THE MOLECULAR AND BIOCHEMICAL BASIS OF NITROGEN TRANSFER BY THE MODEL ARBUSCULAR MYCORRHIZAL FUNGUS RHIZOPHAGUS IRREGULARIS By Taghleb Muhammad Al-Deeb A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Plant Biology-Doctor of Philosophy 2015 ABSTRACT THE MOLECULAR AND BIOCHEMICAL BASIS OF NITROGEN TRANSFER BY THE MODEL ARBUSCULAR MYCORRHIZAL FUNGUS RHIZOPHAGUS IRREGULARIS By Taghleb Muhammad Al-Deeb Plants can increase their effective root length and surface area by investing in symbioses with soil fungi, forming mycorrhizal associations. In mutualistic mode, the plant provides photosynthate (fixed carbon) to the fungus, whereas the fungus provides nutrients to the plant. Arbuscular mycorrhizal (AM) associations are found in 80% of vascular plant families and as a consequence, the AM symbiosis is of tremendous significance to life on this planet, in both natural and agricultural ecosystems. Work in recent years has substantially increased our understanding of nitrogen nutrition in the AM symbiosis. At the molecular level, a working model for nitrogen uptake, metabolism, and transfer has emerged. In this dissertation, mechanisms and genes believed to be responsible for nitrogen flows in the AM symbiosis are described and open questions about the pathway and its regulation are highlighted. Molecular and biochemical experimental approaches were used to investigate these unresolved questions. A compartmented microcosm was developed for aseptic and leakage-free whole-plant mycorrhizal experiments. This was used to monitor S and N uptake by the fungal extraradical mycelium (ERM) and its transfer to host plants. Our results show rapid S and N transfer by ERM to the host plants. Using growth parameter measurements, chlorophyll contents as well as 15N labeling, we conclude that nitrogen transfer from an arbuscular mycorrhizal fungus confers growth benefits on the host plant under nitrogen limiting conditions and that the microcosm system developed will be useful for future work on AM nutrition and metabolism under physiologically relevant conditions. Isotopic labeling time course experiments using different 15N and 13C labeled substrates as well as expression analysis of the expression of key genes were performed using microcosms and an in vitro monoxenic culture system. The results demonstrated the operation of a new pathway of N transfer by AMF to the host via nitrate translocation from the extraradical mycelium to plant roots and shoots. The results also indicate that ornithine is made in the ERM via pyrroline-5-carboxylate and that some of it is broken down in the IRM to glutamate and to a lesser extent to putrescine. Labeling analysis strongly suggests that ornithine is also translocated from the intraradical mycelium to the ERM and is used to make arginine there. Changes in gene expression are consistent with the labeling data on N uptake, metabolism and movement. Gene expression analysis of glutamate dehydrogenase suggests a potential dissimilatory role in the IRM. Copyright by TAGHLEB MUHAMMAD AL-DEEB 2015 ACKNOWLEDGEMENTS I would like to express my warm feeling, deepest appreciation, and deep thanks to my graduate supervisor Prof. Yair Shachar-Hill, for his supervision, encouragement and guidance through this work, his kindness and patience were always encouraging me to go ahead, I couldn’t hope for better advisor. My deep thanks and appreciation to the Professors; Daniel Jones, Frances Trail and Robin Buell for encouragement and kindness to be member of my committee and their help. I would like to thank the mass-spec facility members especially; Prof. Daniel Jones and Lijun Chen for teaching and helping me through my work there. I would like also to thank past and current members of Prof. Shachar-Hill’s lab who have impact on my research including Beth Kasiborski, Chunjie Tian, Rahul Deshpande, Matthew Juergens, Lisa Carey, (Wilson) Chen, Leann Lerie Matta, Mike Pollard, Teresa Clark, Mike Opperman and Bradley Disbrow for their teaching, helping, teasing and encouraging me. My Great and Deep thanks are to the colleagues in the Department of Plant Biology and my friends at Michigan State University. My infinite, great and unlimited deep thanks to my family members; my father, mother, brothers and sisters for their continuous support and encouragement. v TABLE OF CONTENTS LIST OF TABLES .......................................................................................................... viii LIST OF FIGURES .......................................................................................................... ix Chapter 1 ....................................................................................................................... 1 Introduction and literature review .................................................................................... 1 Introduction .................................................................................................................. 2 Morphology, anatomy and life cycle ............................................................................ 3 Phylogeny ................................................................................................................... 5 Plant benefits from AM symbioses .............................................................................. 6 Phosphorous uptake and transfer ........................................................................ 7 Sulfur uptake and transfer ................................................................................... 8 Nitrogen uptake and transfer ............................................................................... 9 The mechanisms of nitrogen transfer and metabolism in the AM symbiosis ............. 10 The pathway of nitrogen movement through the AM symbiosis ........................ 10 The uptake of nitrogen by the ERM of the AM symbiosis .................................. 11 The assimilation of N in the ERM and release of ammonium in the IRM ............ 12 What is the role of AM networks in the uptake and exchange of nitrogen among plants in ecosystems? ......................................................................................................... 14 Movement of N from soil to plants via AMF in plant ecosystems ....................... 14 Colonization by AMF depends on the nutrient status of host plants .................. 16 Nitrogen exchange between plants through mycorrhizal networks and its ecological significance ....................................................................................... 19 Interactions between AMF and symbiotic N2-fixing microbes and their contribution to N accumulation in host plants ................................................................................... 20 Conclusions and perspectives .................................................................................. 21 APPENDIX ................................................................................................................ 24 REFERENCES ......................................................................................................... 32 Chapter 2 ..................................................................................................................... 45 Nitrogen uptake and transfer in an aseptic whole plant-mycorrhizal culture system ..... 45 Abstract ..................................................................................................................... 46 Introduction ............................................................................................................... 48 Experimental procedures .......................................................................................... 52 Chemicals and reagents .................................................................................... 52 Seed sterilization ............................................................................................... 52 Growth conditions .............................................................................................. 52 Experimental design .......................................................................................... 53 Testing the diffusion between compartments and the capillary transfer ............ 54 Extraction and measurement of 35SO4 from mycorrhizal plants ......................... 54 Analysis of mycorrhizal colonization and ERM crossing to the FC .................... 55 vi Growth parameters and chlorophyll content measurements ............................. 55 Extraction, isolation and quantification 15N metabolites ..................................... 56 Liquid chromatography and mass spectrometry (LC-MS) .................................. 56 Results ...................................................................................................................... 58 Discussion ................................................................................................................. 62 APPENDIX ................................................................................................................ 66 REFERENCES ......................................................................................................... 83 Chapter 3 ..................................................................................................................... 88 Evidence for additional pathways in the nitrogen transfer network of arbuscular mycorrhizas .................................................................................................................. 88 Abstract ..................................................................................................................... 89 Introduction ............................................................................................................... 91 Experimental procedures .......................................................................................... 94 Chemicals and reagents .................................................................................... 94 Spore material and mycorrhizal in vitro growth conditions ................................. 94 Isotopic labeling in colonized Ri T-DNA-transformed carrot (Daucus carota clone DCI) roots culture system .................................................................................. 95 15N labeling ............................................................................................. 95 13C labeling ............................................................................................. 95 15N labeling for 15NO measurements in whole plant-mycorrhizal system ......... 96 3 Extraction and isolation of 15N and 13C metabolites ........................................... 96 Liquid chromatography and mass spectrometry (LC-MS) .................................. 97 15N labeled and unlabeled nitrate measurements .............................................. 99 RNA extraction and putative gene fragment isolation ...................................... 100 Quantitative Real-Time PCR measurements ................................................... 100 Results .................................................................................................................... 101 Nitrate movement through the AM symbiosis .................................................. 101 N movement through the AM symbiosis in transformed root system ............... 102 The origin and fate of the N metabolites ........................................................... 104 13C labeling in the ERM and AM roots ............................................................. 104 Gene identification ........................................................................................... 105 Gene expression in response to N addition ..................................................... 106 Discussion ............................................................................................................... 107 APPENDIX .............................................................................................................. 113 REFERENCES ....................................................................................................... 133 Chapter 4 ................................................................................................................... 137 Conclusions and future research ................................................................................ 137 Conclusions ............................................................................................................ 138 Future research ....................................................................................................... 140 REFERENCES ........................................................................................................... 142 vii LIST OF TABLES Table 1-1 Nitrogen metabolic proteins and genes identified in the arbuscular mycorrhizal symbiosis ...................................................................................................................... 29 Table 2-1 Precursor (unlabeled) compounds and product ions for each analyte .......... 81 Table 2-2 Multiple reaction monitoring (MRM) transitions, optimizing source cone voltages, collision cell voltages, and analyte retention time (for 15N labeling experiment) ...................................................................................................................................... 82 Table 3-1 Multiple reaction monitoring (MRM) transitions, optimizing source cone voltages, collision cell voltages, and analyte retention time (for 13C labeling experiment) .................................................................................................................................... 130 Table 3-2 Primers for real-time PCR ........................................................................... 131 Table 3-3 Nitrogen metabolic genes identified in Rhizophagus irregularis in this study .................................................................................................................................... 132 viii LIST OF FIGURES Figure 1-1 Arbuscular mycorrhizal fungal morphologies in tomato (Solanum lycopersicum) roots plant. ............................................................................................. 25 Figure 1-2 Phylogeny of fungi based on SSU rRNA sequences. ................................. 26 Figure 1-3 Taxonomy of the Glomeromycota containing the arbuscular mycorrhizal and related fungi, based on SSU rRNA gene sequences .................................................... 27 Figure 1-4 Working model of N transport and metabolism in the AM symbiosis ........... 28 Figure 1-5 Interactions among arbuscular mycorrhizal fungi, host plants, and N2-fixing bacteria and other soil microbes in nitrogen cycling ...................................................... 29 Figure 2-1 Whole plant-mycorrhizal two-compartment culture system .......................... 67 Figure 2-2 Sulfur transfer from fungus to plant using the whole plant culture system.... 68 Figure 2-3 Colonization of Medicago truncatula by Rhizophagus irregularis ................. 71 Figure 2-4 Seven-week old Medicago truncatula plants growing in the two-compartment system ........................................................................................................................... 72 Figure 2-5 Total chlorophyll content and growth parameters measurements ................ 73 Figure 2-6 Free N metabolite concentrations and 15N labeling in the mycorrhizal whole plant system .................................................................................................................. 76 Figure 3-1 Time course of 15N percentage labeling of nitrate in transformed carrot (Daucus carota) roots colonized by R. irregularis ....................................................... 114 Figure 3-2 The concentrations of 15N labeled nitrate (light grey bars) and unlabeled nitrate (black bars) in the roots and shoots of mycorrhizal M. truncatula plants grown in micosms ...................................................................................................................... 115 Figure 3-3 Nitrogen movement from the ERM to the IRM in mycorrhizal transformed carrot roots .................................................................................................................. 116 Figure 3-4 Time course of the free N metabolites 15N-Isotopomers (mass isomers) in the extraradical mycelium (ERM) after the addition of 4 mM K15NO3 to the ERM .............. 120 ix Figure 3-5 Timing course of 13C labeling of N metabolites in the extraradical mycelium (ERM) and the mycorrhizal roots after the addition of 0.5 mM 13C6 arginine to the fungal ERM compartment. ..................................................................................................... 123 Figure 3-6 The expression of primary N metabolic and transport genes in the arbuscular mycorrhizal symbiosis after the addition of 4 mM KNO3 to the fungal extraradical mycelium (ERM) .......................................................................................................... 128 x Chapter 1 Introduction and literature review 1 Introduction The arbuscular mycorrhizal (AM) symbiosis is the most ancient and widespread plant-microbe mutualism among land plants. Arbuscular mycorrhizas are formed by the roots of over 80% of plant species with fungi of the Glomeromycota. Arbuscular mycorrhizal fungi (AMF) are asexual, obligate symbionts (Smith and Read, 2008) that propagate by forming multinuclear vegetative underground spores and by the growth of the underground mycelium to colonize additional roots. After spore germination, hyphal growth is associated with nuclear divisions as well as the migration of nuclei from the spore (Bianciotto and Bonfante, 1992; Becard and Pfeffer, 1993; Bianciotto et al. 1995). The hyphae use carbohydrate and lipid reserves during pre-symbiotic growth (Becard et al., 1991; Bago et al., 1999). In the absence of a host root, the growth of the mycelia ceases within several days, and spores are capable of multiple germination events. Presymbiotic hyphae respond to signaling molecules in plant root exudates that stimulate branching (Giovannetti et al.,1993; Buee et al. 2000). Akiyama et al., (2005) identified strigolactones in plant root exudates as increasing germ tube metabolism and growth. Strogolactone production is stimulated when plants experience mineral deficiency. Hyphal contact with the root is usually followed by adhesion of the hyphal surface and after 2-3 days, the formation of appressoria followed by root penetration and formation of arbuscules around 2 days later (Brunndrett et al., 1985; Becard and Fortin, 1988; Peterson and Bonfante, 1994). Once the AM symbiosis has been established, the fungal growth proceeds in both the soil (extraradical mycelium) and in the root (intraradical mycelium). 2 AMF have so far proven to be unculturable axenically and appear in nature to be unable to complete their life cycle without forming a symbiosis with plant host roots (Smith and Read, 2008). AMF take up photosynthetically fixed carbon (C) from the plant and convert it into triacylglycerols, which are the main nutritional form of carbon stored and translocated by the fungus for growth and sporulation and utilized during germination (Bago et al. 2000). The intraradical mycelium imports glucose from the roots and converts it into trehalose and glycogen and then to triacylglycerol (Shachar-Hill et al., 1995; Pfeffer et al., 1999). Lipid bodies transport C from the intraradical mycelium (IRM) to extraradical mycelium (ERM) and are utilized for mycelial growth (Bago et al., 2002; Lammers et al., 2001) Morphology, anatomy and life cycle Details of fungal interactions with plant cells and tissues of the root were first described by Gallaud (1905) in which he indicated that AM roots can contain a variety of structures including arbuscules hyphae and vesicles. Gallaud (1905) categorized them into four distinct structural classes; Arum, Paris, hepatic and orchid types. The most common forms are found in AM roots: the Paris-type and Arum-type (Figure 1-1). In Paris type the cortical colonization of the root is characterized by extensive development of intracellular coiled hyphae with arbusculated coils in root cortex cells. These coils spread from cell to cell and do not form significant intercellular hyphae. Paris type morphology occurs in many families of pteridophytes, gymnoserms and angiosperems (Smith and Smith 1997). By contrast, Arum-type is often described in the 3 fast-growing roots systems of crop plants (Brundrett et al., 1990). Although the plant species clearly influences the fungal morphology type, with the same fungus forming Arum morphology on one host species and Paris in another (Smith and Smith, 1997), it has also been shown that different fungal species form different morphologies in the same host plant (Cavagnaro et al., 2001). In these associations the fungus spread rapidly through the root cortex by the growth of intercellular hyphae which extend along well-developed intercellular air spaces. Side branches penetrate the plant cell walls of cortical cells to produce highly branched hyphal structures named arbuscules for their tree-like appearance (Smith and Read, 1997). The arbuscules are formed by dichotomous branching and are transient structures that degenerate and disappear about 2-3 weeks after they develop.The arbuscules invaginate root cortical cells without penetrating the host cell membrane and form a periplasmic space between the root cell and the arbuscular membrane. The host plant cell resumes its normal appearance after the arbuscule within it breaks down. The periplasmic space is believed to be the site of nutrient transfer between the fungus and the plant (Bago, 2000; VanAarle et al., 2005). Outside the root, AMF form mycelia consisting of runner hyphae that branch at a frequency that depends on the nutritional state of the external substrate. From these side branches, arbuscule-like Branched Absorbing Structures (BAS) are formed that are proposed to be involved in nutrient uptake in the soil (Bago, 2000). The BAS are the site of formation of the spores, which are vegetative multinuclear Sporangiospores that develop at the hyphal tips of BAS. A single spore contains several hundred to several thousand haploid nuclei. These spores germinate by forming a germ tube into which nuclei emerge from the spore and proliferate (Becard et al., 1990). Germination, germ 4 tube growth and branching of the pre-symbiotic mycelium is stimulated by plant exudates (Mosse, 1958). Strigolactones have been shown to be major active molecules in these exudates. At the surface of host roots, the presymbiotic hyphae form appressoria, commonly particularly at the sites of lateral root emergence (Smith and Read, 1998), and from which hyphae penetrate the root epidermis to form the intraradical mycelium. Phylogeny It was believed, partly on the basis of their asexual life cycle, that AMF are most closely related to zygomycota but the use of DNA sequences forced a re-evaluation their relationship (Smith and read, 2008). Based on small subunit (SSU) rRNA gene sequences, morphological and ecological characteristics, AMF are now regarded as separate from all other major fungal groups in a monophyletic clade. Consequently they were placed into a new phylum, the Glomeromycota (Schußler et al., 2001; Figure 1-2 & 1-3). About 240 species of glomeromycotan fungi have been described to date based on their spore morphology and molecular phylogenetic data (Schüßler & Walker, 2010; Redecker et al., 2013). Alves de silva et al., (2006) analyzed the large subunit (LSU) ribosomal RNA (rRNA) for AM fungal taxonomy and their data indicated that Archaeosporaceae are a basal group in Glomeromycota, Acaulosporaceae and Gigasporaceae belong to the same clade, while Glomeraceae are polyphyletic. large subunit of RNA polymerase II (RPB1; Redecker and Raab 2006; Stockinger et al., 2014) and β-tubulin gene (Msiska and Morton 2009) were phylogentically analyzed too. 5 The β-tubulin gene phylogeny was similar to the 18S (LSU) rRNA gene phylogeny at the family and species level, but not at the order level (Msiska and Morton 2009). Based on all of these studies, AMF are divided into ten families in the phylum Glomeromycota, order Glomerales. Those families are Gigasporaceae, Glomeraceae, Acaulosporaceae, Diversispora, Paralglomaceae, Geosiphoaceae, Ambisporaceae, Eutrophosphosporaceae, and Arcaesporaceae. Glomus is the largest genus within the phylum with more than 70 species (Redecker and Raab 2006). Several of the Glomus species, most frequently Glomus intraradices (now Rhizophagus irregularis), are commonly studied (Smith & Read, 2008). Based on a molecular analysis of ribosomal DNA and a re-evaluation of an early description of this species, it was recently renamed Rhizophagus irregularis (Krüger et al., 2012) despite the misleading “root eater” title. Tisserant et al., (2013) assembled and annotated the genome of Rhizophagus irregularis from high throughput DNA sequencing in association with transcriptome data. This study provides insight into the capabilities of this fungus and points to genes involved in mycorrhizal symbioses. Plant benefits from AM symbioses Association with AMF is often beneficial to plants by improving their access to nutrients (Smith and Smith, 2011). AMF also enhance pest and disease suppression and improve drought tolerance (Smith and Read, 2008). Among the benefits that AMF provide to host plants, improved phosphorus nutrition has received the most attention. Previous studies, however, indicated that AMF may also be important for a wide variety of nutrients (Smith and Smith, 2011). AMF enhance the uptake of nitrogen 6 (Govindarajulu et al., 2005), sulfur (S, Allen & Shachar-Hill, 2009), zinc (Seres et al., 2006), copper (Toler et al., 2005) and iron (Kim et al., 2010). Hart and Forsythe, (2012) showed that the identity of AMF can influence the uptake of many nutrients but that the magnitude and direction of the nutrient-derived growth response is also affected by host plant characteristics and soil nutrient status. Phosphorous uptake and transfer AMF improve the phosphorous (P) nutrition of host plants (Bolan, 1991). The extraradical mycelium (ERM) explores a larger soil volume than is possible for roots and reaches zones and soil pores that the lateral roots cannot access. The small hyphal diameter (typically several microns) leads to an increased P absorbing surface area (Marschner and Dell, 1994) and estimates of ERM levels around host plants indicate that these reach several meters of hyphal length per gram of soil (Jakobsen et al., 1992). In AMF, polyphosphates (polyP) are formed after P uptake thus lowering internal inorganic P (Pi) concentrations. Organic acids and phosphatases are released by the ERM which increase the availability of P from organic and inorganic sources. In the extraradical hyphae, N is transported as arginine (Govindarajulu et al., 2005; Tian et al., 2010) which may be bound to polyphosphate and therefore be coupled to Pi translocation (Jin et al., 2005). Several studies have shown that AM specific root phosphate transporters are induced in the roots of plants colonized by AMF (Javot et al., 2007; Gomez et al., 2009; Nagy et al., 2009). It was suggested that Pi delivery to cortical cells was necessary for sustaining the symbiosis because in Medicago truncatula mutants affected in the AM-specific Pi transporter 4 gene, the arbuscules 7 accumulated polyphosphate and prematurely degenerated (Javot et al., 2007). Increasing carbohydrate availability stimulates P uptake by the ERM to and translocation to the mycorrhizal roots as well as altering the metabolic and spatial distribution of P within the fungus (Bücking and Shachar-Hill, 2005). Previous studies suggested that there is a cross-talk between P and N nutrition (Blanke et al., 2005, Bonneau et al, 2013). It was found that low P and N fertilization induced a physiological state of plants favorable for AM symbiosis despite their higher P status (Bonneau et al, 2013). Sulfur uptake and transfer Previous studies have reported the effects of AM colonization on the uptake of S (Gray and Gerdemann, 1973; Rhodes and Gerdemann, 1978). Gray and Gerdemann (1973) showed that mycorrhizal colonization in clover (Trifolium pratense) and maize increased 35S uptake compared to nonmycorrhizal plants. Furthermore, Rhodes and Gerdemann (1978) found that mycorrhizal colonization increased 35S uptake in onion compared to nonmycorrhizal plants and they showed that the S uptake was heavily influenced by P nutritional benefits. Allen and Shachar-Hill (2009) showed using in vitro cultures of transformed roots colonized by R. irregularis that the ERM takes up S in the form of sulfate and sulfurcontaining amino acids and transfers it to mycorrhizal roots. They also showed that the root S contents were increased by 25% in a moderate (not growth-limiting) concentration of sulfate. Fifty percent of 35SO -2 4 8 uptake from the fungal compartment was detected in the mycorrhizal roots. Similar quantities of 35S were transferred to mycorrhizal roots whether 35SO42−, [35S]Cys, or [35S]Met was supplied in the fungal compartment (Allen and Shachar-Hill, 2009). Sieh et al., (2013) studied the effect of mycorrhizal colonization on sulfur starvation responses in M. truncatula and they found that colonization reduced S starvation when the plant's phosphate status is high, concluding that mycorrhizal sulfur transfer improves plant S nutrition. Nitrogen uptake and transfer Plant roots and the AM fungal ERM can absorb both nitrate and ammonium from soil (George et al., 1995; Smith and Read, 2008) as well as soluble organic nitrogen which together dominate the soil N pool (Jin et al., 2005, McNeill and Unkovich, 2007). Nitrate and ammonium are mobile under most soil conditions but this can be restricted in dry soil (Tobar et al., 1994, Tinker and Nye, 2000). These observations suggest a possible role for N uptake and transfer, a suggestion that has been strengthened by tracer studies and field observations. Haines and Best (1976) found that colonization of Liquidambar styraciflua by the AM fungus Glomus mosseae retarded the leaching of ammonium and nitrate from soil suggesting that the AMF might be involved in N uptake. When N-15-labelled fertilizer was added to mycorrhizal or non mycorrhizal control pots, Azcon-Aguilar et al., (1993) reported that the 14N/15N ratio was higher in AM onion plants than in uncolonized control plants, suggesting that AMF were able to access soil N that is less available to non mycorrhizal plants. Furthermore, Cliquet and Stewart (1993) demonstrated that 15N translocation from roots to shoots through the xylem was higher in AM plants compared 9 with control plants. Under water-stressed conditions that affect the nitrate availability to roots, the 15N enrichment was four times higher in mycorrhizal than in non-mycorrhizal lettuce plants providing evidence of hyphal transport to the plant of N from labeled nitrate (Tobar et al., 1994). A mechanism of N transfer from the fungus to the plant in the AM symbiosis was proposed (Bago et al., 2001) in which N taken up by the fungus is incorporated into amino acids, translocated from the ERM to the intraradical mycelium (IRM) as arginine, which is broken down to ammonium, that is released to the host root. This model has gained support from 15N and 13C isotopic labeling experiments, measurements of enzymatic activities (Govindarajulu et al., 2005; Jin et al., 2005, Cruz et al., 2007) and from gene expression data (Govindarajulu et al., 2005; Guether et al., 2009; Gomez et al., 2009; Tian et al., 2010). The mechanisms of nitrogen transfer and metabolism in the AM symbiosis The pathway of nitrogen movement through the AM symbiosis The current working model of N transfer from the fungi to the plant in the AM symbiosis was proposed by Bago et al. (2001) based on previous work that demonstrated fungal N uptake and metabolism and implicated amino acids in N handling (Johansen et al., 1996; Bago et al., 1996; 2000). This mechanism involves N uptake up by the fungi which is incorporated into amino acids, translocated from the extraradical mycelium (ERM) to the intraradical mycelium (IRM) as arginine (Arg), but 10 transferred to the plant without C as inorganic N (Govindarajulu et al., 2005; Jin et al., 2005). Gene expression and enzyme activities for plant and fungal proteins involved in the nitrogen metabolic pathway have been analyzed (Table 1-1). Most of the important enzymes and nitrate or ammonium transporters involved in nitrogen metabolism were identified from the AM fungus R. irregularis (Kaldorf et al. 1998; Govindarajulu et al., 2005; Lopez-Pedrosa et al., 2006; Gomez et al., 2009; Tian et al., 2010). Gene expression at the transcriptional and post transcriptional levels and 15N and 13C labeling experiments support the proposed model (Govindarajulu et al., 2005; Cruz et al., 2007; Tian et al., 2010). In Figure 1-4, the current model is illustrated, including the genes identified and the postulated regulation on the transcripts by the metabolites (Tian et al., 2010). The uptake of nitrogen by the ERM of the AM symbiosis A high affinity ammonium transporter has been identified in R. irregularis and characterized by Lopez-Pedrosa et al. (2006), which is involved in the ammonium uptake of ERM. A putative high affinity nitrate transporter which is up-regulated in response to nitrate addition was identified in the ERM of the same AM fungus (Tian et al., 2010). Plant and fungal uptake of macronutrient ions involves low affinity as well as high affinity transporters, and it is expected that low affinity transporters for NO3- and NH4+ will also be identified. AMF can also obtain N from decomposing organic materials and store it in the mycelium (Hodge and Fitter, 2010). Amino acids including Gly, Glu, 11 Pro and Arg can be taken up by the ERM (Hawkins et al., 2000; Jin et al., 2005) and Arg, Gly, Gln and Orn can be taken up by germinating spores (Gachomo et al., 2009). An amino acid permease, GmosAAP1, that can transport proline through a protoncoupled process has been characterized from Glomus mosseae; its expression is transcriptionally upregulated by external amino acids (Cappellazzo et al., 2008). It is likely that AMF express permeases allowing the uptake of other amino acids such as Arg by the fungus from the environment. Indeed, genome-scale analysis of an ectomycorrhizal fungus has highlighted the existence of many N transporters (Lucic et al., 2008). However the uptake rates of amino acids by the ERM measured in AM root cultures are substantially lower than the rates for nitrate, ammonium and urea (Jin et al. 2005, and this work). It also appears that most of the nitrogen taken up by AMF from organic patches in the soil is taken up without carbon (Hodge & Fitter, 2010), in agreement with earlier experiments on the fate of nitrogen and carbon supplied to mycorrhizal plants as labeled peptides (Persson et al., 2003). The assimilation of N in the ERM and release of ammonium in the IRM Kaldorf et al. (1998) reported a partial sequence for a putative nitrate reductase from R. irregularis. The glutamine synthetase/glutamate synthase (GS/GOGAT) pathway is important in the assimilation of ammonium produced from nitrate or taken up directly from the soil. Activities for the GS/GOGAT pathway but not the alternative assimilatory NADP-dependent glutamate dehydrogenase (GDH) were reported by Cliquet and Stewart (1993) in mycorrhizal roots, and the application of a GOGAT 12 inhibitor to extraradical mycelium reduced 15N assimilation. Breuninger et al. (2004) found that GS activity is upregulated in response to N addition in the ERM of R. irregularis and G. mossae. The glutamine synthetase from R. irregularis reported by Breuninger et al. (2004) was recently found to be one of a gene family with at least two members, and this paralog (named GiGS1, Tian et al. 2010) has a high constitutive transcriptional level whereas another member (GiGS2) is actively up-regulated upon exposure to nitrate but has a lower constitutive expression level. This suggests different roles for the two enzymes at different environmental nitrogen levels. Additionally, both of them have low Km values for glutamate, which indicates that the GS/GOGAT pathway could operate even at low nitrogen levels (Tian et al., 2010). Since both GS/GOTAT and NADP-GDH contribute to N assimilation in ectomycorrhizal fungi, more direct measurements in AMF would be desirable to determine if and when NADP-GDH might be involved in the AM symbiosis. Genes for all the reactions of the urea cycle have been identified in R. irregularis. Enzymatic activities associated with the synthesis of Arg, are up-regulated in the ERM after ammonium addition (Cruz et al., 2007). Soon after this, fungal genes for breaking down Arg, including arginase (GiCAR1), urease (GiURE), and ornithine aminotransferase (GiOAT1, 2) are up-regulated in the colonized root tissues (Tian et al., 2010). Enzymatic activity of glutamine synthetase, arginase and urease were shown to be up-regulated in the colonized root in response to N addition to the ERM (Cruz et al., 2007), and later gene expression analyses (Tian et al., 2010, Tisserant et al., 2012) indicate that this activity is fungal. 13 The interface between fungal arbuscules and plant cortical cells is important for P transfer from the fungus to the plant (Harrison, 1999; Pumplin & Harrison, 2009). Localized gene expression analysis indicates that the periarbuscular membrane is also important for ammonium transfer from the fungus to the plant cells in arbuscular mycorrhiza (Javelle et al., 2003; Gomez et al., 2009). From Lotus japonicus, a mycorrhiza-inducible ammonium transporter LjAMT2;2, which is found located in the apoplastic interfacial compartment, was suggested to bind charged ammonium and release uncharged NH3 into the plant cytoplasm (Guether at al., 2009). Furthermore, another mycorrhiza-inducible ammonium transporter GmAMT4.1 from Glycine max was also found to be located in the branch domain of periarbuscular membranes (Kobae et al., 2010). What is the role of AM networks in the uptake and exchange of nitrogen among plants in ecosystems? Movement of N from soil to plants via AMF in plant ecosystems The discovery of an elaborate system for moving N from soil to plant is complemented by demonstrations under controlled experimental conditions of large N fluxes through the symbiosis. In different experiments, at least 21%, 30% and 50 % of the total N present in AM roots came from the fungal ERM in an in vitro mycorrhiza system where mycorrhizal roots had access to N both by direct uptake and via the fungus (Toussaint et al., 2004). Tanaka and Yano (2005) found that 75 % of N measured in the leaves of mycorrhizal maize was taken up by the AM fungus’ ERM. 14 AMF can also obtain nitrogen from organic matter from the soil and transfer it to host plants in significant quantities (Leigh et al., 2009; Hodge and Fitter, 2001; 2010). However, by comparison with P, there are very few reports that directly demonstrate Ndependent enhancement of plant growth by AM colonization due to N transfer from the fungus. In a study by Tanaka and Yano (2005), divided compartments in which nitrogen was available in a compartment to which only the ERM had access were used. The presence or absence of an air gap to prevent diffusion between compartments significantly increased nitrogen contents in host plants when ammonium was used but did not significantly increase plant biomass. The authors concluded from this and 15N labeling results that the improved plant N content was due to direct N transfer by the fungus from soil to plant. However, the apparent absence of N transfer when nitrate was the form of N supplied, contradicts previous and subsequent reports (Jakobsen 1992; Johansen et al., 1993; Tobar et al., 1995; Jin et al., 2005, this study) raising uncertainties about the experimental system used. Although the N content of colonized plants have been observed to be higher than un-colonized plants in a range of other studies (reviewed in He et al., 2003; 2009), these have not distinguished N transfer by the fungus from increased uptake by the plant secondary to improved P status or other changes in the host and/or soil. Indeed, other studies have indicated that the transfer of N does not always confer a net growth benefit to host plants. For example, Reynolds et al. (2005) found that even at low N supply, AM colonization did not increase total N uptake. The extent and significance of N movement through the AM symbiosis under natural conditions is even less clear - partly because of methodological difficulties. 15 Recently, the finding that mycorrhizal fungi discriminate against 15N during nitrogen transfer from soil to host plant has shown a potential application to track N flow in Nlimited ecosystems. The discrimination against transfer will result in relatively enriched rather than 14N 14N 15N by mycorrhizal fungi during nitrogen in plants rather than 15N, while enriched 15N in mycorrhizal fungi. Accordingly, the lower 15N:14N in mycorrhizal plants compared with non-mycorrhizal plants suggests the contribution of N transferred from fungi to host plants (Hobbie & Macko, 2000; Hobbie & Colpaert, 2003; Hobbie et al., 2005). In addition, N transfer from fungi to host plants could be quantified by using 15N:14N values for host plants, mycorrhizal fungi, and soil. However, even though this approach has proved applicable for evaluating nitrogen transfer from fungus to ectomycorrhizal plants in N-limited ecosystems (Hobbie & Colpaert, 2003; Hobbie et al., 2005; Hobbie & Hobbie, 2008), no reliable estimates of the contribution of AMF to plant N contents have yet been made using this approach. This is partly due to uncertainties of interpretation of such analyses and partly because the 15N abundance in AM plants is closer to background than in nitrogen fixing or ectomycorrhizal symbioses. Further improvements for the analysis and interpretation of isotopic discrimination results in model AM systems and natural settings may enable estimates of the extent of N movement from AMF to plants in a range of ecosystem types. Colonization by AMF depends on the nutrient status of host plants Indirect evidence for AMF having a role in N uptake under natural conditions comes from the effect of soil N levels on colonization and growth of AMF. As reported 16 by Bago et al. (2004), the AM fungal morphological and developmental changes subjected to different nutritional conditions, especially nitrogen status, as a strategy to exploit the substrate efficiently. Low nutrient status including low nitrogen in the soil induce the development and growth of AMF (Yoneyama et al., 2007), and AM colonization rates have been shown to be related to N availability in natural and disturbed ecosystems (Egerton-Warburton & Allen, 2000; Jackson et al., 2001; Jia et al., 2004; Blanke et al., 2005). For example, Blanke et al. (2005) reported a negative relationship between percentage root colonization by AMF and both tissue N concentration and N:P ratio in Artemisia vulgaris growing in high P soils and Johnson et al., (2003) found that N fertilization lowered AM colonization in grassland sites with low N:P but not at sites with low P. This inverse relationship between AM colonization rates and nutrient levels is well documented for P in the AM symbiosis and is also consistent with the finding that development of N2-fixing nodules is reduced when N levels in the soil are high (Streeter, 1985). This similarity is consistent with a beneficial role for AMF when N is limiting. Although significant, the effects of N deposition on colonization are generally not so marked as for elevated CO2 or P addition (Constable et al., 2001; Treseder 2004; Gamper et al., 2005). In an in vitro mycorrhizal symbiosis, Olsson et al. (2005; 2010) found that increased availability of N or P to host roots reduced carbon allocation to the fungus and concluded that negative impacts of N high nutrient level on AM abundance are caused by reduced C allocation from the plant when plant requirements are met by direct uptake. However, no colonization rates were reported in that work and it should be followed up in more mycorrhizas. Other evidence suggests that the effects of soil N 17 levels on the colonization of AMF could result in part from the change of fungal morphology which is influenced by the form and availability of inorganic nitrogen (Bago et al., 1996). Furthermore, both root exudates and the growth and metabolism of rhizospheric microorganisms which influence AM colonization are also likely to vary with soil N (Gryndler et al., 2009). Thus, colonization rates may be due to both direct and indirect effects on the plant and fungal partners making it unclear whether the correlation between colonization and N availability is related to a beneficial role for AMF in plant N nutrition. Direct investigation of the importance of plant N status on the regulation of the symbiosis through analyses of exudate composition and plant defense gene expression in response to AMF would be valuable. More recently, the turnover rate of arbuscules, which is accelerated in plants lacking the AM-specific P transporter, was shown to return to normal when N levels were limiting to plant growth (Javot et al., 2011) Thus, isotopic discrimination analyses, the effects of soil N levels on colonization, and the increased N contents of AM plants all point towards a significant role for the symbiosis in N uptake by plants. However, none of these lines of evidence is sufficient to demonstrate unambiguously that N movement through the symbiosis provides a direct growth benefit to host plants, especially under natural conditions. It is interesting in this context that work on P nutrition has shown that a large fraction of plant P can be taken up through AM fungal partners, even in cases when there is no plant growth enhancement or even net increase in P uptake (Smith et al., 2003). This may be the case for N more frequently than for P because the higher mobility of nitrate than 18 phosphate generally gives both partners access to soil N, where P depletion zones can more easily restrict direct access to P by roots. Nitrogen exchange between plants through mycorrhizal networks and its ecological significance Transfer of N among different plants via AMF has been reported in experimental systems indicating that AMF can play an important role in N transfer between plants, especially from legumes to non-legumes (Bethlenfalvay et al., 1991; Frey & Schuepp, 1993; Johansen & Jensen, 1996). Recent work using 15N labeling and natural abundance measurements has reported modest nitrogen transfer between plants via common arbuscular mycorrhizal networks. For example, Jalonen et al. (2009) deduced that 2.5% of the total N of grass was transferred by the common mycorrhizal network from neighboring leguminous trees. Such studies (reviewed by He et al., 2003; 2009), have resulted in estimates of nitrogen transfer between plants through mycorrhizal networks, (especially from N fixing legumes to non-legumes) that range widely, between 0% and 80% of N in recipient plants coming from donor plants. AMF probably also influence the allocation of N at the community level through effects on rhizospheric functioning; plant growth, uptake and release of N; sequestration of N in soil organic matter (Rillig et al., 2001); soil structure, affecting N mobility (Rillig, 2004; Wilson et al., 2009); as well as transport through common mycorrhizal networks. The relative importance of many of these factors in natural, disturbed and agricultural settings has been reviewed by He et al. (2003). Because of the challenges of assigning the relative 19 importance of different mechanisms and the wide variation among findings no clear consensus has yet emerged about the occurrence or significant direct plant-to-plant N transfer via AMF in natural settings. Interactions between AMF and symbiotic N2-fixing microbes and their contribution to N accumulation in host plants Nitrogen-fixing bacteria colonizing mycorrhizal plants play important roles in N metabolism and movement in the symbiosis (Spriggs and Dakora, 2009). Figure 1-5 illustrates the network of nitrogen movement and recycling in ecosystems from a mycorrhizal perspective. It has been found that the N2-fixing activity of the bacteria can be improved when they are inoculated together with AMF because of the more available nutrients for bacteria by AMF in mycorrhizal than non-mycorrhizal plants (Barea et al., 1980; 2005). Indeed, AMF have been found in root nodules (Scheublin et al., 2004) making it possible that nutrients are transferred between the symbionts without being translocated within the plant vasculature. Nitrogen transferred from symbionts to plants is believed to be predominantly in the form of ammonium in both N2-fixing bacteria and AMF (Day et al., 2001; Rosendahl et al., 2001; Govindarajulu et al., 2005, Jin et al. 2005), although uncertain amounts of Ala and/or other amino acids can also be exported by bacteroids (Waters et al., 1998). Photosynthate in the form of sucrose is the major form of carbon translocated from source leaves to symbiotic roots, although neither AMF nor N-fixing bacteria utilize this directly. Sucrose synthase seems to be involved in making C available to the micro20 symbiont in both arbuscular mycorrhizas and nodules - being activated in both tissues (Hohnjec et al., 2003) - although the AM IRM takes up hexose (Shachar-Hill et al., 1995, Solaiman and Saito, 1997) whereas bacteroids take up organic acids produced by their host cells. Thus, N and C exchange with host plants by the two symbioses