor analysis, alpha = 0.05. 111 200 . 180- H § 160- + o '7 140 . U) E U) 120- E 9+ c +- ~.. .2 100- c E O 8 0 so ~ 0 5 a) 60 - 4o - 20 - 0 2d4d 7d 2d4d 7d 2d4d 7d 2d4d 7d 2d4d 7d H20 cvs MET GSH 380 Figure 3-2 Radio-labeled sulfate uptake over time with the addition of sulfur metabolites. Spores were germinated for 2 (2d), 4 (4d), or 7 (7d) days with 7.85 pCi of ”$042- in the presence or absence of 1 mM L-cysteine hydrochloride (CYS), 1 mM L-methionine (MET), 1 mM reduced glutathione (GSH), or 10 mM buthionine sulfoximine (BSO), an inhibitor of GSH synthesis. Tissue was collected and counts fractionated into pools containing but not limited to sulfate (white bars), amino acids (light gray bars), proteins (gray bars), and tissue solubilized cell debris (dark gray bars). Error bars depict standard error of the mean (n = 5). Statistical analyses done by ANOVA single factor analysis, alpha = 0.1. 112 Uptake and assimilation of S from cysteine 35S incorporation from labeled cys and the effect of sulfate, met, GSH, BSO or O-acetylserine (OAS) addition was examined (Fig. 3-3) and the statistical significance of the results determined by ANOVA single factor analysis (p-value < 0.05). GSH addition increased (44 i 11 %) while BSO addition decreased (1 l :l: 6 %) total uptake of [3SS]cys compared to controls. No other treatments significantly altered uptake including OAS, a precursor of cys, either with or without added sulfate. None of the compounds added significantly altered the amino acid or protein fractional content derived from labeled cys. The data also shows no significant synthesis of sulfate or sulfite from cys-derived S (Fig. 3-3). Any counts in this fraction are within the range of errors from contamination of the . . 35 . . . . . . . sulfate fi'actlon by organlc S during sulfate precrpltatlon and countlng maccuracres before and after precipitation. Simultaneous uptake of sulfate and reduced S species An experiment was conducted to determine the degree of preference for the importation and assimilation of organic S sources (cys and met) compared to sulfate (Fig. 3-4). The incorporation of 358 was measured when added as 3SS-sulfate or 35S-amino acids in the presence of unlabeled amino acids and sulfate respectively, or when both . . . . . 35 sulfate and an ammo acrd were labeled. This was compared to the incorporatlon of S added as either sulfate or an amino acid alone. When [3SS]cys and ”$042. were 113 introduced to germinating spores together, the resulting labeling was equivalent to the sum of the labeling from either source individually to within the experimental accuracy. Labeled S from [35$]cys was more readily taken up than from sulfate by 48% (Fig. 3-4), while the opposite was true for met, labeling with which resulted in an 18.7% decrease in total 358 uptake compared to 35SO4-2 labeled samples. Overall labeling from 35SO42- was roughly equivalent whether unlabeled cys or met was added, only changing by 5 :l: 2 %. This finding was similar to ”$042. uptake with cys and met addition after 7 days from figure 2, which did not result in a statistically significant difference (2 i 12%). While the overall uptake was similar, when the label was supplied as [35S]cys compared to 358042., there was a 13 fold increase in the total counts discovered in the solubilized fraction (Fig. 3-4), and an increase in this fraction was also seen in the [35S]CYS/35SO4 samples. Labeled met increased the solubilized fraction by 2.4 fold compared to labeled sulfate. The reason for this fractional increase is not clear, however it shows that inter- conversion of cys and met is limited and suggests that the insoluble fraction contains cys or a derivative thereof. 114 Figure 3-3 Combined uptake of sulfate and reduced S species cys and met. Equal amounts (0. 05 mM) of sodium sulfate and either cys or met were added to 3germinating spores5 with labeling supplied as 8. 8 pCi Na32[3 5SO4] (CSYS/3 5804; MET/3 S04), 8. 2 pCi [3 5S]cys ([3 SSJCYS/SO4), or 8.5 pCi [3 5S]met ([3 5S]MET/SO4). The uptake and incorporation of S was compared between additions of single labeling and double labeling with both Na2[35SO4] and [35S]cys ([3531cvsfssoa), or Na7_[35SO4] and [35S]met ([35S]MET/35SO4). The resulting labeling was chemically separated after 7 days of germination into fractions containing but not limited to sulfate (white bars), amino acids (light gray bars), proteins (medium gray bars), and tissue solubilized residue (dark gray bars) and analyzed by scintillation counting. Error is expressed as standard error of the mean (11 = 4). The total uptake was significantly different for all conditions by ANOVA single factor analysis (alpha = 0.05). 115 Figure 3-3 vowcepmzacm eomcmzfimm «omentmz «Omanamerowmm Swab—mam ”.8365 .2. (MCI L.Bul Bu) uogrelodlooug s 116 The expression of putative S metabolism and transport genes An expressed gene sequence was identified by 454 sequencing of cDNA from extraradical mycelium that has close similarity to 3’-phosphoadenylsulfate reductase (PAPSR). This sequence was amplified from presymbiotic cDNA by PCR and sequenced to confirm its identity. PAPSR catalyzes the first reduction step converting sulfate to sulfite (see introduction). A BLAST search using the nonredundant database at NCBI revealed one conserved PAPSR-specific domain (chl 713; E-value = 5e-33). The protein sequence shares 54% identity with PAPSR from a fungal plant pathogen, Pyrenophora tritici-repentis (XP_001937780; E-value = 9e-70), 51% identity with MET16 (PAPSR) from Aspergillus nidulans (XP_662374; E-value = 4e-65), and 50% identity with PAPSR from Neurospora crassa (XP_964870; E-value = 2e-60), and MET16 from Sacchaomyces cerevisiae (NP_015493; E-value = 4e-59). The analyses of putative sequences encoding a high affinity sulfate perrnease (SUL), sulfate-adenylyl transferase (SAT), cystathionine B-synthase (CBL), and cystathionine y-lyase (CGL) were previously reported (Allen & Shachar-Hill, 2009). Exposure to cys led to both a reduction in SUL expression of 63 +/- 9% after a 6hr induction period (Fig. 3-5) and a reduction of total 358 uptake of 35 i 4% after two days of presymbiotic labeling (Fig. 3-2). In contrast, addition of either met or GSH to germinating spores resulted in increased gene expression for SUL, SAT, and PAPSR, and greatly increased expression of CBS and CGL. CBS expression was increased 3.5 :t 0.8 fold by met and 4.4 :l: 0.8 fold by GSH addition, whereas CGL expression was increased 4.4 i 0.8 fold by met and 4.9 :t 0.2 fold by GSH addition. Although BSO greatly increased uptake of 35SO42- (Fig. 3-2), it had little effect on the expression of any gene 117 analyzed. Changes in the expression of putative gene sequences encoding a high affinity sulfate perrnease (SUL) and the first and third steps of sulfate assimilation, sulfate- adenylyltransferase (SAT), and 3’-phosphoadenylsulfate reductase (PAPSR), were not biologically significant (between 0.5 and 2 fold) for cys, met, or GSH treated samples (Fig. 3-5). 118 s Incorporation (119 mg’1 DW) H20 8042' MET GSH BSO OAS OASISO42' Figure 3-4 [35S]Cysteine uptake and utilization in the presence of common sulfur metabolites. S ores were germinated in a 30°C incubator for 7 days in the presence of 5.25 pCi of [3 S]cys and 0.1 mM sulfate (S04), 1 mM methionine (MET), 1 mM reduced glutathione (GSH), 10 mM buthionine sulfoximine (BSO), 1 mM O-acetyl serine (OAS), 1 mM OAS with 0.1 mM sulfate (OAS/S), or nothing else (H20). Samples were collected and the S species separated by chemical fractionation yielding fractions containing but not limited to sulfate (white bars), amino acids (light gray bars), and proteins (dark gray bars), analyzed by scintillation counting. Error is expressed as standard error (n=3). Stars signify statistical differences (total uptake) from all other treatments by AN OVA single factor analysis (alpha = 0.1). 119 5 4.5. 4‘ § \ a " S s a 31 s s 2 2.5J § IE 2. § 1.5 \ l s \ 0.54 ‘\ N ‘S Figure 3-5 Real-tirne PCR analysis of changes in sulfur assimilation pathway gene expression in response to sulfur metabolites. Spores of G. intraradices were germinated for five days in 2 mM MES (pH 6.0) prior to induction for six hours with 0.1 mM sodium sulfate (8042-), 1 mM L-cysteine hydrochloride (CYS), 1 mM L-methionine (MET), 1 mM reduced glutathione (GSH), or 10 mM buthionine sulfoximine (BSO). Expression as fold change was calculated relative to control samples labeled with only deionized water. Putative sequences of the S assimilation pathway including a high affinity sulfate perrnease (SUL; white), S-adenosyl transferase (SAT; horizontal lines), phosphoadenylsulfate reductase (PAPSR; vertical lines), cytathionine B-synthase (CBS; gray), and cystathionine y-lyase (CGL; down-slashed lines) were analyzed. Values are reported as mean :t SEM (n = 3 biological replicates). 120 Discussion Germinating spores of Glomus intraradices remain metabolically active for approximately two weeks in the absence of host roots before germ tube cytoplasm retreats and spores regain dormancy (Smith & Read, 2008). Exogenously supplied phosphorus, nitrogen, acetate, amino acids, and glucose'are imported if available (Bucking et al., 2008; Lammers et al., 2001; Thomson et al., 1990; Hepper & Jakobsen, 1983; Bago et al., 1999; Gachomo et al., 2009), however little is currently known of the function of this uptake capacity. Imported nutrients and root secreted signaling compounds can affect extension and branching of the germ tube, as well as their metabolic capability (Hepper & J akobsen, 1983; Besserer et al., 2008; Tamasloukht et al., 2003; 2007; Nagahashi et al., 1996). Sulfur is a macronutrient required in quantities similar to phosphorus and its assimilation is regulated similarly among other fimgal species (Marzluf, 1997). Exogenously supplied sulfate is readily imported, reduced, and incorporated by germinating spores of G. intraradices, producing stable proportions of S incorporation into proteins and amino acids. This proportional relationship was unaffected by sulfate concentration (Fig. 3-1), the application of S amino acids, GSH or an inhibitor of GSH synthesis (Fig. 3-2), or total sulfate uptake. These data show that the relative rates of sulfate reduction and incorporation into proteins are constant, and but that sulfate uptake is less closely regulated relative to assimilation. Sulfate assimilation is an energy intensive process and other fiingi control the assimilation of different S forms through strict transcriptional and allosteric regulation of 121 sulfate uptake and reduction so as to preferentially use reduced forms of S when available (Foster et al., 1994; Renosto et al., 1990; Lowe, 1991). In S. cerevisiae for instance, exposure to 0.75 mM met results in a complete absence of sulfate perrnease activity (Breton & Surdin-Kerjan, 1977). In contrast both organic and inorganic S are imported and assimilated by spores of Glomus intraradices throughout germination and germ tube extension with no detectable preference for reduced forms of S (Fig. 3-4). The addition of met and its uptake by presymbiotic G. intraradices had no effect on sulfate uptake. In S. pombe, which lacks the enzymes needed to convert met to cys (Brzywczy et al., 2002), met addition also fails to regulate sulfate uptake, presumably because the demand for cys cannot be met by met. Cysteine initially suppressed uptake by G. intraradices during spore germination although much less than in other fungi, and even this effect diminished during development. One week of exposure to cys as the sole S source led to S incorporation comparable to that from sulfate (Fig. 3-3), so the diminishing effect of cys on sulfate uptake is likely not due to a reduction in the import of cys. The simultaneous uptake and assimilation of sulfate and S amino acids was confirmed for cys and met (Fig. 3-4), and to our knowledge represents the only such case among fungi. Interestingly, the relative rates of incorporation of S from sulfate into amino acid and protein pools are not altered by the simultaneous uptake and incorporation of S from reduced sources (Figs. 3- 2, 3-4), suggesting that assimilation of reduced and oxidized forms of S are regulated separately. Regulation of sulfate perrnease expression controls the flux through the S assimilation pathway in fungi (For a review, see Marzluf 1997). The transcription of genes encoding a high affinity sulfate perrnease in N. crassa, and high and low affinity 122 sulfate perrneases from A. nidulans, P. chrysogenum, and S. cerevisiae is completely suppressed by the availability of reduced S or high concentrations of sulfate (Cherest et al., 1997; Pilsyk et al., 2007 ; Van de Kamp et al., 2000; Ketter and Marzluf, 1988; Natorff et al., 2003). In A. nidulans, a mutant lacking a functional S-adenosyl transferase (SAT) accumulate up to 40 mM sulfate demonstrating that the reduction of sulfate is required for permease regulation (Bradfield et al., 1970), and sulfate perrneases in this species as well as N. crassa are known to be transcriptionally regulated by sulfur metabolite repression mediated through cys (Pilsyk et al., 2007; Ketter et al., 1991). In G. intraradices, the expression of a putative sulfate permease (SUL) was suppressed by the presence of cys. The reduction in SUL expression by cys agreed with the reduction in uptake of sulfate. This is consistent with data on this sequence previously published in which addition of 1 mM cys decreased both uptake of sulfate and expression of SUL by roughly half in the extraradical mycelium during the symbiotic phase of the life cycle (Allen & Shachar-Hill, 2009). However, met addition, while having no effect on sulfate uptake, doubled the expression of SUL. Sulfate uptake and SUL expression were also differently affected when spores were exposed to GSH or BSO (Figs. 3-2, 3-5), suggesting that changes in SUL transcript