(flier: ..., .. : .Ehitmwru . 2:. 1%. E a» g . V . . ‘ A . . _ ‘ . ‘ . _ ”mafia“ .. .. . {i . u. 3% gm: . , . “Ewan? 1.‘ Vt). ‘ THESIS ICHK «lime W 3 l H M W ”Hill 0 6362 = d w LIBRARY Michigan State Unlverslty This is to certify that the dissertation entitled STRUCTURE, FUNCTION, AND REGULATION OF THE EARLY NODULIN GENE ENOD2 FROM LEGUME PLANTS presented by Rujin Chen has been accepted towards fulfillment of the requirements for Ph.D. degree in WU / Major or Date—LaLML MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 PLACE N RETURN BOX to remove this checkout from your record. TO AVOID FINES Mum on or betore date due. F——__———'——————————I'————-—1 ATE DUE DATE DUE DATE DUE ° l J H 1.7 % _L_= l—T—‘l—j MSU Ie An Affirmative Action/Equal Opponunlty lnetltuion STRUCTURE, FUNCTION, AND REGULATION OF THE EARLY NODULIN GENE EN ODZ FROM LEGUME PLANTS By Ruj in Chen A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTORATE OF PHILOSOPHY Department of Biochemistry 1996 ABSTRACT STRUCTURE, FUNCTION, AND REGULATION OF THE EARLY NODULIN GENE EN 0D2 FROM LEGUME PLANTS By Ruj in Chen The early nodulin SrEnodZ gene of the tropical legume Sesbam'a rostrata encodes a putative hydroxyproline-rich cell wall protein, and is expressed exclusively in the parenchyma cells of both stem- and root-nodules. In addition, the SrEnodZ gene can be specifically induced in roots by the plant hormone cytokinin, and this induction occurs at a post-transcriptional level. One focus of this dissertation research is to elucidate the molecular mechanism that controls the nodule parenchyma-specific expression of the SrEnodZ gene using both molecular genetic and biochemical approaches. By analyzing chimeric reporter gene expression in transgenic Lotus corniculatus plants, a detailed characterization of the 5’ and 3’ SrEnodZ flanking sequences has been carried out. These analyses indicate that the SrEnodZ 3’ untranslated region is both required and sufficient for gene expression in nodule parenchyma cells. Another finding is that the 5’ flanking sequence does not appear to contain any significant cis-acting elements for nodule parenchyma-specific expression. This observation is in contrast with the report that the 5’ upstream region of the soybean En0d2 gene directs nodule parenchyma-specific expression, indicating different mechanisms may be involved in regulation of the expression of these two genes. Using an in-vitro protein-RNA UV-cross linking approach, several nodule-specific proteins have been identified to interact with the RNA transcribed from the SrEnodZ 3 ’ untranalated region. Competition experiments using both non-specific and specific competitors suggest that the cis-acting elements responsible for nodule parenchyma-specific expression co-localize with specific protein binding sequences. Cytokinin up—regulation of the EnodZ gene expression is not a common phenomenon among legume plants. To facilitate molecular and genetic analyses of the cytokinin regulation of the EnodZ gene expression, the En0d2 homologue was isolated from the model legume, L. japonicus. Northern blot analyses indicate that the LjEnodZ gene is inducible in roots, but not by cytokinin. Submergence of the L. japonicus roots in water for several hours was sufficient to induce the LjEnodZ gene expression. Further analyses demonstrate that ethylene is responsible for the LjEnodZ gene expression in roots of L. japonicus. To Yihong, June, Rong and my parents iv ACKNOWLEDGMENTS I thank my thesis advisor, Dr. Frans J. de Bruijn, for advice, encouragement and financial support during my dissertation research. I thank my committee members, Drs. Zachary Burton; Frans J. de Bruijn; Pamela Green; Paul Kindel; and Lee McIntosh, for their helpful discussions and advice in the past five years. A special thank is due to Dr. Pamela Green for serving as my major professor. I thank all the members past and present in the de Bruijn’s laboratory for their help and input during my research. Particularly, I thank Krzysztof Szczyglowski for useful discussions and advice, David Silver for sharing information and collaborating on projects related to this research, and Debbie Ganoff for her excellent technical assistance. I also thank colleagues in laboratories of Drs. Pamela Green, Hans Kende, Lee Macintosh, and Natasha Raikhel, for providing expertise and access to their equipments when needed. I thank Drs. David Layzell and Stephen Hunt at Queen’s University of Canada for collaborating on analysis of nitrogenase activities with their non-invasive open-flow system. I thank Dr. David Emerson at the Center for Microbial Ecology of Michigan State University for collaborating on analysis of nodule oxygen with his oxygen microelectrodes. Lastly, I thank my wife Yihong, daughter June and mother-in-law Kong for their love and support. They made this possible and meaningful. TABLE OF CONTENTS LIST OF FIGURES .................................................................................. viii CHAPTER 1 INTRODUCTION ........................................................................................ 1 SECTION 1. Symbiotic Nitrogen Fixation and Nodule ‘ development ..................................................................... 2 SECTION 2. Involvement of mRNA 3’ Untranslated Region (3’ UTR) 1n Control of Gene Expression and Developmental Regulation .............................................. 18 CHAPTER 2 THE NODULE PARENCHYMA-SPECIFIC EXPRESSION OF THE SESBANIA ROS TRA TA EARLY NODULIN GENE SrENODZ IS NIEDIATED BY ITS 3’ UNTRANSLATED REGION (3 ’ UTR) ............. 46 Abstract ................................................................................................ 47 Introduction .......................................................................................... 48 Methods ................................................................................................ 52 Results .................................................................................................. 60 Discussion ............................................................................................ 90 List of References ................................................................................ 98 vi CHAPTER 3 IDENTIFICATION OF CIS-ELEMENTS WITHIN THE 3 ’ UNTRANSLATED REGION OF THE SrENOD2 GENE AND INTERACTING PROTEINS ......................................................... 108 Abstract ............................................................................................ 109 Introduction ...................................................................................... 1 10 Methods ............................................................................................ 1 1 1 Results .............................................................................................. 1 18 Discussion ........................................................................................ 133 List of References ............................................................................ 136 CHAPTER 4 THE LjENODZ GENE IS INDUCIBLE BY ETHYLENE IN ROOTS OF LOTUS JAPONICUS ...................................................... 138 Abstract ............................................................................................ 139 Introduction ...................................................................................... 140 Methods ............................................................................................ 142 Results .............................................................................................. 145 Discussion ........................................................................................ 168 List of References ............................................................................ 171 CHAPTER 5 FUNCTION OF THE ENOD2 PROTEIN IN NODULES ...................... 175 Abstract ............................................................................................ 176 Introduction ...................................................................................... 177 Methods ............................................................................................ 179 Results .............................................................................................. 183 Discussion ........................................................................................ 198 List of References ............................................................................ 203 CHAPTER 6 CONCLUSIONS AND FUTURE PERSPECTIVES .............................. 206 vii LIST OF FIGURES FIGURE l-l. Schematic representation of nodule developmental events and nodulin gene expression ..................................................... 17 FIGURE 2-1. In-situ localization of SrEnod2 gene expression .............. 61 FIGURE 2-2. Structure of gus fusion constructs and determination of GUS activity in transgenic L. corniculatus plants ........................... 65 FIGURE 2-3. Microscopic analysis of the expression of reporter gene constructs in transgenic L. comiculatus plants ............................ 68 FIGURE 2-4. Northern blot analysis of gus reporter gene expression in transgenic nodules harboring the reporter gene constructs 1-5 (Figure 2A) ................................................................... 73 FIGURE 2-5. Structure of gus fusion constructs and determination of GUS activity in transgenic L. comiculatus plants .......................... 77 FIGURE 2-6. Microscopic analysis of the expression of gas fusion constructs in transgenic L. corniculatus plants ..................................... 80 FIGURE 2-7. Structure of SrEnod2 5' deletion reporter gene constructs and GUS activity in transgenic L. corniculatus plants ....... 83 FIGURE 2-8. Structure of SrEnod2 5' deletion reporter gene constructs, enhancer fusions and GUS activity in transgenic L. corniculatus plants ........................................................................... 86 FIGURE 2-9. Structure of SrEnod2 3' deletion reporter gene constructs and determination of GUS activity in transgenic L. comiculatus plants ........................................................................... 89 viii FIGURE 3-1. S1 nuclease protection assay ............................................. 120 FIGURE 3-2. Sequence of the 3’ region of the SrEnod2 gene .............. 122 FIGURE 3-3. Structure of gas fusion constructs and GUS expression in transgenic L. corniculatus .............................................. 125 FIGURE 3-4. RNA-protein UV-cross linking and competition with non-specific competitors ............................................................... 129 FIGURE 3-5. RNA-protein UV-cross linking and competition with specific competitors ...................................................................... 132 FIGURE 4-1. Genomic organization of the Enod2 gene in L. japom'cus plants ................................................................................ 148 FIGURE 4-2. Nucleotide and the deduced amino acid sequences of a partial LjEnodZ cDNA clone ........................................ 150 FIGURE 4-3. The LjEnodZ gene is expressed early during nodule development .............................................................................. 152 FIGURE 4-4. LjEnodZ gene expression in roots of L. japom'cus plants treated with cytokinin ............................................ 155 FIGURE 4-5. LjEnodZ gene expression in roots of L. japonicus plant .................................................................................. 158 FIGURE 4-6. The LjEnodZ gene expression in roots of L. japonicus plants upon ethylene treatment ........................................ 161 FIGURE 4-7. LjEnodZ gene expression in roots of L. japom'cus plants treated with ethylene inhibitors ............................. 164 FIGURE 4-8. LjEnodZ gene expression in roots of L. japonicus plants treated with the protein synthesis inhibitor cycloheximide ....... 167 FIGURE 5-1. Antiscnse Enod2 constructs .............................................. 185 ix FIGURE 5-2. The Enod2 gene expression in L. comiculatus plants ............................................................................ 189 FIGURE 5-3. Nodule fresh weight of transgenic L. corniculatus plants harboring antisensc Enod2 constructs ........................................ 191 FIGURE 5-4. Oxygen microelectrode measurement of 02 concentrations across nodules of transgenic plants expressing antisensc Enod2 transcripts ............................................... 194 FIGURE 5-5. Measurement of nitrogenase activity ............................... 197 FIGURE 5-6. Electron allocation and oxygen limitation coefficients of nitrogenase ................................................................... 200 CHAPTER 1 INTRODUCTION SECTION I. Symbiotic Nitrogen Fixation and Nodule Development A. Symbiotic Nitrogen Fixation Symbiotic nitrogen fixation takes place in the central infected cells of nodules, new plant organs formed on roots of legume and several non- legume plants after infection by soil bacteria of the genera Rhizobium, Bradyrhizobium and Azorhizobium. The bacterial enzyme nitrogenase catalyzes the following reaction: N2 + 8H+ + 8c“ + 16Mg-ATP —) 2NH3 + H2 + l6Mg-ADP + 16Pi (1) In this reaction, atmospheric dinitrogen is reduced to ammonium, which is exported from the infected cells to adjacent non-infected cells and assimilated as a nitrogen-source for the plant host. Plants, in turn, supply the nitrogen-fixing bacteria with carbon sources to provide energy for the nitrogenase-catalyzed reaction. Nitrogenase consists of two components, the homodimeric Fe protein, encoded by the bacterial m‘fH gene, and the tetrameric molybdenum-iron (MoFe) protein, encoded by the nifD and nifK genes (Dixon et a1., 1977; Ruvkun, et a1., 1982). Since the MoFe protein is irreversibly denatured by oxygen (Appleby, 1984), nitrogen fixation requires low oxygen concentration in the infected cells of the nodule. On the other hand, a large amount of energy (ATP) required for the nitrogen- reducing reaction has to be provided via oxidative processes, since rhizobia are generally aerobic bacteria. Thus, there is a high demand for oxygen influx into nodules. Legume plants have developed different strategies to cope with this “oxygen paradox”. A low oxygen concentration in the nodule central nitrogen-fixing tissue is achieved by a combination of high metabolic activities of the bacteroids and the formation of an oxygen diffusion barrier in the periphery of the nodule (nodule parenchyma; Tjepkema and Yocum, 1974; Witty, 1986). In addition, and possibly most importantly, the oxygen carrier protein leghemoglobin provides oxygen buffering activity by transporting oxygen to the actively nitrogen-fixing and respiring bacteroids in the infected cells at a low intracellular free oxygen concentration (< 10 nM; Appleby, 1984). B. Nodule Development Nodule formation initiates with the exchange and recognition of signal molecules between the compatible symbiotic partners (Long et al., 1989; Fisher and Long, 1992; Figure 1-1). Flavonoids and isoflavonoids are signal molecules secreted by the roots of legume plants that serve as chemoattractants as well as inducers of bacterial nodulation (nod) genes (Peters et al., 1986; for a review, see Long, 1989). Expression of the nod genes, in turn, results in the synthesis of lipochitooligosaccharides, known as “Nod factors”, which can induce root hair deformation, cortical cell division and nodule formation at extremely low concentrations (10"8-10"11 M; Lerouge et al., 1990; Truchet et al., 1991; Ardourel et al., 1994; Heidstra et al., 1994; Yang et al., 1994). The structure of Nod factors was first determined in R. meliloti, the microsymbiont of alfalfa (NodRm-l; Lerouge et al., 1990). NodRm-l consists of a backbone of three to five [3-1,4-1inked N-acetylglucosamines bearing a C16 unsaturated fatty acid chain on the non-reducing sugar residue. The structure of Nod factors from many other rhizobia has subsequently been determined (for reviews, see Vijn et al., 1993; Denarie and Cullimore, 1993; Carlson et al., 1995). Now, it is clear that Nod factors generally have a N-acylated chitin oligosaccharide backbone and can have various substitutions on both the reducing and the non-reducing sugar residues. These substitutions likely render Nod factors specific for their host (Vijn et al., 1993; Denarie and Cullimore, 1993; Carlson et al., 1995). 