have common metabolic intermediates, making it tempting to speculate about the relative costs for plants of obtaining N from one or the other. The metabolic cost of converting N2 to ammonium is one and a half reducing equivalents and eight ATP’s per N atom versus four reducing equivalents for nitrate reduction and none for ammonium taken up directly by AMF. Adding the costs of assimilating and then releasing ammonium in AMF in the pathway described earlier and estimating the additional costs of long distance translocation, appear to make N acquisition via nodules significantly less expensive than via arbuscular mycorrhizas. However, a complete accounting should include the respective costs of development and maintenance of the two symbioses, which is much harder to do (Leake et al., 2004; Kaschuk et al., 2009). Conclusions and perspectives The transformation of inorganic nitrogen taken up from the soil into organic form by the fungus and its translocation within the mycelium as arginine and subsequent conversion to ammonium followed by the release of this inorganic N in colonized root tissues seems to constitute the main pathway of nitrogen movement through the AM symbiosis. Many of the molecular mechanisms involved in this pathway have been identified in recent years lending support to the current model and providing detailed 21 information on genes and proteins involved. Future work in this area may be expected to address significant remaining questions about the regulation, undetermined components of the metabolism and transport machinery, and possible additional routes of N movement in both plant and fungal partners. The work of this thesis contributes to this effort. AMF can increase the uptake of N by host plants from the soil under natural and perturbed conditions and nutrient transfer from AMF can account for significant proportions of N in plants in controlled model systems. Plant growth may be improved by this N transfer although the extent to which this N-transfer-dependent growth enhancement occurs outside the laboratory or greenhouse is unclear. Natural abundance isotope fractionation studies indicate that some N is acquired by plants from AMF and it is to be hoped that experts in this methodology will be able to obtain quantitative estimates of N transfer. The effect of elevated soil N levels on suppressing AM colonization rates is also indicative of a beneficial role for the symbiosis in plant N acquisition. This implies control by the plant host through C allocation, for which there is some support or via altered defense and/or signaling mechanisms which have not apparently been explored. The AM symbiosis can increase the exchange of N between plants and common AM mycelial networks have been shown to contribute to this flow. The extent of this AM mediated exchange seems to be quite variable and its ecological implications in natural, agricultural and disturbed systems remain to be determined. More detailed analyses that go beyond demonstrating exchange to quantify net transfer would be valuable in this context. The interaction between AMF and N2-fixing bacteria can increase both N 22 fixation and colonization by AMF (Stancheva et al., 2008), and the interactions in the tripartite association may play a role in the N cycle. The uptake of N through AMF from the soil to the host plants and the exchange of N between plants via the mycorrhizal networks have potential implications for the application of AMF in sustainable agriculture. Horticultural and on-farm experiments have shown that substantial benefits in yields can be realized from inoculation with AMF (Johansson et al. 2004, Artursson et al., 2006). The extent to which N transfer is important in these cases has been little studied, and it is generally believed that P movement is more important. It may be that more attention to this question would increase the range of settings where AMF are usefully applied. The interaction of AMF and beneficial soil bacteria (both N fixing and non-N fixing) contributes to plant fitness and soil quality, which has been argued to be important for a sustainable agricultural development and ecosystems (Jeffries et al., 2003). Exploration of different combinations of AMF and bacterial inocula in greenhouse and on-farm experiments show promise for increased productivity (Zaidi et al., 2003). Optimal combinations will depend on the particular soil, crop, and cultivation methods (Jeffries et al., 2003) and finding them is likely to benefit from investment in systematically identifying AMF strains in a wide range of different environments. 23 APPENDIX 24 APPENDIX A B Figure 1-1 Arbuscular mycorrhizal fungal morphologies in tomato (Solanum lycopersicum) roots plant. A) Arum type morphology formed by Glomus mosseae, showing an intercellular hypha (IH) and arbuscules (A) formed in adjacent host cells. B) Paris type morphology formed by Glomus coronatum, showing a hyphal coil and growth through rather than between adjacent plant cells. Bars indicate 160um. Modified from Caravagnaro et al., (2001). 25 Figure 1-2 Phylogeny of fungi based on SSU rRNA sequences. Thick lines delineate clades supported by bootstrap values above 90%. Modified from Schussler et al,. (2001) 26 Figure 1-3 Taxonomy of the Glomeromycota containing the arbuscular mycorrhizal and related fungi, based on SSU rRNA gene sequences. Thick lines delineate bootstrap support above 95%, lower bootstrap support values are given on the branches. The four orders for the Glomeromycota are shown in tan. Many of the classical, better studied AM species are in the order Glomerales. The families are shown by pink ovals. Rhizophagus irregularis (previously Glomus intraradices) is in the Glomus group A family. Modified from Schussler et al., (2001). 27 - NO3 Glu - NO3 Putrescine + + NH4 NH4 + NH4 + NH4 Glu Orn + NH4 Gln Arg Urea AA + Arg ERM IRM INTERFACIAL APOPLAST SOIL HOST Figure 1-4 Working model of N transport and metabolism in the AM symbiosis. Inorganic N is taken up by the fungal extraradical mycelium (ERM), the nitrogen is then incorporated into arginine (Arg) in the urea cycle which is translocated to the fungal intraradical mycelium (IRM) in colonized root tissues. Arg is broken down to release ammonium which is exported from the fungus and imported by the host into the root cortical cells. In addition, other forms of nitrogen may be transferred from the ERM to the IRM and transferred to or exchanged with the host. Glu, glutamate; Gln, glutamine, Orn, ornithine; AA, amino acid Modified from Tian et al., (2010). 28 Figure 1-5 Interactions among arbuscular mycorrhizal fungi, host plants, and N2-fixing bacteria and other soil microbes in nitrogen cycling. arbuscular mycorrhizal fungi take up inorganic nitrogen from the soil through the extraradical mycelium, assimilate and translocate it into colonized roots where it is transferred to the host. Nitrogen fixed by symbiotic N2-fixing bacteria in root nodules and nitrogen taken up directly by the plant root itself from the soil also contribute to plant nitrogen nutrition. N2-fixing and non-N2-fixing plants are connected by common mycorrhizal networks, allowing exchange and net movement of N between plants. 29 Table 1-1 Nitrogen metabolic proteins and genes identified in the arbuscular mycorrhizal symbiosis. Proteins and/or genes reported Fungal amino acid permease (GmosAAP1 from G. mosseae) Fungal nitrate transporter (GiNT from R. irregularis) Fungal ammonium transporters (GiAMT and GintAMT1 from R. irregularis) Plant ammonium transporters (MtAMT from M. truncatula; LjAMT2;2 from Lotus japonicus; GmAMT4.1 from Glycine max) Fungal (R. irregularis) and Plant (Zea mays) Nitrate reductases. Potential function for N transfer in the AM symbiosis Amino acid acquisition by fungus from the soil. Cappellazzo et al., (2008) characterized a permease that facilitates the uptake of proline and is induced in the presence of several N sources. Nitrate uptake from the soil. Transcript levels upregulated by nitrate addition to the ERM (Tian et al., 2010). Ammonium uptake from soil and transfer to host. Two fungal AMT ammonium transporters reported: one is expressed more highly in the IRM (GiAMT, Govindarajulu et al., 2005) the other (GiAMT1, López-Pedrosa et al., 2006) is induced in the ERM in response to NH4+. Ammonium/ammonia uptake by plant from the host/fungus interface. Plant AMT transporters are induced in mycorrhizal roots of three species (MtAMT, Gomez et al., 2009; LjAMT2;2, Guether et al., 2009; GmAMT4.1, Kobae et al., 2010). LjAMT2;2 and GmAMT4.1 were localized to the periarbuscular membrane of arbusculated cortical cells. Nitrate assimilation in mycorrhizal roots. The mRNA level of maize NR was lower in roots and shoots of mycorrhizal plants than in noncolonized controls, and fungal NR transcripts were localized in the IRM. Suggests that the fungal nitrate reductase assimilated nitrate in AM roots (Kaldorf et al., 1998). 29 Table 1-1 (cont’d) Fungal (R. irregularis) and Plant (Zea mays) Nitrate reductases. Fungal glutamine synthetase (GmGln1 from G. mosseae; GiGln1 (same as GiGS1) and GiGS2 from R. irregularis) and glutamate synthase (GiGluS from R. irregularis) Plant (Daucus carota) glutamine synthetase genes Nitrate assimilation in mycorrhizal roots. The mRNA level of maize NR was lower in roots and shoots of mycorrhizal plants than in noncolonized controls, and fungal NR transcripts were localized in the IRM. Suggests that the fungal nitrate reductase assimilated nitrate in AM roots (Kaldorf et al., 1998). N assimilation in the ERM. Breuninger et al., (2004) identified GS homologs (GiGln1 and GmGln1) in two fungal species expressed in all AM tissues. Activity was elevated after N addition to the ERM. Govindarajulu et al., (2005) reported induction of the R. irregularis gene (GiGS1) in the ERM in response to N. Gomez et al., (2009) reported the expression of GiGS1 in mycorrhizal roots. Tian et al., (2010) found a second GS (GiGS2) in R. irregularis and showed that the two GiGS genes are differently upregulated by N addition to the ERM and have different kinetic properties. Assimilation by plant of fungal-derived N. One of three D. carota GS’s reported by Higashi et al., (1998) is upregulated following the import of N into mycorrhizal roots by the fungus (Tian et al., 2010). 30 Table 1-1 (cont’d) Fungal Arginine synthesis: Carbamoyl-phosphate synthase glutamine chain (GiCPS); Argininosuccinate synthase (GiASS); Arginosuccinate lyase (GiAL); (from R. irregularis) Fungal Arginine breakdown Arginase (GiCAR1); Urease accessory protein (GiUAP); Urease (GiURE); Ornithine aminotransferases (GiOAT1 and GiOAT2); Ornithine decarboxylase (GiODC); from R. irregularis) Arginine synthesis in the ERM. Gomez et al., (2009) reported the expression of GiASS in mycorrhizal roots. Tian et al., (2010) showed that these arginine synthesis genes are induced in the ERM but not in the IRM after N addition to the ERM. N release in the IRM. Govindarajulu et al., (2005) reported higher expression of GiUAP in the IRM than the ERM. Gomez et al., (2009) showed transcription of R. irregularis arginase, and GiOAT1 in cortical cells of mycorrhizal M. truncatula. Tian et al., (2010) reported the upregulation of GiCAR1, GiURE, GiOAT1, GiOAT2 and GiODC in the IRM but not ERM after N addition to the ERM. 31 REFERENCES 32 REFERENCES Akiyama K, Matsuzaki K. Hayashi H. 2005. Plant sesquiterpenes induce hyphal branching in arbuscular mycorrhizal fungi. Nature 435: 824–827. Allen JW, Shachar-Hill Y. 2009. Sulfur transfer through an Arbuscular Mycorrhiza. Plant Physiology 149: 549-560. Alves da Silva G, Lumini E, Costa Maia L, Bonfante P, Bianciotto V. 2006. Phylogenetic analysis of Glomeromycota by partial LSU rDNA sequences. Mycorrhiza 16:183-189. Ames RN, Reid CP, Porter LK, Cambardella C. 1983. Hyphal uptake and transport of nitrogen from two 15N-labelled sources by Glomus mosseae, a vesiculararbuscular mycorrhizal fungus. New Phytologist 95: 381-396. Artursson V, Finlay RD, Jansson JK. 2006. Interactions between arbuscular mycorrhizal fungi and bacteria and their potential for stimulating plant growth. Environmental Microbiology 8: 1-10. Azcón-Aguilar C, Bago B, Barea JM. 1999. Saprophytic growth of AMF.Mycorrhiza: Structure, Function, Molecular Biology and Biotechnology Ed 2 Varma A, Hock B Springer-Verlag, Berlin pp. 391–407. Bago B, Pfeffer PE, Douds DD, Brouillette J, Be´card 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: 263–271. Bago B, Vierheilig H, Piché Y, Ázcon-Aguilar C. 1996. Nitrate depletion and pH changes induced by the extraradical mycelium of the arbuscular mycorrhizal fungus Glomus intraradices grown in monoxenic culture. New Phytologist 133: 273-280. Bago B, Pfeffer P, Shachar-Hill Y. 2000. Carbon metabolism and transport in arbuscular mycorrhizas. Plant Physiology 124: 949-957. Bago B, Pfeffer P, Shachar-Hill Y. 2001. Could the urea cycle be translocating nitrogen in the arbuscular mycorrhizal symbiosis? New Phytologist 149: 4-8. Bago B, Cano C, Azcón-Aguilar C, Samson J, Coughlan AP, Piché Y. 2004. Differential morphogenesis of the extraradical mycelium of an arbuscular mycorrhizal fungus grown monoxenically on spatially heterogeneous culture media. Mycologia 96: 452-462. Bardin S, Dan S, Osteras M, Finan TM. 1996. A phosphate transport system is required for symbiotic nitrogen fixation by Rhizobium meliloti. Journal of Bacteriology 178: 4540-4547. Barea JM, Escudero JL, Azcongdeaguilar C. 1980. Effects of introduced and indigenous VA mycorrhizal fungi on nodulation, growth and nutrition of Medicago 33 sativa in phosphate-fixing soils as affected by fertilizers. Plant and Soil 54: 283296. Barea JM, Pozo MJ, Azco´n R, Azco´n-Aguilar C. 2005. Microbial co-operation in the rhizosphere. Journal of Experimental Botany 56:1761-1778. Bècard G, Fortin JA. 1988. Early events of vesicular-arbuscular mycorrhizal formation on Ri T-DNA transformed roots. New Phytologist 108:211–218. Bècard G, Doner LW, Rolin DB, Douds DD, Pfeffer PE (1991) Identification and quantification of trehalose in vesicular-arbuscular mycorrhizal fungi by in vivo 13C NMR and HPLC analyses. New Phytologist 118: 547–552 Bécard G, Pfeffer, PE. 1993. Status of nuclear division in arbuscular mycorrhizal fungi during in vitro development. Protoplasma, 194: 62-68 Bethlenfalvay GJ, Reyes-Soils MG, Camel SB, Ferrera-Cerrato R. 1991. Nutrient transfer between the root zones of soybean and maize plants connected by a common mycorrhizal mycelium. Physiologia Plantarum 82: 423–432. Bianciotto V, Bonfante P. 1992. Quantification of the nuclear DNA content of two arbuscular mycorrhizal fungi. Mycological Research 96:1071-1076. Bianciotto V, Barbiero G, Bonfante P. 1995. Analysis of the cell cycle in an arbuscular mycorrhizal fungus by flow cytometry and bromodeoxyuridine labelling. Protoplasma 188:161-169. Blanke V, Renker V, Wagner M, Füllner 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: 981-992. Bolan, NS. 1991. A critical review on the role of mycorrhizal fungi in the uptake of phosphorus by plants. Plant and Soil 134: 189-207. Bonneau L, Huguet S, Wipf D, Pauly N, Truong HN. 2013. Combined phosphate and nitrogen limitation generates a nutrient stress transcriptome favorable for arbuscular mycorrhizal symbiosis in Medicago truncatula. New Phytologist 199:188–202 Breuninger M, Trujillo CG, Serrano E, Fischer R, Requena N. 2004. Different nitrogen sources modulate activity but not expression of glutamine synthetase in arbuscular mycorrhizal fungi. Fungal Genetics and Biology 41: 542-552. Brundrett M, Murase G, Kendrick B. 1990. Comparative anatomy of roots and mycorrhizae of common Ontario trees. Canadian Journal of Botany 68: 551–578. Brundrett MC, Piche´ Y, Peterson RL. 1985. A developmental study of the early stages in vesicular-arbuscular mycorrhizal formation. Canadian Journal of Botany 63: 184–194. Bücking 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: 899-912 34 Buee M, Rossignol M, Jauneau A, Ranjeva R, Bècard G. 2000. The pre-symbiotic growth of arbuscular mycorrhizal fungi is induced by a branching factor partially purified from plant root exudates. Molecular Plant Microbe Interaction 13: 693– 698. Cappellazzo G, Lanfranco L, Fitz M, Wipf D, Bonfante P. 2008. Characterization of an amino acid permease from the endomycorrhizal fungus Glomus mosseae. Plant Physiology 147: 429-437. Chalot M, Brun A, Finlay RD, Soderstrom B. 1994. Metabolism of [14C] glutamine by the ectomycorrhizal fungus Paxillus involutus. Microbiology 140: 1641-1649. Chalot M, Blaudez D, Brun A. 2006. Ammonia: a candidate for nitrogen transfer at the mycorrhizal interface. Trends in Plant Science 11: 263-266. Clark RB, Zeto SK. 2000. Mineral acquisition by arbuscular mycorrhizal plants. Journal of Plant Nutrition 23: 867-902. Cliquet JB, Stewart CR. 1993. Ammonium assimilation in Zea mays 1. infected with a vesicular-arbuscular mycorrhizal fungus Glomus fasciculatum. Plant Physiology 101: 865-871. Constable JV, Bassirirad H, Lussenhop J, Zerihun A. 2001. Influence of elevated CO2 and mycorrhizae on nitrogen acquisition: contrasting responses in Pinus taeda and Liquidambar styraciflua. Tree Physiology 21: 83-91. Cruz C, Egsgaard H, Trujillo C, Ambus P, Requena N, Martins-Loucao MA, Jakobsen I. 2007. Enzymatic evidence for the key role of arginine in nitrogen translocation by arbuscular mycorrhizal fungi. Plant Physiology 144: 782-792. Day DA, Poole PS, Tyerman SD, Rosendahl L. 2001. Ammonia and amino acid transport across symbiotic membranes in nitrogen-fixing legume nodules. Cellular and Molecular Life Sciences 58: 61-71. Dürr M, Urech K, Boller T, Wiemken A, Schwencke J, Nagy M. 1979. Sequestration of arginine by polyphosphate in vacuoles of yeast (Saccharomyces cerevisiae). Archives of Microbiology 121: 169-175. Egerton-Warburton LM, Allen EB. 2000. Shifts in arbuscular mycorrhizal fungal communities along an anthropogenic nitrogen deposition gradient. Ecological Applications 10: 484-496. Faure S, Cliquet JB, Thephany G, Boucaud J. 1998. Nitrogen assimilation in Loliun perenne colonized by the arbuscular mycorrhizal fungus Glonus fasciculatum. New Phytologist 138: 411-417. Frey B, Schuepp H. 1993. A role of vesicular–arbuscular (VA) mycorrhizal fungi in facilitating interplant nitrogen transfer. Soil Biology and Biochemistry 25: 651– 658. Gachomo E, Allen JW, Pfeffer PE, Govindarajulu M, Douds DD, Jin HR, 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 184: 399-411. 35 Gallaud J. 1905. Etude sur les mycorrhizes endotrophes. Revue Generale de Botanique 17: 5–48, 66–83, 123–136, 223–249, 313–325, 425–433, 479–500. Gamper H, Hartwig UA, Leuchtmann A. 2005. Mycorrhizas improve nitrogen nutrition of Trifolium repens after 8 yr of selection under elevated atmospheric CO2 partial pressure. New Phytologist 167: 531-542. Gray, LE and Gerdemann, JW. 1973. Uptake of sulphur-35 by vesicular-arbuscular mycorrhizae. Plant Soil 39: 687—689. George E, Marschner H, Jakobsen I. 1995. role of arbuscular mycorrhizal fungi in uptake of phosphorus and nitrogen from soil Critical Reviews in Biotechnology. 15: 257-270. Giovannetti M, Sbrana C, Avio L, Citernesi AS, Logi C. 1993. Differential hyphal morphogenesis in arbuscular mycorrhizal fungi during pre-infection stages. New Phytologist 125:587–593. Gomez SK, Javot H, Deewatthanawong P, Torres-Jerez I, Tang YH, Blancaflor EB, Udvardi MK, Harrison MJ. 2009. Medicago truncatula and Glomus intraradices gene expression in cortical cells harboring arbuscules in the arbuscular mycorrhizal symbiosis. BMC Plant Biology 9: 10. Govindarajulu M, Pfeffer PE, Jin HR, Abubaker J, Douds DD, Allen JW, Bucking H, Lammers PJ, Shachar-Hill Y. 2005. Nitrogen transfer in the arbuscular mycorrhizal symbiosis. Nature 435: 819-823. Gryndler M, Hrselová H, Cajthaml T, Havránková M, Rezácová V, Gryndlerová H, Larsen J. 2009. Influence of soil organic matter decomposition on arbuscular mycorrhizal fungi in terms of asymbiotic hyphal growth and root colonization. Mycorrhiza 19: 255-266. Guether M, Neuhauser B, Balestrini R, Dynowski M, Ludewig U, Bonfante P. 2009. A mycorrhizal-specific ammonium transporter from Lotus japonicus acquires nitrogen released by arbuscular mycorrhizal fungi. Plant Physiology 150: 73-83. Haines, BL and Best GR. 1976. Glomus mosseae, endomycorrhizal with Liquidambar styraciflua L. seedlings retards NO3, NO2, and NH4 nitrogen loss from a temperate forest soil. Plant Soil 45:257–261 Hamel C. 2004. Impact of arbuscular mycorrhizal fungi on N and P cycling in the root zone. Canadian Journal of Soil Science 84: 383-395. Harrison MJ. 1999. Molecular and cellular aspects of the arbuscular mycorrhizal symbiosis. Annual Reviews of Plant Physiology and Plant Molecular Biology 50: 361-389. Hart MM, Forsythe J. 2012. Using arbuscular mycorrhizal fungi to improve the nutrient quality of crops; nutritional benefits in addition to phosphorus. Scientia Horticulturae 148: 206–214. Hawkins HJ, Johansen A, George E. 2000. Uptake and transport of organic and inorganic nitrogen by arbuscular mycorrhizal fungi. Plant and Soil 226: 275–285. 36 He XH, Critchley C, Bledsoe C. 2003. Nitrogen transfer within and between plants through common mycorrhizal networks (CMNs). Critical Reviews in Plant Sciences 22: 531-567. He XH, Xu MG, Qiu GY, Zhou JB. 2009. Use of 15N stable isotope to quantify nitrogen transfer between mycorrhizal plants. Journal of Plant Ecology-UK 2: 107-118. Higashi K, Shiota H, Kamada H. 1998. Patterns of expression of the genes fro glutamine synthetase isoforms during somatic and zygotic embryogenesis in carrot. Plant and Cell Physiology 39: 418-424. Hinnebusch AG. 1988. Mechanisms of gene regulation in the general control of amino acid biosynthesis in Saccharomyces cerevisiae. Microbiological Reviews 52: 248-273. Hobbie EA, Macko SA. 2000. Correlations between foliar d15N and nitrogen concentrations may indicate plant-mycorrhizal interactions. Oecologia 122: 273283. Hobbie EA, Clopaert JV. 2003. Nitrogen availability and colonization by mycorrhizal fungi correlate with nitrogen isotope patterns in plants. New Phytologist 157: 115126. Hobbie EA, Jumpponen A, Trappe J. 2005. Foliar and fungal 15N:14N ratios reflect development of mycorrhizae and nitrogen supply during primary succession: testing analytical models. Oecologia 146: 258-268. Hobbie EA, Hobbie JE. 2008. Natural abundance of 15N in nitrogen-limited forests and tundra can estimate nitrogen cycling through mycorrhizal fungi: a review. Ecosystems 11: 815-830. Hodge A, Campbell CD, Fitter AH. 2001. An arbuscular mycorrhizal fungus accelerates decomposition and acquires nitrogen directly from organic material. Nature 413: 297-299. Hodge A, Fitter AH. 2010. Substantial nitrogen acquisition by arbuscular mycorrhizal fungi from organic material has implications for N cycling. Proceedings of the National Academy of Sciences USA 107: 13754-13759. Hohnjec N, Perlick AM, Puhler A, Kuster H. 2003. The Medicago truncatula sucrose synthase gene MtSucS1 is activated both in the infected region of root nodules and in the cortex of roots colonized by arbuscular mycorrhizal fungi. Molecular Plant-Microbe Interactions 16: 903-915. Jackson LF, Miller D, Smith SE. 2001. Arbuscular mycorrhizal colonization and growth of wild and cultivated lettuce in response to nitrogen and phosphorus. Scientia Horticulturae 94: 205-218. Jackson LE, Burger M, Cavagnaro TR. 2008. Roots, nitrogen transformations, and ecosystem services. Annual Review of Plant Biology 59: 341-363. 37 Jakobsen I; Abbott LK, Robson AD. 1992. external hyphae of vesicular-arbuscular mycorrhizal fungi associated with trifolium-subterraneum l .1. spread of hyphae and phosphorus inflow into roots. New Phytologist 120: 371-380. Jakobsen I, Rosendahl L. 1990. Carbon flow into soil and external hyphae from roots of mycorrhizal cucumber plants. New Phytologist 115: 77-83. Jalonen R, Nygren P, Sierra J. 2009. Transfer of nitrogen from a tropical legume tree to an associated fodder grass via root exudation and common mycelial networks. Plant, Cell and Environment 32: 1366-1376. Javelle A, Andre B, Marini AM, Chalot M. 2003. High-affinity ammonium transporters and nitrogen sensing in mycorrhizas. Molecular Microbiology 47: 411-430. 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 USA 104: 1720– 1725 Javot H, Penmetsa RV, Breuillin F, Bhattarai KK, Noar RD, Gomez SK, Zhang Q, Cook DR, Harrison MJ. 2011. Medicago truncatula mtpt4 mutants reveal a role for nitrogen in the regulation of arbuscule degeneration in arbuscular mycorrhizal symbiosis. Plant Journal. 68(6):954-65. Jeffries P, Gianinazzi S, Perotto S, Turnau K, Barea JM. 2003. The contribution of arbuscular mycorrhizal fungi in sustainable maintenance of plant health and soil fertility. Biology and Fertility of Soils 37: 1-16. Jennings DH. 1995. The Physiology of Fungal Nutrition, Cambridge University Press, Cambridge. Jia Y, Gray VM, Straker CJ. 2004. The influence of Rhizobium and arbuscular mycorrhizal fungi on nitrogen and phosphorus accumulation by Vicia faba. Annals of Botany 94: 251-258. 