levels are not the primary level at which SO42- uptake is regulated in this fungus. There was no increase in the uptake of sulfate when it was supplied in concentrations higher than 0.1 mM, which agrees with the presence of a single high affinity transport system (Fig. 3-1). Single high and low affinity sulfate perrnease genes have been identified in the genomes of S. cerevisiae, N. crassa, and P. chrysogenum (Cherest et al., 1997 ; Ketter et al., 1991; Van de Kamp et al, 2000), and only in A. 123 nidulans has a third sulfate transporter been characterized (Pilsyk et al., 2007). This third transporter is distinct from other known sulfate perrneases, belonging to a separate family of ion perrneases than the highly conserved SulP family in which other sulfate perrneases, and SUL, are part of (Pilsyk et al., 2007). The genome of G. intraradices is smaller than that of other filamentous fungi and from the large amount of unpublished genomic and EST sequence available, there is no evidence that G. intraradices possesses more isoforms of other genes than do other fungi. It is therefore unlikely that there is another high affinity sulfate transporter in the G. intraradices genome, and it seems more probable that the changes in uptake observed are due to post-translational modifications or alloteric controls on perrnease activity. The S assimilation pathway enzymes sulfate adenosyltransferase (SAT) and phosphoadenosyl phosphosulfate reductase (PAPSR), catalyzing the first and third steps of sulfate reduction, are transcriptionally regulated in A. nidulans (Natorff et al., 2003), and S. cerevisiae (Thomas et al., 1990; 1992) by sulfur metabolite repression. In S. cerevisiae, the addition of 2 mM met to the media lowers transcription of the SAT gene by ten fold (Cherest et al., 1985), and 1 mM met completely suppresses the expression of the gene encoding PAPSR (Thomas et al., 1990). In A. nidulans, expression of the gene encoding SAT is strongly repressed by 5 mM met addition (Borges-Walmsley et al., 1995). The expression of putative sequences encoding SAT and PAPSR from G. intraradices germinating spores was different than that of other filngal SAT’s. Spore SAT and PAPSR shared the expression profile of SUL (Fig. 3-5), and it is likely that these three sequences are regulated coordinately as SAT and PAPSR genes are in S. cerevisiae and A. nidulans (Marzluf, 1997). The putative SAT sequence shared the 124 greatest homology with SAT from P. chrysogenum, including a cys residue implicated in the allosteric control of the P. chrysogenum protein (Renosto et al., 1990). Additionally. the G. intraradices putative SAT sequence has homology with a PAPS-binding region in SAT genes from P. chrysogenum, A. nidulans, and A. terreus (Clarke et al., 1997; Schierova et al., 2000). In Aspergillus terreus, the activity of SAT is reduced by 90% after the addition of 1 mM met (Schierova et al., 2000). The addition of met to germinating spores of G. intraradices doubled the expression of SUL, SAT, and PAPSR without changing the uptake or assimilation of sulfate (Figs. 3-2. 3-5). The conversion of met to cys requires the enzymes cystathionine beta synthase (CBS) and cystathionine gamma lyase (CGL), constituting the CT pathway. This capacity varies among fungal species and its absence is associated with a lack of effect of met on transcriptional regulation. In contrast to cys, met addition had no effect on sulfate assimilation in G. intraradices, suggesting an absence of the CT pathway (Fig. 3-2). The presence of putative sequences with homology to CBS and CGL and their expression profiles, however, contradicted this hypothesis, and the expression of these putative genes differs from other fungi. In A. nidulans CBS is not transcriptionally regulated by cys or sulfate (Sienko & Paszewski, 1999) whereas cys suppressed transcript levels in G. intraradices twofold In S. cerevisiae, CGL transcript levels are reduced 15 fold by met addition (Ono et al., 1993) while the G. intraradices transcript level was unaffected by sulfate or cys addition, and exposure to met increased it by more than 4 fold (Fig. 3-5). Glutathione addition yielded similar results to met exposure, although the effects of GSH and met on sulfate and cysuptake were dissimilar (Figs. 3-2, 3-3). Changes in expression of CBS and CGL were, like SUL, not reflected by the 35S uptake data. In S. cerevisiae, 125 CGL transcription is coordinately regulated with a high affinity GSH transporter (GSHl 1) (Hiraishi et al., 2008) through a cis-element linked to the TOR pathway (Miyake et al., 2003), which coordinates cell growth with grth factors and nutrient availability (reviewed by Dann & Thomas, 2006). It is possible that in G. intraradices the CT pathway enzymes are also regulated by more than one signal. While the assimilation of sulfate in relation to reduced S has been well studied, reduced S assimilation itself has not. The addition of precursors of cys synthesis in other firngi, met and O-acetyl serine (OAS), failed to affect incorporation of cys-S by G. intraradices (Fig. 3-3). At least one strain of S. cerevisiae is unable to import OAS (Ono et al., 1996). The failure of OAS addition, with or without sulfate, to alter assimilation of S from cys may also be due to an inability of G. intraradices to import the substance, as OAS also failed to alter sulfate assimilation (unpublished data). Although mct addition up-regulated putative CT pathway genes, it also had no effect on cys-S acquisition, suggesting that either met uptake is reduced in the presence of cys, or that met conversion through the CT pathway is low. Consistent with this possibility, [3SS]cys increased the . . . . . . . 35 .. proportion of radloactlvrty in the insoluble S fractions much more than [ S]met (Flg. 3- 4). In a reverse of their effects on sulfate uptake (Fig. 3-2), GSH increased while BSO decreased cys uptake (Fi g. 3-3), again suggesting that the intracellular concentration of GSH is an important regulator of S uptake and assimilation. Cys-S incorporation was otherwise essentially constitutive. Glutathione is present in millimolar quantities in fungal cells (Penninckx and Jaspers, 1982; Pocsi et al., 2004) and is considered a reservoir of reduced S as its catabolism produces free cys (Miyake et al., 1999). In S. cerevisiae and S. pombe, 126 exogenous GSH produces changes in gene expression and enzymatic activity similar to the addition of cys (Kumar et al., 2003; Wheeler et al., 2002; Brzywczy et al., 2002). Both GSH and cys addition to germinating spores of G. intraradices had a suppressive effect on sulfate uptake which diminished over time (Fig. 3-2), yet had opposite effects on gene regulation (Fig. 3-5). Additionally, glutathione increased the assimilation of cys- S (Fig. 3-3), which, together with the gene expression results, is indicative that GSH effects are not the result of the catabolic breakdown to cys. In support of this, an inhibitor of GSH synthesis, buthionine sulfoximine (BSO), nearly doubled sulfate assimilation while having no effect on the transcript levels of putative S assimilation genes (Fi g. 3-5). Thus it seems that the assimilation of sulfate is negatively correlated with intracellular GSH levels. This should be tested by measuring the intracellular concentration of GSH. In fungi, GSH has a broad range of effects on cellular processes (Pocsi et al., 2004), one example of which is the regulation of vacuolar H+-ATPase by redox state in barley leaves and the bacterium Dictyostelium discoideum (Tavakoli et al., 2001; J eong et al., 2006). Vacuolar and plasma membrane H+-ATPases are also coordinately regulated in S. cerevisiae in response to cytoplasmic pH (Fernandez & Sa-Correia, 2001), which has the potential to impact the activity of sulfate perrnease, a proton syrnporter. Summary The addition of cys led to decreased uptake of radio-labeled sulfate and decreased transcription of putative genes encoding a high affinity sulfate perrnease, CT pathway, sulfate assimilatory pathway enzymes (Fig. 3-5). Unlike other fungi studied with the exception of S. pombe, met addition failed to change sulfate uptake or the transcription of 127 related genes (Fig. 3-2). In S. pombe, the failure of met addition to elicit a regulatory response is due to the lack of a CT pathway converting met to cys, and the observation agrees with the hypothesis that cys is the principal regulatory molecule in filngi (Brzywczy et al., 2002; Schierova et al., 2000). G. intraradices germinating spores, however, possess putative transcripts homologous to the CT pathway genes, CBS and CGL, and these transcripts are up-regulated by met addition suggesting functionality (Fig. 3-5). Cysteine addition had moderate effects in G. intraradices, reducing transcription by two fold compared to similar studies in other fungi demonstrating much more extreme reductions of five to ten fold (Thomas et al., 1990; Cherest et al., 1985; Borges-Walrnsley et al., 1995). Aside from cys effects, sulfate assimilation in G. intraradices germinating spores was independent of transcriptional regulation (Fig. 3-5). These observations, along with the apparent up-regulation of putative S assimilation genes by GSH (Fig. 3-5), demonstrate that G. intrardices germinating spores are unique in the regulation of S assimilation among filngi studied. 128 References Allen JW, Shachar-Hill Y. 2009. Sulfilr transfer through an arbuscular mycorrhiza. Plant Physiology 149(1): 549-560. Bago B, Pfeffer PE, Douds DD, Brouillette J, Becard G, Shachar-Hill Y. 1999. Carbon metabolism in spores of the arbuscular mycorrhizal fungus Glomus intraradices as revealed by nuclear magnetic resonance spectroscopy. Plant Physiology 121(1): 263- 271. Besserer A, Becard G, J auneau A, Roux C, Sejalon-Delmas N. 2008. GR24, a synthetic analog of strigolactones, stimulates the mitosis and growth of the arbuscular mycorrhizal fungus Gigaspora rosea by boosting its energy metabolism. Plant Physiology 148(1): 402-413. Borgeswalmsley MI, Turner G, Bailey AM, Brown J, Lehmbeck J, Clausen IG. 1995. Isolation and characterization of genes for sulfate activation and reduction in Aspergillus-nidulans - implications for evolution of an allosteric control region by gene duplication. Molecular & General Genetics 247(4): 423-429. BradfieI.G, Somerfie.P, Meyn T, Holby M, Babcock D, Bradley D, Segel IH. 1970. Regulation of sulfate transport in filamentous fungi. Plant Physiology 46(5): 720-727. Breton A, Surdinkerjan Y. 1977. Sulfate uptake in Saccharomyces-cerevisiae - biochemical and genetic study. Journal of Bacteriology 132(1): 224—232. Brzywczy J, Natorff R, Sienko M, Paszewski A. 2007. Multiple fungal enzymes possess cysteine synthase activity in vitro. Research in Microbiology 158(5): 428-436. Brzywczy J, Paszewski A. 1993. Role of O-acetylhomoserine sulflrydrylase in sulfur amino-acid synthesis in various yeasts. Yeast 9(12): 1335-1342. Brzywczy J, Sienko M, Kucharska A, Paszewski A. 2002. Sulphur amino acid synthesis in Schizosaccharomyces pombe represents a specific variant of sulphur metabolism in filngi. Yeast 19(1): 29-35. Bucking H, Abuhaker J, Govindarajulu M, Tala M, Pfeffer PE, Nagahashi G, Lammers P, Shachar-Hill Y. 2008. Root exudates stimulate the uptake and metabolism of organic carbon in germinating spores of Glomus intraradices. New Phytologist 180(3): 684-695. Carr GR, Hinkley MA, Letacon F, Hepper CM, Jones MGK, Thomas E. 1985. Improved hyphal growth of 2 species of vesicular arbuscular mycorrhizal fungi in the presence of suspension-cultured plant-cells. New Phytologist 101(3): 417-426. 129 Cherest H, Davidian JC, Thomas D, Benes V, Ansorge W, SurdinKerjan Y. 1997. Molecular characterization of two high affinity sulfate transporters in Saccharomyces cerevisiae. Genetics 145(3): 627-635. Cherest H, Surdinkerjan Y. 1992. Genetic-analysis of a new mutation conferring cysteine auxotrophy in Saccharomyces-cerevisiae - updating of the sulfur metabolism pathway. Genetics 130(1): 51-58. Cherest H, Thao NN, Surdinkerjan Y. 1985. Transcriptional regulation of the MET3- gene of Saccharomyces-cerevisiae. Gene 34(2-3): 269-281. Clarke DL, Newbert RW, Turner G. 1997. Cloning and characterisation of the adenosyl phosphosulphate kinase gene from Aspergillus nidulans. Current Genetics 32(6): 408-412. Cooper KM, Tinker PB. 1978. Translocation and transfer of nutrients in vesicular- arbuscular mycorrhizas .2. Uptake and translocation of phosphorus, zinc and sulfur. New Phytologist 81(1): 43-47. Dann SG, Thomas G. 2006. The amino acid sensitive TOR pathway fi'om yeast to mammals. Febs Letters 580(12): 2821-2829. Dotzler N, Krings M, Taylor TN, Agerer R. 2006. Germination shields in Scutellospora (Glomeromycota : Diversisporales, Gigasporaceae) fi'om the 400 million- year-old Rhynie chert. Mycological Progress 5(3): 178-184. Douds DD, Nagahashi G, Pfeffer PE, Reider C, Kayser WM. 2006. On-farrn production of AM fungus inoculum in mixtures of compost and vermiculite. Bioresource Technology 97(6): 809-818. Fernandes AR, Sa-Correia I. 2001. The activity of plasma membrane H+-ATPase is strongly stimulated during Saccharomyces cerevisiae adaptation to growth under high copper stress, accompanying intracellular acidification. Yeast 18(6): 511-521. Foster BA, Thomas SM, Mahr JA, Renosto F, Patel HC, Segel IH. 1994. Cloning and sequencing of ATP sulfurylase from Penicillium-chrysogenum - identification of a likely allosteric domain. Journal of Biological Chemistry 269(31): 19777-19786. Fu YH, Paietta JV, Mannix DG, Marzluf GA. 1989. CYS-3, the positive-acting sulfur regulatory gene of Neurospora-crassa, encodes a protein with a putative leucine zipper DNA-binding element. Molecular and Cellular Biology 9(3): 1120-1 127. Gachomo E, Allen JW, Pfeffer PE, Govindarajulu M, Douds DD, Jin H, Nagahashi G, Lammers PJ, Shachar—Hill Y, Bucking H. 2009. Germinating spores of Glomus