1 Upon Rhizobium infection, root hairs undergo curling and deformation, and form characteristic structures called Shepherd crooks (Kijne, 1992). Bacteria are entrapped in these structures. A local hydrolysis of the plant cell wall and a subsequent plasma membrane invagination and deposition of new plant cell wall material result in the formation of a tubular structure, the so-called infection thread, by which the bacteria enter the plant (for reviews, see Bauer, 1981; Newcomb, 1981; Brewin, 1991; Kijne, 1992). As the infection threads form, root cortical cells are mitotically reactivated to form the nodule primordium. Infection threads grow toward this primordium and release bacteria into the cytoplasm of nodule primordium cells in a process called endocytosis (V erma, 1992). Concomitant with the infection process, a meristem is formed at the distal part of the primordium and the meristem produces cells that differentiate into the nodule tissue (Nap and Bisseling, 1990; Kijne, 1992). The site of first cortical cell divisions depends on the type of nodule formed upon rhizobial infection. In general, temperate legumes, including pea, alfalfa, clover, form indeterminate nodules, which have a persistent meristem at the apex. As a result of continuous cell divisions, new cells are added to the cortical and central tissues of emerging nodules. Consequently, all tissues of these nodules are of a graded age from the apical meristem to the root attachment point. The first cell divisions in roots of plants forming indeterminate nodules occur in the inner cortical cells (Nap and Bisseling, 1990). The infection threads must traverse the outer cortex before they reach these cells. Before infection thread penetration, the outer cortical cells undergo morphological changes. The nuclei move to the center of the cells, and the microtubules and the cytoplasm rearrange to form radially oriented structures, the cytoplasmic bridges, or pre-infection threads (van Brussel et a1; 1992). The infection threads traverse the outer cortical cells through these radially aligned cytoplasmic bridges. In contrast, tropical legumes, such as soybean, Lotus, and Sesbania, form determinate nodules, whose meristems cease to divide shortly after rhizobial infection (Newcomb, 1981). As a result, development of determinate nodules largely depends on cell elongation and expansion. Within a determinate nodule, cells are at approximately the same developmental stage. In plants forming determinate nodules, first cell divisions occur in the root outer cortex. Subsequently, cell divisions also take place in the pericycle and inner cortex, several cell layers separated from the initial cell divisions. Eventually, the two meristematically active regions merge and give rise to the nodule tissues (Kijne, 1992). C. Gene Expression During Nodule Development During nodule development, a number of plant genes are induced or their expression is differentially enhanced. They are termed nodulin genes (Legocki and Verma, 1980; van Kammen, 1984; Verma and Delauney, 1988). Genes expressed during early stages of nodule development, well before the onset of symbiotic nitrogen fixation, are collectively called early nodulin genes (Enods). Those expressed around the onset of nitrogen fixation are termed late nodulin genes. Although the function of many nodulin genes identified to date remains to be determined, the sequential expression of particular nodulin genes marks the different phases of nodule development (Scheres et al., 1990b; for reviews, see Nap and Bisseling, 1990; Sanchez et al., 1991; Mylona et al., 1995). Additionally, the spatial expression of nodulin genes in nodules clearly suggests their involvement in particular morphological and biochemical aspects of nodule development. Nodulins involved in initial root hair deformation upon rhizobial infection, include RH-42 and RH—44, which were identified by in vitro translation of mRNAs isolated from root hairs (Gloudemans et al., 1989); and a lipid transfer-like protein (LTP), which was isolated from a cDNA library constructed from mRNAs of root hairs infected with rhizobia (Krause et al., 1994). The RH—42 gene is not expressed in root hairs of uninoculated pea plants, but induced in root hairs after infection with R. leguminosarum bv. viciae (Gloudemans et al., 1989). It is possible that the RH—42 is involved in root hair curing. The expression of the pea gene RH- 44 is markedly stimulated in pea root hairs upon rhizobial infection, although this gene is also expressed at a low level in root hairs of uninoculated plants (Gloudemans et al., 1989). The mRNA encoding the LTP is expressed in all tissues, except nodules. However, the transcript level is induced in root hairs 24 hr after application of Nod factors, different plant hormones, or rhizobia (Krause et al., 1994). Two early nodulin genes of pea (Enod5 and Enod12) have been identified that are involved in the infection process (Scheres et al., 1990a). The expression of the Enod12 gene can be detected in root hairs, root cortical cells and nodule cells containing growing infection threads (Scheres et al., 1990a). Interestingly, this gene is expressed not only in cells containing growing infection threads, but also in several cell layers located in front of the infection thread tips, where the cells are preparing for infection thread penetration (Scheres et al., 1990a). The spatial distribution of the Enod5 mRNA in infected pea roots is strikingly different from that of the Enod12 transcript. The Enod5 gene is expressed only in cells containing the growing infection thread (Scheres et al., 1990b). In indeterminate pea nodules, the Enod12 transcript is also detectable in cells adjacent to the distal meristem in the invasion zone, while the En0d5 mRNA reaches a maximum level in the early symbiotic zone (Scheres et al., 1990b). Several nodulin genes are expressed during mitotic reactivation of root cortical cells by rhizobia. They include Enod12 (Scheres et al., 1990a), Gm93 (Kouchi and Hata, 1993), Enod40 (Kouchi and Hata, 1993; Yang et al., 1993; Asad et al., 1994; Crespi et al., 1994; Matvienko et al., 1994), and MtPRP4 (Wilson et al., 1994). These genes are expressed in all cells of the developing nodule primordia. En0d40 gene is particularly interesting, because it is also expressed at an early stage preceding the first cortical cell division in the region of the pericycle opposite to the protoxylem poles (Kouchi and Hata, 1993; Yang et al., 1993; Asad et al., 1994; Matvienko et al., 1994).The Enod40 gene has been shown to be inducible by Nod factors (V ijn et al., 1993; Yang et al., 1993). Expression of antisense Enod40 gene in alfalfa arrests callus growth, while over-expression of the Enod40 gene in alfalfa causes the formation of teratomas, a callus containing developing shoots and short roots (Crespi et al., 1994). Since the Enod40 genes have no significant open reading frame, and in vitro translation of the En0d40 transcripts does not yield any protein product, several groups of investigators have proposed that the Enod40 genes play a role in plant development by acting as RNA (riboregulators; Asad et al., 1994; Crespi et al., 1994; Matvienko et al., 1994). Recently, van de Sande et a1. (1996) have reported evidence to argue that an oligopeptide of about 10 amino acids encoded by the Enod40 transcripts changes the plant cell response to auxin in tobacco protoplasts. An Enod40 homologue was isolated from tobacco, and when expressed in tobacco protoplasts, had an activity identical to the legume Enod40, indicating that the peptides encoded by the Enod40 genes may act as plant growth regulators (Crespi et al., 1996). Several nodulin genes have also been identified during subsequent nodule organogenesis. One of these genes, the Enod2 gene, was first isolated from soybean and pea (Franssen et al., 1987; van de Wiel et al., 1990a). The Enod2 gene encodes a protein with a putative N-terminal signal peptide followed by a sequence containing repeating proline-rich 10 pentapeptide motifs (Franssen et al., 1987). Enod2 homologues have been isolated fiom several other legumes, such as alfalfa (Dickstein et al., 1988), vetch (Moerman etal., 1987), Sesbania (Strittmatter etal., 1989; Dehio and de Bruijn, 1992), common bean (Sanchez et al., 1988), lupin (Szczyglowski and Legocki, 1992), and Lotus japom'cus (R. Chen and R]. de Bruijn, unpublished data). The Enod2 mRNA is present in the inner cortex (nodule parenchyma) of determinate and indeterminate nodules (van de Wiel et al., 1990a), and in similar tissues in empty nodules induced by certain rhizobium strains (van de Wiel et al., 1990b). The exact function of the Enod2 protein has not yet been elucidated. However, since the nodule parenchyma appears to be a physical barrier limiting oxygen diffusion into the infected cells of nodules and the Enod2 transcript (protein) is localized in this tissue, it has been postulated that this putative cell wall protein may contribute to the morphogenesis of the nodule parenchyma and therefore the generation of the oxygen diffusion barrier (Nap and Bisseling, 1990; van de Wiel et al., 1990a; Govers et al., 1990). One of the objectives of this dissertation research was to test this hypothesis by attempting to down- regulate Enod2 gene expression using an antisense approach (see Chapter 5). Following Enod2 gene expression, the late nodulin genes are activated and nitrogen fixation commences. Late nodulins include enzymes involved in ammonia assimilation, such as glutamine synthetase and glutamate 11 synthase (GS, GOGAT, respectively; for reviews, see Atkins, 1987; Miflin and Lea, 1980); carbon metabolism, such as phosphoenolpyruvate carboxylase and sucrose synthase (PEP carboxylase, SS, respectively; for a review, see Atkins, 1987); and amide and ureide biogenesis, such as aspartate aminotransferase and uricase (AAT; uricase II; for a review, see Schubert, 1986). They also include proteins present in the peribacteroid membrane, such as Nodulin- 26 fiom soybean (Fortin et al., 1987; Miao and Verma, 1993; Weaver et al., 1994), leghemoglobins (Lbs; Appleby, 1984), and a putative plant transcription factor, NMH7 (Heard and Dunn, 1995). N- 26 is a nodule-specific ion transporter (Weaver et al., 1994). Leghemoglobins (Lbs) function as oxygen carriers, facilitating Oz diffusion to the nitrogen-fixing bacteroids (Appleby, 1984), and represent the most abundant proteins within nodules. Leghemoglobins are encoded by a multigene family, each member of which is induced at a slightly different time during nodule formation, indicating developmental control of Lb gene expression (Jensen et al., 1988). Different Lbs seem to have distinct affinities for oxygen, that may be important for nodule nitrogen fixing activity in subtle ways (Appleby, 1984; de Bruijn and Schell, 1992). Late nodulin genes are the most well characterized nodule-specific genes, in terms of regulation of gene expression. Nodule-specific expression of the Lb genes, for example, has been studied in detail using reporter gene expression in transgenic plants (de Bruijn et al., 1990; Szabados et al., 12 1990). Using deletions and specific mutations, an organ-specific cis-acting (OSE) element has been identified in the promoter region of a soybean leghemoglobin,lbc3 gene (Stougaard et al., 1987; Ramlov et al., 1993). Similarly, a nodule-infected-cell—specific (NICE) element, which contains three nodule conserved cis-elements (AAAGAT; TTGTCTCTT; and CACCCT), has been identified in the promoter region of a Sesbania rostrata leghemoglobin gene (Srglb3; Szabados et al., 1990; Szczyglowski et al., 1994). D. Hormones as Secondary Messengers As mentioned above, Nod factors stimulate multiple responses in legume plants, including root hair curling and deformation, cortical cell division, induction of nodulin gene expression and finally nodule formation. The molecular basis of Nod factor action is not precisely known. Ehrhardt et a1. (1992) have shown that root hair membranes depolarize rapidly after addition of Nod factors. Allen et a1. (1991) have found that cytoskeletal rearrangements, and changes in vacuole shape occur within root hair cells soon after the addition of Nod factors. Using calcium-sensitive reporter dyes, Ehrhardt et al., (1996) have shown that localized periodic spikes in cytoplasmic calcium levels occur following the application of Nod factors. 