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. Johansen A, Jakobsen I, Jensen ES. 1993. Hyphal transport by a vesiculararbuscular mycorrhizal fungus of N applied to the soil as ammonium or nitrate. Biology and Fertility of Soils 16: 66-70. Johansen A, Finlay RD, Olsson PA. 1996. Nitrogen metabolism of external hyphae of the arbuscular mycorrhizal fungus Glomus intraradices. New Phytologist 133: 705-712. Johansen A, Jensen ES. 1996. Transfer of N and P from intact or decomposing roots of pea to barley interconnected by an arbuscular mycorrhizal fungus. Soil Biology and Biochemistry 28: 73–81. 38 Johansson JF, Paul LR, Finlay RD. 2004. Microbial interactions in the mycorrhizosphere and their significance for sustainable agriculture. FEMS Microbiology Ecology 48: 1-13. Johnson NC, Rowland DL, Corkidi L, Egerton-Warburton LM, Allen EB. 2003. Nitrogen enrichment alters mycorrhizal allocation at five mesic to semiarid grasslands. Ecology 84: 1895-1908. Kaldorf M, Schmelzer E, Bothe H. 1998. Expression of maize and fungal nitrate reductase genes in the arbuscular mycorrhiza. Molecular Plant-Microbe Interactions 11: 439-448. Kaschuk G, Leffelaar PA, Giller KE, Alberton O, Hungria M, Kuyper TW. 2009. Responses of legumes to rhizobia and arbuscular mycorrhizal fungi: A metaanalysis of potential photosynthate limitation of symbioses. Soil Biology and Biochemistry 42: 125- 127. Kim K, Yim W, Trivedi P, Madhaiyan M, Deka Boruah HP, Islam MdR, Lee G, Sa T 2010. Synergistic effects of inoculating arbuscular mycorrhizal fungi and Methylobacterium oryzae strains on growth and nutrient uptake of red pepper (Capsicum annuum L.). Plant and Soil 327: 429–440. Kobae Y, Tamura Y, Takai S, Banba M, Hata S. 2010. Localized expression of arbuscular mycorrhiza-inducible ammonium transporters in soybean. Plant and Cell Physiology 51: 1411-1415. Krüger M, Krüger C, Walker C, Stockinger H, Schüßler A. 2012. Phylogenetic reference data for systematics and phylotaxonomy of arbuscular mycorrhizal fungi from phylum to species level. New Phytologist 193: 970–984. Krusell L, Krause K, Ott T, Desbrosses G, Krämer U, Sato S, Nakamura Y, Tabata S, James EK, Sandal N, Stougaard J, Kawaguchi M, Miyamoto A, Suganuma N, Udvardi M. 2005. The Sulfate Transporter SST1 is crucial for symbiotic nitrogen fixation in Lotus japonicus root nodules. The Plant Cell 17: 1625-1636. Lammers PJ, Jun J, Abubaker J, Arreola R, Gopalan A, Bago B, HernandezSebastia C, Allen JW, Douds DD, Pfeffer PE. et al. 2001. The glyoxylate cycle in an arbuscular mycorrhizal fungus: gene expression and carbon flow. Plant Physiology 127:1287–1298. Leake JR, Johnson D, Donnelly DP, Muckle GE, Boddy L, Read DJ. 2004. Networks of power and influence: the role of mycorrhizal mycelium in controlling plant communities and agroecosystem functioning. Canadian Journal of Botany 8: 1016–1045. 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: 199-207. López-Pedrosa A, González-Guerrero M, Valderas A, Azcón-Aguilar C, Ferrol N. 2006. GintAMT1 encodes a functional high-affinity ammonium transporter that is 39 expressed in the extraradical mycelium of Glomus intraradices. Fungal Genetics and Biology 43: 102-110. Lucic E, Fourrey C, Kohler A, Martin F, Chalot M, Brun-Jacob A. 2008. A gene repertoire for nitrogen transporters in Laccaria bicolor. New Phytologist 180: 343364. Martin F. 1985. 15N-NMR studies of nitrogen assimilation and amino acid biosynthesis in the ectomycorrhizal fungus Cenococcum graniforme. FEBS Letters. 182: 350–354. Marschner H, Dell B. 1994. Nutrient uptake in mycorrhizal symbiosis. Plant and Soil 159: 89–102 McNeill AM and Unkovich MJ. 2007. The Nitrogen Cycle in Terrestrial Ecosystems In “Nutrient Cycling in Terrestrial Ecosystems”, Ed: Z. Rengel and P. Marschner Springer Verlag, pp. 37-64, Berlin Heidelberg, Germany. Morel M, Buée M, Chalot M, Brun A. 2006. NADP-dependent glutamate dehydrogenase: a dispensable function in ectomycorrhizal fungi. New Phytologist 169: 179-190. Mortimer PE, Perez-Fernandez MA, Valentine AJ. 2008. The role of arbuscular mycorrhizal colonization in the carbon and nutrient economy of the tripartite symbiosis with nodulated Phaseolus vulgaris. Soil Biology and Biochemistry 40: 1019-1027. Msiska Z, Morton J. 2009. Phylogenetic analysis of the Glomeromycota by partial βtubulin gene sequences. Mycorrhiza 19: 247–254. Nagy R, Drissner D, Amrhein N, Jakobsen I, Bucher M. 2009. Mycorrhizal phosphate uptake pathway in tomato is phosphorus-repressible and transcriptionally regulated. New Phytologist 181: 950–959 Olsson PA, Burleigh SH, van Aarle IM. 2005. The influence of external nitrogen on carbon allocation to Glomus intraradices in monoxenic arbuscular mycorrhiza. New Phytologist 168: 677-686. Olsson PA, Rahm J, Aliasgharzad N. 2010. Carbon dynamics in mycorrhizal symbioses is linked to carbon costs and phosphorus benefits. FEMS Microbiology Ecology 72: 123-131. Persson J, Högberg P, Ekblad A, Högberg MN, Nordgren A, Näsholm T. 2003. Nitrogen acquisition from inorganic and organic sources by boreal forest plants in the field. Oecologia 137: 252-257. Peterson RL And Bonfante P. 1994. Comparative structure of vesicular-arbuscular mycorrhizas and ectomycorrhizas. Plant and Soil 159:79-88. Pfeffer PE, Douds DD, Bécard G, Shachar-Hill Y. 1999. Carbon uptake and the metabolism and transport of lipids in and arbuscular mycorrhiza. Plant Physiology 120:587–598. 40 Pierleoni R, Vallorani L, Sacconi C, Sisti D, Giomaro G, Stocchi V. 2001. Evaluation of the enzymes involved in primary nitrogen metabolism in Tilia platyphyllos-Tuber borchii ectomycorrhizae. Plant Physiology 39: 1111-1114. Pumplin N, Harrison MJ. 2009. Live-cell imaging reveals periarbuscular membrane domains and organelle location in Medicago truncatula roots during arbuscular mycorrhizal symbiosis. Plant Physiology 151: 809-819. Redecker D, Raab P. 2006. Phylogeny of the Glomeromycota (arbuscular mycorrhizal fungi): recent developments and new gene markers. Mycologia 98: 885–895. Redecker D, Schüßler A, Stockinger H, Stürmer SL, Morton JB, Walker C. 2013. An evidence-based consensus for the classification of arbuscular mycorrhizal fungi (Glomeromycota). Mycorrhiza 23: 515-53. Reich PB, Hobbie SE, Lee T, Ellsworth DS, West JB, Tilman D, Knops JMH, Naeem S, Trost J. 2006. Nitrogen limitation constrains sustainability of ecosystem response to CO2. Nature 440: 922-925. Reynolds HL, Hartley AE, Vogelsang KM, Bever JD, Schultz PA. 2005. Arbuscular mycorrhizal fungi do not enhance nitrogen acquisition and growth of old-field perennials under low nitrogen supply in glasshouse culture. New Phytologist 167: 869-880. Rhodes LH and Gerdemann JW. 1975. Phosphate uptake zones of mycorrhizal and non-mycorrhizal onions. New Phytologist 75: 555-561 Rillig MC, Wright SF, Nichols KA, Schmidt WF, Torn MS. 2001. Large contribution of arbuscular mycorrhizal fungi to soil carbon pools in tropical forest soils. Plant and Soil 233: 167-177. Rillig MC. 2004. Arbuscular mycorrhizae and terrestrial ecosystem processes. Ecology Letters 7: 740-754. Rosendahl L, Mouritzen P, Rudbeck A. 2001. Nitrogen transfer in the interface between the symbionts in pea root nodules. Plant and Soil 230: 31-37. Scheublin TR, Ridgway KP, Young JPW, van der Heijden MGA. 2004. Nonlegumes, legumes, and root nodules harbor different arbuscular mycorrhizal fungal communities. Applied and Environmental Microbiology 70: 6240-6246. Schüßler A, Schwarzott D, Walker C. 2003. A new fungal phylum, the Glomeromycota: phylogeny and evolution. Mycological Research 105:1413– 1421. Schüßler A, Walker C. 2010. The Glomeromycota: a species list with new families and genera. Edinburgh & Kew, UK: The Royal Botanic Garden; Munich, Germany: Botanische Staatssammlung Munich; Oregon, USA: Oregon State University. Seres A, Bakonyi G, Posta K. 2006. Zn uptake by maize under the influence of AMfungi and Collembola Folsomia candia. Ecological Research 21: 692–697. 41 Shachar-Hill Y, Pfeffer PE, Douds D, Osman SF, Doner LW, Ratcliffe RG. 1995. Partitioning of intermediate carbon metabolism in VAM colonized leek. Plant Physiology 108:7–15. Sieh D, Watanabe M, Devers EA, Brueckner F, Hoefgen R, and Krajinski F. 2013. The arbuscular mycorrhizal symbiosis influences sulfur starvation responses of Medicago truncatula. New Phytologist 197: 606–616 Sims AP, Folkes BF. 1964. A kinetic study of the assimilation of 15N-ammonia and the synthesis of amino acids in an exponentially growing culture of Candida utilis. Smith SE, and Read DJ. 1997. Mycorrhizal Symbioses. 2nd edn Academic Press, London. Smith SE, Read DJ. 2008 . Mycorrhizal symbiosis. 3rd edn. Academic Press, London Smith SE, Smith FA. 2011. Roles of arbuscular mycorrhizas in plant nutrition and growth: new paradigms from cellular to ecosystem scales. Annual Review of Plant Biology. 62:227-50. Smith SE, Smith FA, Jakobsen I. 2003. Mycorrhizal fungi can dominate phosphate supply to plants irrespective of growth responses. Plant Physiology 133: 16-20. Solaiman MD, Saito M. 1997. Use of sugars by intraradical hyphae of arbuscular mycorrhizal fungi revealed by radiorespirometry. New Phytologist 136:533–538. Spriggs AC, Dakora FD. 2009. Field assessment of symbiotic N2 fixation in wild and cultivated Cyclopia species in the South African fynbos by 15N natural abundance. Tree Physiology 29: 239-247. Stancheva I, Geneva M, Djonova E, Kaloyanova N, Sichanova M, Boychinova M, Georgiev G. 2008. Response of alfalfa (Medicago sativa L) growth at low accessible phosphorus source to the dual inoculation with mycorrhizal fungi and nitrogen fixing bacteria. General and Applied Plant Physiology 34: 319-326. Stockinger H, Peyret-Guzzon M, Koegel S, Bouffaud ML, Redecker D. 2014. The largest subunit of RNA polymerase II as a new marker gene to study assemblages of arbuscular mycorrhizal fungi in the field. PLoS One. 9(9):e107783. Streeter JG. 1985. Nitrate inhibition of Legume nodule growth and activity I. Long term studies with a continuous supply of nitrate. Plant Physiology 77: 321-324. Tanaka Y, Yano K. 2005. Nitrogen delivery to maize via mycorrhizal hyphae depends on the form of N supplied. Plant, Cell and Environment 28: 1247-1254. ter Schure EG, van Riel NAW, Verrips CT. 2000. The role of ammonia metabolism in nitrogen catabolite repression in Saccharomyces cerevisiae. FEMS Microbiology Reviews 24: 67-83. Tian CJ, Kasiborski B, Koul R, Lammers PE, Bücking H, Shachar-Hill Y. 2010. Regulation of the nitrogen transfer pathway in the arbuscular mycorrhizal 42 symbiosis: gene characterization and the coordination of expression with nitrogen flux. Plant Physiology 153: 1175-1187. Tinker PB, Nye PH. 2000. Solute Movement in the Rhizosphere. Oxford University Press, Oxford. Tisserant E, Kohler A, Dozolme-Seddas P, Balestrini R, Benabdellah K, Colard A, Croll D, Da Silva C, Gomez SK, Koul R, Ferrol N, Fiorilli V, Formey D, Franken P, Helber N, Hijri M, Lanfranco L, Lindquist E, Liu Y, Malbreil M, Morin E, Poulain J, Shapiro H, van Tuinen D, Waschke A, Azcon-Aguilar C, Bécard G, Bonfante P, Harrison MJ, Küster H, Lammers P, Paszkowski U, Requena N, Rensing SA, Roux C, Sanders IR, Shachar-Hill Y, Tuskan G, Young JPW, Gianinazzi-Pearson V, Martin F. 2012. The transcriptome of the arbuscular mycorrhizal fungus Glomus intraradices (DAOM 197198) reveals functional tradeoffs in an obligate symbiont. New Phytologist. 193: 755-769. Tisserant E, Malbreil M, Kuo A, Kohler A, Symeonidi A, Balestrini R, Charron P, Duensing N, Frey NFD, Gianinazzi-Pearson V, Gilbert LB, Handa Y, Herr JR, Hijri M, Koul R, Kawaguchi M, Krajinski F, Lammers PJ, Masclauxm FG, Murat C, Morin E, Ndikumana S, Pagni M, Petitpierre D, Requena N, Rosikiewicz P, Riley R, Saito K, Clemente HS, Shapiro H, Van Tuinen D, Bécard G, Bonfante P, Paszkowski U, Shachar-Hill YY, Tuskan GA, Young PW, Sanders IR, Henrissat B, Rensing SA, Grigoriev IV, Corradi N, Roux C, Martin F. 2013. Genome of an arbuscular mycorrhizal fungus provides insight into the oldest plant symbiosis. Proceedings of the National Academy of Sciences USA 110: 20117-20122. Tobar R, Azcon R, Barea, JM. 1994. improved nitrogen uptake and transport from n-15-labeled nitrate by external hyphae of arbuscular mycorrhiza under waterstressed conditions. New Phytologist 126: 119-122. Tobar RM, Azcón R, Barea JM. 2004. The improvement of plant N acquisition from an ammonium-treated, drought-stressed soil by the fungal symbiont in arbuscular mycorrhizae. Mycorrhiza 4: 105-108. Toler HD, Morton JB, Cumming JR. 2005. Growth and metal accumulation of mycorrhizal sorghum exposed to elevated copper and zinc. Water, Air, and Soil Pollution 164:155–172 Toussaint JP, St-Arnaud M, Charest C. 2004. Nitrogen transfer and assimilation between the arbuscular mycorrhizal fungus Glomus intraradices Schenck & Smith and Ri T-DNA roots of Daucus carota L., in an in vitro compartmented system. Canadian Journal of Microbiology 50: 251-260. Treseder KK. 2004. A meta-analysis of mycorrhizal responses to nitrogen, phosphorus, and atmospheric CO2 in field studies. New Phytologist 164: 347355. van Aarle IM, Cavagnaro TR, Smith SE, Smith FA, Dickson S. 2005. Metabolic activity of Glomus intraradices in Arumand Paris-type arbuscular mycorrhizal colonization. New Phytologist 166:611–618 43 Waters JK, Hughes BL, Purcell LC, Gerhardt KO, Mawhinney TP, Emerich DW. 1998. Alanine, not ammonia, is excreted from N2-fixing soybean nodule bacteroids. Proceedings of the National Academy of Sciences USA 95: 1203812042. Wilson GW, Rice CW, Rillig MC, Springer A, Hartnett DC. 2009. Soil aggregation and carbon sequestration are tightly correlated with the abundance of arbuscular mycorrhizal fungi: results from long-term field experiments. Ecology Letters 12: 452-461. Wang B, Qiu YL. 2006. Phylogenetic distribution and evolution of mycorrhizas in land plants. Mycorrhiza 16: 299-363. Yoneyama K, Xie XN, Kusumoto D, Sekimoto H, Sugimoto Y, Takeuchi Y, Yoneyama K. 2007. Nitrogen deficiency as well as phosphorus deficiency in sorghum promotes the production and exudation of 5-deoxystrigol, the host recognition signal for arbuscular mycorrhizal fungi and root parasites. Planta 227: 125-132. Zaidi A, Khan MS, Amil M. 2003. Interactive effect of rhizotrophic microorganisms on yield and nutrient uptake of chickpea (Cicer arietinum L.). European Journal of Agronomy 19: 15-21. 44 Chapter 2 Nitrogen uptake and transfer in an aseptic whole plant-mycorrhizal culture system 45 Abstract Arbuscular mycorrhizal fungi (AMF) can increase the uptake of N by their host plants and play a significant role in the transformation and movement of N in plant communities. It is widely assumed that N transfer to host plant roots, which has been observed in tracer experiments, can improve host plant growth, but no unambiguous demonstration of this has been reported. To test whether N uptake via an AM fungus can enhance plant growth and fitness, a compartmented microcosm was developed for aseptic and leakage free whole-plant mycorrhizal experiments. This was used to monitor S uptake by the fungal extraradical mycelium (ERM) and its transfer to host plants. Our results indicated rapid S transfer by ERM to the host plants as indicated by high 35S levels in roots and shoots. To assess the contribution of the AM fungal mycelium to plant nitrogen nutrition, N was added to the fungal compartment of mycorrhizal plants. Controls were used to ensure that N transfer was responsible for any benefits (mycorrhizal plants with P but without N addition to the fungal compartment) and that transfer was entirely via the fungus (non-mycorrhizal plants with N added to the empty compartment). Biomass, shoot length, number of pods, chlorophyll content, and 15N labeling were used to investigate N transfer from the fungus as well as nutritional benefits to the plant. Mycorrhizal plants with N available to the fungal ERM had higher weight, longer shoot length, higher number of pods as well as higher chlorophyll content than the controls which showed N deficiency symptoms. Furthermore, mycorrhizal plants with N had high levels of, and fractional 15N labeling in, N metabolites (glutamine, glutamate, pyrroline-5-carboxylate, ornithine and arginine) in both roots and shoots while the control plants showed no labeling and low level of those metabolites. We conclude that N transfer from an AMF confers growth benefits on the host plant under nitrogen limiting conditions and 46 that the microcosm system developed will be useful for future work on AM nutrition and metabolism under physiologically relevant conditions. 47 Introduction Plant-microbial mutualistic associations play an important role in global nutrient cycles as well as in the ecology and physiology of plants (Read and PerezMoreno, 2003). Plant growth depends on mineral nutrients whose availability is frequently limiting. Phosphorous and nitrogen availability most commonly limits plant growth in ecosystems (Reich et al., 2006). Hayman and Mosse, (1971) first documented that the arbuscular mycorrhizal (AM) symbiosis can increase plant growth in P-deficient soils dramatically. Plants gain several benefits from AMF. For example AMF take up and transfer to their host nutrients including phosphorus, nitrogen (Govindarajulu et al., 2005), sulfur (Allen and Shachar-Hill, 2009), and zinc (Clark & Zeto, 2000), improve drought resistance (George et al., 1992), protection from pathogens has also been demonstrated (Newsham et al., 1995). In return, plants translocate photosynthetically fixed carbon to the fungal symbiont (Smith and Read, 2008). While P is generally believed to be the most important nutrient taken up by plants via AM networks, other macronutrients can be moved in large amounts via AMF. After it was shown that mycorrhizal colonization increased plant 35S uptake compared to nonmycorrhizal plants (Rhodes and Gerdemann, 1978; Gray and Gerdemann, 1973 ), Allen and Shachar-Hill, (2009) reported that 50% of 35SO 2 4 uptake by the fungus was detected in the mycorrhizal roots of an in vitro mycorrhizal system. Similar quantities of 35S were transferred to mycorrhizal roots whether 35SO42−, [35S]Cys, or [35S]Met was supplied in the fungal compartment (Allen and ShacharHill, 2009). 48 The AM fungal hyphae take up N from the soil in different forms and transfer it to the plant, however nitrogen is mainly captured by the AM extraradical mycelium in inorganic form (i.e. as nitrate or ammonium) (Bago et al., 1996; Hawkins et al., 2000; Azcon et al., 2001; Hodge et al., 2001; Vazquez et al., 2001). Although nitrogen transfer from AMF to the host plant has been demonstrated (Ames et al., 1983; Johansen et al., 1993; Hawkins et al., 2000; Azcon et al., 2001; Hodge et al., 2001; Vazquez et al., 2001) and a working model of the metabolism and transport processes has been proposed (Bago et al., 2001) and supported ( Jin et al., 2005; Cruz et al., 2007; Tian et al., 2010), nitrogen transfer has never been rigorously demonstrated to confer a growth benefit. Indeed, much of the work in recent years has been on the molecular mechanisms and on ecological aspects in the field. A monoxenic culture system consisting of transformed roots in symbiosis with the mycorrhizal fungus R. irregularis has been used to study the transfer and metabolism of nitrogen (Govindarajulu et al., 2005; Jin et al., 2005; Tian et al., 2010). This transformed root system was used because it is aseptic and provides easy access to the ERM. However, it is not a whole plant system, which casts doubt on the relevance of the plant metabolic results obtained. Unlike phosphate, which is immobile in soils (Gahoonia and Nielsen, 1991) inorganic nitrogen is mobile in most soils which makes it hard to determine the role of nitrogen transfer via AMF on plant growth and reproduction. In order to test whether nitrogen transfer confers a nutritional benefit in plants, an aseptic and leakage –free system is required which will prevent diffusion between compartments and which minimizes physical restrictions on plant growth that are typically seen in closed systems. Such a system should allow rapid and efficient fungal mycelial growth into a fungal compartment separated from the growth 49 zone of the colonized plant roots. This is important so that nutrients can be made available to the fungal extraradical mycelia (ERM) without the plant having access to them other than the fungus. Previous studies reported compartmented pot systems based on the same concept of having a root and fungal compartments (Ames et al., 1983; Frey and Schuepp, 1992; Ma¨der et al., 1993; Schweiger and Jakobsen, 2000; Smith et al., 2000; Jansa et al., 2003; Smith et al., 2003). Several researchers have used different variants of the pot system to demonstrate nutrient transfer in arbuscular mycorrhizal symbionts. Johansen et al., (1993) used containers divided by a fine nylon mesh into a root compartment (RC) and a root-free hyphal compartment (HC) to demonstrate phosphorous nutrient transfer between Subterranean clover (Trifolium subterraneum L. cv. Nuba) and R. irregularis. Nylon mesh was used also in a container system to separate the main soil compartment from the fungal compartment in experiments studying P acquisition by two AMF and its transfer to host plants (Smith et al., 2000; Hodge et al., (2001) found that the arbuscular mycorrhizal symbiosis can enhance the decomposition of, and increase nitrogen capture from, complex organic material in soil. They grew plants of Plantago lanceolata inoculated with the mycorrhizal fungus G. hoi in microcosms in which the compartments were separated by a double layer of 20-µm mesh, which was permeable to hyphae but not roots. In studying nitrogen and phosphorous transfer and regulation, Fellbaum et al., (2014) used a double membrane with an air gap (two sheets of 50-μm nylon mesh with a wire spiral between then) to prevent the diffusion of nutrients from the fungal compartment (FC) to the root compartment (RC), but allow fungal hyphae to cross from the RCs into the FCs has been used. 50 All of these systems have some drawbacks such as lack of control over the presence and growth of other microorganisms which could influence element bioavailability, especially when dealing with organic compounds, and whose levels and behavior is likely to be influenced by the fungal ERM. In other systems without complete isolation via an uninterrupted air gap, diffusion of mobile nutrients or tracers allows direct uptake by the roots due to leakage into the root compartment caused by diffusion or mass-flow due to transpiration. Another system type was used by Dupre´ de Boulois et al. (2006), an arbuscular mycorrhizal–plant (AM–P) in vitro culture system. In this system, the root compartment and hyphal compartments consist of a bi-compartmented Petri plate and a shoot compartment consists of a 50 ml Falcon tube and a membrane filter fixed onto the shoot compartment to allow gas exchange. This system is aseptic but the drawbacks are media drying and substantial physical restrictions on plant growth. Our aim in this study was to overcome the limitations of previous systems by developing an aseptic, leakage-free system that allows longer and more physiologically relevant experiments in a reusable, inexpensive microcosm of modest footprint, with separate reservoirs for supplying nutrients and water independently to the two compartments with sufficient capacity for multi-week long experiments. Here we report such a system. This system was used to investigate growth parameters, chlorophyll content and nitrogen metabolite 15N labeling and levels when N was made available to the fungal ERM at N levels that limited plant growth and reproduction. We observed that nitrogen taken up by the fungal extraradical mycelium and transferred to the plant accounts for much of the nitrogen entering the plant under N-limited conditions and confers growth and reproductive benefits. 51 Experimental procedures Chemicals and reagents Gelzan, (MP Biomedical, Solon, OH) was used for solidification of the minimal (M) media. Radioactive labeled sulfate was obtained as Na2 Biomedicals, (Solon, OH). 