intraradices can use internal and exogenous nitrogen sources for de novo biosynthesis of amino acids. New Phytologist. doi: 10.111l/j.l469-8137.2009.02968.x. 130 Gahoonia TS, Raza S, Nielsen NE. 1994. Phosphorus depletion in the rhizosphere as influenced by soil-moisture. Plant and Soil 159(2): 213-218. Griffith OW, Meister A. 1979. Potent and specific-inhibition of glutathione synthesis by buthionine sulfoximine (s-normal-butyl homocysteine sulfoximine). Journal of Biological Chemistry 254(16): 7558-7560. Hepper CM, Jakobsen I. 1983. Hyphal growth from spores of the mycorrhizal fungus Glomus-caledonius - effect of amino-acids. Soil Biology & Biochemistry 15(1): 55-58. Hepper CM, Warner A. 1983. Role of organic-matter in grth of a vesicular- arbuscular mycorrhizal fungus in soil. Transactions of the British Mycological Society 81(AUG): 155-156. Hepper CM. 1983. Limited independent growth of a vesicular-arbuscular mycorrhizal firngus invitro. New Phytologist 93(4): 537-542. Hepper CM. 1984. Inorganic sulfur nutrition of the vesicular arbuscular mycorrhizal fungus Glomus-caledonium. Soil Biology & Biochemistry 16(6): 669-671. Hiraishi H, Miyake T, Ono B. 2008. Transcriptional regulation of Saccharomyces cerevisiae CYS3 encoding cystathionine gamma-lyase. Current Genetics 53(4): 225-234. Jacobson ES, Metzenberg RL. 1977. Control of arylsulfatase in a serine auxotroph of Neurospora. Journal of Bacteriology 130(3): 1397-1398. Jeong SY, Choi CH, Kim J S, Park SJ, Kang SO. 2006. T hioredoxin reductase is required for grth and regulates entry into culmination of Dictyostelium discoideum. Molecular Microbiology 61(6): 1443-1456. Jun J, J A, Rehrer C, Pfeffer PE, Shachar—Hill Y, Lammers PJ. 2002. Expression in an arbuscular mycorrhizal firngus of genes putatively involved in metabolism, transport, the cytoskeleton and the cell cycle. Plant and Soil 244: 141-148. Ketter J S, Jarai G, Fu YH, Marzluf GA. 1991. Nucleotide-sequence, messenger-ma stability, and dna recognition elements of cys-l4, the structural gene for sulfate perrnease-ii in neurospora—crassa. Biochemistry 30(7): 1780-1787. Ketter JS, Marzluf GA. 1988. Molecular-cloning and analysis of the regulation of cys- 14+, a structural gene of the sulfur regulatory circuit of neurospora-crassa. Molecular and Cellular Biology 8(4): 1504-1508. Kumar C, Sharma R, Bachhawat AK. 2003. Utilization of glutathione as an exogenous sulfur source is independent of gamma-glutamyl transpeptidase in the yeast Saccharomyces cerevisiae: evidence for an alternative gluathione degradation pathway. F ems Microbiology Letters 219(2): 187-194. 131 Kuras L, Rouillon A, Lee T, Barbey R, Tyers M, Thomas D. 2002. Dual regulation of the Met4 transcription factor by ubiquitin-dependent degradation and inhibition of promoter recruitment. Molecular Cell 10(1): 69-80. Lammers PJ, Jun J, Abuhaker J, Arreola R, Gopalan A, Bago B, Hernandez- Sebastia C, Allen JW, Douds DD, Pfeffer PE, Shachar—Hill Y. 2001. The glyoxylate cycle in an arbuscular mycorrhizal fungus. Carbon flux and gene expression. Plant Physiology 127(3): 1287-1298. Lewandowska M, Paszewski A. 1987. Regulation of sulfur arnino-acids metabolic enzymes in Cephalosporium-acremonium strains differing in antibiotic production. Acta Microbiologica Polonica 36(1-2): 39-45. Li XL, George E, Marschner H. 1991. Extension of the phosphorus depletion zone in VA-mycorrhizal white clover in a calcareous soil. Plant and Soil 136(1): 41-48. Livak KJ, Schmittgen TD. 2001. Analysis of relative gene expression data using real- time quantitative PCR and the 2(-Delta Delta C(T)). Methods 25: 402-408. Lowe G. 1991. Mechanisms of sulfate activation and transfer. Philosophical Transactions of the Royal Society of London Series B-Biological Sciences 332(1263): 141-148. Marzluf GA, Metzenberg RI. 1968. Positive control by CYS-3 locus in regulation of sulfur metabolism in Neurospora. Journal of Molecular Biology 33(2): 423-437. Marzluf GA. 1997. Molecular genetics of sulfilr assimilation in filamentous fungi and yeast. Annual Review of Microbiology 51: 73-96. Miyake T, Hiraishi H, Sammoto H, Ono BI. 2003. Involvement of the VDE homing endonuclease and raparnycin in regulation of the Saccharomyces cerevisiae GSHll gene encoding the high affinity glutathione transporter. Journal of Biological Chemistry 278(41): 39632-39636. Miyake T, Sammoto H, Kanayama M, Tomochika K, Shinoda S, Ono B. 1999. Role of the sulphate assimilation pathway in utilization of glutathione as a sulphur source by Saccharomyces cerevisiae. Yeast 15(14): 1449-1457. Morzycka E, Paszewski A. 1979. 2 pathways of cysteine biosynthesis in Saccharomycopsis-lipolytica. F ebs Letters 101(1): 97-100. Nagahashi G, Douds DD, Abney GD. 1996. Phosphorus amendment inhibits hyphal branching of the VAM fungus Gigaspora margarita directly and indirectly through its effect on root exudation. Mycorrhiza 6(5): 403-408. 132 Natorff R, Sienko M, Brzywczy J, Paszewski A. 2003. The Aspergillus nidulans metR gene encodes a bZIP protein which activates transcription of sulphur metabolism genes. Molecular Microbiology 49(4): 1081-1094. Ono B, Hazu T, Yoshida S, Kawato T, Shinoda S, Brszczy J, Paszewski A. 1999. Cysteine biosynthesis in Saccharomyces cerevisiae: a new outlook on pathway and regulation. Yeast 15(13): 1365-1375. Ono B, Kijima K, Ishii N, Kawato T, Matsuda A, Paszewski A, Shinoda S. 1996. Regulation of sulphate assimilation in Saccharomyces cerevisiae. Yeast 12(11): 1153- 1 162. Ono Bl, Ishii N, Naito K, Miyoshi SI, Shinoda S, Yamamoto S, Ohmori S. 1993. Cystathionine gamma-lyase of Saccharomyces-cerevisiae - structural gene and cystathionine gamma-synthase activity. Yeast 9(4): 389-397. Ono BI, Naito K, Shirahige Y], Yamamoto M. 1991. Regulation of cystathionine gamma-lyase in Saccharomyces-cerevisiae. Yeast 7(8): 843-848. Paszewski A, Grabski J. 1974. Regulation of s-amino acids biosynthesis in Aspergillus- nidulans - role of cysteine and-or homocysteine as regulatory effectors. Molecular & General Genetics 132(4): 307-320. Penninckx MJ, Jaspers CJ. 1982. On the role of glutathione in microorganisms. Bulletin De L Institut Pasteur 80(4): 291-301. Pilsyk S, Natorff R, Sienko M, Paszewski A. 2007. Sulfate transport in Aspergillus nidulans: A novel gene encoding alternative sulfate transporter. Fungal Genetics and Biology 44(8): 715-725. Piotrowska M. 1993. Candida-valida - a yeast lacking the reverse transsulphuration ' pathway. Acta Microbiologica Polonica 42(1): 35-39. Pocsi I, Prade RA, Penninckx MJ 2004. Glutathione, altruistic metabolite in fungi. Advances in Microbial Physiologr 49: 1-76. Renosto F, Martin RL, Wailes LM, Daley LA, Segel IH. 1990. Regulation of inorganic sulfate activation in filamentous fungi - allosteric inhibition of atp sulfurylase by 3'- phosphoadenosine-S'-phosphosulfate. Journal of Biological Chemistry 265(18): 10300- 10308. Schierova M, Linka M, Pazoutova S. 2000. Sulfate assimilation in Aspergillus terreus: analysis of genes encoding ATP-sulfurylase and PAPS-reductase. Current Genetics 38(3): 126-131. 133 Sienko M, Paszewski A. 1999. The metG gene of Aspergillus nidulans encoding cystathionine beta-lyase: cloning and analysis. Current Genetics 35(6): 638-646. Smith SE, Read DJ. 2007. Mycorrhizal symbiosis, 3rd edn. London, UK. Tamasloukht M, Sejalon-Delmas N, Kluever A, J auneau A, Roux C, Becard G, Franken P. 2003. Root factors induce mitochondrial-related gene expression and fungal respiration during the developmental switch fiom asymbiosis to presymbiosis in the arbuscular mycorrhizal fungus Gigaspora rosea. Plant Physiology 131(3): 1468-1478. Tamasloukht M, Waschke A, Franken P. 2007. Root exudate-stimulated RNA accumulation in the arbuscular mycorrhizal fungus Gigaspora rosea. Soil Biology & Biochemistry 39(7): 1824-1827. Tavakoli N, Kluge C, Golldack D, Mimura T, Dietz KJ. 2001. Reversible redox control of plant vacuolar H+-ATPase activity is related to disulfide bridge formation in subunit E as well as subunit A. Plant Journal 28(1): 51-59. Thomas D, Barbey R, Surdinkerjan Y. 1990. Gene-enzyme relationship in the sulfate assimilation pathway of Saccharomyces-cerevisiae - study of the 3'- phosphoadenylylsulfate reductase structural gene. Journal of Biological Chemistry 265(26): 15518-15524. Thomas D, Jacquemin I, Surdinkerjan Y. 1992. Met4, a leucine zipper protein, and centromere-binding factor-i are both required for transcriptional activation of sulfur metabolism in Saccharomyces-cerevisiae. Molecular and Cellular Biology 12(4): 1719- 1 727. Thomson BD, Clarkson DT, Brain P. 1990. Kinetics of phosphorus uptake by the germ-tubes of the vesicular arbuscular mycorrhizal fungus, Gigaspora-margarita. New Phytologist 116(4): 647-653. van de Kamp M, Schuurs TA, Vos A, van der Lende TR, Konings WN, Driessen AJM. 2000. Sulfur regulation of the sulfate transporter genes sutA and sutB in Penicillium chrysogenum. Applied and Environmental Microbiology 66(10): 4536-4538. Wang B, Qiu YL. 2006. Phylogenetic distribution and evolution of mycorrhizas in land plants. Mycorrhiza 16(5): 299-363. Wang ZY, Kelly JM, Kovar JL. 2004. In situ dynamics of phosphorus in the rhizosphere solution of five species. Journal of Environmental Quality 33(4): 1387-1392. Wheeler GL, Quinn KA, Perrone G, Dawes IW, Grant CM. 2002. Glutathione regulates the expression of gamma-glutamylcysteine synthetase via the Met4 transcription factor. Molecular Microbiology 46(2): 545-556. 134 Zhou D, White RH. 1991. Transsulfuration in archaebacteria. Journal of Bacteriology 173(10): 3250-3251. 135 Chapter 4 The Relationship between N and S Assimilation in Germinating Spores of Glomus intraradices 136 Abstract The relationship between the assimilation of two macronutrients, N and S, was analyzed in germinating spores of the arbuscular mycorrhizal fungi Glomus intraradices . . 35 2- . by measuring the effect of varlous N sources on S04 uptake, sequestratlon, and assimilation into amino acid and protein pools. Additionally, 15N-labeling of free amino acids by GC-MS and HPLC based amino acid concentration analyses were used to determine the effect of various S sources on N assimilation. Changes in putative S uptake and assimilation genes associated with exogenous ammonia availability were measured . . . . . . 35 usrng quantitative real-tlme PCR and the results analyzed in relation to S-uptake data. All N sources tested increased sulfate uptake by a minimum of two fold. Ammonia and . . 2- urea led to the greatest increases in 804 uptake, and were also preferred N sources as . . . . 15 . reflected by lncreased free amino ac1d concentratlons and N-labellng compared to supplementation with nitrate, Gln, and Arg. Ammonia addition failed to increase sulfate uptake in the presence of reduced glutathione (GSH), which also halved the concentrations of amino acids related to N assimilation and doubled that of Leu, which is a regulator of S assimilation in other organisms. Cysteine, methionone, and 0- acetylserine addition failed to counteract the effect of armnonia addition. When S was supplied to germinating spores as cysteine, ammonia addition had no effect on assimilation in the presence of absence of GSH, whereas GSH increased assimilation by 30%. The transcription of putative high affinity sulfate perrnease, sulfate adenosyl transferase, phosphoadenosylephosphosulfate reductase, cystathionine beta-synthase, and 137 cystathionine gamma-lyase was analyzed in relation to labeling data and found to be unrelated. While N assimilation is not altered by S addition, S assimilation is increased with N addition in proportion to general increases in amino acid concentrations from de novo synthesis. 138 Introduction In fungi, N assimilation is tightly controlled on a transcriptional level and involves a global positive acting transcriptional factor and a host of associated pathway specific regulatory proteins (Marzluf, 1993). Orthologs of the global transcriptional factor responsible for the amino acid control circuit (gC) in yeasts and cross pathway control circuits (CPC) in filamentous fungi have been characterized in Neurospora crassa (Nit2; Chiang & Marzluf, 1995), Aspergillus nidulans (AreA/NirA; Okamoto et al., 1993) and Saccharomyces cerevisiae (Gcn4/Gln3; Sosa et al., 2003; Roussou et al. , 1988). Additionally, orthologs have been identified in the basidiomycete biotrophic plant pathogen Ustilago maydis (McCann & Snetselaar, 2008), the ascomycete ectomycorrhizal truffle Tuber borchii (TbNrel; Guescini et al., 2009), the yeast Hansula polymorpha (Siverio et al., 2002), as well as several other Aspergilli (Braus et al., 2006), which show a high degree of conservation. Amino acid control in S. cerevisiae is initiated by the binding of a sensor kinase, Gcn2p, to uncharged tRNAs, resulting in a phosphorylation cascade, the general reduction of the translation initiation rate, and the increase in the transcription of Gcn4-controlled sequences (Dong et al., 2000). Both AreA and Nit2 contain putative glutamine (gln) binding sites, disruption of which results in a loss of sensitivity to N repression (Marzluf, 1993). Intracellular gln concentrations in A. nidulans are inversely related to the activity of AreAp, suggesting that the cross pathway control regulatory system senses gln (Berger et al., 2008). There are direct and indirect associations between N and S assimilation on both the enzymatic and transcriptional level. The Gcn4p/Nit2 proteins directly or indirectly 139 regulate the expression of hundreds of genes (Tian et al, 2007; Natarajan et al., 2001), including genes involved with S amino acid synthesis and sulfate reduction pathways. In S. cerevisiae, several of these genes have known Gcn4 promoter sequences including methionine synthase, sulfite reductase, 3’-phospho-5’-adenylylsulfate (PAPS) reductase, and the positive transcriptional activator Met4 (Thomas & Surdin-Kelj an, 1997). Interestingly, the only known transcriptional regulator of Met4, which is not regulated ny S source, is Gcn4p (Mountain et al., 1993). Additionally, an analysis of conserved direct targets of Gcn4 and orthologs in S. cerevisiae, C. albicans, and N. crassa using genomic and transcriptomic data by Tian and co-workers (2007) revealed ten directly regulated genes, one of which is part of the transsulfuration pathway converting cysteine to methionine, cystathionine B-synthase. 