13 In plants forming indeterminate nodules, cortical cell divisions occur in the inner cortex, and Yang et a1. (1994) have shown that only narrow rows of cortical cells located opposite protoxylem poles, are activated and express a histone H4 gene. At the time of H4 gene induction, the infection thread tips, the sites where Nod factors are released, are still in the epidermis. This indicates that Nod factors act at distance. Nod factors may directly activate the cortical cell division or indirectly affect them through endogenous signal molecules. Plant hormones have been implicated in nodule development since Thimann (193 6) first hypothesized that the plant growth hormone auxin was involved in pea root nodulation (for a recent review, see Hirsch, 1992). Subsequently, the plant hormone cytokinin was found to induce cortical cell divisions that resembled the initial response of roots to rhizobial infection (Arora et al., 1959; Libbenga and Bogers, 1974). Libbenga et a1. (1973) determined that inner cortical cells opposite the xylem poles were stimulated to divide in devascularized explants after treatment with auxin and cytokinin. Furthermore, they found that an ethanol extract from the stele could replace cytokinin in promoting cell division. A substance responsible for this activity, the so-called stele factor (uridine), has since been isolated and is thought to be released from the protoxylem poles to promote susceptibility to the cortical cells located opposite the protoxylem poles (Smit et al., 1993; 1995). 14 More recently, both cytokinin (Cooper and Long, 1994) and compounds that block polar auxin transport (Hirsch, et al., 1989) have been shown to induce the formation of nodule-like structures in which several early nodulin genes, including Enod2, are expressed. Additionally, the Nod factor-inducible ENOD40 gene has been shown to have plant hormone-like activity when expressed in alfalfa or tobacco (Crespi et al., 1994; van de Sande, 1996). It is possible that ENOD40 expression in the pericycle of legume roots causes a change in the cytokinin/auxin ratio in the cortical cells that triggers the initiation of cell division (Mylona et al., 1995). Support for the notion that plant hormones play a role in nodule development also comes from studies which demonstrate that cytokinin can specifically induce early nodulin Enod2 gene expression in the roots of S. rostrata (SrEnod2; Dehio and de Bruijn, 1992; Silver et al., 1996), and in roots of alfalfa (Hirsch et al., 1993). However, the mechanism by which cytokinin regulates early nodulin gene expression and the exact role of early nodulins in nodule development remain to be established. One objective of this dissertation research was to understand how the SrEnod2 gene expression is controlled developmentally and cell specifically, and to identify cis-elements responsible for this type of regulation. Chapter 2 presents the results of this research. The data described in this chapter will be directly compared with those obtained in a separate study in our laboratory on the cytokinin regulation of Enod2 gene 15 expression. One of the future goals will be to dissect the signal transduction pathways leading to the developmental, cell-specific, and cytokinin regulation of the Enod2 gene expression by isolating trans-acting factors that interact with cis-elements identified here. Towards this goal, Chapter 3 describes some of the initial effort to identify putative protein factors interacting with these cis-elements using in vitro RNA-protein UV-cross linking strategies. To further our understanding of plant hormone regulation of Enod2 gene expression, we also characterized the hormone response of the Enod2 gene from L. japonicus (see Chapter 4). Surprisingly, we found that ethylene, instead of cytokinin, induces the LjEnodZ gene expression, indicating that Enod2 genes from different legume plants might be regulated differentially. Lastly, we employed a genetic approach in an attempt to down-regulate Enod2 gene expression for the purpose of understanding the function of the Enod2 protein during nodule development (see Chapter 5). Since the results presented in Chapter 2 strongly suggest that the 3’ untranslated region (3’ UTR) of the SrEnod2 gene mediates the developmental and cell-specific expression of this gene, in the following section, I will review the current literature on what is known about 3’ UTR’s in control of gene expression and developmental regulation. 16 LEGEND FOR FIGURE 1-1. Schematic representation of nodule developmental events and nodulin gene expression. The developmental stages are shown on the lefi and the corresponding nodule development events are represented in the center. On the right, expression of some early and late nodulin genes is listed sequentially. Nodule types are represented in the middle (INDETERMINATE: M= meristem; IZ= infection; ES= early symbiotic; LS= late symbiotic; and SZ= senescing zones; VB= vascular bundle; NP= nodule parenchyma. DETERMINATE: IC= infected; NC= uninfected cells; VB= vascular bundle; NP= nodule parenchyma.) Major metabolic processes (bottom) are schematically represented as linked pathways of carbon, nitrogen, and oxygen metabolism. (Modified from Sandrez et al., 1991). Developmental Stages Nodule Development Expression of Nodulin Genes Preinfection it Infection and Nodule Formation Nodule Function and Maintainance i Q] Root Exudatcs ‘_ Nod Factors T 6:? Initial Induction Early Nodulins l Late Nodulins Scncsccncc Enhanced Exprcssion RH-44 Rh-42 Enod12 Enod5 Enod40 Gm93 MtPRP4 Enod2 N-26 Lbs NMH7 GS Uricase PEPcase GOGAT 18 SECTION H. Involvement of mRNA 3' Untranslated Regions (3' UTRs) in Control of Gene Expression and Developmental Regulation Eukaryotic mRNAs, once transcribed in the nucleus, are processed, transported into the cytoplasm, bound by ribosomes, and translated. The accumulation of a particular mRNA is determined by the cumulative effects of its rates of synthesis and decay. Thus, post-transcriptional events such as mRNA turnover contribute a great deal to the regulation of gene expression. The time required for a mRNA to reach a new steady-state level, following a change in its rate of synthesis, has been shown to be directly proportional to its half-life (Hargrove and Schmidt, 1989). Therefore, the induction and repression of genes encoding unstable mRNAs are more rapid than those of genes encoding stable mRNAs. It has also been noted that protein and mRNA turnover are often coordinately regulated (Hargrove and Schmidt, 1989). This allows rapid regulation of steady-state protein levels after transcriptional activation of a particular gene. Control of mRNA turnover has been extensively studied in mammalian cells (for recent reviews, see Peltz et al., 1991; Ross, 1995, 1996; Decker and Parker, 1994), yeast (for a review, see Caponigro and Parker, 1996), and in plants (Green, 1993; Sullivan and Green, 1993; Abler and Green, 1996). Since in our studies, we found that the 3’ UTR of the SrEnod2 gene is responsible for the nodule l9 parenchyma-specific expression of this gene, this section will focus on 3' UTRs and their involvement in the processes that control gene expression. Of course, 3' UTRs are not the only sequences that are involved in the control of gene expression, and therefore other sequences that are important will also be briefly mentioned, wherever appropriate. A. Instability Elements Localized in 3' UTRs The half-lives of specific eukaryotic mRNAs vary over a large range. The basis for this difference is determined, to a large extent, by specific sequences within each mRNA. A number of elements that affect mRNA decay rates have been identified. One general class of elements consists of sequences that destabilize mRNAs. These instability elements have been identified in yeast, mammalian cells and plants. Instability elements have also been found in the 3' UTRs of many unstable mRNAs. Examples include the yeast WAZ (Muhlrad and Parker, 1992) mRNA; mammalian c- fos (Chen et al., 1994), c-myc (Jones and Cole, 1987), and granulocyte- macrophage colony-stimulating factor (GM-CSF; Shaw and Kamen, 1986) transcripts; as well as the plant SAUR (Small-Auxin-Up-RNA) transcript (Gil and Green, 1996). The WAZ gene encodes the mating pheromone, a-factor, which needs to be rapidly removed from the cell after mating-type switching a-Ierrick et 20 al., 1990). Extensive molecular genetic analyses indicate that the 3' UTR contains sequences that specify rapid mRNA decay. Moreover, these sequences within the 3’ UTR appear to be redundant, since various deletions and point mutations do not affect decay rate dramatically whereas double mutations or large deletions have a pronounced effect on slowing transcript decay (Muhlrad et al., 1992). Two repeating motifs, Y6-8CAU (where Y=C or U) and U(A/C)AUUUAUU, have been found to be involved in stimulating rapid mRNA deadenylation and subsequent decay. Similar instability elements have been identified in the 3’ UTRs of the c-fos, c-myc and GM—CSF transcripts. The mechanism by which these instability elements increase the rate of deadenylation and subsequent mRNA degradation is not known. However, the fact that these instability elements within 3' UTRs are often redundant and are located close to the poly(A) tail suggests that these sequences may provide multiple binding sites for factors that stimulate mRNA deadenylation, which is the first rate- limiting step in mRNA degradation of many eukaryotic mRNAs. Instability elements have also been identified in the coding regions of the c—fos, c-myc and many other unstable transcripts (Ross, 1995, 1996; Decker and Parker, 1994). Other types of instability determinants have also been identified. For examples, the stem-loop structure at the 3’ end of a cell cycle-regulated histone mRNA has been found to be necessary and sufficient for rapid 21 mRNA decay at the end of S phase (Pandey and Marzluff, 1987). The ion- responsive-elements (IRES) at the 3’ UTR of the human transferrin receptor (hTR) gene have been found to be responsible for the iron-dependent feedback regulation of the transcript (Owen and Kfihn, 1987). A long range stem-loop structure at the 3’ UTR of mammalian insulin-like grth factor II (IGF-II) has been identified to be an endonuclease cleavage site and a mRNA stability determinant (Nielsen and Christiansen, 1992). In plant cells, several different instability elements have been shown to affect half-life of transcripts. The first one is the highly conserved DST element, derived from the 3' UTRs of the unstable SAUR transcripts of soybean and other plants (McClure et al., 1989; Newman et al., 1993; for a review, see Green, 1993). It has been shown that a dimer of the DST sequences, when inserted into the 3' UTRs of two different reporter genes is sufficient to promote rapid degradation of the reporter transcripts in stably transformed tobacco cells, as well as in transgenic tobacco plants (Newman etaL,l993) The second instability element found to destabilize plant gene transcripts is a synthetic sequence consisting of multiple copies of a AUUUA element (Ohme-Takagi et al., 1993), that is a repeating sequence found in the 3’ UTRs of many unstable mammalian transcripts, including o- fos and GM-CSF (Peltz et al., 1991). Insertion of this sequence into the 3' 22 UTRs of reporter genes stimulates rapid degradation of the reporter transcripts in stably transformed tobacco cells (Ohme-Takagi et al., 1993). B. Stability Elements Localized in 3’ UTRs It is known that mRNA instability elements are responsible for rapid decay of unstable transcripts. However, it is unclear whether stable transcripts contain stability determinants or are stable by default. Recently, the 3 ’ UTR of the human a2-globin gene has been found to contain erythroid cell- specific mRNA stability elements (Weiss and Liebhaber, 1995), indicating that at least some stable transcripts contain stability elements. A base-pair scanning mutagenesis in the 3’ UTR of the aZ-globin gene has revealed that three C-rich regions are required for the longevity of the aZ-globin gene in erythroid cells. Mutations that allow the translating ribosome to read through into the aZ-globin 3’ untranslated region target the mutant mRNAs for accelerated turnover in erythroid cells, but not in non-erythroid cells, indicating a cell type-specific half-life regulation of this transcript (Weiss and Liebhaber, 1994). 23 C. Messenger RNA Intracellular Localization Signals in 3’ UTRs Although very little is known about intracellular localization of mRNAs in plants, a lot of information is available for mammalian systems. It has been proposed that intracellular localization of specific maternal mRNAs is a general mechanism for protein targeting that probably occurs in all polarized cell types (see St Johnston, 1995). Localization of mRNAs allows specific proteins to be synthesized in the sub-cellular location where they are required and prevents their expression in areas where they are not. In Drosophila melanogaster oocytes, several mRNAs have been found to be localized to different locations, and these transcripts play a variety of roles in the specification of the embryonic body plan (see Ding and Lipshitz, 1993). For examples, the oskar mRNA localizes to the posterior pole and plays a role in defining anteroposterior polarity by determining where the pole plasm forms (Ephrussi and Lehmann, 1992). Similarly, gurken mRNA localizes to the dorsal anterior corner of the oocyte and defines the dorsoventral axis of the embryo, by being translated to produce a transforming growth factor on (TGFoc)-like protein (Neuman-Silberberg and Schiipbach, 1993). On the other hand, bicoid and nanos mRNAs are anchored at the anterior and posterior poles, respectively, and act as localized sources of diffusible proteins that spread toward the center of the embryo. The resulting opposite protein gradients specify the pattern and 24 polarity of the embryo (St Johnston and Nusslein-Volhard, 1992). In Xenopus laevis oocytes, like in its Drosophila counterpart, maternal messengers are also localized to different sites (Kloc et al., 1993). A common feature of all the localized mRNAs identified so far is that the cis-acting sequences required for localization reside in the 3' UTRs of transcripts and are typically long. One of the best characterized localization signal derives from the bicoid mRNA. It is 625 nt long and located within the bicoid 3' UTR, of which a 53 base sequence, called the BLEl element, is required and sufficient to direct localization of the bicoid mRNA to the anterior of the oocyte (Macdonald et al., 1993). A 400 nt region overlapping the BLEl elements is required for the anchoring of the mRNA in the anterior cytoplasm, once released there. The latter step requires an association of the 400 nt sequence with the staufen protein to prevent diffusion of the RNA (Ferrandon et al., 1994). Therefore, the bicoid localization signal contains at least two distinct, but overlapping, sequence elements that mediate different steps in its localization. Interestingly, the same staufen protein is also required for the translocation of oskar mRNA from anterior to posterior of the oocyte. The localization of the staufen protein switches from the posterior pole to the anterior pole soon after fertilization, thus facilitating translocation of two differently localized transcripts (Ferrandon et al., 1994). The Staufen protein contains five copies of a dsRNA-binding