15N-Labeled 35SO 4 from MP nitrate was obtained as K15NO3 from (MP Biomedicals, (Solon, OH)). Seed sterilization Medicago truncatula seeds were scarified with concentrated sulfuric acid for 5-10 minutes. After that, seeds were rinsed with sterile water 4-5 times. Surface sterilization was done in concentrated Clorox for 2 minutes, then seeds were rinsed 8 times with water. Growth conditions Seeds were germinated at 22.5 °C for 48 hours on M-medium (Fortin et al., 2002) following cold treatment at 4 °C for 36 hours. After that, M. truncatula seedlings and old fragments of Daucus carota roots (Ri T-DNA transformed) colonized with R. irregularis were transferred to the root compartments filled with bacto:perlite (2:1) soils in the two-compartment system (Figure 2-1). Alternatively, the fungal compartment was filled with medium grain sand. This system was constructed using 50 µm woven double mesh glued to metal frame with an air gap to separate the two compartments and prevent the crossing of roots, yet allow fungal 52 crossover into the fungal compartment (RCs). The lower compartments were used as reservoir to supply both compartments with 0.5x Hoagland’s solution (Hoagland and Arnon, 1950). Linen cotton rope wicks were used for capillary transfer of Hoagland’s solution to both compartments. The plants were grown in a growth chamber under the following conditions: 16 h photoperiod, 22.5°C, photosynthetically active radiation of 200 µmol m-2 s-1, and 30% humidity. Experimental design Three experiments were conducted to demonstrate the role of arbuscular mycorrhizal in nutrition transfer. Whole plant two compartments system were used to test the transfer of 35SO 4 from AMF to the host plants. After growing plants for 5 weeks, plants were deprived of sulfur for a week, then 100 µCi of 35SO 4 was added directly to the fungal compartment for a week. Plant leaves were sampled 1, 3, 5, 7 days after addition of 35SO 4. Plant shoots and roots were collected to measure radioactivity. Two controls were used, mycorrhizal colonized plants with no added and non-mycorrhizal plants with 35SO 4 35SO 4 added to FC to test leakage. Five biological replicates were used for each. The second experiment was conducted to demonstrate the nutritional benefit of arbuscular mycorrhizal colonization on N transfer. After growing plants for five weeks, plants were deprived from N for 4 days then 10 mM KNO3 was added to the reservoir of fungal compartment for two weeks. Plant shoots and roots were collected for growth metrics and chlorophyll measurements. Two controls were used, 53 mycorrhizal colonized plants with no N added and non-mycorrhizal plants with 10 mM KNO3 added to the FC to test leakage. Total of 10 biological replicates in two different experiments were used. The third experiment was done to test the uptake and transfer of 15N in the whole plant system. After growing plants for 5 weeks, plants were deprived from N for 4 days then 10 mM K15NO3 was added to the reservoir of the fungal compartment for two weeks. Plant shoots and roots were collected for N metabolite analysis. Two controls were used, mycorrhizal colonized plants with no 15N added and non- mycorrhizal plants with 10 mM K15NO3 added to FC reservoir to test leakage. Total of 10 biological replicates in two different experiments were used. Testing the diffusion between compartments and the capillary transfer In order to make sure that there was no diffusion of nutrients from the FC into the RC, 35SO 4 was added to FC reservoir, Aliquots of the soil and the reservoir solution of the RC were collected (3, 5, 7 days after adding 35SO 4). Aliquots of the sand of FC were collected to investigate the rate of nutrients transfer using linen cotton rope wicks. The 35S content was extracted from all samples and measured by liquid scintillation counting. Extraction and measurement of 35SO4 from mycorrhizal plants Plant leaves were sampled 1, 3, 5, 7 days after addition of 35SO 4 to the FC. Also plant roots and shoots were sampled after 7 days. 100 mg leaves of each time point as well as 100 mg of each plants roots and shoots(stems and leaves) were 54 ground in a mortar and pestle with a 100mg acid washed sand and extracted three times with a mixture of cold methanol : water (70 : 20). then the solution were vortexed for 5 min. While keeping particulates suspended, a 1-mL aliquot of the solution was transferred to a microcentrifuge tube and centrifuged and 0.5 mL of the supernatant solution was counted in scintillation counter after adding to 5 mL of BioSafe II (MP Biomedicals) scintillation cocktail. Analysis of mycorrhizal colonization and ERM crossing to the FC Fraction colonization were checked for plants 1, 2, 3 and 4 weeks after transferring the seedlings with the spores to the system. Three biological replicates were used for each time point. Thirty root fragments of each biological replicates were stained with trypan blue and the percentage root length colonized by R. irregularis was estimated using the gridline intersect method (Newman, 1966). In order to check crossing, ERM from the FCs of 5 week old plants were collected, cleared in KOH, and stained with Trypan blue. Growth parameters and chlorophyll content measurements Experimental plants were collected two weeks after adding KNO3 to the FC reservoir. The two control plants were collected. The plants were weighed and the length of the shoots were measured. The fruits were collected and counted. Leaf chlorophyll was extracted using 80% spectrophotometer according to Ni et al., (2009). 55 acetone and quantified using Extraction, isolation and quantification of 15N metabolites Three hundred mg of each plants roots and shoots were ground in a mortar and pestle with a 100mg acid washed sand and extracted three times with a mixture of methanol : chloroform : water (12 : 5 : 3, v/v/v). Methylene chloride and water were added to the extraction solution to facilitate the separation of chloroform and the methanol–water phases. The methanol–water phase containing the amino acids (AAs) was collected and evaporated in a rotary evaporator at 50°C, and the residues containing the AAs were dissolved in 1 ml of 0.01 M HCl and loaded onto a cation exchange column (0.3 ml of DOWEX 50 X8-200– hydrogen form; Sigma-Aldrich, St Louis, MO, USA), which was previously washed with 1 M NH4OH, deionized H2O and 1 M HCl, and followed by deionized H2O. The neutral compounds, principally carbohydrates washed off the column with 5 ml of water and the free amino acids were eluted with 5 ml of 1 M NH4OH (Bengtsson & Odham, 1979). This eluent was collected and dried then resuspended in 70 µl of Milli-Q water. N metabolite levels and labeling were measured using Liquid chromatography (LC)-MS analyses. Liquid chromatography and mass spectrometry (LC-MS) A Waters (Milford, MA) Quattro micro mass spectrometer coupled to a Shimadzu (Columbia, MD) LC-20AD HPLC system and SIL-5000 autosampler was used. A Waters Symmetry C18 column (2.1 × 100 mm, 3 μm particle size) was used with column oven temperature at 30 °C. The injection volume was 10 μL, and the HPLC flow rate was 0.3 mL/min using 1 mM perfluroheptanoic acid in a water/acetonitrile (A/B) gradient at ambient temperature, The intial gradient 56 (A/B)=99/1, held until 2 minutes, followed by a linear gradient to 60/40 at 4 minutes, held at 60/40 until 8 minutes, followed by a ramp to 99/1 at 8.01 Mass spectra were acquired using electrospray ionization in positive ion mode and MRM. The capillary voltage, extractor voltage, and rf lens setting were set at 3.17 kV, 4 V, and 0.3, respectively. The flow rates of cone gas and desolvation gas were 20 and 400 L/h, respectively. The source temperature and desolvation temperature were 110 and 350 °C, respectively. Collision-induced dissociation employed argon as collision gas at a manifold pressure of 2 × 10-3mbar, and collision energies and source cone potentials were optimized for each transition using Waters QuanOptimize software. This method was composed of two ESI+ functions (0−1.8 and 1.8−6.0 min) covering full run time to allow for adequate dwell time for each analyte. Data were acquired with MassLynx 4.0 and processed for calibration and for quantification of the analytes with QuanLynx software. Table (2-1) of precursor and product ions for each compound is given in appendix. To determine concentrations, standard curves were run for authentic unlabeled standards of each compound in the table above with concentrations range from 0 to 100 μM together with 10uM Phenylalanine-D8 as internal standard, which was also added to each biological sample during extraction to control for recovery efficiency and as internal standard. Integrated peak areas were obtained from daughter ion chromatograms using the MassLynx 4.0. Total levels for each compound were obtained from the sum of all the labeled and unlabeled molecules. For 15N labeled samples, each product ion mass isomer for each precursor mass isomer was quantified. Thus for example for 15N labeling in Arginine the precursor ions have mass-to-charge ratios (m/z) of 175 (15N0 = M+0), 176 (15N1 = 57 M+1), 177 (15N2 = M+2), 178 (15N3 = M+3), and 179 (15N4 = M+4) with the product ion CH6N3+ have m/z of 60 (15N0 = M+0), 60 and 61 (15N1 = M+1), 61 and 62 (15N2 = M+2), 62 and 63 (15N3 = M+3), and 63 (15N4=M+4) (see Table 2-1 and 2-2) . Results To establish conditions for studying N transfer between the AM fungus and the host plants, a whole plant mycorrhizal two-compartment culture system was developed (Figure 2-1). The final model has three compartments. The top compartment allows room for shoot growth and is separated from the outside atmosphere using a 0.22 µM pore size breathe-EASIER membrane across the entire top of the microcosm to allow for gas exchange, including water vapor, and to maintain the sterility of the system. The central compartment contains soil or other solid medium for root and fungal growth. This is divided into two subcompartments: a soil and a sand sub-compartment separated by two layers of metal woven mesh (Dutch weave 40 µm exclusion size) that are glued together with autoclavable epoxy to a metal frame. The divider is fixed vertically in the middle of the central compartment using autoclavable silicone rubber. This creates a robust air gap of ~1.5mm and an area for hyphal penetration that spans most of the area of the divider. A seedling is inoculated with R. irregularis spores and grown in the soil compartment. The woven mesh was found to allow the fungal ERM to cross from plant to fungal compartment at a high frequency while preventing plant roots from doing so. By sticking two stiff mesh sheets to a metal frame to form the divider, the air gap consistently prevented diffusion of nutrients between compartments, which occurs for diffusible nutrients or tracers when single mesh barriers are used. To 58 create air gaps, flexible double mesh barriers were tested with smaller area dividers to prevent the mesh sheets from touching. These had low hyphal connectivities between compartments. Using a course mesh as a spacer between the rootexcluding mesh layers can facilitate diffusion between compartments by forming a continuous diffusion pathway. The design shown in Figure 2-1 prevents leakage between compartments as indicated by the absence of 35 SO42- in the plant compartment when this diffusible tracer was added to the distal compartment of nonmycorrhizal plants (Figure 2-2A). The third, lower compartment contains liquid reservoirs to supply the root and fungal compartments with water and nutrients via wicks. The reservoirs in the lower compartment were separated by attaching a 50100 mL glass beaker to the floor using autoclavable silicone gel under the fungal compartment (FC) while the rest of the chamber was used as a root compartment (RC) reservoir. In order to supply each compartment of nutrients from reservoirs, different kinds of ropes were tested as wicks. Some of them allowed too little capillary transfer capacity, resulting in water stress for the plants while other materials allowed too much transfer, saturating the RC and FC and causing diffusional tracer leakage between compartments and excessive soil moisture that reduced plant growth. Linen cotton rope wicks were found to provide efficient but not excessive transfer of water and nutrients as indicated by rapid 35SO 24 movement into the fungal (sand) compartment (Figure 2-2A) without leakage into the root compartment or soil water saturation. In order to eliminate complications arising from the uptake and metabolism of nitrogen, carbon, or other tracers by other microorganisms, the maintenance of aseptic conditions over several weeks was required. By selection of appropriate materials for wicks, dividers, and adhesives, the microcosms can be autoclaved after assembly, with the easy breath membranes 59 being added under sterile conditions (laminar flow hood). The sterility of the system was investigated using periodic sampling of all compartments and inoculation onto enrichment media (trypticase soy agar and potato dextrose agar) for bacteria and fungi. This culture system was found to be reproducibly aseptic with no growth of any microbes seen on the rich media throughout the time frame of the experiments. The level of colonization per root length of M. truncatula plants (Figure 2-3) reached approximately 90% for internal hyphae and 62% for arbuscules and vesicles in four-week old plants. These high and reproducible colonization levels and establishment rate (most plants were substantially colonized within a week) allow observations of nutrient exchange between the fungal mycelium and the plant roots to be interpreted without uncertainties due to variable or slow colonization rates. We used the culture system described above to monitor S uptake by the fungal extraradical mycelium and its transfer to host plants. By supplying 35SO 4 to the FC, It was found that leaves had measurable 35S after one day and reaching over 60 fmol/g (FW) counts after 7 days (Figure 2-2B). Furthermore, 35S levels were measured in whole plant roots and shoots after 7 days and compared with two controls (mycorrhizal plants with no plants with 35SO 4 35SO 4 added to the FC and non-mycorrhizal provided to the FC) (Figure 2-2C) . High shoots were found. However, plant shoots had higher 35S 35S levels of roots and level than roots. The two control roots and shoots showed no counts above background (comparable with the background counts) proving that this culture system had no leakage between compartments. By adding nitrogen to the fungal compartment reservoir (FCR) of the mycorrhizal plants, it was found that nitrogen transfer confers growth benefit as 60 plants were larger and greener with multiple stems compared with the other two controls (mycorrhizal plants with nutrients but no nitrogen added to the FCR and non-mycorrhizal plants with N provided to the FCR, see Figure 2-4). In the two controls, plants stopped growing and senesced. Shoots dried out and had low numbers of leaves as many dehisced with most of the residual leaves were visibly chlorotic (a symptom of nitrogen deficiency, Figure 2-4). Chlorophyll content of mycorrhizal plants (2-5) with nitrogen supplied to the FCR were significantly higher than for plants from the two control treatments (p<0.01). In the two controls, no nitrogen was available to the plant through the fungus and this presumably led to chlorophyll degradation, and/or reduced biosynthesis (Figure 2-5A). Mycorrhizal plants with nitrogen added to their fungal compartment reservoirs had significantly longer shoots, higher biomasses and higher numbers of fruits compared with the two control treatments (Figure 2-5 B,C, and D) (p<0.01). Mycorrhizal plants with nitrogen added to the fungal compartment reservoir had significantly higher levels of the soluble nitrogen-containing metabolites glutamate, glutamine, carboxylate, ornithine and arginine (p<0.05) as well as a significant 15 pyrroline-5N percentage labeling in both root and shoot tissues (p<0.01) . This confirms that nitrogen transfer from the fungal compartment via the fungal mycelium drives substantial fluxes through N metabolism (Figure 2-6). The control plants showed low levels of N metabolite and no detectable 15N labeling, consistent with the 35S results that indicated no significant nutrient movement between compartments not connected by fungal mycelium. Mycorrhizal plants with P and other nutrients but not N supplied to the FCR had similar N metabolite levels and growth parameters to non-mycorrhizal plants. Thus, the nutritional benefit to plants is due to direct fungal-mediated N transfer. The high percentage labeling of intermediary metabolites in shoots as well 61 as roots demonstrates directly, in a way that the use of transformed roots cannot, that N transfer to mycorrhizal roots benefits the N status of the whole plant. Discussion Many of the detailed studies of mycorrhizal metabolism and nutrient transfer (e.g. Bago et al., 1996, Jin et al., 2005, Govindarajulu et al., 2005) were performed using transformed roots cultured on bicompartmental petri plates as the model mycorrhizal system (St. Arnaud et al., 1996). This system was used because of ease of handling, maintenance of sterility, the advantages of defined media and ease of ERM tissue isolation. However, these studies did not investigate nutrient transfer in whole-plant mycorrhizal systems which are likely to have different physiological characteristics and regulatory dynamics. The studies of N nutrition and metabolism in the AM symbiosis that have been reported using whole plants were not conducted under aseptic conditions (e.g. Johansen et al., 1993, Hodge et al., 2001; Fellbaum et al., 2014). Although such studies have been important in establishing for example the contribution of AMF to the mobilization of soil N (Tobar et al. 1995, Hodge et al. 2001) and are physiologically substantially more realistic than transformed root systems, the impractibility of maintaining sterility complicates the interpretation of metabolic and translocation experimental results. In this study, we developed a sterile and leakagefree (Figure 2-2A) whole plant two-compartment culture system (Figure 2-1) allowing the exclusion of other microbes and avoidance of diffusion between compartments. Low cost, easy to use, reproducible and autoclavable, these microcosms make feasible investigations of mycorrhizal transport, gene expression, metabolism and 62 nutrition that have hitherto been challenging or impossible. In particular, the sterility of this system will enable the study of the role of organic C, P, N, and S compounds in AM symbiosis. The culture system provided M. truncatula plants with optimized levels of nutrients and water, including low phosphorous levels to stimulate colonization. Within two weeks of seedling planting, the plants were associated with higher levels of colonization and numbers of arbuscules (Figure 2-3) than is commonly observed, and significantly, also with well-developed external hyphal mycelia of R. irregularis. This allows experiments on uptake and transfer to be conducted on plants that, while mature, are not senescent or growth-limited by physical space constraints. The uptake and transfer of 35SO42- by the fungal partner to the host plant was demonstrated in the microcosms (Figure 2-2B) with host plants after one day. Shoots had higher 35S 35S detected in the leaves of concentration than roots (Figure 2- 2C) which is consistent with the fact that plants transport sulfate to the aerial parts, where the majority is stored as a vacuolar sulfate pool or metabolized in reductive sulfur assimilation (Kataoka et al.,2004). We conclude that sulfate is transferred by R. irregularis to host plants, as has been reported for AM transformed roots (Allen and Shachar-Hill, 2009). Allen and Shachar-Hill (2009) demonstrated the transfer of sulfate by R. irregularis to transformed roots and found that the fungus can uptake and transfer reduced forms of S at rates comparable to sulfate. Since 95% of S is in organic form in soil (Tabatabai, 1986; Scherer, 2001), this observation points to a wider role for AMF for S plant nutrition in nature. This example also highlights the need for aseptic whole plant AM experiments to determine whether this capacity may be significant for S nutrition of AM plants. 63 It has been reported that AMF are able to substantially increase the uptake of N by host plant roots (Ames et al., 1983; Johansen et al., 1993; Bago et al., 1996; Johansen et al., 1996; Tobar et al., 2004; Hodge et al., 2010). Nevertheless, the literature lacks a demonstration of growth or reproductive benefits to plants of N transfer from AMF. The microcosm system developed in this study provides sufficient room for plant growth with minimal physical restriction, sufficient gas exchange, and high rates of hyphal crossing to examine this important question. We observed that N transfer by AMF conferred growth and reproductive benefits to the host plant. After supplying the fungal compartment with N, the mycorrhizal plants appeared healthier, greener and larger than the controls. Mycorrhizal plants with N had longer shoots, higher biomass and increased chlorophyll contents compared to the controls. In contrast, control plants showed N deficiency symptoms. N deficiency in plants results in a breakdown of chlorophyll (Gaude et al., 2007) and also affects the abundance of thylakoid membranes in chloroplasts (Malavolta et al., 2004). Furthermore, N deficiency causes severe consequences for N and C metabolism (Wang et al., 2003). 15N labeling and concentrations of N metabolites (glutamine, glutamate, pyrroline-5-carboxylate, ornithine, and arginine) showed that 15N taken up by the ERM in the FC arrives in significant amounts (p<0.01) at the roots and shoots of host plants (Figure 2-6). Control plants have low levels of N metabolites and no 15N labeling, proving that there is no diffusion between compartments. It has been reported that N deficiency affects the abundance of amino acids in plants (Scheible et al., 2004). 64 The high abundance of N metabolites and mycorrhizal plants after adding 15NO 3 15 N percentage labeling in coincides with N transfer (model in which Inorganic N is taken up by the fungal ERM, and assimilated by GS-GOGAT system, raising the levels of glutamate and glutamine then nitrogen is incorporated into arginine in the urea cycle which is translocated to the fungal IRM in colonized root tissues and then broken down into ornithine and urea that is in turn is broken down to release ammonium which is exported from the fungus and imported by the host into the root cortical cells. Plants then assimilate ammonium to produce free amino acids (Tian et al., 2010) and as a consequence, the 15N labeled metabolites in the root are in fungal IRM and plant roots. The presence of high levels of 15N labeled metabolites in shoots indicated a significant amount of N being transferred to the host plants. 65 APPENDIX 66 APPENDIX A B 0.22um gas permeable Double mesh barrier with air Fungal compartmen Wick FC Reservoir Plant compartment Wick RC Reservoir Figure 2-1 Whole plant-mycorrhizal two-compartment culture system. A) Diagrammatic representation of the culture system that is composed of two compartments: root (RC) and fungal compartments (FC). RC and FC are separated by double woven mesh glued to metal, thus air gap is created between the two layers of mesh. The upper compartment is sealed with breathe-EASIER membrane and the lower compartments are used as reservoirs; root compartment reservoir and fungal compartment reservoir. Nutrients were transferred using linen cotton wicks. B) Four-week old Medicago truncatula plant growing in the system. 67 A 30 0.1 d A Fungal compartment 0.09 25 0.08 Root compartment 0.07 20 15 fmol/g fmol/g 0.06 c Root compartment reservoir 0.05 0.04 0.03 10 ee ee eeee 0.02 5 a b 0.01 0 1 day 0 1 day 3 days 5 days 7 days 5 7 3 days days days Figure 2-2 Sulfur transfer from fungus to plant using the whole plant culture system. A) 35SO4 concentration (fmol/g) in root compartment soil and reservoir at the days 1, 3, 5, and 7 days after applying 35SO4 to fungal compartment reservoir to investigate diffusion between compartments. 35SO4 concentration in fungal compartment sand wa measured to investigate the rate of transfer of nutrients by the cotton wicks. The right panel is the same figure as the left but with lower maximum limits to show smaller values. Values are reported as mean ± SEM (n=3 biological replicates). The transfer of 35SO4 was significantly increasing with time (ANOVA single factor analysis (alpha=0.05). the 35SO4 concentration in root compartment soil and reservoir was comparable with the background. B) 35S concentration (fmol/g (FW)) in leaves at 1, 3, 5, and 7 days after applying 35SO4 to Fungal compartment. Values are reported as mean ± SEM (n=5 biological replicates). C) 35S concentration (fmol/g (FW)) in whole plant shoots (above ground stem and leaves) and roots at 7 days after applying 35SO4 to Fungal compartment. Values are reported as mean ± SEM (n=5 biological replicates). Statistical analyses were done by ANOVA single factor analysis, alpha = 0.05. Letters (a, b, ...) above the bar graphs designate statistically significant difference between means. 