0n the level of metabolism, the synthesis of met is directly related to threonine (thr) and indirectly related to aspartate (asp) and branched chain amino acid synthesis (Thomas & Surdin-Kerjan, 1997). General amino acid control over transcription is known to be required for life cycle completion and survival in several fungi. The expression of AreA/Nit2/Gcn4 are required for complete virulence of A. funigatus in immuno-suppressed mice, morphogenesis in C. albicans, and cell-cell or cell-surface adhesion in S. cerevisiae (Braus et al., 2006; Yin et al., 2004; Braus et al., 2003; Kleinschmidt et al., 2005). The TbNrel gene is also specifically upregulated in the T uber borchii symbiotic state, demonstrating a possible role for cross pathway control in a mycorrhizal symbiosis (Guescini et al., 2009). The widespread use of the endomycorrhizal symbiosis in agriculture is currently restricted by difficulties obtaining large scale inoculum, a consequence of their obligatory symbiotic nature. The study of nutrient regulation may 140 uncover the limitations on the transition from the pre-symbiotic to symbiotic state in arbuscular mycorrhizal fungi, enabling their large scale agricultural use. In Glomus intraradices pre-gerrninating spores both N and S are imported and utilized, though exogenous sources are unnecessary for germination (Gachomo et al, 2009; Smith & Read, 2008). The present study examined the inter-relationship between N and S metabolism in arbuscular mycorrhizal fimgi using germinating spores of Glomus . . 35 . intraradices as a model system. Labeling With S revealed a dramatic response of increased sulfate uptake and assimilation when N sources, especially ammonia, were ' supplied. The addition of glutathione (GSH) counteracted the effects of ammonia on sulfate uptake whereas cys and met had no suppressive effects. The addition of a specific inhibitor of GSH synthesis, buthionine sulfoximine increased sulfate uptake with and without available ammonia. The concentrations of free amino acids related to ammonia assimilation were halved when GSH was supplied exogenously. The data suggests that intracellular GSH concentrations supercede the affects of N source availability on sulfate acquisition by altering the assimilation of ammonia. Gene expression analysis of putative sulfate assimilation genes failed to correlate with sulfate uptake data, and transcriptional regulation therefore does not explain the control of sulfate assimilation by ammonia. Finally, a model is presented in interpretation of the results of the study and discussed. 141 Experimental Procedures Amino acid quantification Approximately 25,000 spores (Premier Tech Biotechnologies, Quebec, Canada) were germinated per sample at 30°C in sterile 12 well petri plates containing lmL of M medium (St.Amaud et al., 1996) modified to remove N (Govindarajalu et al., 2005) and sucrose. Four samples were analyzed for each condition for each experiment (Figures 4- 1, 4-2, and 4-8). Additional to the modified M medium, MilliQ water, 2 mM urea, 4 mM NH4C1, 4 mM KNO3, 2 mM L-glutamine, or 2 mM L-arginine hydrochloride (Si gma- Aldrich, St. Louis, M0) were added as 0.05 mL of filtered stock solutions to study the effect of exogenous nitrogen sources (Figure 4-2). Germinating spores were collected at 0, 7, or 14 days (Figure 4-1) by sieving and immediate freezen in liquid nitrogen, lyophilized and weighed. Measurements of the effects of GSH addition on amino acid concentration were made by adding 2 mM MES (pH 6.0) with or without 1 mM GSH. Free amino acids were extracted by first disintegrating spores and germ tubes in 0.1 mL of a 0.01 N HCl solution using a bead mill (Retsch MM301) set to 30 Hz and two 3 mM stainless steel beads for 4 min. The complete disintegration was confirmed by analyzing small aliquots of five randomly chosen samples under a dissecting microscope. An additional 0.4 mL of 0.01 N HCl was added followed by 10 min of vortexing. Approximately 2500 pmols of norleucine was added as an internal standard prior to extraction. Chloroform (0.5 mL) was added followed by 5 min of further vortexing, after which the chloroform, aqueous solution, and cell debris were separated by centrifugation. The supematants were collected and lyophilized. Dried samples were re-dissolved in 0.02 142 mL of 0.02 N HCl , 0.06 ml of 0.1 M sodium borate buffer (pH 10.0), and 0.02 mL of AccQ-FluorTM reagent (6-aminoquinolyl-N-hydroxysuccinimidylcarbarnate; Waters, Milford, MA, USA). Free amino acids were analyzed from aliquots of 0.01 or 0.02 mL of each sample by HPLC-fluorescence using a Waters 2695 Alliance HPLC (Waters, Milford, MA) equipped with a 37°C column oven, a Waters 2474 Fluorescence Detector, and Empower software (Waters, Milford, MA). A four buffer system was used for HPLC: A and B, sodium acetate buffer (Waters, Milford, MA) adjusted to pH 5.7 and pH 6.8 with phosphoric acid; C, acetonitrile; and D, methanol:water (1:9). 35S-Iabeling experiments Approximately 5x103 sterile spores of Glomus intraradices (Premier Tech Biotechnologies, Reviere-du-Loup, Québec, Canada) per sample were washed on a 300 mesh sieve, collected, and blended together with cold solidified sterile Gel-gro agar (MP Biomedicals, Solon, OH) using a Waring blender to form a slurry of uniform spore density, and aliquots of 2 mL were transferred to the wells of sterile 12 well polystyrene plates. The germination slurry contained only 2 mM MES buffer (pH 6.0), 2 mM CaClz, and spores, and were maintained at 4°C for at least one week prior to labeling with either . 35 . 35 . . . 16.5 pCi of Na2[ S04] or 9.3 pCl of L-[ S]cystelne (MP Biomedicals, Solon, OH). Labeled sulfate was supplied with 0.1 mM NaZSO4, and 35S-cysteine with 0.1. mM unlabeled cysteine as carrier molecules. Additionally, 1 mM O-acetylserine, 1 mM reduced glutathione, 1 mM cysteine, 1 mM methionine, or 10 mM buthionine sulfoximine were added with or without 4 mM NH4Cl (Figures 4-5 and 4-6). For the 143 analysis of N source effects on sulfate uptake (Figure 4-3), N was supplied as 4 mM KN03, 2 mM L-glutamine, 2 mM L-arginine, 2 mM urea, or 4 mM NH4C1. Spores were germinated for seven days after labeling in a 30°C incubator. Samples were collected and S-containing compounds extracted and fractionated into different biochemical pools exactly as described in Chapter 3. 15N-labeling experiments . . . 15 . 15 Experiments exarnrnmg percent N-labelrng used 4 mM NH4C1 or 4 mM K15N03 as N sources in Gel-gro gellan (MP Biomedicals, Solon, OH), 2 mM CaClz, and 2 mM MES (pH 6.0) as previously described (Chapter 3). All sources of 15N were 98% labeled. Approximately 1x104 sterile spores of Glomus intraradices (Premier Tech Biotechnologies, Reviere-du-Loup, Québec, Canada) were germinated per sample in a 30°C incubator and in the presence of 15N label prior to collection after one week. Additionally, 0.1 mM Na2S04, 1 mM L-cysteine hydrochloride, 1 mM reduced glutathione, or 10 mM buthionine sulfoximine were added at the time of labeling. Samples were collected by adding them to 100 mL of cold 10 mM sodium citrate buffer (pH 6.0) and sieving the solution two minutes later on a 300 mesh wire sieve, followed immediately by liquid N2 freezing and lyophilization. Amino acids were extracted by pulverizing each sample in 0.1 mL of 0.01 M HCl solution with two 3 mm stainless steel beads using a bead mill (Retsch MM301) oscilating at 30Hz for four minutes (Allen and Shachar-Hill, 2009). The amino acid pool was extracted by the addition of 0.4 mL of 0.01 144 M HCl followed by moderate vortexing for 10 min. Proteins and non-polar compounds were removed by the addition of 0.5 mL of CHC13 followed by 2 min of moderate vortexing. Samples were centrifilged at 14,000 g for 10 min and the supematants collected. The final solutions were dried under gaseous N2 streams, heated to 60°C on a heat block. The dried amino acid extracts were reconstituted in 0.1 mL of MilliQ water, transferred to GC-MS vial inserts, and lyophilized in preparation of derivatization. Amino acid samples were derivatized in pyridine: N-tert-butyldimethylsilyl- trifluoroacetamide (MTBSTFA) (1:1) for 1 hr at 55°C prior and l pL injected at 280°C using an Agilent 7683 series injector (Santa Clara, CA, USA) into an Agilent GC-MS system (6890N Netwrok GC System and 5973 inert Mass Selective Detector) equipped with a J &W Scientific 30m DB5-ms column (Agilent Technologies, Santa Clara, CA, USA). The oven temperature was maintained at 100°C for 4 min, ramped to 200°C at 5°C/min, then to 300°C at 10°C/min where it was held for 2min. Amino acid standards were used to identify GC peaks and mass fragments. RNA extraction and gene expression measurements RNA was extracted from frozen tissue as described in previous chapters. Briefly, tissue for extraction was oscillated in 1.5 mL screw-cap microcentrifuge tubes at 30 Hz in the presence of 100uL extraction buffer (RN easy Plant Mini Kit, Qiagen, Valencia, CA) and two 3 mm stainless steel beads using a Retsch MM301 bead mill three times for two minutes each time. Five 2 min homogenation steps were done in total and the samples were refrozen in liquid nitrogen between steps. Aliquots of 2 pL were examined microscopically for complete spore breakage. Extractions using an RNeasy Plant Mini 145 Kit (Qiagen, Valencia, CA) were followed by treatment with RNase-free DNase (Turbo DNA-free; Ambion, Austin, TX), after which aliquots were measured by UV absorbance to determine RNA concentration and approximate purity. Equal quantities of RNA were used to synthesize cDNA with SuperScript II reverse transcriptase (Invitro gen corp., Carlsbad, CA), and negative controls were generated with parallel sample sets which did not contain the enzyme. Sequences of putative sulfate assimilation-related genes were obtained by BLAST batch analyses using homologous sequences of conserved regions of each gene of interest as previously described (chapters 2 & 3), and these sequences were verified by PCR amplification and sequencing using an ABI PRISM® 3100 Genetic Analyzer (Applied Biosystems, CA, USA). Primer sequences used (Primer Express software, Applied Biosystems, Austin, TX; [DT, Syntegen, Skokie, IL) are identical to those listed in chapter 3. Gene expression measurements were analyzed as previously reported in chapter 3. 146 Results Free amino acid concentrations during germination Free amino acid (FAA) concentrations were measured at three time points, 0, 7, and 14 days (Figure 4-1). The total N content in the FAA pool decreased by 27 i 4% between 0 and 7 days of germination, 72 i 14% of that amount was due to a fall in asp content. The total N content of the FAA pool did not significantly change significantly between days 7 and 14 by AN OVA single factor analysis (alpha = 0.05). Aspartate represented almost half (41 :l: 0.7%) of the total N in this pool before germination, dropping to 20 :t 2% by day 14. On day 14 the levels of five FAA’s, asn, gly, leu, pro, and gln, had increased and three FAA’s, ala, asp, and ile, had decreased more than two fold compared to the levels at day 0. Fewer than half of the FAA concentrations changed significantly between days 7 and 14 (ANOVA; alpha = 0.05), and only two out of nineteen, glycine and leucine, showed greater than two fold differences. With few exceptions FAA concentrations were stable after 7 days of germination, and experiments were therefore conducted over one week time periods. 147 Figure 4-1. Time course of free amino acid concentration in germinating spores of Glomus intraradices. Spores were germinated in deionized water for 0 (white bars), 7 (down slashed bars), and 14 (gray bars) days prior to the extraction of free amino acids and quantification by HPLC analysis, 11 = 4 i SEM. 148 Figure 4-1 (0 ‘— V' N o co co ‘1- N (MCI l.6ui slowu) uoirelruaouoo pioe 01.1va 149 Ala Arg Asn Asp Cys Gln Glu Gly His Met Ile Leu Lys Phe Pro Ser 1hr Tyr Val Figure 4-2 Free amino acid concentrations in the presence of exogenous N sources. The free amino acid composition of extracts of Glomus intraradices spores germinated for 7 days with no nitrogen source (white bars), 4mM KN O3 (horizontally striped bars), 2mM Gln (light gray bars), 2mM Arg (up slashed bars), 2mM urea (dark gray bars), or 4mM NH4C1 (down slashed bars). The charts represent concentrations (A) above or (B) below 2 nmols mg.1 DW, and (C) the total N concentration under each condition derived from the sum of amino acid concentrations adjusted for the number of N atoms per molecule, n = 4 i SEM. 150 Figure 4-2 Z . 7, ;‘.:‘:::‘:"'""“"’""‘"‘"""""""""::::::§ ~ , “1* ~~~771-r~ r--~--*---~~--~~~.~~~-i RR\\\\\\\\\\‘ .\\\\\\\\\Q~' x\\\\\\ Z _ r Mi :3. \\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ :i;7////////////////4 m l l I I I l I ' IO 0 ID 0 ID 0 m M N N F 1- '0 O < (MCI l_Bru slowu) suoirerrueouoo prov ourruv 151 Asn Asp Gln Glu Gly His Pro Ser Arg Figure 4-2 (cont’d) t-\\\\\\\\\\\\‘ §§§§§§§§§:3I.\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\‘ ‘7/4 - 2\\\\\\\\\\\\\\\\\\2\\\\\\\\‘ {xx/(e pt it k\\\\\\\\\\\\\\\\\\\\\\\\\\\\ I’I/I/A - S\\\\V :§\\\\\\\\\\\\\\\\\\\\\\\‘ 3 5 2 15- 5 In (MCI L_Iiru sioruu) suoireuueouoo prov ouiulv 152 Cys Met Ile Leu Lys Phe Thr Tyr Val Ala ~. '. -.‘.: ...._'-;‘- I ;. .5 3"».3:12.22‘...'?‘-;‘A‘.‘;t."..r.'-.'-_‘_‘.