motif and binds to dsRNA 25 in vitro (St Johnston et al., 1992). Since the binding is not sequence-specific in vitro, it is possible that staufen recognizes the secondary structure of bicoid and oskar RNAs, rather than their sequence (St Johnston et al., 1992) Several other mechanisms responsible for RNA localization within a cell have been identified. One of them uses spatial control of mRNA stability. This is a simple but inefficient way to localize RNA. The localization of hsp83 RNA to the posterior pole of the Drosophila embryo involves a localized protection from mRNA degradation in the posterior pole plasm (Ding et al., 1993). On the other hand, the localization of hunchback RNA in the anterior involves a localized destabilization in the posterior due to a translational repression of the hunchbach RNA by nanos (Curtis et al., 1995). D. Messenger RNA 3’-End Formation In eukaryotic cells, mRNA 3’-end formation is an important process in the regulation of gene expression. This process includes cleavage of transcripts and addition of a tract of poly(A). In general, polyadenylated RNAs are more stable than their unpolyadenylated counterparts. The mechanism that governs the 3’-end formation of mammalian mRNAs has been well characterized (Proudfoot, 1991; Wahle et al., 1992). The hexanucleotide 26 AAUAAA, the highly conserved poly(A) signal, is located 10—30 nt upstream of the poly(A) site. In addition to the poly(A) signal, a less conserved downstream GU-rich sequence is also required for efficient 3 ’- end formation. At least four protein complexes are involved in the 3’-end processing. The cleavage polyadenylation-specific factor (CPSF) recognizes the poly(A) signal, while the cleavage stimulation factor (CstF) recognizes the downstream GU-rich sequence. The cooperative binding of these two factors to the precursor mRNA brings in cleavage factors (Cfs) and poly(A) polymerase (PAP) for the cleavage and polyadenylation of the precursor mRNA, respectively. The mechanism that controls the 3’-end formation of transcripts encoded by plant nuclear genes has started to be elucidated (for reviews, see Hunt, 1994; Wu et al., 1995). It is apparent that the poly(A) signal is not conserved near poly(A) sites in many plant nuclear genes. A systematic search for the AAUAAA signal near poly(A) sites indicates that about 1/3 of genes surveyed contain the AAUAAA sequence, about 1/2 of genes contain 4-5 of 6 base pair matches for this sequence and about 15% of genes do not have any AAUAAA-like sequence (Joshi, 1987). Additionally, the 3’ untranslated regions (3’UTRs) of most plant nuclear genes tend to contain multiple AATAAA-like and GT-rich sequences. Unlike their mammalian counterparts that require downstream GU-rich sequence for efficient 3’-end processing, most plant mRNAs require less 27 conserved, upstream sequences (Hunt, 1994). Removal of these sequences results in a drastic decrease in polyadenylation. In some plant mRNAs, downstream regions are required for efficient polyadenylation. These include rch and ocs mRNAs (Hunt and MacDonald, 1989; Macdonald et al., 1991). Using a functional analysis, An et a1. (1989), have demonstrated that the 3' flanking region of the potato wound-inducible proteinase inhibitor 11 gene that contains the poly(A) signal and a downstream GT—rich sequence is essential for wound-inducible reporter gene expression. E. Polyadenylation and Translational Control of Gene Expression Controlling gene expression by regulating translation allows a cell to respond to environmental cues more quickly than de novo transcription permits. Translation can be regulated by modulating several factors, including translation initiation and the length of the poly(A) tail. Since ribosomal binding is most sensitive to secondary structures in the 5' UTR, the sequence of the 5' UTR determines the intrinsic rate of translation initiation of an mRNA. However, in vivo and in vitro evidence support the idea that the poly(A) tail stimulates the rate of translational initiation of a mRNA. Recently a yeast poly(A) binding protein (Pablp)- poly(A) complex has been found to recruit the 40S ribosomal subunit to the mRNA, indicating that the 5' and 3' ends of a mRNA physically interact (Tarun and 28 Sachs, 1995). Recently, poly(A) tails of most mammalian mRNAs have been found to be modulated in cytoplasm by cytoplasmic polyadenylation machinery (Simon et al., 1992). Cytoplasmic polyadenylation requires two sequence elements within 3' UTRs: a U-rich cytoplasmic polyadenylation element (CPE) and the nuclear polyadenylation sequence AAUAAA (Wormington, 1993). Precise spacing of a CPE element relative to the AAUAAA sequence has been found to be important for developmental control of polyadenylation (Simon et al., 1992). Other specific translational regulatory sequences distant from the CPEs have been identified in the 3' UTRs of a number of developmentally regulatory genes. One of them is the 10 random repeats of a pyrimidine-rich motif (R sequence) located in the 3' UTR of the Lox gene, which encodes the 15-lipoxygenase enzyme required for the breakdown of internal membranes in mature reticulocytes (Ostareck- Lederer et al., 1994). A 48 KDa protein has been identified to specifically bind to these repeats and this binding is necessary and sufficient for translational repression of a heterologous mRNA containing the R sequences in an in vitro assay (Ostareck-Lederer et al., 1994). F. A Regulatory Role for 3' UTRs in Growth and Differentiation Using a genetic complementation approach, three genes, troponin I, .tropomyosin, and a-cardiac actin, have been identified that activate muscle- 29 specific promoters in a differentiation-deficient mutant (NMU 2) (Rastinej ad and Blau, 1993). This activity was mapped to the 3' UTRs of these genes and was found not to be limited to myogenic cells, since the 3' UTRs also promoted differentiation of wild type muscle cells, and suppressed proliferation of fibroblasts (Rastinejad and Blau, 1993). The mechanism by which these 3' UTRs promote differentiation and inhibit growth is not known. However, it is speculated that 3' UTRs of certain differentiation- specific RNAs are trans-acting regulators that modulate cell division and therefore differentiation. The phenomenon that 3' UTRs modulate growth and differentiation has been further exemplified by the observation that the 3' UTR of tropomyosin can efficiently suppress mouse rhabdomyosarcomas, induced by transplanting the NMU2 cell line into muscle cells of adult mouse hindlimbs (Rastinejad et al., 1993). G. SUMMARY Symbiotic nitrogen fixation takes place in the infected cells of legume nodules. This process requires a large input of energy and limitation of the oxygen cOncentration, presenting the oxygen paradox. Legume nodules have developed strategies, such as formation of diffusion barriers, production of a large amount of the oxygen carrier (leghemoglobin), and active respiration of bacteroids. A further detailed understanding of how 30 these various processes are coordinated during nitrogen fixation will be necessary before one could even contemplate transferring the symbiotic nitrogen-fixation system to non-nodulated agronomically important crop plants, such as rice (dc Bruijn et al., 1995). Nodule organogenesis is triggered by Nod factors. Nod factors are hypothesized as plant growth regulators, since they cause root hair deformation and stimulate cell divisions. Nodule organogenesis provides a model to study signal molecules, their perception and action. Nodules provide a rich source of specifically induced plant and bacterial genes to study gene regulation. Some of these genes are regulated by Nod factors and plant hormones. They may prove to serve as excellent tools to study the corresponding signal transduction pathways and this, in turn, will help to understand how nodule organogenesis utilizes and integrates these various pathways.- Regulation of gene expression occurs at different levels. Post- transcriptional regulation is clearly very important for modulating gene expression. It allows a cell to rapidly adjust gene expression to accommodate developmental and environmental cues. This type of regulation occurs both in the nucleus and cytoplasm. Steps involved include 5' capping and 3' polyadenylation in the nucleus, transport to cytoplasm, poly(A) tail shortening and lengthening in the cytoplasm, degradation and stabilization, intracellular localization, and translation. Each step involves 31 the recognition of cis-elements and their corresponding trans-acting factors. Some of these steps are connected and coordinately regulated. 3' untranslated regions of mRNAs have been implicated in various processes of gene regulation and 3' UTRs of some structural genes have been found to exert profound effects on grth and differentiation by acting as trans- acting regulators (riboregulators). In this dissertation research (chapter 2), I have identified the 3’ UTR of the early nodulin SrEnod2 gene as the mediator for the tissue-specific expression. To our knowledge, this type of regulation is the first example among plant nuclear genes clearly demonstrating that the 3 ’ UTR regulates tissue-specific expression. 32 LIST OF REFERENCES Abler, M.L., and Green, PJ. (1996). Control of mRNA stability in higher plants. (in press) Allen, T., Raja, S., and Dunn, K., (1991). Cells expressing Enod2 show differential spatial organization during the development of alfalfa root nodules. Mol. Plant-Microbe Interact. 2, 139-146. An, G., Mitra, A., Choi, H.K., Costa, M.A., An, K., Thornburg, R.W., and Ryan, CA. (1989). Functional analysis of the 3’ control region of the potato wound-inducible proteinase inhibitor II gene. Plant Cell 1, 1 1 5-122. Appleby, CA. (1984). Leghemoglobin and Rhizobium respiration. Annu. Rev. Plant Physiol. 35, 443-478. Ardourel, M., Demont, N., Debelle, F., Maillet, F., dc Billy, F., Prome, J.C., Denarie, J., and Truchet, G. (1994). Rhizobium meliloti lipooligosaccharide nodulation factors: Different structural requirements for bacteria entry into target root hair cells and induction of plant symbiotic developmental responses. Plant Cell 6, 1357-1374. Arora, N., Skoog, F., and Allen, O.N. (1959). Kinetin-induced pseudonodules on tobacco roots. Am J. Bot. 46, 610-613. Asad, S., Fang, Y., Wycofl', KL, and Hirsch, A.M. (1994). Isolation and characterization of cDNA and genomic clones of MsENOD40: Transcripts are detected in meristematic cells of alfalfa. Protoplasma 183, 10-23. Atkins, GA. (1987). Metabolism and translocation of fixed nitrogen in the nodulated legume. Plant Soil 100, 157- 1 69. Bauer, W.D. (1981). Infection of legumes by rhizobia. Annu. Rev. Plant Physiol. 32, 407-449. 33 Brewin, NJ. (1991). Development of the legume root nodule. Annu. Rev. Cell Biol. 7, 191-226. Caponigro, G., and Parker, R. (1996). Mechanisms and control of mRNA turnover in Saccharomyces cerevisiae. Microbiol. Rev. 60, 233-249. Carlson, R.W., Price, N.P.J., and Stacey, G. (1995). The biosynthesis of rhizobial lipo-oligosaccharide nodulation signal molecules. Mol. Plant- Microbe Interact. 7, 684-695. Chen, C.-Y., Chen, T.-M., and Shyu, A.-B. (1994). Interplay of two functionally and structurally distinct domains of the c-fos AU-rich element specifies its mRNA-destabilizing function. Mol. Cell. Biol. 14, 4 1 6-426. Cooper, J.B., and Long, SR. (1994). Morphogenetic rescue of Rhizobium meliloti nodulation mutants by trans-zeatin secretion. Plant Cell 6, 215- 225. Crespi, M.D., Jurkevitch, E., Poiret, M., d’Aubenton-Carafa, Y., Petrovics, G., Kondorosi, E., and Kondorosi, A. (1994). enod40, a gene expressed during nodule organogenesis, codes for a non-translatable RNA involved in plant growth. EMBO J. 13, 5099-5112. Curtis, D., Apfeld, J., and Lehmann, R. (1995). Nanos is an evolutionarily conserved organizer of anterior-posterior polarity. Development 121, 1899-1910. de Bruijn, F.J., Szabados, L., and Schell, J. (1990). Chimeric genes and transgenic plants are used to study the regulation of genes involved in symbiotic plant-microbe interactions. Dev. Genet. 11, 182-196. de Bruijn, F.J., and Schell, J. (1992). Regulation of plant genes specifically induced in developing and mature nitrogen-fixing nodules: cis-Acting elements and trans-acting factors. In Control of Plant Gene Expression, D.P.S. Verma, ed. (Boca Raton, FL: CRC Press), pp. 241- 258. 34 de Bruijn, F.J., Jing, Y, and Dazzo, EB. (1995). Potential and pitfalls of trying to extend symbiotic interactions of nitrogen-fixing organisms to presently non-nodulated plants, such as rice. Plant and Soil 174, 225-240. Decker, C.J., and Parker, R. (1994). Mechanisms of mRNA degradation in eukaryotes. Trends Biochem. Sci. 19, 336-340. Dehio, C., and de Bruijn, FJ. (1992). The early nodulin gene SrEnod2 fi'om Sesbania rostrata is inducible by cytokinin. Plant J. 2, 117-128. Denarie, J., and Cullimore, J. (1993). Lipo-oligosaccharide nodulation factors: A minireview new class of signaling molecules mediating recognition and morphogenesis. Cell 74, 951-954. Dickstein, R.T., Bisseling, T., Reinhold, V.N. and Ausubel, EM. (1988). Expression of nodule-specific genes in alfalfa root nodules blocked at an early stage of nodule development. Genes Dev. 2, 677-678. Ding, D., and Lipshitz, H.D. (1993). Localized RNAs and their functions. Bioessays 15, 651-658. Ding, D., Parkhurst, S.M., Halsell, SR, and Lipshitz, H.D. (1993). Dynamic hsp83 RNA localization during Drosophila oogenesis and embryogenesis. Mol. Cell. Biol. 13, 3773-3781. Dixon, R., Kennedy, C., Kondorosi, A., Krishnapillai, V., and Merrick, M. (1977). Complementation analysis of Klebsiella pneumoniae mutants defective in nitrogen fixation. Mol. Gen. Genet. 157, 189-198. Ehrhardt, D.W., Atkinson, E.M., and Long, SR. (1992). Depolarization of alfalfa root hair membrane potential by Rhizobium meliloti Nod factors. Science 256, 998-1000. Ehrhardt, D.W., Wais, R., and Long, SR. (1996). Calcium spiking in plant root hairs responding to Rhizobium nodulation signals. Cell 85, 673 -68 1 . 35 Ephrussi, A., and Lehmann, R. (1992). Induction of germ cell formation by oskar. Nature 358, 387-392. Ferrandon, D., Elphick, L., Nusslein-Volhard, C., and St Johnston, D. (1994). Staufen protein associates with the 3' UTR of bicoid mRNA to form particles that move in a microtubule-dependent manner. Cell 79, 122 1- 1 232. Fisher, RE, and Long, SR. (1992). Rhizobium-plant signal exchange. Nature 357, 655-660. Fortin, M.G., Morrison, N.A., and Verma, D.P.S. (1987). Nodulin-26, a peribacteroid membrane nodulin is expressed independently of the development of peribacteroid compartment. Nucleic Acids Res. 15, 813- 824. Franssen, H.J., Nap, J.