68 Figure 2-2 (cont’d) B d 80 70 fmol/g (FW) leaves 60 50 c 40 30 b 20 a 10 0 1 day 3 days 5 days 69 7 days Figure 2-2 (cont’d) C 110 b 100 90 80 70 a 60 fmol/g (FW) 50 40 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 c 0.1 c c c Roots Shoots 0 Roots Shoots Myc plants + 35S Roots Shoots nonmyc plants +35S myc plants -35S 70 A b 1 0.9 b Fractional Colonization (internal hypha) 0.8 0.7 0.6 a 0.5 0.4 0.3 0.2 0.1 0 2wks 3wks 4wks B 0.8 b Fractional colonization (Arbuscules and vesicles) 0.7 b 0.6 0.5 a 0.4 0.3 0.2 0.1 0 2wks 3wks 4wks Figure 2-3 Colonization of Medicago truncatula by R. irregularis. Plants were raised in the two-compartments system under low phosphate and in the presence of old Daucus carota colonized roots. Values are fraction of root length associated with (A) internal hyphae and (B) arbuscules and vesicles. Values are reported as mean ± SEM (n=3 biological replicates (30 technical replicates)). Statistical analyses were done by ANOVA single factor analysis, alpha = 0.05. Letters (a, b, ...) above the bar graphs designate statistically significant difference between means. 71 Non-myc plants Myc plants +N Myc plants –N Figure 2-4 Seven-week old Medicago truncatula plants growing in the twocompartment system. Mycorrhizal plants with KNO3 (Myc plants +N) added to the fungal compartment reservoir for two weeks. Two controls were used; nonmycorrhizal (non-myc) plant with KNO3 added to fungal compartment reservoir to check for leakage and mycorrhizal plants with no N added (Myc –N). 72 A 0.9 Total chlorophyll content (mg/g) 0.8 a 0.7 0.6 0.5 0.4 0.3 b 0.2 b 0.1 0 Figure 2-5 Total chlorophyll content and growth parameter measurements. 10mM KNO3 was added to the fungal compartment of Mycorrhizal plants (Myc plants +N) reservoir for two weeks. Two controls were used; non-mycorrhizal (non-myc) plant with KNO3 added to fungal compartment reservoir to check for leakage and mycorrhizal plants with no N added (Myc –N). A) Total chlorophyll content, B) number of pods, C) Fresh weight, and D) shoot length were measured. Values are reported as mean ± SEM (n=10 biological replicates). Statistical analyses were done by ANOVA single factor analysis, alpha = 0.05. Letters (a, b, ...) above the bar graphs designate statistically significant difference between means. 73 Figure 2-5 (cont’d) B 4.5 a 4 Fresh weight (g) 3.5 3 2.5 b b 2 1.5 1 0.5 0 C 12 a Number of pods 10 8 b b 6 4 2 0 74 Figure 2-5 (cont’d) D 35 Shoot length (cm) 30 25 a b b 20 15 10 5 0 75 A Glutamate Unlabeled N labeled 15 15 Non-mycorrhizal plant + N Mycorrhizal plant +15N 120 450 nmol/g (FW) 350 300 250 * 100 nmol/g (FW) * 400 80 60 40 20 200 0 150 Root Shoot 100 50 Mycorrhizal plant with no 0 Root 15 N Shoot 120 nmol/g (FW) 100 80 60 40 20 0 Root Shoot Figure 2-6 Free N metabolite concentrations and 15N labeling in the whole mycorrhizal plant system. Total of 10 mM K15NO3 was added to the fungal compartment reservoir of mycorrhizal plants (Myc plants +N) for two weeks. Two controls: non-mycorrhizal plant with 15N added nonplant compartment to the and mycorrhizal plants with no 15N added. N metabolites are glutamate (A), glutamine (B), pyrroline-5-carboxylate (C), ornithine (D) and arginine (E). The labeling was calculated by adding the concentrations of all 15N labeled isotopomers for each N metabolite. Values are reported as mean ± SEM (n=10 biological replicates). Statistical analyses were done by ANOVA single factor analysis, alpha = 0.01. Stars show statistically significant 15N labeling (p<0.01). 76 Figure 2-6 (cont’d) B Glutamine Unlabeled N labeled 15 Non-mycorrhizal plant +15N 50 Mycorrhizal Plant +15N 40 * 200 * 30 20 10 150 0 Root Shoot 100 50 50 Mycorrhizal plant with no 15N 45 40 0 Root Shoot nmol/g (FW) nmol/g (FW) 250 nmol/g (FW) 300 35 30 25 20 15 10 5 0 Root 77 Shoot Figure 2-6 (cont’d) C Pyrroline-5-Carboxylate Unlabeled N labeled 15 Non-mycorrhizal plant +15N Mycorrhizal plant +15N 7 6 * nmol/g (FW) 25 20 15 * 10 nmol/g (FW) 30 Mycorrhizal plant with no 15N 5 4 3 2 1 5 0 Root 0 Shoot Shoot Mycorrhizal plant with no 15N 5 4 nmol/g (FW) Root 3 2 1 0 Root 78 Shoot Figure 2-6 (cont’d) D Ornithine Unlabeled N labeled 15 Non-mycorrhizal plant +15N 70 300 Mycorrhizal plant +15N * 200 50 40 30 20 10 150 0 100 Root * 50 70 0 Shoot Mycorrhizal plant with no 15N 60 Root Shoot nmol/g (FW) nmol/g (FW) 250 nmol/g (FW) 60 50 40 30 20 10 0 Root 79 Shoot Figure 2-6 (cont’d) E Arginine Unlabeled N labeled 15 14 Non-mycorrhizal plant +15N 12 nmol/g (FW) Mycorrhizal plant +15N 160 140 10 8 6 4 2 * 100 0 80 60 Root * 14 Shoot Mycorrhizal plant with no 15N 40 12 20 10 0 Root nmol/g (FW) nmol/g (FW) 120 Shoot 8 6 4 2 0 Root 80 Shoot Table 2-1 Precursor (unlabeled) compounds and product ions for each analyte. Compound Precursor compound molecular formula Daughter ion(s) quantified Molecular formula Arginine C6H14N4O2 CH6N3+ Glutamate C5H9NO4 C4H6NO+ Glutamine C5H10N2O3 C5H7NO3+ Ornithine C5H12N2O2 C4H8N+ Putrescine C4H12N2 C4H10N+ Pyrroline-5-Carboxylate C5H7NO2 C4H6N+ 81 Table 2-2 Multiple reaction monitoring (MRM) transitions, optimizing source cone voltages, collision cell voltages, and analyte retention time (for 15N labeling experiment). Compound Arginine Glutamate Glutamine Ornithine Putrescine Pyrroline 5Carboxylate Phenylalanine -D8 Precursor ion> product ion (m/z) M+0: 175.1>60 M+1: 176.1>60 M+1: 176.1>61 M+2: 177.1>62 M+2: 177.1>62 M+3: 178.1>62 M+3:178.1>63 M+4:179.1>63 M+0: 148.02>83.84 M+1: 149.02>84.84 M+0: 147.06>129.95 M+1: 148.06>129.95 M+1: 148.06>130.95 M+2: 149.06>130.95 M+0: 133.1>70 M+1: 134.1>70 M+1: 134.1>71 M+2: 135.1>71 M+0: 89.03>71.79 M+1: 90.03>71.79 M+1: 90.03>72.79 M+2: 91.03>72.79 M+0: 113.98>67.75 M+1: 114.98>68.75 174.04>128.06 Cone Voltage (V) 22 Collision Voltage (V) 22 Retention time (min) Function no. 6.08 1 22 16 1.23 2 16 10 1.07 2 16 16 5.52 2 50 10 6.05 2 16 10 1.15 2 18 15 5.26 2 82 REFERENCES 83 REFERENCES Allen JW, Shachar-Hill Y. 2009. Sulfur Mycorrhiza. Plant Physiology 149: 549-560. transfer through an Arbuscular Ames RN, Reid CP, Porter LK, Cambardella C. 1983. Hyphal uptake and transport of nitrogen from two 15N-labelled sources by Glomus mosseae, a vesicular-arbuscular mycorrhizal fungus. New Phytologist 95: 381-396. Azcón R., Ruíz-Lozano JM, Rodríguez R. 2001. Differential contribution of arbuscular mycorrhizal fungi to plant nitrate uptake (15N) under increasing N supply to the soil. Canadian Journal of Botany 79: 1175-1180. Bago B, Vierheilig H, Piché Y, Ázcon-Aguilar C. 1996. Nitrate depletion and pH changes induced by the extraradical mycelium of the arbuscular mycorrhizal fungus Glomus intraradices grown in monoxenic culture. New Phytologist 133: 273-280. Bago B, Pfeffer P, Shachar-Hill Y. 2001. Could the urea cycle be translocating nitrogen in the arbuscular mycorrhizal symbiosis? New Phytologist 149: 4-8. Bengtsson G, Odham G .1979. A micromethod for the analysis of free amino acids by gas chromatography and its applications to biological systems. Analytical Biochemistry 92: 426-441. Dupré de Boulois H, Voets L, Delvaux B, Jakobsen I, Declerck S. 2006. Transport of radiocaesium by arbuscular mycorrhizal fungi to Medicago truncatula under in vitro conditions. Environmental Microbiology 8(11): 1926-1934. Clark RB, Zeto SK. 2000. Mineral acquisition by arbuscular mycorrhizal plants. Journal of Plant Nutrition 23: 867-902. Cruz C, Egsgaard H, Trujillo C, Ambus P, Requena N, Martins-Loucao MA, Jakobsen I. 2007. Enzymatic evidence for the key role of arginine in nitrogen translocation by arbuscular mycorrhizal fungi. Plant Physiology 144: 782-792. Fellbaum CR, Mensah JA, Cloos AJ, Strahan GE, Pfeffer PE, Kiers ET, Bücking H. 2014. Fungal nutrient allocation in common mycorrhizal networks is regulated by the carbon source strength of individual host plants. New Phytologist 203(2):646-656. Frey B, Schuepp H. 1993. A role of vesicular–arbuscular (VA) mycorrhizal fungi in facilitating interplant nitrogen transfer. Soil Biology and Biochemistry 25: 651–658. Gahoonia TS, Nielsen NE. 1991. A method to study rhizosphere processes in thin soil layers of different proximity to roots. Plant and Soil 135: 143–146. Govindarajulu M, Pfeffer PE, Jin HR, Abubaker J, Douds DD, Allen JW, Bucking H, Lammers PJ, Shachar-Hill Y. 2005. Nitrogen transfer in the arbuscular mycorrhizal symbiosis. Nature 435: 819-823. 84 Gaude N, Brehelin C, Tischendorf G, Kessler F, Dörmann P. 2007. Nitrogen deficiency in Arabidopsis affects galactolipid composition and gene expression and results in accumulation of fatty acid phytyl esters. Plant Journal 49:729–739. George E, Haussler K, Vetterlein D, Gorgus E, Marschner H. 1992. Water nutrient translocation by hyphae of Glomus mosseae. Canadian Journal of Botany 70:2130– 2137 Gray, LE and Gerdemann, JW. 1973. Uptake of sulphur-35 by vesicular-arbuscular mycorrhizae. Plant Soil 39: 687—689. Hawkins HJ, Johansen A, George E. 2000. Uptake and transport of organic and inorganic nitrogen by arbuscular mycorrhizal fungi. Plant and Soil 226: 275–285. Hayman DS, Mosse B. 1971. Plant growth responses to vesicular arbuscular mycorrhiza. I. Growth of Endogone-inoculated plants in phosphate-deficient soils. New Phytologist 70:19–27 Hoagland DR, Arnon DI. 1950. The water-culture method for growing plants without soil. California Agricultural Experiment Station Circular 347:1-32. Hodge A, Campbell CD, Fitter AH. 2001. An arbuscular mycorrhizal fungus accelerates decomposition and acquires nitrogen directly from organic material. Nature 413: 297-299. Hodge A, Helgason T, Fiter AH. 2010. Nutritional ecology of arbuscular mycorrhizal fungi. Fungal Ecology 3:267-273. Jansa J, Mozafar A, Frossard E. 2003. Long-distance transport of P and Zn through the hyphae of an arbuscular mycorrhizal fungus in symbiosis with maize. Agronomie 23 (5-6): 481-488 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. Johansen A, Jakobsen I, Jensen ES. 1993. Hyphal transport by a vesiculararbuscular mycorrhizal fungus of N applied to the soil as ammonium or nitrate. Biology and Fertility of Soils 16: 66-70. Johansen A, Finlay RD, Olsson PA. 1996. Nitrogen metabolism of external hyphae of the arbuscular mycorrhizal fungus Glomus intraradices. New Phytologist 133: 705712. Kataoka T, Hayashi N, Yamaya T, Takahashi H. 2004. Root-to-shoot transport of sulfate in Arabidopsis: Evidence for the role of SULTR3;5 as a component of lowaffinity sulfate transport system in the root vasculature. Plant Physiology 136: 4198– 4204. 85 Mäder P, Vierheilig H, Alt M, Wiemken A, 1993. Boundaries between soil compartments formed by microporous hydrophobic membranes (GORE-TEX®) can be crossed by vesicular-arbuscular mycorrhizal fungi but not by ions in the soil solution. Plant Soil 159:201-206 Malavolta E, Nogueira NG, Heinrichs R, Rodriguez ENV, Guerra E, de Oliveira SC, Cabra CP. 2004. Evaluation of nutritional status of the cotton plant with respect to nitrogen. Communications in Soil Science and Plant Analysis 35: 1007-1019. Newman E. 1966. A method of estimating the total length of root in a sample. Journal of Applied Ecology 3: 139-145. Newsham KK, Fitter AH, Watterson AR.1995. Arbuscular mycorrhiza protect an annual grass from root pathogenic fungi in the field. Journal of Ecology 83:991– 1000. Ni Z, Kim ED, Ha M, Lackey E, Liu J, Zhang Y, Sun Q, Chen ZJ.2009. Altered circadian rhythms regulate growth vigour in hybrids and allopolyploids. Nature 457:327-331. Read DJ, and Perez-Moreno J. 2003. Mycorrhizas and nutrient cycling in ecosystems — a journey towards relevance? New Phytologist 157: 475–492. Reich PB, Hobbie SE, Lee T, Ellsworth DS, West JB, Tilman D, Knops JMH, Naeem S, Trost J. 2006. Nitrogen limitation constrains sustainability of ecosystem response to CO2. Nature 440: 922-925. Rhodes LH and Gerdemann JW. 1975. Phosphate uptake zones of mycorrhizal and non-mycorrhizal onions. New Phytologist 75: 555-561. Scheible WR, Morcuende R, Czechowski T, Fritz C, Osuna D, PalaciosRojas N, Schindelasch D, Thimm O, Udvardi MK, Stitt M. 2004. Genome-wide reprogramming of primary and secondary metabolism, protein synthesis, cellular growth processes, and the regulatory infrastructure of Arabidopsis in response to nitrogen. Plant Physiology 136: 2483–2499 Scherer HW. 2001. Sulphur in crop production. European Journal of Agronomy 14: 81– 111. Schweiger P, Jakobsen I, 2000. Laboratory and field methods for measurement of hyphal uptake of nutrients in soil. Plant Soil 226(2): 237-244. Smith FA, Jakobsen I, Smith SE. 2000. Spatial differences in acquisition of soil phosphate between two arbuscular mycorrhizal fungi in symbiosis with Medicago trunculata. New Phytologist 147: 357-366. Smith SE, Read DJ. 2008 . Mycorrhizal symbiosis. 3rd edn. Academic Press, London 86 Smith SE, Smith FA, Jakobsen I. 2003. Mycorrhizal fungi can dominate phosphate supply to plants irrespective of growth responses. Plant Physiology 133: 16-20. St-Arnaud M, Hamel C, Vimard B, Caron M, Fortin JA. 1996. Enhanced hyphal growth 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, pp. 207–226. Tian CJ, Kasiborski B, Koul R, Lammers PE, Bücking H, Shachar-Hill Y. 2010. Regulation of the nitrogen transfer pathway in the arbuscular mycorrhizal symbiosis: gene characterization and the coordination of expression with nitrogen flux. Plant Physiology 153: 1175-1187. Tobar RM, Azcón R, Barea JM. 2004. The improvement of plant N acquisition from an ammonium-treated, drought-stressed soil by the fungal symbiont in arbuscular mycorrhizae. Mycorrhiza 4: 105-108. Vázquez MM, Barea JM, Azcón R. 2001. Impact of soil nitrogen concentration on Glomus spp.-Sinorhizobium interactions as affecting growth, nitrate reductase activity and protein content of Medicago sativa. Biology and Fertility of Soils 34: 57– 63. Wang R, Okamoto M, Xing X, Crawford NM. 2003. Microarray analysis of the nitrate response in Arabidopsis roots and shoots reveals over 1000 rapidly responding genes and new linkages to glucose, trehalose-6-phosphate, iron, and sulfate metabolism. Plant Physiology 132: 556–567. 87 Chapter 3 Evidence for additional pathways in the nitrogen transfer network of arbuscular mycorrhizas 88 Abstract In recent years, our understanding of nitrogen nutrition in the arbuscular mycorrhizal (AM) symbiosis has substantially increased. Work at the ecological level has highlighted interactions among soil nitrogen, mycorrhizal fungi, and plant communities; and at the molecular level a working model for nitrogen uptake, metabolism, and transfer has been established. However, significant questions remain about potential additional N fluxes through the metabolic network, including nitrate transport and the origins and fate of ornithine. In 15N labeling experiments, using cultured mycorrhizal transformed roots as well as with mycorrhizal plant microcosms, we observed that nitrate taken up by the AM fungal extraradical mycelium (ERM) is translocated to the intraradical mycelium (IRM) and transferred to the host roots. Labeled nitrate was measured in substantial quantities in shoots, showing translocation within the plant after transfer from the fungus. After providing 15N labeled nitrate or 13C 6 arginine to the fungal ERM, the levels and labeling of metabolic intermediates as well as the expression of fungal N metabolism genes were measured over a 3-day time course in the ERM and colonized roots. The results are consistent with ornithine biosynthesis in the ERM via pyrroline-5-carboxylate and indicate that that some ornithine is broken down in the IRM to glutamate and to a lesser extent to putrescine. The timing of ornithine labeling in the IRM and ERM indicates that a significant proportion of ornithine in the ERM is derived from arginine breakdown in the IRM and translocated back to ERM where it is recycled to make arginine. The results extend the working model of the N metabolic and transport 89 network in ways that point to significantly greater flexibility and energetic efficiency by the fungal partner than appeared possible in the network structure. 90 Introduction More than 80% of land plant species are mycorrhizal. Of the different mycorrhizal types, the arbuscular mycorrhizal symbiosis is the predominant one (Wang & Qiu, 2006). Arbuscular mycorrhizal fungi (AMF) take up and transfer nutrients including P, N, S, and Zn to their hosts (Clark & Zeto, 2000; Hamel, 2004; He et al. 2005). In return, AMF depend on fixed carbon received from host roots, which can consume up to 20% of photosynthate (Jakobsen & Rosendahl, 1990). Nitrogen availability limits plant growth in many ecosystems (Reich et al., 2006; Jackson et al., 2008; Marschner, 1995). The arbuscular mycorrhiza creates a well distributed and extensive absorption network, which has been shown to be able to substantially increase the uptake of N by host plant roots (Ames et al., 1983; Johansen et al., 1993; Bago et al., 1996; Johansen et al., 1996; Tobar et al., 2004; Hodge et al., 2010). In the current model for N transfer in the symbiosis (Bago et al., 2001, Govindarajulu et al. 2005), nitrogen taken up by the ERM is assimilated via the GS/GOGAT pathway and used to synthesize arginine, which is translocated to the intraradical mycelium (IRM), and broken down to ammonium, which is released to the host inside the colonized root without nutritionally significant amounts of carbon (see Figure 1-5 in Chapter 1). Since it was proposed by Bago et al. (2001), substantial biochemical and molecular biological evidence has accumulated to support this scheme (Govindarajulu et al., 2005, Cruz et al., 2007; Tian et al., 2010). However, important questions remain to be answered about the mechanisms and regulation of nitrogen handling and about N transfer from the fungus to the host plant. 91 In the current model, arginine translocation is responsible for all or almost all N movement from ERM to IRM (Bago et al., 2001, Tian et al., 2010)), but a possible role for nitrate in translocation is unknown. Kaldorf et al. (1998) reported that the expression of plant nitrate reductase is lower in mycorrhizal than non-mycorrhizal maize plants, suggesting that N transfer is predominantly not in this form. However, Faure et al. (1998) reported that nitrate reductase is increased in leaves of mycorrhizal plants independently of P status, and this was interpreted as indicating that nitrate is transferred from fungus to plant. Direct assessment of nitrate movement within the fungus and/or between IRM and the plant would be valuable in answering this question, and has not been addressed in previous studies that tracked 15N and 13C labeling using LC/MS methods that do not detect inorganic nitrogen molecules. Likewise, the role if any of ammonium in the translocation of N within the fungus is not known although it can accumulate to millimolar levels in some fungi (Jennings, 1995), and Chalot et al. (2006) suggested that ammonium might be taken up into vesicles within the mycorrhizal fungi and released to the host by exocytosis. Although AM, as well as ectomycorrhizal fungi, have the enzymes of the GS/GOGAT pathway for N assimilation (Chalot et al., 1994; Johansen et al., 1996; Tian et al. 2010), the role, if any, of glutamate dehydrogenase (GDH) in nitrogen assimilation in the ERM of the arbuscular mycorrhizal is still unknown. In Tilia platyphyllos-Tuber borchii ectomycorrhizae, N assimilation in the mycelium is the combined function of NADPH dependent GDH and GS (Pierleoni et al., 2001). However, Morel et al. (2005) found that NADP-GDH is dispensable for ammonium assimilation by ECM fungi. The expression of a putative GDH gene from R. irregularis was reported (Govindarajulu et 92 al., 2005) although this now appears not to be a GDH (Tian and Shachar-Hill unpublished) and another gene was annotated with this function (Tisserant et al. 2012, 2013) but none of the published data provide evidence for its function. The breakdown of ornithine (a product of arginine hydrolysis) in the IRM yields glutamate, apparently via ornithine aminotransferase, and ornithine may also be converted to putrescine via ornithine decarboxylase, but the importance and fates of these products is unclear. Glutamate in the IRM might be broken down to release ammonium or metabolized by amino-transferase reactions to produce other amino acids. Also, the origin of ornithine in ERM is still not well understood, although enzymes that can be involved in its synthesis, including ornithine amino-transferase and other urea cycle enzymes are upregulated in the ERM upon N addition. Specifically, since carbon is not transferred to the plant upon arginine breakdown (Jin et al. 2005) in the IRM, the carbon skeleton may return to the ERM for re-use in ornithine and arginine or their precursors. The expression of N transport and metabolism genes in fungi is commonly regulated by the levels of nitrogenous metabolites such as Arg, Gln and Orn (nitrogen metabolite control, Hinnebusch, 1988; ter Schure et al., 2000) and the levels of amino acids in presymbiotic AMF tissue are increased substantially in the presence of inorganic N (Gachomo et al., 2009). However R. irregularis, which is the primary model AM fungus in molecular genetics and biochemical studies, does not down-regulate the genes for Arg synthesis in the ERM when intracellular Arg levels are high, as occurs for example in S. cerevisiae suggesting either that Arg is sequestered, or perhaps that host demands for N are communicated to maintain N flow. Thus studying the gene 93 expression of the key genes of the pathway coupled with 15N labeling can help to understand how the genes involved in nitrogen metabolism and transport are regulated (Jin et al., 2010). In this study, isotopic labeling time course experiments using different 13C 15N and substrates were performed and the levels and labeling patterns of putrescine, pyrroline-5-carboxylate and nitrate in addition to the amino acids glutamate, glutamine, ornithine and arginine were measured. The identification and expression time courses of several new N metabolic enzyme transcripts were measured over the same time course. The results extend our understanding of the N transfer network, including the origin and fate of key intermediates and shed light on its regulation. Experimental procedures Chemicals and reagents Gelzan, (MP Biomedical, Solon, OH) was used for solidification of M media. 