-'__£ 25.41: '2‘" .. ‘l n ': vl '- h. : The effects of N sources on free amino acid concentrations The reduction of sulfate and nitrate are energy intensive processes and the oxidative state of either nutrient may have an effect on the assimilation of the other through the consumption of reducing equivalents. Control of the S assimilation pathway has also been directly linked to general amino acid control in S. cerevisiae (Mountain et al., 1993). Organic (arg and gln), reduced inorganic (NH4+ and urea), or oxidized inorganic (N 03-) N were therefore supplied to germinating spores to determine the plasticity of the FAA content (Figure 4-2). Inorganic N sources had equivalent effects on the FAA pool overall, although there were differences in responses of glu, tyr, his, and pro. The total N content in the FAA pool increased by more than 6 fold compared to control samples with ammonia, urea, or nitrate addition; the response to these N sources was not statistically distinct by ANOVA single factor analysis (Figure 4-2). In fungi, gln concentrations modulate the amino acid control circuit (gC) in yeasts and cross pathway control circuits (CPC) in filamentous fungi by directly binding to a global transcriptional regulatory protein (Marzluf, 1993; Berger et al., 2008), and arg is a putative form of N transfer to the hosts of AM filngal symbioses (Govindarajale et al., 2005). In contrast to the inorganic N sources, supplying arg or gln led to only 2.5 i 0.5 fold or 1.2 i 0.2 fold increases in total N. However, an analysis of percent labeling by 15N revealed that 15N03- labeled amino acids 43 i 5% less on average than comparable amounts of 15 + . . . . . . . . NH4 (Figure 4-3), reveallng that the utllrzatlon of external nitrate is different than that of reduced inorganic N. The internal source of N which makes up the difference in the FAA total N content was not identified. 154 300- D 250- % 200 bl g c E. c 150' + .2 E D §1m- B a B c 0 E U) H20 N03' GLN ARG URE NH4* Figure 4-3 Influence of N sources on 35SO42- uptake, reduction, and allocation. Spores were germinated in the presence of 0.1 mM Na2SO4 and 16.5 pCi of Naz[3SSO4] for seven days. Prior to germination at 30°C, 4 mM KN03 (N03), and NH4C1 (NH4), 2 mM urea (URE), glutamine (GLN), and arginine (ARG) were added. Samples were collected and 3 S separated by chemical fractionation yielding fractions containing but not limited to sulfate (white bars), amino acids (light gray bars), and proteins (dark gray bars), analyzed by scintillation counting. Error is expressed as standard error of the mean from the analysis of five samples per condition, and letters ANOVA single factor analyses (alpha = 0.05) between fractions and total uptake (above white bars). 155 The effects of N sources on sulfate uptake and assimilation The addition of different N sources revealed a general stimulation of S incorporation into all fractions tested, disproportionately increasing sulfate sequestration in some cases. While gln and arg addition did not change the amount of S incorporation into the protein pool compared to nitrate addition, the introduction of urea and ammonia produced significant increases in this S pool (Figure 4-3). Urea increased protein S by 37% more than nitrate, which increased protein S by 2.1 +/- 0.24 fold compared to controls germinated without N. Ammonia addition doubled protein S incorporation compared to nitrate, which caused an increase of the protein fraction of 4.1 +/- 0.24 fold over controls, and also led to a disproportionately large amount of sequestered sulfate (Figure 4-3). The type of N added had the greatest effect on the sulfate fractions. Urea addition led to an increase in the sulfate fraction of 6.1 +/- 0.08 fold compared to 35 . . . . . . controls, % less than ammonia addition. Sulfate fractions from urea and gly addltlons were not statistically different by AN OVA analysis. Other treatments led to a doubling of the sulfate fraction compared to controls and 2-3 fold increases in S incorporation into the protein and amino acid (Figure 4-3). The addition of ammonia, however, led to a disproportionate 9.4 +/- 0.2 fold increase in the sulfate fraction, the largest change measured (Figure 4-3). Compared to the absence of N, nitrate, gln, arg, gly, urea, or ammonia addition led to total increases in S incorporation of 2.31 +/- 0.08, 2.03 +/- 0.21, 2.46 +/- 0.10, 4.61 +/- 0.26, 4.41 +/- 0.08, and 6.66 +/- 0.16 fold, respectively. 156 Figure 4-4 Effect of SO42- (A) and S metabolites (B) on 15N-labeling of free amino acids. (A) Spores were germinated at 30°C for 7 days in the presence of 4 mM K[15NO3] either alone (white bars) or with 0.1 mM Na2S04 (down slashed bars); or with 4 mM [ISNH4]CI either alone (light gray bars) or with 0.1 mM Na2804 (up slashed bars). (B) Spores were germinated at 30°C for 7 days in the presence of 4 mM [ISNH4]C1 and either no S source (white bars), or 0.1 mM NaZSO4 (down slashed bars), 1 mM L- cysteine (light gray bars), 1 mM reduced glutathione (up slashed bars), or 10 mM buthionine 157 Figure 4-4 7/////////////////////////////////////z '6 — :\\\\\\\\\\\\\\\\\\\\\\s > W/////////////////////////////////////////////////A ; l- h m U) o S a. ‘63 E 3 at .1 2b 0 E (D 2 (D :7//////////////////// g ;\\\\\\\\\\\\\\' 0 I. WWW/”WWW; % < 2 < l f T I T I I T I T :'l O o o o o o o o 2 a) 8 l~ to to v to 8 ,_ o < (%) Burlaqe'l N9, 158 Figure 4-4 (cont’d) . . =3€5${{:7////////////////////////// 100 ‘ (%) Burleqe'l N9, 159 Interactions between sulfate and nitrogen assimilation pathways The interaction of N and S metabolism was examined in two experiments. In the first, the effects of S on N metabolism was examined by supplying 15N nitrate or 15N ammonium with or without SO42' or S metabolites and analyzing the 15N-labeling of free amino acids (Figure 4-4). In the second experiment the effect of N on the regulation of S assimilation was examined through an analysis of 35SO42- labeling in different S fractions when S metabolites were added in the presence or absence of ammonia (Figure 4-5). Reduced glutathione altered both the labeling of FAA’s from 15N-ammonia and also the effect of ammonia on sulfate uptake. When supplied with 15N-ammonia, the average labeling after one week of germination was 74.5 d: 12.6% for all amino acids measured (Figure 4-5). Among the compounds examined, only reduced glutathione (GSH) significantly altered percent labeling, reducing it modestly in all amino acids examined. The largest reductions were in val, leu, met, and tyr which were 31.5 i 5.1%, 27.2 :t 10.3%, 22.7 d: 9.5%, and 12.9 i 4.1% less labeled than the average of other treatments, respectively. The experiments revealed little effect of sulfate, cys, or a specific inhibitor of GSH synthesis, buthionine sulfoximine (BSO). The effect of ammonia addition (which stimulates S assimilation, Figure 4-3) was further analyzed by the introduction of 1 mM 0-acetyl serine (oas), the immediate precursor of cysteine biosynthesis in A. nidulans and N. crassa (Marzluf, 1997). S uptake and assimilation was not significantly different when oas was added (Figure 4-4A). The simultaneous addition of oas and 2 mM ammonia reduced S incorporation into sulfate (34.4 +/- 3.3%), amino acids (16.5 +/- 0.5%), and the solubilized fraction (31.6 +/- 4.4%). 160 There was no significant effect of ammonia on S incorporation into the protein fractions (ANOVA; alpha = 0.05). In contrast, the addition of 1 mM GSH sharply reduced total and fractional S incorporation in the presence or absence of ammonia (Figure 4-4B). GSH addition reduced S incorporation by 88.5 +/- 4.5%. Introducing 2 mM ammonia with gsh failed to significantly alter either fractional or total S incorporation compared to GSH alone. Glutathione addition in relation to sulfate assimilation and free amino acid concentrations An inhibitor of GSH synthesis, buthionine sulfoximine (BSO) was used to determine if the effects of GSH addition were mediated via the intracellular GSH concentration (Figure 4-4B). Inhibiting GSH synthesis with 10 mM 880 resulted in significant changes in S incorporation only in the amino acid fraction, which was increased by 3.1 +/- 0.5 fold. The addition of 2 mM ammonia (Figures 4-4A and B) led to less S incorporation than 4 mM (Figure 4-3). Both figures 4A and 4B show a more than 3 fold increase in total S incorporation (3.3 +/- 0.2 fold in both cases) compared to the 6.7 +/- 0.2 fold enhancement shown in Figure 4-3. To examine the effects of N and GSH on the uptake and incorporation of organic S, [3SS]cys was supplied in the presence or absence of ammonia and/or GSH. Sulfate was not added to the germination solution and the 0.1 mM cys added was the only S source available. Ammonia (2 mM) had no effect on incorporation alone or with GSH addition (1 mM). The application of GSH resulted in three changes: a 71 +/- 11% reduction in amino acid fraction radioactive counts, a 2.1 +/- 0.21 fold increase in protein fraction 161 counts, and an increase of 56 +/- 17% in tissue solubilized fraction counts (Figure 4-5). The possibility of conversion of cys to sulfate was also analyzed by measuring the sulfate fractions, which contained less than 1.5% of the total radioactivity under all conditions. This low level is within the error intrinsic to the barium sulfate precipitation procedure. Since glutathione was the only compound tested that affected both ammonia (Figure 4-5) and sulfate (Figure 4-6) assimilation, we analyzed its effects on N metabolism further. Exposure to GSH in the absence of an N source led to decreased concentrations of glu (2.3 i 0.1 fold), gln (1.9 d: 0.2 fold), asp (2 i 0.1 fold), asn (1.8 :I: 0.2 fold), and arg (1.5 i 0.1 fold) compared to germination in pure water, while most of the FAA concentrations measured did not change (Figure 4-7). Aside from ala, which showed a slight increase, the only FAA to increase in concentration was leu, which more than doubled (2.1 :t 0.2 fold) in the presence of GSH. 162 35o . F 300 ‘ 250 ‘ S Incorporation (ng S mg" DW) N- M N- N+ N- N+ N. m N- N+ N- M H20 OAS CYS MET GSH 380 Figure 4-5 35S-sulfate uptake, reduction, and allocation in the presence of ammonia and S metabolites. Spores were germinated for 7 days at 30°C with 0.1 mM Na2S04 labeled with 16.53 pCi of ”$042- per sample and 1 mM O-acetylserine (OAS), 1 mM reduced glutathione (GSH), 1 mM L-cysteine hydrochloride (CYS), 1 mM L-methionine (MET), or 10 mM buthionine sulfoximine (BSO) ill the absence (N-) or presence (N+) of 4 mM ammonium chloride. Tissue was collected and S fractionated into pools containing sulfate (white bars), amino acids (light gray bars), proteins (gray bars), and solubilized cell debris (dark gray bars), n = 4 i SEM. 163 350 - B 300~ E _ o 250 b: E m 200— U! E. C .2 H . g 150 o O. h o O 5 1001 to 50- 0.. H20 NH4“ GSH NH4+IGSH Figure 4-6 Effects of ammonia and glutathione (GSH) on [3SS]Cysteine uptake and utilization. Spores were germinated for 7 days in a 30°C incubator with 0.1 mM L- cysteine labeled with 9.3 pCi of [35S]Cys in the presence of water alone (H20), 4 mM ammonia (NH4+), 1 mM reduced glutathione (GSH), or 1 mM GSH with 4 mM NH4+ (GSH/NH4+). Samples were collected and the S species separated by chemical fiactionation yielding fractions containing but not limited to sulfate (white bars), amino acids (light gray bars), proteins (medium gray bars), and tissue solubilized cellular debris (dark gray bars), analyzed by scintillation counting, n = 6 :1: SEM. 164 Figure 4-7 Change in amino acid concentrations from glutathione addition. The free amino acid composition of spores germinated for 7 days in water (gray bars) or with 1 mM reduced glutathione (up slashed bars) were evaluated by HPLC, n = 4 i SEM. 165 It llifi ll if. f Figure 4-7 ('0 N ‘- o (MCI L.tSur slowu) uolrelrueouoo prov oulwv 166 Ile Leu Lys Phe Pro Ser Thr Tyr Val Ala Arg Asn Asp Glu Gln Gly His Figure 4-8 Transcriptional changes of putative sulfur assimilation pathway genes after NH4+ addition. Spores of G. intraradices were germinated for five days in 2 mM MES (pH 6.0) prior to induction for six hours with 4 mM NH4C1 (up slashed bars) or a combination of 4 mM NH4C1 and 0.1 mM Na2S04 (down slashed bars). Expression was calculated relative to control samples given only deionized water and reported as fold change. Putative sequences of the S assimilation pathway including a high affinity sulfate perrnease (SUL), S-adenosyl transferase (SAT), phosphoadenylsulfate reductase (PAPSR), cytathionine B-synthase (CBS), and cystathionine y-lyase (CGL) were analyzed, 11 = 3 i SEM. 167 Figure 4-8 5 r / f /////// Z V 7 ’/ . efiueuo Pl0:I 168 SAT PAPSR CBS CGL SUL Changes in gene expression of putative sulfate assimilation genes in relation to nitrogen addition Genes from Glomus intraradices germinating spores were identified from 454 sequencing data based on homology to fungal sequences encoding a high affinity sulfate transporter (SUL), sulfate adenylytransferase (SAT), 3’-phosphoadenylsulfate reductase (PAPSR), and the enzymes making up the reverse transsulfuration pathway (CT pathway) cystathionine B-synthase (CBS) and cystathionine y-lyase (CGL) (see chapters 2 & 3). The expression of these genes was analyzed after a six hour induction by the addition of ammonia to spores germinated for seven days prior (Figure 4-8). Changes in expression more than two fold greater or lower than two fold lesser than control samples were considered biologically significant. None of the sequences were biologically significantly down-regulated, and only CBS, which controls the condensation of homocysteine and serine to form the intermediate cystathionine, was significantly up-regulated (3 :t 0.58 fold). The second step of the pathway converting cystathionine to cysteine and or- ketobutyric acid governed by CGL was not up-regulated (1 i 0.23 fold). The addition of sulfate with ammonia had the opposite effect on the expression of CBS, down-regulating it by 60 a: 4% (0.41 :t 0.04 fold; Figure 4-8), and also halved expression of CGL (0.45 :l: 0.13 fold). The high affinity sulfate perrnease putative sequence tested (SUL) was generally down-regulated by ammonia addition, and the fold change from sulfate addition with the ammonia was not statistically different (ANOVA; alpha = 0.05) from the addition of ammonia alone (Figure 4-8). Expression of SAT and PAPSR was not statistically significantly different from 1 as measured by ANOVA single factor analysis (alpha = 0.05), with or without sulfate addition. In general there were few changes in 169 gene expression of the sequences tested due to ammonia addition, and the reduction of expression from ammonia/sulfate addition does not clearly correlate to the increase in sulfate uptake and assimilation under these conditions (Figure 4-3). 