-P., Gloudemans, T., Stickema, W., van Dam, H., Govers, F., Louweres, J., van Kammen, A., and Bisseling, T. (1987). Characterization of cDNA for nodulin-75 of soybean: a gene product involved in early stages of root nodule development. Proc. Natl. Acad. Sci. USA 84, 4495-4499. Gil, P., and Green, P.J. (1996). Multiple regions of the Arabidopsis SA UR- AC1 gene control transcript abundance: the 3’ untranslated region functions as an mRNA instability determinant. EMBO J. 15, 1678-1686. Gloudemans, T., and Bisseling, T. (1989). Plant gene expression in early stages of Rhizobium-legume symbiosis. Plant Sci. 65, 1-4. Govers, F., Franssen, H.J., Pieterse, C., Wilmer, J., and Bisseling, T. (1990). Function and regulation of the early nodulin gene ENODZ. In Genetic Engineering of Crop Plants, G.W. Lycett and D.W. Grierson, eds (London: Butterworths), pp. 259-269. Green, P.J. (1993). Control of mRNA stability in higher plants. Plant Physiol. 102, 1065-1070. 36 Hargrove, J.L., and Schmidt, EH. (1989). The role of mRNA and protein stability in gene expression. F ASEB J. 3, 2360-2370. Heard, J., and Dunn, K. (1995). Symbiotic induction of a MADS-box gene during development of alfalfa root nodules. Proc. Natl. Acad. Sci. USA 92, 5273-5277. Heidstra, R., Geurts, R., Franssen, H., Spanik, H.P., van Kammen, A., and Bisseling, T. (1994). Root hair deformation activity of nodulation factors and their fate on Vicia sativa. Plant Physiol. 105, 787-797. Herrick, D., Parker, R., and Jacobson, A. (1990). Identification and comparison of stable and unstable mRNAs in Saccharomyces cerevisiae. Mol. Cell. Biol. 10, 2269-2284. Hirsch, A.M., Bhuvaneswari, T.V., Torrey, J .G., and Bisseling, T. (1989). Early nodulin genes are induced in alfalfa root outgrowths elicited by auxin transport inhibitors. Proc. Natl. Acad. Sci. USA 86, 1244-1248. Hirsch, A.M. (1992). Developmental biology of legume nodulation. New Phytol. 122, 211-237. Hirsch, A.M., Assad, S., Fang, Y., Wycoff, K., and Loebler, M. (1993). Molecular interactions during nodule development. In New Horizons in Nitrogen Fixation, R. Palacios, J. Moira and WE. Newton, eds (Dordrecht: Kluwer Academic Publishers), pp. 291-296. Hunt, A.G., and MacDonald, M.H. (1989). Deletion analysis of the polyadenylation signal for a pea ribulose- l ,5-bisphosphate carboxylase small-subunit gene. Plant Mol. Biol. 13, 125-138. Hunt, A.G. (1994). Messenger RNA 3’ end formation in plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 45, 47-60. Jensen, E.Q., Stougaard, J., Jorgensen, J.E., Sandal, N., and de Bruijn, FJ. (1988). Regulation of nodule-specific plant genes. In Nitrogen 37 Fixation: Hundred Years After, H., Bothe, F.J., de Bruijn, and WE. Newton, eds (Stuttgart: Gustav Fisher), pp. 605-609. Jones, T.R., and Cole, MD. (1987). Rapid cytoplasmic turnover of c-myc mRNA: requirement of the 3’ untranslated sequences. Mol. Cell. Biol. 7, 45 13-4521 . Joshi, CF. (1987). Putative polyadenylation signals in nuclear genes of higher plants: a compilation and analysis. Nucleic Acids Res. 15, 9627- 9640. Kij ne, J .W. (1992). The Rhizobium infection process. In Biological Nitrogen Fixation, G., Stacey, R.H., Burris, and H.J., Evans, eds (New York: Chapman and Hall), pp. 349-398. Kloc, M., Spohr, G., and Etkin, LB. (1993). Translocation of repetitive RNA sequences with the germ plasm in Xenopus oocytes. Science 262, 1 712-1 714. Kouchi, H., and Hata, S. (1993). Isolation and characterization of novel cDNAs representing genes expressed at early stages of soybean nodule development. Mol. Gen. Genet. 238, 106-1 19. Krause, A., Sigrist, C.J., Dehning, 1., Sommer, H., and Broughton, W.J. (1994). Accumulation of transcripts encoding a lipid transfer-like protein during deformation of nodulation-competent Vigna unguiculata root hairs.Mol. Plant Microbe Interact. 7, 41 1-418. Legocki, RR, and Verma, D.P.S. (1980). Identification of “nodule- specific” host proteins (nodulins) involved in the development of rhizobium-legume symbiosis. Cell 20, 153-163. Lerouge, P., Roche, P., Faucher, C., Maillet, F., Truchet, G., Prome, J.C.,. and Denarie, J. (1990). Symbiotic host-specificity of Rhizobium meliloti is determined by a sulfated and acylated glucosamine oligosaccharide signal. Nature 344, 781-784. 38 Libbenga, K.R., Van Iren, F., Bogers, R.J., and Schraag-Lamers, M.F. (1973).The role of hormones and gradients in the initiation of cortex proliferation and nodule formation in Pisum sativum L. Planta 114, 29- 39. Libbenga, K.R., and Bogers, RJ. (1974). Root-nodule morphogenesis. In The Biology of Nitrogen Fixation, A., Quispel, ed (Amsterdam: North- Holland Publishing Co.). pp. 430-472. Long, SR. (1989). Rhizobium-legume nodulation. Life together in the underground. Cell 56, 203-214. MacDonald, M.H., Mogen, B.D., and Hunt, A.G. (1991). Characterization of the polyadenylation signal from the T-DNA-encoded octopine synthase gene. Nucleic Acids Res. 19, 5575-5581.] Macdonald, P.M., Kerr, K., Smith, J.L., and Leask, A. (1993). RNA regulatory element BLEl directs the early steps of bicoid mRNA localization. Development 1 l 8, 1 23 3- 1 243. Matvienko, M., van de Sande, K., Yang, W.C., van Kammen, A., Bisseling, T., Franssen, H. (1994). Comparison of soybean and pea ENOD40 cDNA clones representing genes expressed during both early and late stages of nodule development. Plant Mol. Biol. 26, 487-493. McClure, B.A., Hagen, G., Brown, C.S., Gee, M.A., Guifoyle, T.J. (1989). Transcription, organization, and sequence of an auxin-regulated gene cluster in soybean. Plant Cell 1, 229-239. Miao, G.-H., and Verma, D.P.S. (1993). Soybean nodulin-26 gene encoding a channel protein is expressed only in the infected cells of nodules and is regulated differently in roots of homologous and heterologous plants. Plant Cell 5, 781-794. Miflin, B.J., and Lea, P.J. (1980). Ammonia assimilation. In The Biochemistry of Plants: A Comprehensive Treaties, Vol 5, B.J., Miflin, ed (New York: Academic), pp. 169-202. 39 Moerman, M., Nap, J.P., Govers, F., Schilperoort, R., and van Kammen, A. (1987). Rhizobium nod genes are involved in the induction of two early nodulin genes in Vicia sativa root nodules. Plant Mol. Biol. 9, 1 71 - 1 79. Muhlrad, D., and Parker, R. (1992). Mutations affecting stability and deadenylation of the yeast WAZ transcript. Genes Dev. 6, 2100-2111. Mylona, P., Pawlowski, K., and Bisseling, T. (1995). Symbiotic Nitrogen Fixation. Plant Cell 7, 869-885. Nielsen, F.C., and Christiansen, J. (1992). Endonucleolysis in the turnover of insulin-like growth factor II mRNA. J. Biol. Chem. 267, 19404-19411. Nap, J.-P., and Bisseling, T. (1990). Developmental biology of a plant- prokaryote symbiosis: the legume root nodule. Science 250, 948-954. Neuman-Silberberg, F., and Schiipbach, T. (1993). The Drosophila dorsoventral patterning gene gurken produces a dorsally localized RNA and encodes a TGFa-like protein. Cell 75, 165-174. Newcomb, W. (1981). Nodule morphogenesis and differentiation. In Biology of Rhizobiaceae, K.L., Giles and A.G., Atherly, eds (New York: Academic Press), pp. 247-297. Newman, T.C., Ohme-Takagi, M., Taylor, C.B., and Green, P.J. (1993). DST sequences, highly conserved among plant SA UR genes, target reporter transcripts for rapid degradation in tobacco. Plant Cell 5, 701- 714. Ohme—Takagi, M., Taylor, C.B., Newman, T.C., and Green, P.J. (1993). The effect of sequences with high AU content on mRNA stability in tobacco. Proc. Natl. Acad. Sci. USA 90, 1181 1-1 1815. Ostareck-Lederer, A., Ostareck, D.H., Standart, N., and Thiele, B.J. (1994). Translation of 15-lipoxygenase mRNA is inhibited by a protein 40 that binds to a repeated sequences in the 3' untranslated region. EMBO J. 13, 1476-1481. Owen, D., and Kiihn, LC. (1987). Noncoding 3’ sequences of the transferrin receptor gene are required for mRNA regulation by iron. EMBO J. 6, 1287-1293. Pandey, N.B., and Marzluff, W.F. (1987). The stem-loop structure at the 3’ end of histone mRNA is necessary and sufficient for regulation of histone mRNA stability. Mol. Cell. Biol. 7, 4557-4559. Peltz, S.W., Brewer, G., Bernstein, P., Hart, P.A., and Ross, J. (1991). Regulation of mRNA turnover in eukaryotic cells. Crit. Rev. Euk. Gene Exp. 1, 99-126. Peters, N .K., Frost, J.W., and Long, S.R. (1986). A plant flavone, luteolin, induces expression of Rhizobium meliloti nodulation genes. Science 233, 977-980. Proudfoot, NJ. (1991). Poly(A) signals. Cell 64, 671-674. Ramlov, K.R., Laursen, N.B., Stougaard, J., and Marcker, K.A. (1993). Site-directed mutagenesis of the organ-specific element in the soybean leghemoglobin le3 gene promoter. Plant J. 4, 577-580. Rastinejad, F., and Blau, H.M. (1993). Genetic complementation reveals a novel regulatory role for 3' untranslated regions in growth and differentiation. Cell 72, 903-917. Rastinejad, F., Conboy, M.J., Rando, T.A., and Blau, H.M. (1993). Tumor suppression by RNA from the 3 ’ untranslated region of alpha- tropomyosin. Cell 75, 1 107-1 1 17. Ross, J. (1995). mRNA stability in mammalian cells. Microbiol. Rev. 59, 423-450. Ross, J. (1996). Control of messenger RNA stability in higher eukaryotes. TIG 12,171-175. 41 Ruvkun, G.B., Sundaresan, V., and Ausubel, EM. (1982). Directed transposon Tn5 mutagenesis and complementation analysis of Rhizobium meliloti symbiotic nitrogen fixation genes. Cell 29, 551-559. Sanchez, F., Quinto, C., Vazquez, M., De las Penas, A., Cevallos, M.A. (1988). The symbiotic association between Phaseolus vulgaris and Rhizobium phaseoli. In Molecular Genetics of Plant-Microbe Interactions, R., Palacios, and D.P.S. Verma, eds (St. Paul, MN: APS Press), pp. 370-375. Sanchez, F., Padilla, J.E., Perez, H., and Lara, M. (1991). Control of nodulin genes in root-nodule development and metabolism. Annu. Rev. Plant Physiol. Plant Mol. Biol. 42, 507-528. Shaw, G., and Kamen, R. (1986). A conserved AU sequence from the 3' untranslated region of GM-CSF mRNA mediates selective mRNA degradation. Cell 46, 659-667. Scheres, B., van de Wiel, C., Zalensky, a., Horvath, b., Spaink, H.P., van Eck, H., Zwartkruis, F., Walters, Aa.-M., Gloudemans, T., van kammen, A., and Bisseling, T. (1990a). The ENODIZ gene product is involved in the infection process during pea—Rhizobium interaction. Cell 60, 281-294. Scheres, B., van Engelen, F., van der Knaap, E., van de Wiel, C., van kammen, A., and Bisseling, T. (1990b). Sequential induction of nodulin gene expression in the developing pea nodule. Plant Cell 2, 687-7 00. Schubert, KR. (1986). Products of biological nitrogen fixation in higher plants: Synthesis, transport, and metabolism. Annu. Rev. Plant Physiol. 37, 539-574. Silver, D., Pinaev, A., Chen, R, and de Bruijn, FJ. (1996). Post- transcriptional regulation of the Sesbanz'a rostrata early nodulin gene SrEnod2 by cytokinin. Plant Physiol. 112, 559-567. 42 Simon, R., Tassan, J.-P., and Richter, J.D. (1992). Translational control by poly(A) elongation in Xenopus development: differential repression and enhancement by a novel cytoplasmic polyadenylation element. Genes Dev. 6, 2580-2591. Smit, G., van Brussel, T.A.N., and Kijne, J.W. (1993). Inactivation of a root factor by ineffective Rhizobium: A molecular key to autoregulation of nodulation in Pisum sativum. In New Horizons in Nitrogen Fixation, R. Palacios, J. Mora, and WE. Newton, eds (Dordrecht: Kluwer Academic Publishers), p. 371. Smit, G., de Koster, C.C., Schripsema, J., Spaink, H.P., van Brussel, A.A., and Kijne, J.W. (1995). Uridine, a cell division factor in pea roots. Plant Mol. Biol. 29, 869-873. St Johnston, D., Brown, N.H., Gall, J.G., and Jantsch, M. (1992). A conserved double-stranded RNA-binding domain. Proc. Natl. Acad. Sci. USA 89, 10979-10983. St Johnston, D., and N usslein-Volhard, C. (1992). The origin of pattern and polarity in the Drosophila embryo. Cell 68, 201-219. St Johnston, D. (1995). The intracellular localization of messenger RNAs. Cell 81, 161-170. Stougaard, J., Sandal, N.N., Gron, A., Kiihle, A., and Marcker, K.A. (1987). 5’-Analysis of the soybean leghemoglobin lbc3 gene: Regulatory elements required for promoter activity and organ specificity. EMBO J. 6, 3565-3569. Strittmatter, g., Chia, T.-F., Trinh, T.H., Katagiri, F., Kuhlemeier, C., and Chua, NH. (1989). Characterization of nodule-specific cDNA clones from Sesbania rostrata and expression of the corresponding genes during initial stages of stem nodules and root nodules formation. Mol. Plant-Microbe Interact. 2, 122-127. 43 Sullivan, M.L., and Green, P.J. (1993). Post-transcriptional regulation of nuclear-encoded genes in higher plants: the roles of mRNA stability and translation. Plant Mol. Biol. 23, 1091-1104. Szabados, L., Ratet, P., Grunenberg, B., and de Bruijn, FJ. (1990). Functional analysis of the Sesbania rostrata leghemoglobin glb3 gene 5'- upstream region in transgenic Lotus corniculatus and Nicotiana tobacum plants. Plant Cell 2, 973-986. Szczyglowski, K., and Legocki, A.B. (1990). Isolation and nucleotide sequence of cDNA clone encoding nodule-specific (hydroxy) proline- rich protein LEnodZ from yellow lupin. Plant Mol. Biol. 15, 361-363. Szczyglowski, K., Szabados, L., Fujimoto, S.F., Silver, D., and de Bruijn, FJ. (1994). Site-specific mutagenesis of the nodule-infected cell expression (NICE) element and the AT-rich element ATRE-BSZ* of the Sesbania rostrata leghemoglobin glb3 promoter. Plant Cell 6, 317-332. Tarun, S.Z., and Sachs, A.B. (1995). A common function for mRNA 5' and 3' ends in translation initiation in yeast. Genes Dev. 9, 2997-3007. Thimann, K.V. (1936). On the physiology of the formation of nodules on legume roots. Proc. Natl. Acad. Sci. USA 22, 511-515. Tjepkema, J.D., and Yocum, GS. (1974). Measurement of oxygen partial pressure within soybean nodules by oxygen microelectrodes. Planta 119, 3 5 1-3 60. Truchet, G., Roche, P., Lerouge, P., Vasse, J., Camut, S., de Billy, F., Prome, J.C. and Denaire, J. (1991). Sulphated lipooligosaccharide signals from Rhizobium meliloti elicit root nodule organogenesis in alfalfa. Nature 351, 670-673. van Brussel, A.A.N., Bakhuizen, R., van Spronsen, F.C., Spaink, H.P., Tak, T., Lugtenberg, B.J.J., and Kijne, J.W. (1992). Induction of pre- infection thread structures in the leguminous host plant by mitogenic lipo-oligosaccharides of Rhizobium. Science 257, 70-72. 44 van de Sande, K., Pawlowski, K., Czaja. 