13C- argiinine and 15N labeled potassium nitrate were obtained from MP Biomedicals, (Solon, OH). Spore material and mycorrhizal in vitro growth conditions The spore material of R. irregularis (DAOM 181602) was purchased from Premier Tech Biotechnologies in units of 106 and was stored at 4°C until further use. Ri T-DNA-transformed carrot (Daucus carota clone DCI) roots were grown at 25°C in modified medium (Bécard and Fortin, 1988) with 3.5 g L−1 Phytagel (Sigma) using the 94 split-plate method of St-Arnaud et al. (1996). The roots and fungus were allowed to proliferate on both sides of bicompartmented petri plates at 25°C until the fungal ERM was well developed (approximately 6 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 with no nitrogen. The fungal ERM typically grew over the barrier within 2 week of the transfer, colonizing the empty compartment. Root growth over the barrier after transplantation was prevented by pruning. Isotopic labeling in colonized Ri T-DNA-transformed carrot (Daucus carota clone DCI) roots culture system 15N labeling The medium of the root compartment was modified to limit the nitrogen concentration to 1 mM. The ERM was allowed to cross over the divider into the fungal compartment that contained modified medium with no N or Sucrose added and only 2 g L−1 Phytagel. After 2-3 weeks, the ERM were supplied with 4 mM KNO3 for gene expression studies or K15NO3 to determine the labeling percentage of the free N metabolites and 15NO 3 in the IRM part and plant tissues. The colonized roots and ERM samples were collected after 0, 2, 4, 8, 16, 24, and 72 h, rinsed with sterilized water, and immediately frozen in liquid N and stored at −80°C 13C labeling [13C6]arginine (0.5 mM) was supplied to ERM. ERM and AM roots samples were collected after 0, 3, 6, 12, 24, and 48 h for metabolite analysis, rinsed with sterilized water, immediately frozen in liquid N and stored at −80°C 95 15N labeling for 15NO3 measurements in whole plant-mycorrhizal system Medicago truncatula plants were grown for 5 weeks in two-compartment culture system (growth condition and the culture system mentioned in the methods of Chapter 2) after that plants were deprived from N for 4 days at which time 10 mM K15NO3 was added to the reservoir of fungal compartment for two weeks. Plant shoots and roots were collected for N metabolite analysis. Two controls were used, mycorrhizal colonized plants with no 15N added and non-mycorrhizal plants with 10 mM K15NO3 added to fungal compartment (FC) reservoir to test leakage. Total of 10 biological replicates in two different experiments were used. Extraction and isolation of 15N and 13C metabolites A total of 300 mg of AM roots (or plant roots and shoots in whole plant system) were ground in a mortar and pestle with a pinch of acid-washed sand and extracted three times with a mixture of methanol:chloroform:water (12:5:3, v/v/v). Methylene chloride and water were added to the extraction solution to facilitate the separation of chloroform and the methanol–water phases. The methanol–water phase containing the amino acids (AAs) was collected and evaporated in a rotary evaporator at 50°C, and the residue containing the AAs and N metabolites were dissolved in 1 ml of 0.01 M HCl. Lyophilized fungal mycelium was pulverized with two 3-mm stainless steel beads using a bead mill (Retsch MM301). Cold methanol:water (70:30, 0.2ml) was added to aid in disruption. The samples were shaken at 30 Hz for 4 min, and 2-μL samples were analyzed by dissecting microscope to ensure that hyphae and any spores had been 96 broken. After disruption, 0.8 mL of cold methanol:water (70:30) was added, and the sample was vortexed for 5 min. Samples were then centrifuged and the supernatants collected. The cold aqueous methanol extraction was repeated twice more using 1 mL each time and the supernatants pooled. Then the methanol-water mixer were evaporated in a rotary evaporator at 50°C, and the residue containing the AAs and N metabolites were dissolved in 1 ml of 0.01 M HCl. ERM and AM root samples in 1 ml of 0.01 M HCl were loaded onto a cation exchange column (0.3 ml of DOWEX 50 X8-200– hydrogen form; Sigma-Aldrich, St Louis, MO, USA), which was previously washed with 1 M NH4OH, deionized H2O and 1 M HCl, and followed by deionized H2O. after loading the sample, the columns were washed 5 times with water so the neutral compounds and anions were eluted so that they were collected for nitrate measurements and the free amino acids were eluted with 5 ml of 1 M NH4OH (Bengtsson & Odham, 1979). This eluent was collected and dried then resuspended in 70 µl of Milli-Q water. Liquid chromatography and mass spectrometry (LC-MS) A Waters (Milford, MA) Quattro micro mass spectrometer coupled to a Shimadzu (Columbia, MD) LC-20AD HPLC system and SIL-5000 autosampler was used. A Waters Symmetry C18 column (2.1 × 100 mm, 3 μm particle size) was used with column oven temperature at 30 °C. The injection volume was 10 μL, and the HPLC flow rate was 0.3 mL/min using 1 mM perfluroheptanoic acid in a water/acetonitrile (A/B) gradient at ambient temperature, The intial gradient (A/B)=99/1, held until 2 minutes, 97 followed by a linear gradient to 60/40 at 4 minutes, held at 60/40 until 8 minutes, followed by a ramp to 99/1 at 8.01. Mass spectra were acquired using electrospray ionization in positive ion mode and MRM. The capillary voltage, extractor voltage, and rf lens setting were set at 3.17 kV, 4 V, and 0.3, respectively. The flow rates of cone gas and desolvation gas were 20 and 400 L/h, respectively. The source temperature and desolvation temperature were 110 and 350 °C, respectively. Collision-induced dissociation employed argon as collision gas at a manifold pressure of 2 × 10-3mbar, and collision energies and source cone potentials were optimized for each transition using Waters QuanOptimize software. This method was composed of two ESI+ functions (0−1.8 and 1.8−6.0 min) covering full run time to allow for adequate dwell time for each analyte. Data were acquired with MassLynx 4.0 and processed for calibration and for quantification of the analytes with QuanLynx software. A table (2-1) of (unlabeled) precursor and product ions for each compound is given in chapter 2. To determine concentrations, standard curves were run for authentic unlabeled standards of each compound in the table above with concentrations range from 0 to 100 μM together with 10uM Phenylalanine-D8 as internal standard, which was also added to each biological sample during extraction to control for recovery efficiency and as internal standard. Integrated peak areas were obtained from daughter ion chromatograms using the MassLynx 4.0. Total levels for each compound were obtained from the sum of all the labeled and unlabeled molecules. For 15N and 13C labeled samples, each product ion mass isomer for each precursor mass isomer was quantified. Thus for example for 15N labeling in Arginine the precursor ions have mass-to-charge ratios (m/z) of 175 (15N0 = M+0), 176 (15N1 = 98 M1), 177 (15N2 = M+2), 178 (15N3 = M+3), and 179 (15N4 = M+4) with the product ion CH6N3+ have m/z of 60 (15N0 = M+0), 60 and 61 (15N1 = M+1), 61 and 62 (15N2 = M+2), 62 and 63 (15N3 = M+3), and 63 (15N4=M+4) (see Table 1-2, 2-2, 3-1) . The proportion of N atoms which were 15N labeled was calculated ( calculation is described in Figure 3-3) to shows how much of the total N in that metabolite pools is new. All calculations was done after correction for natural abundance, mainly 13C. 15N labeled and unlabeled nitrate measurements 3200 QTRAP® LC/MS/MS coupled to a Shimadzu (Columbia, MD) LC-20ADvp HPLC system and SIL-HTC autosampler was used. ZIC®-pHILIC column (50 x 2.1 mm, 5 µm particle size) was used with column oven temperature at 40 °C. The injection volume was 1 μL, and the flow rate was 0.2 mL/min using water/acetonitrile (A/B) gradient at ambient temperature. The initial gradient (A/B)=10/90, held until 2 minutes, followed by a linear gradient to 30/70 at 3 minutes, held at 30/70 until 4 minutes, followed by a ramp to 10/90 at 4.01. Mass spectra were acquired using turbo VTM ion source and multiple ion monitoring scan type in negative ion mode. Masses of 62 (unlabeled nitrate) and 63 (15N labeled nitrate) were measured at retention time of 1.3 min. To determine the concentration of authentic 15N 15N labeled and unlabeled nitrate, standard curve were run for labeled nitrate standard with concentration range from 0 to 100 μM. Data were processed for calibration and for quantification of the analytes with QTRAP analyst software. 99 RNA extraction and putative gene fragment isolation Sequences of nitrite reductase, ornithine transcarbamylase, pyrroline-5- carboxylate dehydrogenase, and glutamate dehydrogenase gene fragments were identified from EST sequences obtained by high throughput RNAseq of R. irregularis transcriptome performed at MSU (Tian et al. 2010). To confirm their identity total RNA was extracted using the RNeasy Plant Mini kit (Qiagen, Valencia, CA) from R. irregularis germinating spore tissue that was disrupted using a bead mill, followed by DNA removal using RNase-free DNase (Turbo DNA-free; Ambion, Austin, TX). cDNA was synthesized using SuperScript II reverse transcriptase (Invitrogen, Carlsbad, CA). Primer sets were developed from EST sequence and designed using the primer3 website (http://frodo.wi.mit.edu/primer3/). PCR was performed and the products were separated by agarose gel electrophoresis then extracted using the QIAquick Gel Extraction kit (Qiagen) and sequenced at the MSU RTSF on an ABI PRISM 3100 Genetic Analyzer (Applied Biosystems, CA, USA). Sequences were compared with the R. irregularis genomic database to confirm identity. Quantitative Real-Time PCR measurements Mycorrhizal split plates were grown until the fungal compartment was approximately one-half colonized. To the fungal compartment, 1 ml of sterile KNO3 solution was applied to give a final concentration of 4 mM KNO3. Plates were incubated for 0, 2, 4, 8, 16, 24 and 72 h before tissue from 9-12 plates was collected and immediately frozen in liquid N. RNA was extracted and converted to cDNA as described 100 above. The initial quantitative real-time PCR (qRT-PCR) reaction mixture containing primers at a concentration of 300 nM and 1 ng of cDNA template. The PCR reactions were monitored using an ABI Prism 7900 HT Sequence Detection system (Applied Biosystems, CA, USA) with the following cycling program: 50°C for 2 min, 95°C for 10 min, followed by 40 cycles of 95°C for 15 s, and 60°C for 1 min. Power SYBR Green 2Step Master mix (Applied Biosystems, CA, USA) was used for all real-time PCR assays. The ΔΔCT, 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 normalize relative gene expression data as described by Govindarajulu et al. (2005). Primers of the four genes above and previously used primers (Table 3-1) and the closest homology of the newly identified genes are listed in Table 3-2. Results Nitrate movement through the AM symbiosis The fungal ERM of transformed carrot roots was exposed to 4 mM K15NO3 in the fungal compartment of split plate in vitro mycorrhizal cultures (Fortin et al. 1996) so that the ERM but not the roots had access to labeled nitrate. The levels of labeled and unlabeled nitrate (15NO3- and 14NO -) 3 in the mycorrhizal roots were measured over the next 3 d using LC-MS/MS and the percentage labeling was calculated by dividing the concentration of 15N nitrate by the total concentration of labeled plus unlabeled nitrate. The percentage labeling of 15NO 3 increased rapidly in extracts of AM roots reaching 101 approximately 20% within 8 h (Figure 3-1). The percentage labeling of 15NO 3 reached 50% after 24 h with no significant increase between 24 and 72 h (Figure 3-1). The mycorrhizal carrot roots contained 1-2% fungal biomass, so it is uncertain whether the labeled nitrate remained in the fungal tissue or was transferred to plant cells. To assess transfer of nitrate from the fungus to the host plant, an aseptic microcosm (see Chapter 2) system was used to perform a similar experiments with mycorrhizal M. truncatula plants. Levels of 15NO 3 and 14NO 3 were measured in root and shoot tissues after the ERM in the fungal compartment was exposed to 10 mM K15NO3 for two weeks. The mycorrhizal plants showed high 15NO 3 concentrations in both roots and shoots (Figure 3-2). The control plants (non-mycorrhizal plants with the second compartment or mycorrhizal plants with no 15NO 3 15N 15N added to added) showed no labeling and low levels of 14NO3 – consistent with the N deprivation pretreatment. N movement through the AM symbiosis in transformed root system The levels and labeling of the free N metabolites glutamine, glutamate, ornithine, arginine, pyrroline-5-carboxylate, and putrescine in the ERM and AM roots were measured using LC-MS/MS after the ERM was exposed to 4 mM K15NO3. The fraction of N in each metabolite pool that was labelled with 15N 15N is expressed as the amount of in molecules of that metabolite divided by the total amount of N (both labeled and unlabeled) in molecules of that metabolite. The amount of 15N in a metabolite pool was calculated from the sum of the MS signals from each mass isomer multiplied by the number of 15N atoms in that isomer: (0xM0 + 1xM1 + 2xM2 …), and the total nitrogen in 102 the molecule was calculated as the sum of all the mass isomers multiplied by the number of N atoms in the one metabolite molecule (one for glutamate, two for ornithine, four for arginine etc). Fractional labeling levels are shown as a function of time after 15N addition in Figure 3-3. Significant labeling levels were observed in free N metabolites within 2 h after 15N addition, and within 8 h about 40% of the arginine molecules were 15N labeled in the ERM. Pyrroline-5-carboxylate, glutamate, glutamine, ornithine also became rapidly labeled within the ERM, consistent with their serving as precursors for arginine biosynthesis. After 24 h, more than 25% of N in glutamate, glutamine, ornithine and arginine molecules were labeled within the ERM. On the other hand, about 25% of N in pyrroline-5-carboxylate molecules were labeled after 24 h within the ERM and the labeling levels rose in a hyperbolic manner indicating saturation of the labeling levels significantly below full labeling. No 15N labeling of putrescine was detectable in the ERM. After 72 hours more than 80% labeling of the amino acid molecules contained one or more 15N atoms. In the colonized roots (IRM plus plant cells), the labeling of metabolites showed a lag of ~2h and was slower than in the ERM over the first 8 h. The fractional labeling of N in arginine molecules reached approximately 15% after 8 h, and rose to about 80% by 72h. A total of 30% of N in ornithine molecules were labeled within the colonized roots at 16 h. The other metabolites showed similar patterns of AM roots. The 15N 15N fractional labeling in the fractional labeling of arginine and ornithine and other metabolites is consistent with arginine was transport from ERM to the IRM and its breakdown there via arginase. 15N- putrescine was detectable in IRM after 8 h and reaching more than 15% fractional labeling after 72 h (Figure 3-3). 103 The origin and fate of the N metabolites In order to study the origin and fate of the metabolites in the ERM, isotopomers (chemically identical molecules containing different numbers of 15N 15N atoms in different positions) of glutamate, ornithine and arginine were measured using intact molecular ions and fragment ions to obtain the number of 15N atoms and information on positional labeling from mass spectra. M+1 isotopomers were detectable first and higher mass isotopomer levels increased after that. M+2 ornithine molecules were detected at significant levels (p<0.01) after 8 h while the fully labeled M+4 arginine mass isomer was detectable after 4 h (Figure 3-4). 13C labeling in the ERM and AM roots The levels and 13C labeling of free N metabolites (glutamate, pyrroline-5- carboxylate, glutamine, ornithine, arginine and putrescine) in the ERM and AM roots of mycorrhizal transformed carrot roots were measured after the ERM was exposed to 0.5 mM [13C6]arginine. [13C6]arginine was found in the ERM within 3 h (Figure 3-5E). [13C5]Arginine was detected in the ERM at 12 h and then rapidly increased. On the other hand, glutamine 13C , 5 glutamate 13C , 5 putrescine 13C 4 were not detectable at any time in the ERM. Ornithine and pyrroline-5-carboxylate 13C 5 13C 5 was detectable in the ERM at 12 h and then rapidly increased (Figure 3-5C). In the AM roots, [13C6]arginine appeared first at 3 h and its level rapidly reached over 30 nmol/g at 12 h. No significant increase of [13C6]arginine levels were observed between 12 and 24 h, levels then doubled between 24 and 48 h. [13C5]Arginine was 104 detectable at 24h in AM roots and reached over 50 nmol/g. [13C5]Arginine level rose between 24 and 48 h (Figure 3-5E). Glutamine was not detectable at any time. Ornithine 13C5 was detectable in AM roots within 3 h while glutamate 13C5, and pyrroline5-carboxylate 13C 5 and putrescine were detectable in AM roots after 6 h and completely absent in ERM 13C 4 was detectable after 12h in AM roots but not detectable in ERM (Figure 3-5). Gene identification Based on sequence data previously deposited in public databases (GenBank), sequences from R. irregularis for four putative enzymes of N metabolism were identified. Partial transcarbamylase coding (OTC), sequences of nitrite pyrroline-5-carboxylate reductase (NiR), dehydrogenase ornithine (P5CD) and glutamate dehydrogenase (GDH) were obtained that show high sequence similarities to known genes involved in N uptake and metabolism in fungi and bacteria (Table 3-3). Nitrite reductase was identified with 83% similarity at the amino acid level to the Laccaria bicolor NADPH-nitrite reductase (EC 1.7.1.4), an assimilatory nitrite reductase which catalyzes the NADPH-dependent formation of ammonia from nitrite. Ornithine carbamoyltransferase was identified with 81% similarity to the Cryptococcus gattii (WM276) Ornithine carbamoyltransferase, mitochondrial precursor (OTCase; EC 2.1.3.3) which catalyze catalyzes the reaction between carbamoyl phosphate and ornithine to form citrulline and phosphate. Pyrroline-5-carboxylate dehydrogenase was identified with 63% similarity to Cryptococcus neoformans var. neoformans 105 JEC21 1-pyrroline-5-carboxylate dehydrogenase (EC 1.5.1.12) which catalyze the reversible conversion of pyrroline-5-carboxylate to glutamate. Furthermore, glutamate dehydrogenase was identified with 63% similarity with glutamate dehydrogenase NAD+ (EC 1.4.1.2) of Sphingobacterium spiritivorum (ATCC 33861), which catalyzes the reversible oxidative deamination of glutamate to alpha-ketoglutarate and ammonia. Gene expression in response to N addition Based on the current model and previous results (Cruz et al., 2007; Tian et al., 2010) we hypothesized that the expression of the genes involved in N movement and metabolism from soil through the fungus and into the host is temporally and spatially coordinated with the flux of N. The transcriptional levels for eight genes from R. irregularis (4 of them identified here, the others identified and functionally confirmed previously) were measured in ERM and IRM tissues by quantitative real-time PCR. The transcript levels of the putative assimilatory nitrite reductase (NiR) and ornithine transcarbamylase (OTC) increased strongly in the ERM beginning by 2 h after 4 mM KNO3 was added to the fungal compartment. NiR and OTC were highly upregulated after 4 h of nitrate addition. By contrast ornithine aminotransferase 1 (OAT1) and pyrroline-5-carboxylase (P5CD) were modestly upregulated and reached over 2folds expression after 4 h in the ERM. Arginase (CAR1), ornithine aminotransferase 2 (OAT2), glutamate dehydrogenase (GDH), and ornithine decarboxylase (ODC) transcript levels in the ERM were little affected by the supply of nitrate. By contrast, the expression of CAR1, OAT1, OAT2, P5CD and GDH was 106 substantially up-regulated in the IRM within 24 h. The transcript level of ODC increased modestly in the IRM by 24 h. NiR transcript level was low in the IRM and not upregulated over the time course. Discussion The current working model of N transfer from the fungi to the plant in the AM symbiosis was proposed by Bago et al. (2001) based on previous work that demonstrated fungal N uptake and metabolism and implicated amino acids in N handling (Johansen et al., 1996; Bago et al., 1996; 2000). Supporting evidence for the model (Govindarajulu et al., 2005, Cruz et at., 2007, Jin et al., 2005, Tian et al., 2010) has led to its widespread acceptance (Smith and Read 2008, Parniske et al., 2009, He et al., 2011). However, the network is incomplete on the question of the fate of carbon translocated from ERM to IRM in the form of arginine and the potential remains for other forms of N besides arginine to be involved in N movement from ERM to the IRM and roots. The detection of high 15N labeling of nitrate in AM carrot roots after 15N nitrate was supplied to the ERM (Figure 3-1) indicates that not all of the nitrate taken up by the ERM is reduced and assimilated there. Because total nitrate levels are not high, it may be that the nitrate translocated to the colonized roots remains in the IRM. To address this question, we conducted gene expression analysis of the fungal putative assimilatory nitrite reductase in the colonized transformed roots and ERM as well as a 15N nitrate labeling experiment using the whole plant microcosm described in Chapter 2. NiR 107 expression is upregulated in the ERM but not IRM in response to nitrate addition (Figure 3-6) suggesting that the nitrate that reaches the IRM is not converted to ammonia. Tian et al., (2010) reported that other genes involved in N assimilation, such as GS/GOGAT were not upregulated in the IRM. The results of the whole plant experiment showed that the roots and shoots both had high 15N nitrate labeling levels while control treatments showed no labeling (Figure 3-2). We conclude that nitrate also contributes to N transfer by AMF to host plants so that arginine synthesis and translocation within the fungus followed by its breakdown and the transfer of ammonium to the plant is not the only route for N transfer in AM symbiosis. After supplying K15NO3 to the ERM, rapid 15N labeling was found in N metabolites in the ERM and in the AM roots (Figure 3-3). This is consistent with previous observations of amino acid labeling by Tian et al. (2010). The increase of the ornithine 15N fractional labeling in the ERM and in the mycorrhizal roots coincided with the increase of labeled arginine (Figure 3-3 C,D), which is consistent with their close location in the metabolic network. These results are consistent with previous observations (Jin et al. 2008) and with the synthesis of arginine in ERM and its breakdown in the IRM as previously proposed (Bago et al. 2001). Neither the data shown in Figures 3-1 - 3-3, nor previous studies provide a complete picture of the origins of ornithine in the ERM or its fate in the IRM (Govindarajulu et al., 2005; Jin et al., 2005; Cruz et al., 2007). Ornithine released during arginine breakdown in the IRM might be stored, transferred to the host, broken down or translocated to the ERM. Govindarajulu et al., (2005) reported that ornithine was neither stored in substantial quantities in the IRM nor is transferred to the host in detectable 108 amounts. Breakdown of ornithine in the IRM with the release of ammonium would seem to be more efficient for N transfer to the host compared with its translocation back to the ERM, which would return half of the N in arginine to its original location. However, the simultaneous de novo synthesis of ornithine in the ERM and its catabolism in the IRM also consumes energy and carbon. The presence of 15N labeled pyrroline-5-carboxylate as well as 15N glutamate in the ERM (Figure 3-3) is consistent with de novo ornithine biosynthesis, since pyrroline5-carboxylate is the intermediate metabolite for the formation of ornithine from glutamate via P5CD. Gene expression analysis of OAT1 and P5CD showed 3-fold increases in gene expression for these genes 4 h after adding nitrate to the ERM. Both of those enzymes catalyze the reversible interconversion of ornithine and glutamate. However, OAT1 and P5CD were not highly upregulated in the ERM compared with the IRM (Figure 3-6). Enzyme activity analyses will be required to determine which of the routes is more active. The absence of detectable levels of 15N putrescine in the ERM is consistent with the gene expression analysis of ODC, which was not upregulated in the ERM after N addition. By contrast, significant levels of 15N putrescine were found in the IRM later during the time course, which coincided with an approximately 3-fold ODC gene expression increase in IRM 24 h after nitrate addition to the ERM (Figure 3-6). Significant amounts of 15N glutamate and 15N pyrroline-5-carboxylate were found in the IRM coinciding with the upregulation of OAT1, OAT2 and P5CD. Based on the levels and fractional labeling of putrescine, pyrroline-5-carboxylate and glutamate as well as OAT1, OAT2 and P5CD gene expression analysis, we conclude that significant 109 amounts of ornithine are converted to glutamate and that lower quantities are converted to putrescine in the IRM. The levels of 15N isotopomers of glutamate, ornithine and arginine were measured in the ERM after adding K15NO3 to the ERM (Figure 3-4). The increase in M+1 (Non R group N(NR) ornithine in the ERM coincided with the increase in M+1 15N- glutamate is consistent with the formation of ornithine from glutamate. M+2 ornithine levels increased with the increase of 15N arginine isotopomers containing 2 or more atoms and these increased more rapidly when of 13C 6 15N arginine M+4 levels rise. Application arginine to the ERM resulted in detectable level of arginine ornithine 13C 5 15N 13C 6 within 3 h while was detectable in the ERM only after 12 h (Figure 3-5). The absence of ornithine 13C5 before 12 h and the appearance of ornithine 13C5 in AM roots indicate that arginine was broken down in the IRM but not in the ERM (consistent with low arginase expression there). The absence in the ERM of 13C labeled N metabolites other than ornithine indicates that ornithine is not broken down in the ERM. Several observations are consistent with translocation of ornithine from IRM to ERM. The appearance of 13C 5 ornithine in the ERM is substantially delayed compared to the IRM. This could be due to either the return of 13C 5 ornithine from IRM to ERM or to a delayed induction of arginase in the ERM. The total levels of arginine, ornithine and other amino acids in the ERM are not affected by arginine uptake (which is much slower than the uptake of inorganic N substrates) so that a delayed induction of arginase activity in the ERM is improbable as an explanation. Indeed arginase expression and activity were not induced in the ERM by N addition, (this study, Tian et al., 2010 and Cruz et al,. 2007). 