170 Discussion During spore germination exogenous nitrogen sources were readily taken up and used to synthesize amino acids; they also increased the uptake, reduction and incorporation of exogenous sulfate. Stable isotope labeling and amino acid concentration data demonstrated a preference for ammonia and urea, and both of these reduced N sources also led to larger increases sulfate uptake than the organic and oxidized inorganic N sources (Figure 4-3). Urea is catabolized to, and nitrate is reduced to, ammonia before assimilation, and their effects are likely reflective of the efficiency of their uptake and conversion to ammonia. The results of these experiments suggest a correlation between the amount of N assimilated and the uptake of sulfate. The addition of organic N as gln or arg, however, seems to be in opposition to a direct correlation between N and S assimilation, as neither increased the total N content of the amino acid pool as much as nitrate addition (Figure 4-2A), and both led to similar increases in sulfate assimilation compared to nitrate addition (Figure 4-3). Additionally, gln is imported by germinating spores of G. intraradices at ten times the rate of arg (Gachomo et al., 2009). Glutamine is known to have a regulatory role in fungal gene expression. The global positive acting transcriptional factors controlling cross pathway control circuits (CPC) in A. nidulans (AreA) and N. crassa (N it2) have gln binding sites, and in A. nidulans the concentration of gin negatively correlates with the activity of AreA (Marzluf, 1993; Berger et al., 2008). The concentration of gln in the FAA pool, however, does not correlate with the amount of sulfate uptake, although there is a slight increase from urea or ammonia assimilation (Figure 4-2A). The data is therefore suggestive, but inconclusive, that gln l7l and arg addition led to a regulatory effect on sulfate assimilation unrelated to both the total N in the FAA pool and the concentration of specific amino acids. In Aspergillus nidulans, a result of biotin deficiency is an increase in ammonia assimilation, which leads to a 70% increase in protein content as well as higher concentrations of glu, asp, leu, and met (Desai, 1979). Data presented in this study are consistent with the hypothesis that ammonia availability influences protein synthesis in Glomus intraradices as well. As the largest N pool, the rate of protein synthesis or degradation likely determines the intracellular concentration of F AA’s. If these rates . . . . 15 . . . remain static, an increase in % N labeling of an ammo ac1d from an exogenous N source . . . . . . . . . 15 + wrll lead to an equivalent relative increase 1n lts concentratlon. The addltlon of NH4 , however, resulted in 30% more labeling in the free amino acid (FAA) pool than did adding 15N03' (Figure 4-4), while ammonia addition led to an increase of only 12% in the total N in the FAA pool relative to nitrate addition (Figure 4-2C). Since 35S labeling results indicate a nearly twofold increase in the protein-S pool when ammonia is added compared to nitrate, it is likely that ammonia increases protein synthesis (Figure 4-9), leading to less total N in the FAA pool than expected from the 1SN-labeling data. The addition of reduced glutathione (GSH) counteracted the increased sulfate uptake measured when spores were germinated with ammonia whereas the addition of buthionine sulfoximine (BSO), a specific inhibitor of the first enzyme of GSH synthesis, y-glutamylcysteine synthetase, induced sulfate uptake and assimilation (Figure 4-3). These data repeat the finding of the previous chapter on GSH and B30 addition at seven days in the absence of ammonia, as well as the data on cys and met addition (Chapter 3). 172 In contrast to GSH, addition of cys slightly increased ammonia stimulation of sulfate uptake (Figure 4-5). Since cys is readily imported by germinating spores (Figure 4-6), GSH does not appear to affect sulfate uptake by contributing to the cys pool through its catabolism. In S. cerevisiae, GSH is readily imported and utilized as a source of cys (Elskens et al., 1991; Kumar et al., 2003; Miyake et al., 1999). Addition of 10 mM BSO has been shown to reduce the intracellular GSH concentration by 76% in the yeast Kluyveromyces lactis (Coulon et al., 2007) and 70% in S. cerevisiae (Rossi et al., 1997). The intracellular concentration of GSH in G. intraradices was not measured directly, and therefore the effects of GSH and BSO addition on GSH concentration in the germinating spores is uncertain, although it seems plausible that both have their effects either through changes in the free GSH levels or indirectly through their effects on FAA levels. The concentrations of the amino acids that are synthesized from glu, gln and arg either directly or through transarnination reactions involving glu, asp, and asn, were reduced by roughly half with GSH addition (Figure 4-7). Glutamate and asp concentrations have been shown to specifically increase in A. nidulans in response to increased assimilation of ammonia (Desai, 1979), as expected fiom N assimilation through the GS/GoGAT cycle. Both gln and arg addition increased the concentration of glu by more than two fold (Figure 4-2A), however, preliminary evidence suggests that the addition of glu has no effect on sulfate uptake with or without N addition (data not shown). Based on these data, the concentration of glu is likely to be symptomatic of the efficiency of N assimilation into the FAA pool rather than being directly involved in sulfate uptake. Glutathione may therefore effect sulfate assimilation indirectly by reducing ammonia assimilation in this fungus. 173 In contrast to other amino acids measured under GSH exposure, leu accumulated to more than double the concentration found in control samples (Figure 4-5). In S. cerevisiae, genes encoding two enzymes in the leu biosynthetic pathway, isopropylmalate isomerase (LE U1 ) and beta-isopropylmalate dehydrogenase (LE U2) are transcriptionally suppressed by the regulatory protein Leu3p or activated by Leu3p complexed with the pathway intermediate alpha-isopropylmalate (Kohlhaw, 2003). The influence of Leu3p extends to seven genes including GDHI (Kohlhaw, 2003; Hu et al., 1995), which encodes an NADP+-dependent glutamate dehydrogenase, an integral enzyme in the assimilation of ammonia. All of the amino acid concentrations reduced when GSH is present are either synthesized from or transaminated by glu (Figure 4-7). Leucine accumulation is possibly symptomatic of how GSH exposure negates the hyper- accumulation of sulfate when ammonia is added to germinating spores, by affecting leu pathway intermediate concentrations, and is suggestive that a similar regulatory system exists in G. intraradices as S. cerevisiae and N. crassa with regard to Leu3p (Baichwal et al., 1983; Gross, 1965). Additionally in S. cerevisiae, S-adenosylrnethionine transport to the vacuole is greatly reduced by leu (Murphy & Spence, 1972), which may increase the concentration of the cytosolic pool and affect S assimilation. Exposure to BSO leads to a reduction in GSH (Rossi et al., 1997; Coulon et al., 2007), which may leave the germinating spores vulnerable to oxidative damage, as has been shown in S. cerevisiae (Izawa et al., 1995; Grant et al., 1996). Also in S. cerevisiae, transcriptional suppression of the S assimilatory pathway is known to be superseded by . . ++ . . . . . . oxrdatlve stress. Under Cd stress, the ubrqumatlon controlling the actrvrty of the positive regulatory protein met4p is removed, up-regulating S assimilatory pathway 174 enzymes and increasing GSH synthesis (Barbey et al., 2005). This filrther illustrates the possibility that S assimilation is regulated independently by the oxidative state of the germ tube. However, not enough data is presented here to draw any definite conclusions about the mechanism by which BSO exerts its regulatory effects, or about its relationship to N—mediated effects on sulfate assimilation. By contrast with sulfate assimilation, uptake of reduced S in the form of cys was unaffected by ammonia and was increased by GSH addition, suggesting that the assimilation of exogenous cys is not regulated by the same mechanism as sulfate utilization. The increase in cys utilization with GSH addition has been previously demonstrated (Chapter 3), and was repeated here. A single cys transporter was recently shown to be the primary route of entry into S. cerevisiae cells (Kaur & Bachhawat, 2007), and its regulation has not been studied. To date, little other evidence exists concerning cys transport in filngi. Although ammonia addition increased sulfate uptake and assimilation, it failed to alter the transcription of putative sulfate perrnease or primary S assimilatory pathway genes after six hours of induction (Figure 4-8). Ammonia added without sulfate led to a three fold increase of a putative sequence encoding cystathionine B-synthase (CBS), the first step of the reverse trans-sulfuration pathway (CT pathway), however sulfate added in equal quantity as in uptake experiments reduced the expression of this gene by roughly six fold in comparison (chapter 3). The data for gene expression in the presence of both sulfate and ammonia is similar to data obtained from cys addition (Chapter 3), and may indicate the results of a short term increase of cys concentration. The data is clear that six hours of incubation with ammonia and sulfate does not result in transcriptional changes 175 that would account for the stimulation of sulfate uptake. Further analyses are necessary to determine the relationship between N source availability and transcriptional regulation of the S assimilatory pathway at later time points to form meaningful conclusions of the interaction between N and S assimilation pathway regulation. 176 BSO GSH 804 NH4 N03 URE BSO GSH -I® 69 N03 URE GD FAA . > PRO Figure 4-9 Working model of the regulation of sulfate assimilation in germinating spores of Glomus intraradices. Sulfate uptake was increased by exogenous ammonia (NH4) and buthionine sulfoximine (BSO). Nitrate (N 03) and urea (URE) likely affect S metabolism through their conversion to ammonia. Reduced glutathione (GSH) antagonized the stimulatory effect of ammonia on sulfate uptake, reduced ammonia assimilation through glu, and has been previously shown to decrease importation of sulfate. Ammonia also increased both the assimilation of sulfate into free amino acids (FAA) and their incorporation into protein (PRO). There is evidence that ammonia generally increases the flux from the free amino acid to protein pool as well. 177 Sufficient evidence was obtained from this study to make an initial model of the relationship between N assimilation and the regulation of the S assimilatory pathway (Figure 4-9). The addition of N sources increased sulfate uptake, assimilation, and sequestration. Additionally, an analysis of the 15N labeling of the FAA pool in relation to . . . . 35 . . . . the concentration of ammo ac1ds, coupled wrth S-labelrng evrdence, 1ndlcates that ammonia increased the synthesis of proteins. Glutathione had opposing effects on the ability of the ftmgus to use exogenous sources of oxidized and organic S, slightly decreasing sulfate assimilation and increasing uptake of cys. It was the only compound tested which antagonized the stimulation of sulfate acquisition fiom ammonia addition. The addition of BSO increased uptake and assimilation of sulfate alone and in an ammonia background. The data overall reveals a close relationship between N and S assimilation, which may be mediated by FAA concentrations. 