1., Wieneke, U., Schell, J., Schmidt, J., Walden, R., Matvienko, M., Wellink, J., van Kammen, A., Franssen, H., and Bisseling, T. (1996). Modification of phytohormone response by a peptide encoded by Enod40 of legumes and a nonlegume. Science 273, 370-373. van de Wiel, C., Scheres, B., Franssen, H., van Lierop, M.J., van Lammeren, A., van Kammen, A., and Bisseling, T. (1990a). The early nodulin transcript ENODZ is located in the nodule parenchyma (inner cortex) of pea and soybean root nodules. EMBO J. 9, 1-7. van de Wiel, C., Norris, H.J., Bochenek, B., Dickstein, R., Bisseling, T., and Hirsch, M.A. (1990b). Nodulin gene expression and ENOD2 localization in effective, nitrogen-fixing and ineffective, bacteria-free nodules of alfalfa. Plant Cell 2, 1009-1017. van Kammen, A. (1984). Suggested nomenclature for plant genes involved in nodulation and symbiosis. Plant Mol. Biol. Rep. 2, 43-45. Verma, D.P.S. and Delauney, AJ. (1988). Root nodule symbiosis: nodulins and nodulin genes. In Temporal and Spatial Regulation of Plant Genes, R.B., Goldberg, and D.P.S., Verma, eds (New York: Springer- Verlag), pp. 169-199. Verma, D.P.S. (1992). Signals in root nodule organogenesis and endocytosis of Rhizobium. Plant Cell 4, 373-3 82. Vijn, I., das Neves, L., van Kammen, A., Franssen, H., and Bisseling, T. (1993). Nod factors and nodulation in plants. Science 260, 1764-1765. Wahle, E., Keller, W. (1992). The biochemistry of 3’-end cleavage and polyadenylation of messenger RNA precursors. Annu. Rev. Biochem. 61, 419-440. Weaver, C.B., Shower, N.H., Louis, CE, and Roberts, D.M. (1994). Nodulin-26, a nodule specific symbiosome membrane protein from soybean, is an ion channel. J. Biol. Chem. 269, 17858-17862. 45 Weiss, I.M., and Liebhaber, S.A. (1995). Erythroid cell-specific mRNA stability elements in the a2-globin 3’ nontranslated region. Mol. Cell. Biol. 15, 2457-2465. Wilson, R.C., Long, F., Maruoka, E.M., and Cooper, J.B. (1994). A new proline-rich early nodulin from Medicago truncatula is highly expressed in nodule meristematic cells. Plant Cell 6, 1265-1275. Witty, J.F., Minchin, F.R., Skot, L., and Sheehy, J.E. (1986). Nitrogen fixation and oxygen in legume nodules. Oxford Surv. Plant Mol. Cell Biol. 3, 275-314. Wormington, M. (1993). Poly(A) and translation: development control. Curr. Opin. Cell Biol. 5, 950-954. Wu, L., Ueda, T., and Messing, J. (1995). The formation of mRNA 3 ’- ends in plants. Plant J. 8, 323-329. Yang, W.-C., de Blank, C., Meskiene, 1., Hirt, H., Bakker, J., van Kammen, A., Franssen, H., and Bisseling, T. (1994). Rhizobium Nod factors reactivate the cell cycle during infection and nodule primordium formation, but the cycle is only completed in primordium formation. Plant Cell 6, 1415-1426. CHAPTER 2 THE NODULE PARENCHYMA-SPECIFIC EXPRESSION OF THE SESBANIA ROST RA TA EARLY NODULIN GENE SrENODZ IS MEDIATED BY ITS 3' UNTRANSLATED REGION (3’UTR) This section has been submitted to the Plant Cell for publication. Reference: Chen, R., Silver, D., and de Bruijn, F .J . (1996). The nodule parenchyma-specific expression of the Sesbania rostrata early nodulin gene A SrEnod2 is mediated by its 3' untranslated region (3’UTR). 46 47 ABSTRACT The early nodulin Enod2 gene encodes a putative hydroxyproline-rich cell wall protein, and is expressed exclusively in the nodule parenchyma cell layer. The latter suggests that the Enod2 protein may contribute to the special morphological features of the nodule parenchyma and to the creation of an oxygen diffusion barrier (Nap and Bisseling, 1990, Science 250, 948- 954). We have shown that the Enod2 gene of the stem-nodulating legume Sesbania rostrata (SrEnod2) can be specifically induced in roots by the plant hormone cytokinin (Dehio and de Bruijn, 1992, Plant J. 2, 117-128), and that this induction occurs at a posttranscriptional level (Silver et al., 1996, Plant Physiol. 112, 559-567). Here, we have characterized cis- determinant(s) in the SrEnod2 locus responsible for nodule parenchyma- specific expression and report that the 3 ’ untranslated region (3' UTR) of the SrEnod2 gene is both required and sufficient for directing chimeric reporter gene expression in the nodule parenchyma of transgenic Lotus corniculatus plants. By carrying out a detailed deletion analysis of the 5’ and 3’ SrEnod2 regions, we delimited the minimal promoter of the SrEnod2 gene and revealed that the 5’ flanking sequences do not appear to be essential for nodule parenchyma-specific expression. This finding contrasts with the report that the 5’ upstream region of the soybean Enod2 gene directs nodule . parenchyma-specific expression (Lauridsen et al., 1993, Plant J. 3, 483- 48 492), indicating different mechanisms may be involved in regulation of the expression of these two genes. To our knowledge, this is also the first demonstration that cis-element(s) specific for tissue-specific expression are located within the 3 ’ UTR of a plant nuclear gene. INTRODUCTION The Enod2 gene is an early nodulin gene that was first identified by differential screening of a soybean root nodule cDNA library (F ranssen et al., 1987). Isolation of Enod2 homologues from other legumes, and subsequent DNA sequence comparisons, have shown that the Enod2 gene is highly conserved among leguminous plants (Franssen et al., 1987; Dickstein et al., 1988; Govers et al., 1990; van de Wiel et al., 1990a; Szczyglowski and Legocki, 1990; Dehio and dc Bruijn, 1992). The amino acid sequence derived from the soybean Enod2 cDNA clone revealed that the Enod2 protein is very proline-rich and is mainly composed of two repeating pentapeptides, ProProI-IisGluLys and ProProGluTyrGln (Franssen et al., 1987; van de Wiel et al., 1990a). The Enod2 proteins of other legumes appear to be composed of the same or similar repeating pentapeptides (Dickstein et al., 1988; van de Wiel et al., 1990a; Szczyglowski and Legocki, 1990; Dehio and de Bruijn 1992). A putative signal peptide is present at the N-terminus of all of the Enod2 proteins identified thus far, 49 and the amino acid sequences of both pea and soybean Enod2 proteins highly resemble a hydroxyproline-rich cell-wall protein identified in soybean seed cell walls (Averyhart—Fullard et al., 1988; van de Wiel, et al., 1990a). Therefore, it has been proposed that the Enod2 protein is a cell wall protein (van de Wiel et al., 1990a). The Enod2 gene is expressed during early stages of nodule organogenesis (Franssen et al., 1987 ; van de Wiel et al., 1990a; Dehio and de Bruijn 1992). It has been shown that this early nodulin gene is also expressed in empty nodules induced by certain Rhizobium strains on the roots of soybean and alfalfa (Finan et al., 1985; Franssen et al., 1987; Dickstein et al., 1988). These empty nodules contain neither infection threads nor intracellular bacteroids. Expression of the Enod2 gene has also been demonstrated in “pseudo-nodules” induced by auxin transport inhibitors (Hirsch et al., 1989; van de Wiel et al., 1990b) or by Rhizobium nodulation-deficient (N od‘) mutants secreting the cytokinin trans-zeatin (Cooper and Long, 1994). The expression of the Enod2 gene in these empty/pseudo-nodules strongly suggests that the Enod2 protein is not involved in the infection process per se, but may play a role in nodule ontogeny. The finding that the Enod2 gene is expressed in alfalfa pseudo-nodules induced by auxin transport inhibitors (Hirsch et al., 1989) extends the observations made several decades ago suggesting that plant hormones are 50 involved in the induction of nodules on plant roots (Thimann, 1936; Arora et al., 1959; Libbenga et al., 1973; reviewed by Hirsch, 1992; Verma, 1992). In fact, plant hormones clearly appear to be involved in the expression of the Enod2 gene; Dehio and de Bruijn (1992) first demonstrated that the expression of the Enod2 gene of Sesbania rostrata is specifically up- regulated in roots after the application of cytokinin. Similar results have been reported for the alfalfa Enod2 gene (Hirsch et al., 1993), although the size of the cytokinin-enhanced Enod2 transcript in roots is larger than that observed in alfalfa nodules. In-situ hybridization studies have demonstrated that the soybean Enod2 gene is expressed in the nodule parenchyma and in the nodule tissue surrounding the vascular bundle that connects the nodule to the root central cylinder (van de Wiel et al., 1990a). Low levels of the Enod2 transcript have also been found in the nodule endodermis and the nodule outer-cortical cells directly adjacent to the endodermis (van de Wiel et al., 1990a). The nodule parenchyma is composed of several layers of cells, with fewer and smaller intercellular spaces than most other cortical cells (van de Wiel et al., 1990a). The compact structure of these cell layers has suggested to various investigators that they may be involved in forming an oxygen diffusion barrier (Tjepkema and Yocum, 1974; Witty et al., 1986; Nap and Bisseling, 1990; van de Wiel et al., 1990a). Since the Enod2 protein is localized in the same cell layers, it has been hypothesized to contribute to the special 51 morphology of the nodule parenchyma and therefore to the formation of the oxygen barrier (Nap and Bisseling, 1990; van de Wiel et al., 1990a). However, no experimental evidence supporting this hypothesis exists. Recently, considerable efforts have been made to understand how Enod2 gene expression is controlled in developing soybean and S. rostrata nodules and in roots of S. rostrata plants after cytokinin treatment (Lauridsen et al., 1993; de Bruijn et al., 1994; Silver et al., 1996). We have shown that the cytokinin-mediated up-regulation of the SrEnod2 gene in uninfected roots of S. rostrata occurs mainly in cytoplasm at the post- transcriptional level, requires protein synthesis, and appears to involve protein phosphorylation (Silver et al., 1996). We have also reported that a 3- kb 5’ upstream region of the SrEnod2 gene failed to consistently direct the expression of the uidA reporter gene (gus; Jefferson et al., 1987) in the nodule parenchyma of transgenic Lotus corniculatus plants (de Bruijn et al., 1994). In contrast, a 3-kb 5’ region of the soybean Enod2(B) gene [GmEnod2(B)] has been shown to direct reporter gene expression in the nodule parenchyma of transgenic L. corniculatus plants (Lauridsen et al., 1993). Here, we describe the detailed analysis of DNA sequence determinants in the SrEnod2 locus responsible for tissue-specific expression in nodules. We conclude from this analysis that, although the 5' proximal region of the SrEnod2 gene is required for gene expression, the 3’ untranslated region (3 ’ UTR) of the SrEnod2 gene contains DNA sequence 52 determinants responsible and sufficient for nodule parenchyma-specific gene expression. METHODS Plant growth and nodulation Sesbania rostrata seeds were surface-sterilized and germinated on petri dishes in the dark at 28°C for one day, as described by Pawlowski et al., (1991). Seedlings were grown for three weeks in autoclaved soil mix (MetroMixzsand; 2:1) in growth chambers with an 18-hr light (28°C) and 6-hr dark (22°C) cycle. Plants were inoculated with a two-day-old culture of Azorhizobium caulinodans ORSS71 (Dreyfus and Dommergues, 1981) either in pots for root nodulation or on stems for stem nodulation. Six-day- old root nodules and stem nodules were collected and processed (see below). Lotus corniculatus transgenic plants were grown for one week in autoclaved soil mix (vermiculitezsandzmetromix; 10:10:1) in growth chambers with a 16-hr light (24°C) and 8-hr dark (18°C) cycle. Plants were inoculated with a three-day-old culture of Rhizobium lotz' NZP2037 (Pankhurst, 1983) and grown for 21 days. Plant tissues (nodules, infected roots, and leaves plus stems) were collected and stored in liquid nitrogen. 53 In-situ RNA localization In-situ RNA localization studies were carried out essentially as described by Cox and Goldberg (1988) and Van de Wiel et al. (1990a), except that a non- radioactive approach was used according to De Block and Debrouwer (1993). Stern and root nodules were fixed in a 10 mM sodium phosphate buffer (pH 6.8) containing 4% paraformaldehyde, 0.25% glutaraldehyde, and 100 mM NaCl at room temperature, dehydrated in a graded ethanol and xylene series, and embedded in paraplast X-tra (Fisher Scientific, F air Lawn, NJ). Nodule sections (7 um thick) were cut and attached to superfrost/plus microscopic slides (Fisher Scientific, Pittsburgh, PA). After removal of the paraplast with xylene and rehydration through an ethanol series, sections were treated with proteinase K (5 ug/ml) in 100 mM Tris- HCl (pH 7.5); 50 mM EDTA for 30 min at 37°C and with 0.25% acetic anhydride in 100 mM triethanolamine (pH 8.0) for 10 min at room temperature. Slides were subsequently dehydrated through an ethanol series and dried and hybridized with digoxigenin-labeled sense and antisense RNA probes generated by transcribing the coding region of the SrEnod2 gene (Dehio and de Bruijn, 1992) with T3 and T7 RNA polymerase, according to the manufacturer's instructions (Boehringer Mannheim GmbH, Mannheim, Germany). 54 Hybridization and washing were essentially carried out according to Cox and Goldberg (1988) and De Block and Debrouwer (1993). Sections were dehydrated via an ethanol series and mounted with poly-mount (Polysciences Inc, Warrington, PA). Sections were photographed using an Axiophot microscope (Zeiss, Oberkochen, Germany). Isolation of total RNA and poly(A)+ RNA Total RNA was isolated according to the hot phenol method of Verwoerd et al., (1989), with the following modifications. Briefly, plant material (up to 1 gm) was ground in liquid nitrogen to a fine powder and then transferred into a 15-ml tube containing 4 ml of preheated (65°C) RNA isolation buffer (Hall et al., 1978), and 4 ml of preheated (65°C) phenol. Samples were vortexed vigorously for 5 min, 4 ml of chloroform was added, the mixture was vortexed again for 15 min and centrifuged at 10,000 rpm for 20 min. The supernatant was extracted once with an equal volume of chloroform and centrifuged again. RNA in the supernatant was precipitated at 10,000 rpm for 30 min after adding 1/4 volume of 8 M LiCl and 1/10 volume of 1.6 M KCl and an overnight incubation on ice. The RNA pellet was resuspended in 400 pl of DEPC-treated water and extracted with an equal volume of chloroform until a clear interface was formed. RNA was then 55 precipitated and resuspended in DEPC-HzO. The RNA concentration was determined photometrically. Poly(A)+ RNA was prepared by passing total RNA through oligo—dT columns according to the manufacturer's instructions (5 prime—>3 prime, Boulder, CO). Northern analysis was performed according to standard procedures (Maniatis et al., 1982). Briefly, Poly(A)+ RNA was loaded on a 1% formaldehyde agarose gel and separated at 4°C. The gel was soaked in 10x SSC for 30 min, to remove formaldehyde, and subsequently blotted to a 0.22 um nitrocellulose membrane (NitroPlus, Micron Separation Inc, Westborough, MA). The membrane was rinsed briefly in DEPC-HzO and baked at 80°C for 2 hr under vacuum. Prehybridization was carried out at 51°C overnight in a solution containing 5x SSC, 10x Denhardt's solution, 0.1% SDS, 0.1M KPO4, pH 6.8, and 100 ug/ml salmon sperm (ss) DNA. Hybridization reactions were carried out at 65°C for 18 hr in a solution containing 1x SSC, 10x Denhardt's solution, 0.1M KPO4 (pH 6.8), 100ug/ml ssDNA, 10% dextran sulphate, 50% formamide. After hybridization, the