13C 5 arginine, which is made from 110 13C 5 ornithine, was detected in the ERM much later than 13C 5 of 13C 6 arginine, which is consistent with the delayed appearance of ornithine in the ERM as a precursor not a product of 13C 5 13C 5 arginine. The appearance arginine in ERM preceded its appearance in the IRM, consistent with the unidirectional translocation of arginine from ERM to IRM (Figure 3-5). The fractional labeling of ornithine in the IRM is higher as well as rising faster than that in the ERM (since carbon is not transferred to the plant, ornithine in the plant cells is unlabeled so that the measured fractional labeling of ornithine in the AM roots represents a still higher labeling level in the IRM). This is necessary if ornithine in the IRM is the precursor of ornithine in the ERM. The presence of glutamate putrescine 13C 4 13C , 5 pyrroline-5-carboxylate 13C 5 and indicated that some ornithine is being converted to them in the IRM of AM roots. The absence in the ERM of 13C labeled glutamate, glutamine, pyrroline-5- carboxylate or putrescine despite their accumulation in the AM roots indicates that none of these metabolites returns to the ERM from the IRM. The possible role of glutamate dehydrogenase in N movement is still unknown. Cliquet and Stewart (1993) reported the activities of the GS/GOGAT pathway but not assimilatory NADP-dependent glutamate dehydrogenase (GDH) in mycorrhizal roots, and the application of a GOGAT inhibitor to extraradical mycelium reduced 15N assimilation. A putative GDH gene was identified and it is most probably dissimilatory GDH. The gene expression of GDH in slightly upregulated in ERM and highly unregulated in IRM indicating that GDH breaks down glutamate to release more ammonium to the host in IRM. Within 2 h of nitrate addition to the ERM, the expression of the OTC transcript was upregulated in the ERM, coinciding with the building up of arginine since OTC is 111 part of urea cycle. On the other hand, OTC was not upregulated in the IRM, which is consistent with the previously reported down regulation of other arginine biosynthesis genes in the IRM (Tian et al., 2010). The up regulation in the IRM of the expression of fungal genes involved in arginine and ornithine breakdown (Arginase: CAR1, OAT1, OAT2 as well as P5CD and ODC) coincides with the arrival of 15N label and a rise in arginine levels within AM roots. Thus, it is likely that this is the signal for regulatory gene expression in the IRM. 112 APPENDIX 113 APPENDIX 70 50 40 30 15NO 3 Percentage labeling 60 20 10 0 0 10 20 30 40 50 60 70 80 Time (h) Figure 3-1 Time course of 15N percentage labeling of nitrate in transformed carrot (Daucus carota) roots colonized by R. irregularis . 15N labled and unlabeld nitrate were measured after supplying 4 mM K15NO3 to the fungal extraradical mycelium (ERM) compartment in divided petri dishes. Values are reported as mean ± SEM (n=3 biological replicates). 114 15 Unlabeled nitrate 15 N labeled nitrate Non-mycorrhizal plant + N nmol/g (FW) 2 15 Mycorrhizal plant + N 20 * 1.5 1 0.5 nmol/g(FW) 15 * 0 Root 10 Shoot Mycorrhizal plant with no 5 Root nmol/g (FW) 0 Shoot 15 N 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 Root Shoot Figure 3-2 The concentrations of 15N labeled nitrate (light grey bars) and unlabeled nitrate (black bars) in the roots and shoots of mycorrhizal M. truncatula plants grown in micosms. Mycorhizal plant +15N: 10mM K15NO3 was supplied to the fungal compartment reservoir (FCR) for two weeks before tissue collection and analysis. Nonmycorrhizal plant +15N: non-mycorrhizal plants supplied with 10mM 15N nitrate in the FCR– controlling for non-mycorrhizal N levels and for non-fungally mediated N label movement. Mycorrhizal plant with no 15N: mycorrhizal plants with no 15N added – controlling for N deprived N levels in myc plants and for natural abundance isotope levels. Values are reported as mean ± SEM (n=10 biological replicates). Statistical analyses were done by ANOVA single factor analysis, alpha = 0.01. Stars show statistically significant 15N labeling (p<0.01). 115 A 15N-Glutamate ERM 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0.8 0.7 Fractional labeling Fractional labeling AM roots 0.6 0.5 0.4 0.3 0.2 0.1 0 0 8 16 24 32 40 48 56 64 72 0 Time (h) 8 16 24 32 40 48 56 64 72 Time (h) Figure 3-3 Nitrogen movement from the ERM to the IRM in mycorrhizal transformed carrot roots . These resullts are based on the timing of labeling in N metabolites in the ERM and AM roots after the addition of 4 mM K15NO3 to the fungal ERM. The fraction of N in each metabolite pool that is labelled with 15N is expressed as the amount of 15N in molecules of that metabolite divided by the total amount of N (both labeled and unlabeled) in molecules of that metabolite. The amount of 15N in a metabolite pool is calculated from the sum of the MS signals from each mass isomer multiplied by the number of 15N atoms in that isomer: (0xM0 + 1xM1 + 2xM2 …), and the total nitrogen in the molecule is calculated as the sum of all the mass isomers multiplied by the number of N atoms in the one metabolite molecule (one for glutamate, two for ornithine, four for arginine etc). Fractional labeling of A) glutamate, B) pyrroline5-carboxylate, C) glutamine, D) ornithine, E) arginine in ERM and AM roots, and F) putrescine in AM roots only as 15N-labeled putrescine was not detectable in ERM. Means and standard error of the means of three replicates. 116 Figure 3-3 (cont’d) B 15N-1-Pyrroline-5-carboxylate ERM AM roots 0.55 0.5 0.45 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 0.5 0.45 Fractional labeling Fractional labeling 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 0 8 0 16 24 32 40 48 56 64 72 8 16 24 32 40 48 56 64 72 Time (h) Time (h) C 15N-Glutamine 0.7 0.7 ERM AM roots 0.6 0.5 0.5 Fractional labeling Fractional labeling 0.6 0.4 0.3 0.2 0.1 0.4 0.3 0.2 0.1 0 0 0 8 16 24 32 40 48 56 64 72 0 Time (h) 8 16 24 32 40 48 56 64 72 Time (h) 117 Figure 3-3 (cont’d) D 15N–Ornithine AM roots 0.6 ERM 0.5 Fractional labeling E Fractional labeling 0.6 0.4 0.3 0.2 0.1 0.5 0.4 0.3 0.2 0.1 15N Arginine 0 0 0 8 0 16 24 32 40 48 56 64 72 8 16 24 32 40 48 56 64 72 Time (h) Time (h) E 15N-Arginine 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0.7 AM roots 0.6 Fractional labeling Fractional labeling ERM 0.5 0.4 0.3 0.2 0.1 0 8 0 16 24 32 40 48 56 64 72 0 Time (h) 8 16 24 32 40 48 56 64 72 Time (h) 118 Figure 3-3 (cont’d) F 15N-Putrescine AM roots Fractional labeling 0.25 0.2 0.15 0.1 0.05 0 0 8 16 24 32 40 Time (h) 119 48 56 64 72 A Glutamate (M+1) 15N isotopomer 160 140 nmol/g (DW) 120 100 80 60 40 20 0 0 10 20 30 40 Time (h) 50 60 70 80 Figure 3-4 Time course of the free N metabolites 15N-Isotopomers (mass isomers) in the extraradical mycelium (ERM) after the addition of 4 mM K15NO3 to the ERM. M+1, singly labeled, M+2 doubly labeled, etc. position of label within a molecule is indicated with inset chemical structures. The levels of 15N isotopomers of glutamate (A) ornithine (B) and arginine (C). Means and standard errors of means of three replicates. R, nitrogen is in the R-group; NR, nitrogen is not in the R-group; G, nitrogen is in the Guanidine group; NG, nitrogen is not in the Guanidine group. 120 Figure 3-4 (cont’d) B 15N Onithine isotopomers 1600 1400 nmol/g (DW) 1200 1000 M+1 (NR) 800 M+1 (R) 600 M+2 400 200 0 0 10 20 30 40 Time (h) 121 50 60 70 80 Figure 3-4 (cont’d) C NG 15N Arginine isotopomers G 3250 M+1 NG 3000 M+1 G 2750 M+2(NG&G) 2500 M+2 (2G) M+3 (2G&1NG) 2000 nmol/g (DW) nmol/g (DW) 2250 M+4 1750 1500 1250 1000 750 500 250 0 0 8 16 24 32 40 48 56 64 72 M+1 NG 200 190 180 170 160 150 140 130 120 110 100 90 80 70 60 50 40 30 20 10 0 M+1 G M+2(NG&G) M+2 (2G) M+3 (2G&1NG) M+4 0 80 2 4 6 8 10 12 14 16 18 Time (h) Time (h) 122 A 60 Glutamate 13C5 50 nmol/g (FW) 40 30 ERM AM roots 20 10 0 0 6 12 18 24 30 Time (h) 36 42 48 Figure 3-5 Timing course of 13C labeling of N metabolites in the extraradical mycelium (ERM) and the mycorrhizal roots after the addition of 0.5 mM 13C6 arginine to the fungal ERM compartment. Levels of A) glutamate 13C5, B) pyrroline-5-carboxylate 13C5, C) ornithine 13C5 (the right panel is the earlier time points of the left panel), D) arginine 13C6 (the right panel is the earlier time points of the left panel) and arginine 13C5 , and E) putrescine 13C4. Symbols show the means and standard errors of the means of three biological replicates. 123 Figure 3-5 (cont’d) B Pyrroline-5-carboxylate 13C5 45 nmol/g (FW) 40 35 30 25 ERM 20 AM roots 15 10 5 0 0 6 12 18 24 30 Time (h) 124 36 42 48 Figure 3-5 (cont’d) C Ornithine 13C5 1800 ERM AM roots 1500 Ornithine 300 13C 5 nmol/g (FW) nmol/g (FW) 250 1200 900 600 300 200 150 100 50 0 0 6 12 18 24 30 Time(h) 36 42 48 0 0 125 6 12 18 24 30 Time (h) 36 42 48 Figure 3-5 (cont’d) Arginine 13C 6 13C 6 nmol/g (FW) Arginine 10000 9000 8000 7000 6000 5000 4000 3000 2000 1000 0 ERM AM roots 0 6 12 18 24 30 36 42 48 Time (h) 900 1200 1100 1000 900 800 700 600 500 400 300 200 100 0 ERM AM roots 0 6 12 18 24 Time (h) Arginine 800 13C 5 700 nmol/g (FW) nmol/g (FW) D 600 500 ERM 400 AM roos 300 200 100 0 0 6 12 18 24 30 Time (h) 126 36 42 48 30 36 42 48 Figure 3-5 (cont’d) E 12 Putrescine 13C4 nmol/g (FW) 10 8 6 ERM AM roots 4 2 0 0h 3h 6h 12h Time 127 24h 48h A ERM 27 24 Relative expression 21 18 15 12 9 6 3 NiR OTC CAR1 OAT1 OAT2 P5CD GDH 2h 4h 8h 24h 72h 2h 4h 8h 24h 72h 2h 4h 8h 24h 72h 2h 4h 8h 24h 72h 2h 4h 8h 24h 72h 2h 4h 8h 24h 72h 2h 4h 8h 24h 72h 2h 4h 8h 24h 72h 0 ODC Figure 3-6 The expression of primary N metabolic and transport genes in the arbuscular mycorrhizal symbiosis after the addition of 4 mM KNO3 to the fungal extraradical mycelium (ERM). Bars are as follows: 2 h (white bars), 4 h (gray bars), 8 h (hatched bars), 24 h ( black bars), and 72 h (striped bars). Gene expression of nitrite reductase (NiR), ornithine transcarbamylase (OTC), arginase (CAR1), ornithine aminotransferase 1 and 2 (OAT1, 2), pyrroline-5carboxylate dehydrogenase (P5CD), glutamate dehydrogenase (GDH), and ornithine decarboxylase (ODC) was measured by quantitative real-time PCR. Gene expression was measured in ERM (A) and intraradical mycelium (IRM) (B) with fungal ribosomal protein S4 gene as the reference. Means and standard errors of means of three biological replicates and each had three technical replicates. 128 NiR OTC CAR1 OAT1 129 OAT2 P5CD GDH 2h 4h 8h 24h 72h 2h 4h 8h 24h 72h 2h 4h 8h 24h 72h 2h 4h 8h 24h 72h 24 2h 4h 8h 24h 72h 2h 4h 8h 24h 72h 2h 4h 8h 24h 72h 2h 4h 8h 24h 72h Relative expression Figure 3-6 (cont’d) B IRM 21 18 15 12 9 6 3 0 ODC Table 3-1 Multiple reaction monitoring (MRM) transitions, optimizing source cone voltages, collision cell voltages, and analyte retention time (for 13C labeling experiment). Compound Precursor 13 ion> ( C isotopomers) product ion (m/z) Arginine M+0: 175.1>60 M+5: 180.1>60 M+6: 181.1>61 Glutamate M+0: 148.02>83.84 M+5: 153.02>87.84 Glutamine M+0: 147.06>129.95 M+5: 152.06>134.95 Ornithine M+0: 133.1>70 M+5:138.1>74 Putrescine M+0: 89.03>71.79 M+4:93.03>75.79 Pyrroline-5M+0: 113.98>67.75 Carboxylate M+5: 118.98>71.75 Cone Voltage (V) Collision Voltage (V) Retention time (min) Function no. 22 22 6.08 1, 3 22 16 1.23 2, 3 16 10 1.07 2, 3 16 16 5.52 2, 3 50 10 6.05 2 16 10 1.15 2, 3 130 Table 3-2 Primers for real-time PCR. Primer Forward (5’-3’) Reverse (5’-3’) Nitrite reductase (GiNiR) Ornithine transcarbamylase (OTC) 1-Pyrroline-5carboxylate dehydrogenase (P5CD) Glutamate dehydrogenase (GDH) Ornithine aminotransferase 1 (GiOAT1) Ornithine aminotransferase 2 (GiOAT2) S4 ribosomal protein (GiSR4) CCAGCTATACGCGTCA ATTTT GCTCAACGTATAAAAG ATTTTGCTG AGGCGTAATTTCA CCTCCAG CATTAGTGCATCA ATAACGGCTA TTAAGACCGGTCCTCC TGAA AAACTTGGGCTTC CTGCTTT This study TTCCCTTTACACCATAA TACACACC CTGCGCCAGATAT GGGTACT This study GGTTCGAGCGGATATT GTCATAC AGGACTGCTGATA TTGGGTAAACG Tian et al., (2010) CGGGTAAGATGCTTTG TCAAGA Tian et al., (2010) AAGCCGCCTACGTGTC GTT GCCTGAAAGTGCT TTACCAAGTATAA C AACAGGTGGTAGA AATATGGGAAG TGATGCGGTGAATCCT AAGAGA TTGATTGCGTTACCAA AAATGG GATCAAGTGCATC AACGTCAAAG TCGAAATACAACC AGTCACCAAGA Arginase (GiCAR1) Ornithine decarboxylase (GiODC) 131 Referenc e This study This study Govindara julu et al., (2005) Tian et al., (2010) Tian et al., (2010) Table 3-3 Nitrogen metabolic genes identified in Rhizophagus irregularis in this study. Genes (encodes Closest the enzymes homolog below) (Nucleotides similarity) Blastn Accession: Nitrite reductase ZP_01854725 (Planctomyces (NiR) maris DSM 8797) Identities=51% Accession: Glutamate Q54KB7 (Dictyostelium dehydrogenase discoideum) (GDH) Identities= 61% Accession: DQ662599 Ornithine trans(Neocallimastix carbamylase frontalis) (OTC) Identities=76% Closest homolog R. irregularis genome (amino acids reference** similarity) (gene model name) Blastx Accession: ZP_01854725 (Planctomyces maris DSM 8797) Identities=56% Accession: EFA84685 (Polysphondylium pallidum PN500) Identities= 61% Accession: KFH69100 (Mortierella verticillata NRRL 6337) Identities= 68% Accession: Accession: KFH71447 KFH71447 Pyrroline-5(Mortierella (Mortierella carboxylate verticillata NRRL verticillata NRRL dehydrogenase 6337) 6337) (P5CD) Identities= 67% Identities= 67% ** http://genome.jgi-psf.org/Gloin1/Gloin1.home.html 132 CE150538_1056 e_gw1.12044.3.1 (accession: EXX78633) MIX6106_366_99 (accession: EXX65553) fgenesh1_kg.1256_#_3 _#_ACTTGA_L001_R1 _(paired)_contig_4977 REFERENCES 133 REFERENCES Ames RN, Reid CP, Porter LK, Cambardella C. 1983. Hyphal uptake and transport of nitrogen from two 15N-labelled sources by Glomus mosseae, a vesicular-arbuscular mycorrhizal fungus. New Phytologist 95: 381-396. Bago B, Vierheilig H, Piché Y, Ázcon-Aguilar C. 1996. Nitrate depletion and pH changes induced by the extraradical mycelium of the arbuscular mycorrhizal fungus Glomus intraradices grown in monoxenic culture. New Phytologist 133: 273-280. Bago B, Pfeffer P, Shachar-Hill Y. 2000. Carbon metabolism and transport in arbuscular mycorrhizas. Plant Physiology 124: 949-957. Bago B, Pfeffer P, Shachar-Hill Y. 2001. Could the urea cycle be translocating nitrogen in the arbuscular mycorrhizal symbiosis? New Phytologist 149: 4-8. Bècard G, Fortin JA. 1988. Early events of vesicular-arbuscular mycorrhizal formation on Ri T-DNA transformed roots. New Phytologist 108:211–218. Bengtsson G, Odham G .1979. A micromethod for the analysis of free amino acids by gas chromatography and its applications to biological systems. Analytical Biochemistry 92: 426-441. Benitez T, Farrar L. 1980. Induction of arginase and ornithine transaminase in the fission yeast Saccharomyces pombe. Journal of Bacteriology 144: 836–839 Chalot M, Brun A, Finlay RD, Soderstrom B. 1994. Metabolism of [14C] glutamine by the ectomycorrhizal fungus Paxillus involutus. Microbiology 140: 1641-1649. Chalot M, Blaudez D, Brun A. 2006. Ammonia: a candidate for nitrogen transfer at the mycorrhizal interface. Trends in Plant Science 11: 263-266. Clark RB, Zeto SK. 2000. Mineral acquisition by arbuscular mycorrhizal plants. Journal of Plant Nutrition 23: 867-902. Cliquet JB, Stewart CR. 1993. Ammonium assimilation in Zea mays 1. infected with a vesicular-arbuscular mycorrhizal fungus Glomus fasciculatum. Plant Physiology 101: 865-871. Cruz C, Egsgaard H, Trujillo C, Ambus P, Requena N, Martins-Loucao MA, Jakobsen I. 2007. Enzymatic evidence for the key role of arginine in nitrogen translocation by arbuscular mycorrhizal fungi. Plant Physiology 144: 782-792. Faure S, Cliquet JB, Thephany G, Boucaud J. 1998. Nitrogen assimilation in Loliun perenne colonized by the arbuscular mycorrhizal fungus Glonus fasciculatum. New Phytologist 138: 411-417. 134 Gachomo E, Allen JW, Pfeffer PE, Govindarajulu M, Douds DD, Jin HR, 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 184: 399-411. Govindarajulu M, Pfeffer PE, Jin HR, Abubaker J, Douds DD, Allen JW, Bucking H, Lammers PJ, Shachar-Hill Y. 2005. Nitrogen transfer in the arbuscular mycorrhizal symbiosis. Nature 435: 819-823. Hamel C. 2004. Impact of arbuscular mycorrhizal fungi on N and P cycling in the root zone. Canadian Journal of Soil Science 84: 383-395. Hinnebusch AG. 1988. Mechanisms of gene regulation in the general control of amino acid biosynthesis in Saccharomyces cerevisiae. Microbiological Reviews 52: 248273. Hodge A, Fitter AH. 2010. Substantial nitrogen acquisition by arbuscular mycorrhizal fungi from organic material has implications for N cycling. Proceedings of the National Academy of Sciences USA 107: 13754-13759. Jackson LE, Burger M, Cavagnaro TR. 2008. Roots, nitrogen transformations, and ecosystem services. Annual Review of Plant Biology 59: 341-363. Jakobsen I, Rosendahl L. 1990. Carbon flow into soil and external hyphae from roots of mycorrhizal cucumber plants. New Phytologist 115: 77-83. Jennings DH. 1995. The Physiology of Fungal Nutrition. Cambridge University Press, Cambridge, UK. 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. Johansen A, Jakobsen I, Jensen ES. 1993. Hyphal transport by a vesiculararbuscular mycorrhizal fungus of N applied to the soil as ammonium or nitrate. Biology and Fertility of Soils 16: 66-70. Johansen A, Finlay RD, Olsson PA. 1996. Nitrogen metabolism of external hyphae of the arbuscular mycorrhizal fungus Glomus intraradices. New Phytologist 133: 705712. Kaldorf M, Schmelzer E, Bothe H. 1998. Expression of maize and fungal nitrate reductase genes in the arbuscular mycorrhiza. Molecular Plant-Microbe Interactions 11: 439-448. Morel M, Buée M, Chalot M, Brun A. 2006. NADP-dependent glutamate dehydrogenase: a dispensable function in ectomycorrhizal fungi. New Phytologist 169: 179-190. 135 Pierleoni R, Vallorani L, Sacconi C, Sisti D, Giomaro G, Stocchi V. 2001. Evaluation of the enzymes involved in primary nitrogen metabolism in Tilia platyphyllos-Tuber borchii ectomycorrhizae. Plant Physiology 39: 1111-1114. Reich PB, Hobbie SE, Lee T, Ellsworth DS, West JB, Tilman D, Knops JMH, Naeem S, Trost J. 2006. Nitrogen limitation constrains sustainability of ecosystem response to CO2. Nature 440: 922-925. St-Arnaud M, Hamel C, Vimard B, Caron M, Fortin JA. 1996. Enhanced hyphal growth 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. ter Schure EG, van Riel NAW, Verrips CT. 2000. The role of ammonia metabolism in nitrogen catabolite repression in Saccharomyces cerevisiae. FEMS Microbiology Reviews 24: 67-83. Tian CJ, Kasiborski B, Koul R, Lammers PE, Bücking H, Shachar-Hill Y. 2010. Regulation of the nitrogen transfer pathway in the arbuscular mycorrhizal symbiosis: gene characterization and the coordination of expression with nitrogen flux. Plant Physiology 153: 1175-1187. Tobar RM, Azcón R, Barea JM. 2004. The improvement of plant N acquisition from an ammonium-treated, drought-stressed soil by the fungal symbiont in arbuscular mycorrhizae. Mycorrhiza 4: 105-108. Wang B, Qiu YL. 2006. Phylogenetic distribution and evolution of mycorrhizas in land plants. Mycorrhiza 16: 299-363. 136 Chapter 4 Conclusions and future research 137 Conclusions Nitrogen nutrition has been a topic of considerable interest and growing importance in mycorrhizal research. AMF can increase the uptake of N by host plants from the soil under natural and perturbed conditions, and transfer from AMF can account for significant proportions of N in plants in controlled model systems. However, our knowledge of the nutritional importance of, as well as of the enzymes and transporters involved in, N transfer is still limited. Furthermore, important questions remain to be answered about the mechanism and regulation of nitrogen handling and its transfer from fungus to the host plant. In this study, I highlighted recent findings about nitrogen transfer and metabolism in AM symbiosis and report my findings concerning the remaining unanswered questions. We developed a sterile and leakage-free whole plant two-compartment culture system allowing the exclusion of other microbes and the avoidance of nutrient diffusion between compartments. Low cost, easy to use, reproducible and autoclavable, these microcosms make feasible investigations of mycorrhizal transport, gene expression, metabolisms and nutrition that have hitherto been challenging or impossible. The uptake and transfer of 35SO42- by the fungal partner to the host plant was demonstrated in the microcosms. Based on our data, we conclude that sulfate is transferred by R. irregularis to host plants leaves within a day as has been reported for AM transformed roots (Allen and Shachar-Hill, 2009). Using this microcosm to study the role of N transfer by AMF to host plants indicated, that N transfer by AMF conferred growth and reproductive benefits to the host plant. 15N 15N isotopic labeling results showed increased levels of N metabolites with high percentage labeling in mycorrhizal plants after adding 138 15NO 3. Labeling time course data are consistent with the current N transfer model in which inorganic N is taken up by the fungal ERM and assimilated via the GS-GOGAT pathway raising the levels of glutamate and glutamine. Nitrogen is then incorporated into arginine via enzymes of the urea cycle and is translocated to the fungal IRM in colonized root tissues and broken down into ornithine and urea that is in turn is broken down to release ammonium which is exported from the fungus and imported by the host into the root cortical cells. The presence of high levels of amount of 15N 15N labeled metabolites in shoots indicated that a significant being transferred to the host plants. One of the major questions in N transfer in the AM symbiosis was the role, if any, of nitrate translocation. Based on 15N labeling experiments in microcosms (described in Chapter 2) and monoxenic culture system (mycorrhizal transformed roots) we conclude that arginine translocation is not the only pathway for N movement and transfer in AM symbiosis, and that nitrate is also directly transferred by the AM fungus to host plants. Gene expression of a putative fungal nitrite reductase indicated that a significant amount of nitrate was assimilated in the ERM but not IRM, which is consistent with the labeling and the revised N transfer model. Investigating the origin and fate of ornithine was another objective of this study. The presence of 15N labeled pyrroline-5-carboxylate as well as 15N-glutamate in the ERM suggests a role for these metabolites in making ornithine there since pyrroline-5carboxylate is a biosynthetic intermediate between glutamate and ornithine. This conclusion was also supported by gene expression analysis of OAT1 and P5CD which showed a 3-fold increase in gene expression 4h after adding nitrate to ERM. The presence of 15N glutamate, 15N pyrroline-5-carboxylate and low levels of 139 15N putrescine in 15N labeling experiments as well as of glutamate and putrescine 13C 4 13C , 5 pyrroline-5-carboxylate in the IRM after labeling the ERM with arginine 13C 6 13C 5 coincided with the upregulation of OAT1, OAT2 and P5CD. This suggests that ornithine is converted to glutamate and to a lesser extent to putrescine but that a significant proportion is translocated back to the ERM. Together, with the translocation and transfer of nitrate, this additional flux in the N transfer network provides increased flexibility and potential for improved efficiency. The expression of a putative GDH gene is slightly upregulated in the ERM and highly upregulated in IRM indicating that GDH may play a dissimilatory role in breaking down glutamate to release more ammonium to the host. Future research In this dissertation, I report experimental data on the metabolism and transfer of N from the fungus to the host plant in the AM symbiosis and its role plant growth. We extended our understanding of the N transfer network, including the origins and fate of key intermediates and shed light on its regulation. However, there are some unresolved questions that need investigation. The way that arginine is transferred from the ERM to the IRM is still not well understood. Passive diffusion is too slow for the long distances from ERM to IRM. It has been found that arginine is bound to polyphosphate in ectomycorrhizal symbiosis (Martin, 1985), and polyphosphate is believed to move in vesicles or tubular vacuoles (Dürr et al., 1979). Accordingly, where polyphosphate from and how arginine is bond with polyphosphate are the key points for illuminating the transfer of arginine. 140 R. irregularis does not down-regulate the genes for arginine synthesis in the ERM when intracellular arginine levels are high, as occurs for example in Saccharomyces cerevisiae suggesting either that arginine is sequestered, or perhaps that host demands for N are communicated to maintain N flow. How the genes involved in nitrogen metabolism and transport are regulated need further studies. Studying the enzymatic activity of the putative genes which were identified to date, will confirm their role in N transfer. Enzyme assays of glutamate dehydrogenase will clarify its role in N transfer as well as its specificity and regulation. On the other hand, the possibility of transforming R. irregularis shall lead to a better understanding of the role of genes in nutrients transfer through studying knockout mutants. Cross talk between C and N as well as N and P are not well understood and need further investigation. Furthermore, the incorporation of Medicago truncatula mutants impaired in N assimilation pathway genes would help reveal the extent to which AMF contribute to plant N metabolism and the possibility of plant-to-plant N transfer. 141 REFERENCES 142 REFERENCES Allen JW, Shachar-Hill Y. 2009. Sulfur Mycorrhiza. Plant Physiology 149: 549-560. transfer through an Arbuscular Martin F. 1985. 15N-NMR studies of nitrogen assimilation and amino acid biosynthesis in the ectomycorrhizal fungus Cenococcum graniforme. FEBS Letters. 182: 350– 354. Dürr M, Urech K, Boller T, Wiemken A, Schwencke J, Nagy M. 1979. Sequestration of arginine by polyphosphate in vacuoles of yeast (Saccharomyces cerevisiae). Archives of Microbiology 121: 169-175. 143