178 References Allen JW, Shachar—Hill Y. 2009. Sulfur transfer through an arbuscular mycorrhiza. Plant Physiology 149(1): 549-560. Baichwal VR, Cunningham TS, Gatzek PR, Kohlhaw GB. 1983. Leucine biosynthesis in yeast - identification of 2 genes (leu4, leu5) that affect alpha-isopropylmalate synthase activity and evidence that leul and leu2 gene-expression is controlled by alpha- isopropylmalate and the product of a regulatory gene. Current Genetics 7(5): 369-377. Barbey R, Baudouin-Cornu P, Lee TA, Rouillon A, Zarzov P, Tyers M, Thomas D. 2005. Inducible dissociation of SCFMet30 ubiquitin ligase mediates a rapid transcriptional response to cadmium. Embo Journal 24(3): 521-532. Berger H, Basheer A, Bock S, Reyes-Dominguez Y, Dalik T, Altmann F, Strauss J. 2008. Dissecting individual steps of nitrogen transcription factor cooperation in the Aspergillus nidulans nitrate cluster. Molecular Microbiology 69(6): 1385-1398. Braus GH, Sasse C, Krappmann S. 2006. Amino acid acquisition, cross-pathway control, and virulence in Aspergillus. Medical Mycology 44: S91-S94. Chiang TY, Marzluf GA. 1995. Binding-affinity and functional-significance of nit2 and nit4 binding-sites in the promoter of the highly regulated nit-3 gene, which encodes nitrate reductase in neurospora-crassa. Journal of Bacteriology 177(21): 6093-6099. Coulon J, Matoub L, Dossot M, Marchand S, Bartosz G, Leroy P. 2007. Potential relationship between glutathione metabolism and flocculation in the yeast Kluyveromyces lactis. Fems Yeast Research 7(1): 93-101. Desai JD. 1979. Factors affecting protein-synthesis during biotin deficiency in aspergillus-nidulans. Folia Microbiologica 24(5): 379-&. Dong JS, Qiu HF, Garcia-Barrio M, Anderson J, Hinnebusch AG. 2000. Uncharged tRNA activates GCN2 by displacing the protein kinase moiety from a bipartite tRNA- Binding domain. Molecular Cell 6(2): 269-279. Elskens MT, Jaspers CJ, Penninckx MJ. 1991. Glutathione as an endogenous sulfur source in the yeast Saccharomyces-cerevisiae. Journal of General Microbiology 137: 637-644. Gachomo E, Allen JW, Pfeffer PE, Govindarajulu M, Douds DD, Jin H, Nagahashi G, Lammers PJ, Shachar-Hill Y, Bucking H. 2009. Germinating spores of Glomus intraradices can use internal and exogenous nitrogen sources for de novo biosynthesis of amino acids. New Phytologist. doi: 10.1111/j.1469-8137.2009.02968.x. 179 Govindarajulu M, Pfeffer PE, Jin HR, Abuhaker J, Douds DD, Allen JW, Bucking H, Lammers PJ, Shachar-Hill Y. 2005. Nitrogen transfer in the arbuscular mycorrhizal symbiosis. Nature 435(7043): 819-823. Grant CM, MacIver FH, Dawes IW. 1996. Glutathione is an essential metabolite required for resistance to oxidative stress in the yeast Saccharomyces cerevisiae. Current Genetics 29(6): 51 1-515. Gross SR. 1965. Regulation of synthesis of leucine biosynthetic enzymes in neurospora. Proceedings of the National Academy of Sciences of the United States of America 54(6): 1 53 8-&. Guescini M, Stocchi L, Sisti D, Zeppa S, Polidori E, Ceccaroli P, Saltarelli R, Stocchi V. 2009. Characterization and mRNA expression profile of the TbNrel gene of the ectomycorrhizal fungus Tuber borchii. Current Genetics 55(1): 59-68. Hu YM, Cooper TG, Kohlhaw GB. 1995. The saccharomyces-cerevisiae Leu3 protein activates expression of gdhl , a key gene in nitrogen assimilation. Molecular and Cellular Biology 15(1): 52-57. Izawa S, Inoue Y, Kimura A. 1995. Oxidative stress-response in yeast - effect of glutathione on adaptation to hydro gen-peroxide stress in Saccharomyces-cerevisiae. F ebs Letters 368(1): 73-76. Kaur J, Bachhawat AK. 2007. thlp, a novel, high-affinity, cysteine-specific transporter from the yeast Saccharomyces cerevisiae. Genetics 176(2): 877-890. Kleinschmidt M, Grundmann O, Bluthgen N, Mosch HU, Braus GH. 2005. Transcriptional profiling of Saccharomyces cerevisiae cells under adhesion-inducing conditions. Molecular Genetics and Genomics 273(5): 382-393. Kohlhaw GB. 2003. Leucine biosynthesis in fungi: Entering metabolism through the back door. Microbiology and Molecular Biology Reviews 67(1): 1-+. Kumar C, Sharma R, Bachhawat AK. 2003. Utilization of glutathione as an exogenous sulfur source is independent of gamma-glutamyl transpeptidase in the yeast Saccharomyces cerevisiae: evidence for an alternative gluathione degradation pathway. F ems Microbiology Letters 219(2): 187-194. Livak KJ, Schmittgen TD. 2001. Analysis of relative gene expression data using real- time quantitative PCR and the 2(-Delta Delta C(T)). Methods 25: 402-408. Marzluf GA. 1993. Regulation of sulfur and nitrogen-metabolism in filamentous fungi. Annual Review of Microbiology 47 : 31-55. 180 Marzluf GA. 1997. Molecular genetics of sulfur assimilation in filamentous fungi and yeast. Annual Review of Microbiology 51: 73-96. McCann MP, Snetselaar KM. 2008. A genome-based analysis of amino acid metabolism in the biotrophic plant pathogen Ustilago maydis. Fungal Genetics and Biology 45: S77-S87. Miyake T, Sammoto H, Kanayama M, Tomochika K, Shinoda S, Ono B. 1999. Role of the sulphate assimilation pathway in utilization of glutathione as a sulphur source by Saccharomyces cerevisiae. Yeast 15(14): 1449-1457. Mountain HA, Bystrom AS, Korch C. 1993. The general amino-acid control regulates met4, which encodes a methionine-pathway-specific transcriptional activator of saccharomyces-cerevisiae. Molecular Microbiology 7(2): 215-228. Murphy JT, Spence KD. 1972. Transport of s-adenosylmethionine in Saccharomyces- cerevisiae. Journal of Bacteriology 109(2): 499-5 08. Natarajan K, Meyer MR, Jackson BM, Slade D, Roberts C, Hinnebusch AG, Marton MJ. 2001. Transcriptional profiling shows that Gcn4p is a master regulator of gene expression during amino acid starvation in yeast. Molecular and Cellular Biology 21(13): 4347-4368. Okamoto PM, Garrett RH, Marzluf GA. 1993. Molecular characterization of conventional and new repeat-induced mutants of nit-3, the structural gene that encodes nitrate reductase in neurospora-crassa. Molecular & General Genetics 238(1-2): 81-90. Rossi C, Poli P, Candi A, Buschini A. 1997. Modulation of mitomycin C mutagenicity on Saccharomyces cerevisiae by glutathione, cytochrome p-450, and mitochondria interactions. Mutation Research-Genetic Toxicology and Environmental Mutagenesis 390(1-2): 113-120. Roussou I, T hireos G, Hauge BM. 1988. Transcriptional-translational regulatory circuit in Saccharomyces-cerevisiae which involves the gcn4 transcriptional activator and the gcn2 protein-kinase. Molecular and Cellular Biology 8(5): 2132-2139. Siverio JM. 2002. Assimilation of nitrate by yeasts. F ems Microbiology Reviews 26(3): 277-284. Smith SE, Read DJ. 2007. Mycorrhizal symbiosis, 3rd edn. London, UK. Sosa E, Aranda C, Riego L, Valenzuela L, DeLuna A, Cantu JM, Gonzalez A. 2003. Gcn4 negatively regulates expression of genes subjected to nitrogen catabolite repression. Biochemical and Biophysical Research Communications 310(4): 1175-1180. 181 StArnaud M, Hamel C, Vimard B, Caron M, Fortin JA. 1996. Enhanced hyphal growth and spore production of the arbuscular mycorrhizal filngus Glomus intraradices in an in vitro system in the absence of host roots. Mycological Research 100: 328-332. Thomas D, SurdinKerjan Y. 1997. Metabolism of sulfur amino acids in Saccharomyces cerevisiae. Microbiology and Molecular Biology Reviews 61(4): 503-532. Tian CG, Kasuga T, Sachs MS, Glass NL. 2007. Transcriptional profiling of cross pathway control in Neurospora crassa and comparative analysis of the Gcn4 and CPCl regulons. Eukaryotic Cell 6(6): 1018-1029. Yin ZK, Stead D, Selway L, Walker J, Riba-Garcia I, Mclnerney T, Gaskell S, Oliver SG, Cash P, Brown AJP. 2004. Proteomic response to amino acid starvation in Candida albicans and Saccharomyces cerevisiae. Proteomics 4(8): 2425-2436. 182 Chapter 5 Conclusions and Future Research 183 Introduction Global atmospheric sulfur emissions increased by more than 3000% between 1860 and 1990 (Lefohn et al., 1999), representing approximately 0.0003% of terrestrial plant evolutionary time (Welhnan et al., 2003). In the absence of anthropogenic sources, plants and other organisms evolved complex regulatory mechanisms for the attainment and assimilation of sulfur from the environment. Additionally, terrestrial plant life evolved in conjunction with arbuscular mycorrhizal fungi (AMF) since the Devonian Period (400 MYA; Pirozynsky & Malloch, 1975). The symbiosis assists plant roots in obtaining otherwise unavailable soil nutrient resources (Smith & Read, 2008) and is today found in an estimated 80% of land plants (Wang & Qiu, 2006). Although the mycorrhizal symbiosis is widely considered to be a defining factor in plant phosphorus nutrition (Cress et al., 1979; Murdock et al., 1967), and several studies point towards an important role in plant nitrogen acquisition (Govindaraj alu et al., 2005; Hodge et al., 2001), the role of mycorrhizal symbioses in sulfur acquisition has not been as well characterized, and the assimilation of sulfur by mycorrhizal fungi themselves not analyzed at all. The removal of sulfur from fossil fuel sources prior to combustion has reduced sulfilr emissions in the United States by more than 50% between 1980 to 2000 (Baumgardner et al., 2002), increasingly leading to sulfur deficiency symptoms in crop plants. Since mycorrhizae constitute the dominant soil symbiosis, information of their role in plant S nutrition is important for assessing the impact of the rise and fall of environmental sulfur concentrations. In this thesis, three major aspects of mycorrhizal S acquisition were analyzed using the AMF Glomus intraradices: their role in plant S 184 nutrition, their assimilation of S in the absence of the host, and the regulation of uptake in comparison to other known fungal systems. AMF and plant S nutrition The role of AMF in plant S nutrition has been little and inconclusively studied. A lack of basic measurements make it impossible to even assess whether or not AMF transfer S to plants in physiologically relevant amounts. Therefore, the research outlined in chapter 2 was to provide reliable, basic measurements of S transfer through an AMF. . 35 . . . . These measurements were achieved through S tracer studies usrng a monoxenlc, bl- compartmental petri dish system containing a mycorrhiza established between Glomus intradarices and transformed Daucus carota roots (StAmaud et al., 1986). This system allows the fungal mycelium to separate itself from the colonized roots, and enables the isolated labeling of fungus and host. Measuring in parallel experiments the uptake of 35SO42- by mycorrhizal roots and the transfer of ”$042. from the fungal sides, the transfer of S through the symbiosis was quantified in a non-growth limiting sulfate . . . 35 . . concentration 1n the context of dlrect root uptake. The amount of S supplied to roots in this condition amounted to 25% of total root uptake during the labeling period. The percentage of a nutrient obtained from mycorrhizal transfer compared to direct uptake has only been measured in one other study. Mycorrhizal transfer of P has been measured for flax, medic, and tomato plants, and constitutes close to 100% of the total plant uptake during the labeling period (Smith et al., 2003), leading the authors to speculate that AMF 185 directly affect direct root uptake of P. Smith an co-workers (2003) used pot- grown plants, which generally obtain higher percentages of mycorrhzial root length, whereas colonization in the monoxenic split plate system is normally much lower. Given this, whether or not AMF directly affect the ability of roots to import sulfate is inconclusive based on the data obtained as much of the direct 35S uptake was likely by non- mycorrhizal root sections. Additionally, whether effects on direct root uptake mediated by AMF are systemic or localized to colonized root sections, or if photosynthetic tissue has a role, are all unknowns. However, the 25% of total uptake measured is clearly of physiological importance. Additional to this measurement, the functional importance of fungal S supply to the plant was also assessed by tacking 35$ transferred through the .AMF to roots grown in limiting or non-limiting S concentrations. Transfer under non- limiting conditions was halved compared to low S concentrations, suggesting that the transfer amount is relative to root S status and not constitutive. This implies that AMF may play a functional role in the acquisition of S for plant hosts. Having assessed the regulation of sulfate transfer through an AMF from the perspective of the plant S status, a further goal of the research outlined in chapter 2 was to assess the regulation of uptake by the fimgus itself. To achieve this, S metabolites known to have regulatory effects on sulfate assimilation in other fungi were exogenously supplied to isolated mycorrhizal mycelium and 358042- transfer tracked. Methionine (met) and cysteine (cys) reduced transfer by 26% and 45%, respectively, over a one month labeling period. Also due to cys addition, quantitative real-time PCR measurements of G. intarardices extra-radical mycelium revealed a roughly four fold 186 decrease in the expression of a putative high affinity sulfate permease gene, suggesting this may be the cause of the lower transfer rate for labeled sulfate. Uptake of ”$042. by the fungus was not significantly reduced by met and was halved by cys addition, in contrast to other fungi. Equivalent concentrations of cys or met completely suppress the expression of sulfate assimilation genes in S. cerevisiae, A. nidulans, and N. crassa (Kuras & Thomas, 1995; Breton & Surdin Keljan, 1977; Ketter & Marzluf, 1988; Ketter et al., 1991; Grynberg et al., 2001). Sulfate was shown to be transferred through an AMF symbiosis in physiologically significant quantities which are inversely related to sulfate availability to host roots. Additionally, the transfer of reduced S sources cysteine and methionine was demonstrated, which could provide plants an alternate option to their dependency on re- mineralized sulfate from organic material by microorganisms in the soil. Interestingly, cysteine, and to a lesser extent methionine, made available to the extra-radical fungal mycelium reduced transfer of sulfate to the roots. Since these amino acids are transferred in a reduced state to host roots, whether or not this is due to regulation by the fungus was investigated by measuring gene expression changes from exogenous cys application. There was a roughly four fold reduction in mRNA concentration for a putative high affinity sulfate perrnease gene corresponding to the reduction in uptake and transfer of sulfate. Other putative pathway genes were not down-regulated significantly, suggesting that control of sulfate transfer to the host may be on the level of extra-radical mycelial uptake. 