membrane was washed at 65°C in 2x SSC and 1x SSC for 20 min each and finally in 0.2x SSC for 40 min. Blots were exposed to X- ray film at -75°C. 56 Chimeric reporter gene construction A 5.3-kb EcoRI fragment, which contains a 1.9-kb 5' region, a l-kb coding region, and a 2.4-kb 3' downstream region of the SrEnod2 gene was subcloned from the genomic clone ICDl (Dehio and de Bruijn, 1992) into the pBluescript KS- vector (Stratagene, La Jolla, CA). Restriction sites (BamHI and SalI) were introduced at the translational initiation codon and at the 3' end of the putative signal peptide sequence of the SrEnod2 gene, using site-directed mutagenesis as described by Kunkel et a1. (1985). The oligonucleotide primers used for introducing the BamI-II and SalI sites were (5')GAGTAGTGTAGAGAQGATCCTTCTATCTAT(3') and (5')GGCTCA TAGTAATI‘AGTCGAQACTGGAGO'), respectively. The underlined nucleotide sequences correspond to the restriction endonuclease recognition sites and the highlighted nucleotides correspond to the mutations introduced. A 1.9-kb EcoRI-BamHI fragment, which contains the entire 5' untranslated region and upstream region of the SrEnod2 gene, was inserted in front of the gus coding region carried on plasmid pB1101.1 (Clontech Laboratories, Palo Alto, CA). The EcoRI end of the fragment was filled in by Klenow before cloning. The resulting binary vector was used to generate constructs 1 and 2 (Figure 2-2A). Similarly, a 1.98-kb EcoRI-Sall fragment, which contains the 75-bp signal peptide sequences in addition to the 5' sequences of the SrEnod2 gene, was inserted in front of the gus coding 57 region of plasmid pBIlOl.2, and the resulting vector was used to generate construct 5. To construct a transcriptional fusion between the 1.98-kb 5’ sequences of the SrEnod2 gene and the gus coding region, a second-round mutagenesis was performed on the 5.3-kb EcoRI fragment containing the introduced SalI site with an oligonucleotide carrying a mutation in the translational initiation codon (ATG to GGA). The resulting 1.98-kb EcoRI- (BamHI)-SalI fragment was excised and inserted in front of the gus coding region carried on plasmid pBIlOl.1 This vector was used to generate constructs 3 and 4 (Figure 2-2A). A restriction site (SacI) was introduced immediately downstream of the translational stop codon (TAA) of the SrEnod2 gene using the oligonucleotide primer (5')GTAGTGGTAGTGGTTGGAGQTCTTAATTT TI'ITTGG(3'). A 2.4-kb SacI-EcoRI fragment, containing the entire 3' untranslated and downstream region of the SrEnod2 gene, was inserted into pBIlOl.1 to generate constructs 1, 3, and 5 (Figure 2-2A). Constructs 6 and 7 were generated by inserting a 3.4-kb HindIII-Sacl fragment of plasmid pLPl4 (Szczyglowski et al., 1994), containing a 1.4-kb 5’ sequence of the Srglb3 gene and the gus coding region, into pBIlOl.1 and pBIlOl-3'SrEnod2, respectively. Constructs 8, 9, 10, and 11 were generated in a similar fashion by inserting a 2.1-kb BamHI-SacI fragment of pLP70 containing the 96-bp minimal CaMV35S promoter (-90 bp to +6 bp) and the gus coding region; a 2.15-kb BamHI-Sacl fragment of pLP62, 58 containing the 150-bp minimal nopaline synthase promoter (-150 bp to +1 bp) and the gus coding region, into the BamHI-Sacl sites of pBIlOl.1 and pBI 101-3'SrEnod2, respectively. Several 5’ deletion fragments were generated by restriction enzyme digestion of the 1.9—kb EcoRI-BamHI fragment. XbaI, Sau3A, RsaI, EcoRV, DraI, MspI Hinfl, AluI, SspI and Tsp were used to generate deletions ending at positions -1628, -1180, -854, -684, -464, -384, -306, -191, -115, and -50 bp, relative to the transcriptional start point. Several 3’ deletion fragments were generated by nested deletions using the Erase-a- Base kit (Promega, Madison, W1). A XhoI linker was inserted into the EcoRI site at the end of the 3’ sequences of the SrEnod2 gene in construct 1. The 3’ deletion fragments were then inserted into the SacI-XhoI sites of the modified construct 1. Regeneration of transgenic Lotus comiculatus plants Transgenic Lotus corniculatus cv Rodeo plants were generated according to Szabados et al. (1990). Binary vectors were conjugally transferred into the Agrobacterium rhizogenes strain A4 (Tempe and Casse-Delbart, 1989). Regenerated plants were transferred to a growth chamber (Conviron, Asheville, NC) and nodulated as described (Szabados et al., 1990). 59 Quantification of GUS enzymatic activity GUS enzymatic activity was quantified according to Jefferson et a1. (1987). The units of GUS activity used were picomoles of 4-methylumbelliferone (MU) produced per minute per milligram of protein. The concentration of protein in the extract was determined by the Bradford assay (Bradford, 1976). Fluorometric assays were performed using a fluorescence spectrophotometer (Model F-2000; Hitachi, Tokyo, Japan). Histochemical staining of GUS enzymatic activity GUS enzymatic activity was analyzed histochemically according to the procedure described by Jefferson (1987), with the following modifications. Hand-cut nodule sections were incubated at 37° in a solution containing 50 mM Na2PO4 and NaHPO4, pH 7, 1 mM ferri-lferro-cyanide and 20% methanol and 1 mM X-Gluc (5-bromo-4-chloro-3-indolyl B-D—glucuronide; Gold Biotechnology Inc., St. Louis, MO). Stained sections were then embedded in historesin (Reichert-Jung; Cambridge Instruments, Heidelberg, Germany), as described by De Block and DeBrouwer (1992), or in paraplast X—Tra, as described by Cox and Goldberg (1988). Sections of 10 um-thick were generated and examined by dark— and light-field microscopy using an Axiophot microscope. 60 RESULTS Localization of the SrEnod2 transcript in the parenchyma of S. rostrata stern and root nodules The tropical legume S. rostrata forms nodules on both roots and stems after infection with Azorhizobium caulinodans ORSS71 (Dreyfus and Dommergues, 1981; de Bruijn, 1989). The formation of stem nodules is induced following a crack-entry mediated infection of A. caulinodans ORSS71 at the base of dormant root primordia, which are present in rows along the stem (Tsien et al., 1983; de Bruijn, 1989). Root nodule organogenesis in S. rostrata appears to be intermediate between the indeterminate and determinate processes of nodule formation (Ndoye et al., 1994). These observations suggest that nodule formation in S. rostrata, as in many other tropical legumes, represents a primitive type of nodule organogenesis (Ndoye et al., 1994), and prompted us to study the tissue- specific expression of the SrEnod2 gene in S. rostrata stem and root nodules. The results, presented in Figure 2-1, show that the SrEnod2 gene is expressed in the nodule parenchyma cells of both stem and root nodules (Figure 2-1A and B). In both types of nodules, hybridization signals are also observed in cells immediately adjacent to the nodule vascular tissue (Figure 2-1C and D). Moreover, in stem nodules, hybridization signals are also 61 LEGEND FOR FIGURE 2-1. In-situ localization of SrEnod2 gene expression. Sections (7-um) of S. rostrata root and stem nodules harvested 6 days after inoculation with A. coulinodans ORSS71, were hybridized to a digoxigenin- labeled antisense RNA probe derived from the SrEnod2 coding region. Pink and black colors indicate hybridization signals. No signals were observed in neighboring sections hybridized with the sense probe (data not shown). (A) Section of a root nodule. (B) Section of a stem nodule. (C) Magnification of the section shown in (A). (D) Magnification of the section shown in (B). CT, central tissue; NP, nodule parenchyma; VB, vascular bundle. The arrowhead indicates expression in the nodule endodermis. 62 63 detected in cell layers corresponding to the nodule endodermis (Figure 2-1D, arrowhead). Expression of the SrEnod2 gene was not observed in other nodule cells. Therefore, the expression pattern of the SrEnod2 gene in the parenchyma cells of S. rostrata stem and root nodules appears to be similar to that of the Enod2 genes of other legumes (van de Wiel et al., 1990a) The 3' flanking region of the SrEnod2 gene is required for gene expression in the nodule parenchyma To identify DNA sequence determinants responsible for nodule parenchyma-specific expression of the SrEnod2 gene, a chimeric reporter gene fusion (gus; Jefferson, 1987; Jefferson et al., 1987) was constructed, in which the 5' and 3' flanking regions of the SrEnod2 gene were connected to the coding region of the gus reporter gene (5'SrEnod2-gus-3'SrEnod2; Figure 2-2A, construct 1). A construct containing the 3’ terminator region of the nopaline synthase gene (3’nos) was used as a control (5'SrEnod2-gus- 3'nos; Figure 2-2A, construct 2). These two constructs were introduced into L. corniculatus plants by, Agrobacterium rhizogenes-mediated plant transformation (Szabados et al., 1990), and their presence in transgenic plants was verified by Southern blotting (data not shown). Expression of the 64 LEGEND FOR FIGURE 2-2. Structure of gus fusion constructs and determination of GUS activity in transgenic L. corniculatus plants. (A) Structure of gus fusion constructs. The construct numbers are listed. The boxes in black and white represent gus, SrEnod2 and nos gene sequences, respectively. The thin open box represents a 75-bp putative signal peptide sequence. The vertical lines and wavy arrows represent the transcriptional start site of the SrEnod2 gene. The vertical bars attached to the boxes represent the translational start of the corresponding genes. + and - designate positive and negative staining results, respectively. (B) Quantification of GUS activity in transgenic L. corniculatus plants. The data shown represent mean values in units of pmol 4-methylumbelliferone (MU) per minute per mg protein. The error bar represents the standard deviation of the mean. The number of transgenic plants tested is indicated with an n. GUS ACTIVITY (STAININ G) B 1 1 :munnmnnmmnmnu- I no d u le 2 m root 3 ——l 4 2 1 5 n=31 0 50 100 150 200 250 300 350 GUS activity (pmol MU/min/mg protein) 66 gus reporter gene was histochemically and quantitatively determined in tissues of transgenic plants 21 days after Rhizobium inoculation. Histochemical staining revealed that gas reporter gene expression was visible as a ring-like pattern in cross sections of mature nodules expressing the 5'SrEnod2-gus-3'SrEnod2 construct (Figure 2-3B). No gus reporter gene expression could be detected in leaves, stems, or uninfected roots of the transgenic plants (Figure 2-2A). However, weak reporter gene expression could be detected in regions near the root tips of some transgenic plants, but only after Rhizobium infection (Figure 2-3B). When the 3’nos region was present instead of the SrEnod2 3’ region, no reporter gene expression could be detected in any tissues of the transgenic plants (Figures 2-2; and 2-3A). Dark-field microscopy revealed that the expression of the S'SrEnodZ- gus-3'SrEnod2 was localized in the nodule parenchyma and vascular tissues (Figure 2-3D). A low level of GUS activity was also detected in the outer cortex of the nodule. The histochemical staining pattern of the reporter gene expression correlates well with the in-situ Enod2 mRNA hybridization pattern found in S. rostrata nodules (Figure 2-1) and those observed for other legume plants (van de Wiel, 1990a). The only exception to this correlation is that our reporter gene construct was also expressed in the nodule vascular tissue. Dark-field microscopy confirmed that no GUS 67 LEGEND FOR FIGURE 2-3. Microscopic analysis of the expression of reporter gene constructs in transgenic L. corniculatus plants. (A) Hand-cut sections of transgenic nodules harboring 5’SrEnod2-gus- 3’nos construct (Figure 2-2A, construct 2). (B) Hand-cut sections of transgenic nodules harboring the 5’SrEnod2-gus- 3’SrEnod2 construct (Figure 2A, construct 1). GUS activity (blue color) can be seen in the nodule parenchyma and in regions near the root tip of selected infected roots. (C) Dark-field microscopy of a IO-um section of one of the transgenic nodules shown in (A). (D) Dark-field microscopy of a IO-um section of one of the transgenic nodules shown in (B). GUS activity (red color) can be seen in the nodule parenchyma, Vascular, and outer cortical tissues. CT, central tissue; E, endodermis; NP, nodule parenchyma; OC, outer cortex; and VB, vascular tissue. I 7’ ‘. '7' J 68 69 activity was detectable in nodules of transgenic plants harboring the 5'SrEnod2-gus-3'nos construct (Figure 2-3C). The histochemical staining analysis revealed limited variations in the expression of the reporter gene. Although 12 out of 17 transgenic plants harboring 5'SrEnod2-gus-3'SrEnod2 (Figure 2-2A, construct 1) had significant reporter gene expression in nodules and 5 plants were inactive, one plant had a high level of reporter gene expression in both nodules and infected roots (data not shown). A variation in reporter gene expression pattern was also observed in one out of 18 transgenic plants harboring the 5'SrEnod2-gus-3'nos construct, which had a low level of gus expression in the nodule vascular tissue (data not shown). It is likely that these variations are due to "position effects" (Odell et al., 1985; Sanders et al., 1987; Benfey et al., 1989), and were therefore excluded from further analysis. GUS enzymatic activity in nodules and infected roots of transgenic plants was quantified using a fluorometric assay. As shown in Figure 2-2B, GUS activity was detected only in nodules of transgenic plants harboring the 5'SrEnod2-gus-3'SrEnod2 construct. No GUS activity was detected in nodules of transgenic plants harboring the 5'SrEnod2-gus-3'nos construct, confirming the results of the histochemical staining analysis. These results strongly suggest that the 3' region of the SrEnod2 gene is required for tissue-specific reporter gene expression in transgenic plants. 