187 Uptake and assimilation of S by AMF The use of pre-symbiotic fungal mycelium for the study of the sulfate assimilation pathway reduces the time needed for an experiment from a minimum of two months to one week as well as the variability inherent to split plate studies due to differences in colonization time and extent within the roots. It is also necessary to remove the influence of the host roots on the regulation of fungal uptake and transfer to assess fungal-specific regulation. The study of S metabolism in pre-symbiotic AMF is therefore an important component in understanding the regulation of S transfer in the symbiotic phase. To this end, two lines of research were conducted using germinating spores: the regulation of S assimilation and the interaction between N and S regulation. Regulation of sulfate assimilation and the synthesis of cysteine is well studied in S. cerevisiae, A. nidulans, and N. crassa (for a review see Marzluf, 1997). Symbiotic transfer of nutrients, however, requires uptake of those nutrients when uptake by other fungi would likely be suppressed. Since fungal regulatory systems for S assimilation and metabolism are well conserved in fungi, differences between AMF and other fungal systems are possibly specific to the symbiosis and may therefore provide valuable information concerning the transition from pro-symbiotic to symbiotic fungal states. The ability of germinating spores to import sulfate and the fractionation profiles of the incorporation of imported sulfate into various metabolic pools was analyzed and the results compared with data from other fungi. Exposure to unlabeled and labeled cys and met, glutathione (GSH) and buthionine sulfoximine (BSO) was analyzed. In contrast to other fungi, Glomus intraradices did not show a preference for reduced or oxidized S, but 188 imported an assimilated both sources simultaneously and to a relatively equal extent. The addition of met failed to reduce sulfate uptake at all, and cys reduces uptake by a third at the most, far less suppression than in other fungi (Grynberg et al., 2001; Marzluf, 1997). The symbiotic nature of AMF may reduce the sensitivity of the regulatory system for sulfate assimilation. Additionally, based on the data it is likely that the intracellular concentration of GSH is negatively correlated with the importation of sulfate, as GSH addition initially reduced uptake by more than two thirds, while the addition of B80, a specific inhibitor of GSH synthesis, more than doubled it. Although GSH is considered to be a reservoir for reduced S in fungi (Miyake et al., 1999), increased incorporation from BSO addition suggests that the effect of GSH/BSO is not due to GSH catabolism. In support of this argument, the exogenous addition of GSH does not decrease %15N labeling of cys from ammonia, and actually increases [35$]cys uptake by about a third. Vacuolar H+-ATPase in S. cerevisiae is regulated by redox state (Fernandez et al., 2003), altering transport of S metabolites to and from the vacuole. This may provide an explanation for the effect of GSH/BSO, however the full determination of this aspect of S assimilation in AMF is beyond the scope of this thesis. The assimilation of N and S are linked through a set of orthologous positive acting global transcriptional factors controlling the amino acid control circuit (gC) in yeasts, or the cross pathway control circuits (CPC) in filamentous fungi (Chiang & Marzluf, 1995; Sosa et al., 2003; Okamoto et al., 1993). In S. cerevisiae, expression of the S regulatory protein, Met4p, is not controlled by S metabolite concentrations, but through the gC (Mountain et al., 1993). The addition of various sources of N uniformly led to a minimum two fold increase in sulfate uptake. Ammonia addition had the most 189 pronounced effect of the N sources, increasing sulfate uptake by six fold. This was also the most pronounced effect of any exogenously supplied substance, suggesting that the assimilation of N, and subsequent demand for S-amino acids, controls sulfate assimilation to a greater extent than S metabolite availability in this fungus. 111 support of this, the combinatory effect of adding S metabolites and ammonia was equivalent to the addition of ammonia alone with the exception of GSH, which completely negated the ammonia effect. An analysis of free amino acid (FAA) concentrations and %15N labeling in relation to N addition demonstrated a likely positive correlation between ammonia availability and protein synthesis. This is in agreement with the hypothesis that sulfate assimilation is demand driven. Further analysis of the mechanism of the GSH mediated reduction in sulfate assimilation by FAA concentration measurements revealed that GSH addition roughly halved the concentrations of amino acids associated directly with ammonia assimilation. This supports a hypothesis that GSH reduces sulfate assimilation by countermanding ammonia assimilation by an unknown mechanism, although a speculative mechanism based on control at the level of gene expression was discussed in chapter 4. The analysis of transcriptional regulation of putative genes involved in sulfate assimilation demonstrated extreme differences between the AMF studied and other fungi. Additionally, changes in gene expression failed to correlate in most cases with uptake data. The most apparent difference is the induction of pathway genes by met addition in AMF, compared to their complete suppression in other fungi under similar conditions (Natorff et al., 2003; Cherest et al., 1985; Borges-Walmsley et al., 1995; Marzluf, 1997; Schierova et al., 2000). Exposure to met and GSH had very similar effects 190 of gene expression, increasing the transcript levels of the enzymes of the reverse- transsulfiiration pathway, cytathionine B-synthase and cystathionine y-lyase, by more than four fold. In comparison, met had no effect on sulfate uptake, and GSH reduced uptake of sulfate by two thirds. The expression of the putative high affinity sulfate perrnease gene did, however, coincide with uptake measurements. Exposure to cys led to a halving of transcript levels and a reduction of uptake of a third after two days of labeling. The addition of ammonia failed to affect gene expression of putative pathway enzymes in most cases, and certainly failed to account for the dramatic increases in sulfate uptake and assimilation under similar conditions. These data do not support the hypothesis that transcriptional regulation in AMF is controlled by an orthologous transcriptional regulator responding to amino acid starvation, as in S. cerevisiae (Mountain et al., 1993), however uptake data does. Gene expression data obtained demonstrate that AMF may have a distinct transcriptional regulatory environment among the fungi with respect to sulfate uptake and assimilation. However, the incongruity with uptake data requires further examination of these findings, possibly through yeast complementation experiments where the comparison between AMF genes and those of S. cerevisiae are more direct. Future research The thesis presented was designed to provide key and conclusive data on S assimilation and transfer by AM fungi which can be used as a basis for further research. The transport and reduction of sulfate, as well as the synthesis of cysteine, are well 191 conserved and studied pathways in fungi. Therefore, much of the future research on this subject lies in the comparison between pathway elements fi'om AMF, S. cerevisiae, A. nidulans, and N. crassa. In the short term, obtaining full length sequence information for all of the genes identified to this point would allow for complementation studied of known S. cerevisiae pathway mutants to determine if, and the extent to which, fimctional and regulatory homologies exist between the two systems. Unreported in this thesis, putative sequences obtained from Glomus intraradices germinating spore 454 sequencing data of both a positive and a negative transcriptional regulatory gene were discovered. Complementation studies of these genes would immediately reveal the correlations existing between AMF and other fungi in transcriptional regulation of the sulfate assimilation pathway. Additionally, the over-expression of these genes in a yeast background would make possible kinetic analyses of the pathway components, which could be compared to previously determined flmgal enzyme kinetics. The importance of S metabolic regulation to the AMF symbiosis is a topic of interest with respect to the root-less propagation of the firngus. Regulation of the S assimilation genes is connected to N assimilation and nutrient status in fungi, and therefore may provide an indicator of important steps in elucidating the genetic basis of the changes in life cycle stages. A study analyzing the expression patterns of sulfate assimilatory pathway genes in relation to life cycle stage would be a good first step in this regard, and may lead to a global regulatory cycle. Finally, the incorporation of mutant Medicago truncatula plants impaired in sulfate assimilatory pathway genes would help determine more definitively to what extent AMF contribute to plant S metabolism. A mutant with impairment in sulfate 192 transport, for instance, could directly demonstrate the extent to which AMF can replace plant uptake by roots. A selective screening of S assimilation mutants for the ability to create a functional AMF symbiosis also has the potential for discovery of any S-related genes required for the establishment of a functional endomycorrhiza. 193 References Baumgardner RE, Lavery TF, Rogers CM, Isil SS. 2002. Estimates of the atmospheric deposition of sulfirr and nitrogen species: Clean Air Status and Trends Network, 1990- 2000. Environmental Science & Technology 36(12): 2614-2629. Borges-Walmsley MI, Turner G, Bailey AM, Brown J, Lehmbeck J, Clausen IG. 1995. Isolation and characterization of genes for sulfate activation and reduction in aspergillus-nidulans - implications for evolution of an allosteric control region by gene duplication. Molecular & General Genetics 247(4): 423-429. Breton A, Surdinkerjan Y. 1977. Sulfate uptake in saccharomyces-cerevisiae - biochemical and genetic study. Journal of Bacteriology 132(1): 224-232. Cherest H, Thao NN, Surdinkerjan Y. 1985. Transcriptional regulation of the met3- gene of saccharomyces-cerevisiae. Gene 34(2-3): 269-281. Chiang TY, Marzluf GA. 1995. Binding-affinity and functional-significance of nit2 and nit4 binding-sites in the promoter of the highly regulated nit-3 gene, which encodes nitrate reductase in neurospora-crassa. Journal of Bacteriology 177(21): 6093-6099. Cress WA, Throneberry GO, Lindsey DL. 1979. Kinetics of phosphorus absorption by mycorrhizal and non-mycorrhizal tomato roots. Plant Physiology 64(3): 484-487. Govindarajulu M, Pfeffer PE, Jin HR, Abuhaker J, Douds DD, Allen JW, Bucking H, Lammers PJ, Shachar—Hill Y. 2005. Nitrogen transfer in the arbuscular mycorrhizal symbiosis. Nature 435(7043): 819-823. Grynberg M, Piotrowska M, Pizzinini E, Turner G, Paszewski A. 2001. The Aspergillus nidulans metE gene is regulated by a second system independent fi'om sulphur metabolite repression. Biochimica Et Biophysica Acta-Gene Structure and Expression 1519(1-2): 78-84. Hodge A, Campbell CD, Fitter AH. 2001. An arbuscular mycorrhizal fungus accelerates decomposition and acquires nitrogen directly from organic material. Nature 413(6853): 297-299. Ketter JS, J arai G, Fu YH, Marzluf GA. 1991. Nucleotide-sequence, messenger-ma stability, and dna recognition elements of cys-14, the structural gene for sulfate perrnease-ii in neurospora-crassa. Biochemistry 30(7): 1780-1787. Ketter JS, Marzluf GA. 1988. Molecular-cloning and analysis of the regulation of cys- l4+, a structural gene of the sulfur regulatory circuit of neurospora-crassa. Molecular and Cellular Biology 8(4): 1504-1508. 194 Kuras L, Thomas D. 1995. Functional-analysis of met4, a yeast transcriptional activator responsive to s-adenosylmethionine. Molecular and Cellular Biology 15(1): 208-216. Lefohn AS, Husar JD, Husar RB. 1999. Estimating historical anthropogenic global sulfur emission patterns for the period 1850-1990. Atmospheric Environment 33(21): 3435-3444. Marzluf GA. 1997. Molecular genetics of sulfur assimilation in filamentous fungi and yeast. Annual Review of Microbiology 51: 73-96. Miyake T, Sammoto H, Kanayama M, Tomochika K, Shinoda S, Ono B. 1999. Role of the sulphate assimilation pathway in utilization of glutathione as a sulphur source by Saccharomyces cerevisiae. Yeast 15(14): 1449-1457. Mountain HA, Bystrom AS, Korch C. 1993. The general amino-acid control regulates met4, which encodes a methionine-pathway-specific transcriptional activator of saccharomyces-cerevisiae. Molecular Microbiology 7(2): 215-228. Natorff R, Sienko M, Brzywczy J, Paszewski A. 2003. The Aspergillus nidulans metR gene encodes a bZIP protein which activates transcription of sulphur metabolism genes. Molecular Microbiology 49(4): 1081-1094. Okamoto PM, Garrett RH, Marzluf GA. 1993. Molecular characterization of conventional and new repeat-induced mutants of nit-3, the structural gene that encodes nitrate reductase in neurospora-crassa. Molecular & General Genetics 238(1-2): 81-90. Pirozynski KA, Malloch DW. 1975. Origin of land plants - matter of mycotropism. Biosystems 6(3): 153-164. Schierova M, Linka M, Pazoutova S. 2000. Sulfate assimilation in Aspergillus terreus: analysis of genes encoding ATP-sulfurylase and PAPS-reductase. Current Genetics 38(3): 126-131. Smith SE, Smith FA, J akobsen I. 2003. Mycorrhizal fungi can dominate phosphate supply to plants irrespective of growth responses. Plant Physiology 133(1): 16-20. Sosa E, Aranda C, Riego L, Valenzuela L, DeLuna A, Cantu JM, Gonzalez A. 2003. Gcn4 negatively regulates expression of genes subjected to nitrogen catabolite repression. Biochemical and Biophysical Research Communications 310(4): 1175-1180. StArnaud M, Hamel C, Vimard B, Caron M, Fortin JA. 1996. Enhanced hyphal grth and spore production of the arbuscular mycorrhizal fungus Glomus intraradices in an in vitro system in the absence of host roots. Mycological Research 100: 328-332. Wang B, Qiu YL. 2006. Phylogenetic distribution and evolution of mycorrhizas in land plants. Mycorrhiza 16(5): 299-363. 195 Wellman CH, Richardson JB. 1993. Terrestrial plant microfossils from silurian inliers of the midland valley of scotland. Palaeontology 36: 155-193. 196 TATE UNIVERSI 62 ”Willi 293 030 Illllllllllllllfs 9541