70 In previous studies examining the interaction of DNA-binding proteins (trans-acting factors) with the SrEnod2 locus, a specific protein-DNA interaction with sequences encoding the putative SrEnod2 leader peptide was identified (A. Goel and F.J. de Bruijn, unpublished observation). Therefore, we also examined the role of the SrEnod2 putative signal peptide sequence in modulating gus reporter gene expression. A 75-bp nucleotide sequence encoding the N-terminal putative signal peptide of 25 amino acids (Dehio and de Bruijn, 1992) was fused to the truncated SrEnod2 5’ flanking region previously used (Figure 2-2A, constructs 1 and 2). Both in-frame translational and transcriptional fusions to the gas coding region were constructed (see Methods). The transcriptional fusions (Figure 2-2A, constructs 3 and 4) displayed a pattern and level of reporter gene expression similar to the original truncated constructs, lacking the leader sequence (Figure 2-2). The putative signal peptide sequence, therefore, does not appear to influence the expression of the reporter gene, when fused to it in a transcriptional configuration. In transgenic plants harboring the in-frame translational fusion (Figure 2-2A, construct 5), however, the gus reporter gene was not expressed (Figure 2-2B). Since the reporter gene transcript could be detected in nodules of transgenic plants harboring this gene fusion, albeit at a lower level than that observed with the transcriptional fusion (see below), the lack of GUS activity may be due to poor accumulation of GUS protein or to protein inactivation due to N-linked glycosylation in the 71 glycosylation in the endoplasmic reticulum during secretion, because of the presence of the putative leader sequence (Farrell and Beachy, 1990). Gus reporter gene activities correlate with steady state gus mRNA levels in transgenic nodules Northern blot analysis was performed to determine whether the gus reporter gene activities observed correlate with the accumulation levels of the reporter gene transcript. As shown in Figure 2-4, nodule poly(A)+ RNA isolated from a pool of three independent transgenic plants harboring constructs containing the SrEnod2 3’ region, strongly hybridized to the riboprobe prepared from the gus coding region (Figure 2-4, lanes 1, 3, and 5). The message detected was approximately 2.3 kb in length, as expected (Dehio and de Bruijn, 1992; R. Chen and F. J. de Bruijn, unpublished data). In contrast, nodule poly(A)+ RNA isolated from a pool of three independent transgenic plants harboring constructs containing the 3’ nos terminator did not hybridize to the gus probe (Figure 2-4, lanes 2 and 4). The same blot was stripped and reprobed with an eIF 4A gene probe (Taylor et al., 1993) to standardize loading. The signal peptide sequence of the SrEnod2 gene did not influence the accumulation of the reporter gene transcript when fused in a transcriptional configuration to the gus coding region (Figure 2-4, lanes 1 and 3). The gus 72 LEGEND FOR FIGURE 2-4. Northern blot analysis of gus reporter gene expression in transgenic nodules harboring the reporter gene constructs 1-5 (Figure 2A). The lane numbers correspond to the construct numbers. Two micrograms of poly(A)+ RNA were loaded in each lane and hybridized with a 32P-1abe1ed riboprobe derived from the gus coding region. The same blot was stripped and rehybridized with a 32P-labeled eIF 4A probe (Taylor et al., 1993) as a loading control. 73 74 reporter gene transcript also accumulated in nodules harboring the translational gene fusion (Figure 2-2A, construct 5), although the level of gus mRNA accumulation was reduced to approximately 50% of that of the transcriptional gene fusion (Figure 2-4, lanes 3 and 5). These data clearly show that the accumulation of the reporter gene transcript correlates well with the gus reporter activity detected in transgenic plants. Additionally, they reveal that the reporter gene transcript accumulates in transgenic plants harboring the constructs containing the SrEnod2 3 ’ region, but not the 3' nos terminator. The 3' region of the SrEnod2 gene contains sequence determinants sufficient for expression in the nodule parenchyma To further examine the role of the SrEnod2 3' region in directing nodule parenchyma-specific expression, additional chimeric reporter gene fusions were constructed containing heterologous promoters, e. g. the leghemoglobin (Srglb3) promoter from S. rostrata (Szabados et al., 1990; Szczyglowski et al., 1994), the nopaline synthase minimal promoter (-150 bp to +1 bp; Depicker et al., 1982; Ebert et al., 1987), and the CaMV3SS minimal promoter (-90 bp to +6 bp; domain A; Franck et al., 1980; Pietrzak et al., 1986; Benfey et al., 1989). The resulting constructs (see Figure 2-5A) 75 were introduced into L. corniculatus, and the expression of the gus gene was determined histochemically and quantitatively. It has previously been shown that a 5'Srglb3-gus-3'nos construct (Figure 2-5A, construct 6) is expressed exclusively in the infected cells of transgenic nodules (Szabados et al., 1990; Szczyglowski et al., 1994; see also Figure 2-6A). However, when the 3’ nos terminator was replaced by the SrEnod2 3’ region (Figure 2-5A, construct 7), GUS activity was detected not only in the infected cells of transgenic nodules, but also in the nodule parenchyma cells, the vascular tissue, and the outer cortex (Figures 2-6B and D). Dark-field microscopy revealed that GUS expression directed by the 5'Srglb3-gus-3'SrEnod2 construct could also be detected in the central uninfected (interstitial) cells of transgenic nodules (Figure 2-6D), but could not be detected in transgenic stem, leaf, and root tissues (Figure 2- 5). This expression pattern suggests an additive effect of the leghemoglobin promoter and the SrEnod2 3 ’ region. In addition, the average expression level of the 5'Srglb3-gus-3'SrEnod2 construct was consistently higher than the 5'Srglb3-gus-3'nos construct (Figure 2-5B). The 5'CaMV35S(-90/+6)-gus-3'nos construct (Figure 2-5A, construct 8) did not direct detectable levels of gus expression in either nodules or roots of transgenic plants (Figures 2-5, and 2-6E). However, when the 3’nos terminator was replaced by the 3’ SrEnod2 region (Figure 2-5A, construct 9), gus expression was detected in the nodule parenchyma, vascular 76 LEGEND FOR FIGURE 2-5. Structure of gus fusion constructs and determination of GUS activity in transgenic L. corniculatus plants. (A) Structure of gus fusion constructs. The construct numbers are listed on the left. The open box with angled lines represents the S. rostrata leghemoglobin gene promoter (5’Srglb3; 1.4 kb). The open box with horizontal lines represents the CaMV35S minimal promoter (CaMV35S; - 90 to +6 bp). The open box with vertical lines represents the nopaline synthase minimal promoter (Pnos; -150 to +1 bp). For other symbols used see the legend of Figure 2A. (B) Quantification of GUS activity in transgenic L. corniculatus plants. The data represent the mean values. The error bars represent the standard deviation of the mean. The construct numbers are listed at the bottom. 77 A GUS ACTIVITY (STAINING) Nodule ........... is? ,an CONSERUCTS on... "mg“ cortex ""°‘ 6 -l400bp m + - - - 7 -ltomip 4+ + + - s CaMV35S(-90/+6bp) - - - - 9 CaMV35S(—90/+6bp) : - + - + 10 Pnos(-l50/+lbp) I - _ - + 11 Pnos(-l50/+Ibp) I - + + +1- B 4000 80 E 3500 70 .71 3:22? J I, '0 L .mdule 5 3000 mi , 17* ’ ,, L: g 3 2500 so -. .2 E § E 2000 40 - U) E 1500 30 _ D O 1000 20 _ g 500 10- 5’ 0 0 - 78 bundles, and tissues surrounding the vascular bundle connecting the nodule to the root central cylinder (Figures 2-6F and G). gus expression was also detected in root vascular tissues, but not in stems or leaves of transgenic plants (data not shown). Fluorometric measurements confirmed this observation (Figure 2-5B). The 5'nos(-150/+1)-gus-3'nos construct (Figure 2-5A, Construct 10) did not direct detectable levels of gus expression in nodules, but was weakly expressed in the root tissues of transgenic plants (Figure 2-SB). Again, when the 3’ nos terminator was replaced by the 3’ SrEnod2 region (Figure 2-5A, construct 11), gus expression could be readily detected in the nodule parenchyma, vascular bundles, and outer cortex, as well as in root vascular tissues and cortical cells (Figures 2-6H and I). No staining was found in stems or leaves (data not shown). GUS activity measurements confirmed these observations (Figure 2-5B). These data reveal that the 3' region of the SrEnod2 gene can extend the expression of the gus reporter gene to the parenchyma and vascular tissues of transgenic plants, when fused to heterologous promoters, indicating that the SrEnod2 3' region contains sequence determinants necessary and sufficient for directing reporter gene expression in the nodule parenchyma and vascular tissues. Moreover, these data reveal that the SrEnod2 3' region neither inhibits reporter gene expression in those tissues where the respective 5' (promoter) regions are normally expressed, nor extends 79 LEGEND FOR FIGURE 2-6. Microscopic analysis of the expression of gus fusion constructs in transgenic L. corniculatus plants. Staining for GUS activity was carried out overnight, or for 1 hr as indicated. (A) Hand-cut section of a transgenic nodule harboring the 5’Srglb3-gus-3 ’nos construct (Figure 2-5A, construct 6). GUS activity (blue color) can be seen in the nodule central tissue (CT). Staining was carried out for 1 hr. (B) Hand-cut section of a transgenic nodule harboring the 5’Srglb3-gus-3 ’SrEnod2 construct (Figure 2-5A, construct 7). GUS activity (blue color) can be seen in the nodule central tissue, as well as in surrounding tissues. Staining was carried out for 1 hr. (C) Dark-field microscopy of the nodule section shown in (A). GUS activity (blue color) can be seen in the infected cells of the nodule central tissue. (D) Dark-field microscopy of the nodule section shown in (B). GUS activity (blue and red colors) can be seen in the nodule central tissue (CT), as well as in the nodule parenchyma (NP), vascular bundles (VB), and outer cortical tissue (0C). (E) Dark-field microscopy of a transgenic nodule section harboring the 5’CaMV35S(-90/+6)-gus-3’nos construct (Figure 2-5A, construct 8). (F) and (G) Dark-field microscopy of a proximal and a distal view of a transgenic nodule section, harboring the 5’CaMV3SS(-90/+6)-gus-3’SrEnodZ construct (Figure 2-5A, construct 9). GUS activity can be seen in the nodule parenchyma (NP), vascular tissue bundles (VB) and the root central cylinder (CC). No GUS activity is observed in the root cortex. (H) and (I) Dark-field microscopy of a proximal and a distal view of a transgenic nodule section harboring the 5’nos(-150/+1)-gus-3’SrEnod2 construct (Figure 2- 5A, construct 11). GUS activity can be seen in the nodule parenchyma (NP), vascular bundles (VB) and outer cortical tissue (OC), as well as in the root central cylinder (CC) and in the root cortex (RC). (J) Dark-field microscopy of a transgenic nodule section harboring the CaMV35S (-150/-90)-5'SrEnod2(-115/+23)-gus-3'SrEnod2 construct (Figure 2-8, construct 22). GUS activity can be seen in the nodule parenchyma (NP), the nodule vascular bundles (VB) and the nodule central tissue (CT). (K) and (L) Dark-field microscopy of nodule and root-sections harboring the CaMV35S(-150/-90)-5 'SrEnod2(-50/+23)-gus-3'SrEnod2 construct (Figure 2-8, construct 23). GUS activity can be seen in the nodule parenchyma (NP), vascular bundles (VB) and central tissue (CT), as well as in the root central cylinder (CC) and cortex (RC). Staining was carried out for 1 hr. 80 81 reporter gene expression to tissues where the 5’ (promoter) regions used are not expressed. Rather, the SrEnod2 3' region appears to enhance reporter gene expression in those tissues to which the 5’ regions direct expression. It is possible that the same regulatory element in the SrEnod2 3' region is responsible for both nodule parenchyma-specific expression and general enhancement of gene expression. Alternatively, separate mechanisms (cis- elements) may be responsible for the observed alterations in gene expression. Delimitation of sequences in the 5’ proximal region of the SrEnod2 gene required for detectable levels of gene expression A series of 5’93’ unidirectional deletions in the 5' SrEnod2 region was constructed and fused to the gus coding region, as well as to the 3' region of the SrEnod2 gene (Figure 2-7A). The resulting constructs were introduced into L. corniculatus plants and reporter gene expression was analyzed histochemically and quantitatively. GUS activity measurement revealed that constructs with deletions up to position -191, relative to the transcriptional start site, directed reporter gene expression in the same fashion as the full 5'SrEnod2-gus-3'SrEnod2 construct (Figure 2-2A, construct 1), namely to the nodule parenchyma and vascular tissues (Figure 2-7A and B). Only the level of GUS enzymatic activity varied slightly (Figure 2-7B). These data 82 LEGEND FOR FIGURE 2-7. Structure of SrEnod2 5' deletion reporter gene constructs and GUS activity in transgenic L. corniculatus plants. (A) Structure of gus fusion constructs. The construct numbers are listed on the left. The base pair (bp) numbers indicate the end point of the 5' deletion relative to the start of transcription. The vertical bars attached to the boxes represent the translational start of the corresponding genes. The interrupted lines downstream of the gus reporter gene represent the 2.4-kb 3' flanking sequences of the SrEnod2 gene. + and - represent positive and negative GUS staining activity, respectively. For other designations, see the legend of Figure 2A. (B) Quantification of GUS activity in transgenic L. corniculatus plants. The data represent the mean values. The error bar represents the standard deviation of the mean. The number of transgenic plants tested is indicated with an n. 83 A GUS ACTIVITY (STAINING) ........ @14— ads “3% was. CONSERUCT S dunes m tissues 1 - + - - 12 - + - - 13 - + - - 14 - + - - 15 - + - - l6 - + - - 17 - + - - 18 - + - - 19 .. + .. - 20 - + - - B T .noduic -—-t .root 4 Dstcm 15 $0111" ".9 ,6 _—I,—._—-- 3 l7 __— — r20 — i— .. ——‘ 19 .“ “_ n=12 20 _ — "=9 r 50 100 150 200 250 300 GUS activity (pmol MU/min/mg protein) 84 suggest that sequences up to position -191 are not required for nodule parenchyma-specific reporter gene expression. A deletion down to position -115 (Figure 2-7A, construct 20), was found to lead to a significant reduction of reporter gene expression in nodules (Figure 2-7B), although a limited degree of nodule enhanced expression was maintained. Histochemical GUS staining revealed that this construct retained gus gene expression in the nodule parenchyma and vascular tissues (Figure 2-7A; data not shown). Therefore, we suggest that the sequences down to position -115 may not be absolutely necessary for nodule parenchyma-specific expression. A construct with a deletion up to position -50 in the 5’ SrEnod2 region was found to be inactive in transgenic plants (Figure 2-8). To test whether the lack of reporter gene expression was due to a deletion of tissue-specific cis-element(s) or to a deletion of proximal promoter element(s), the transcriptional enhancer from the CaMV3SS promoter (-150 to -90 bp, part of domain B; Benfey et al., 1989) was firsed in cis to the -50 bp fiagment (Figure 2-8, construct 23). GUS activity measurement in transgenic plants harboring this construct revealed that the reporter gene was strongly expressed in nodule, root, and stem tissues (Figure 2-8). Within nodules, the reporter gene was highly expressed in the nodule parenchyma and vascular tissues, and to a lower level in the infected tissue (Figure 2-6K). Within roots, the gus gene was expressed in the central cylinder and cortical 85 LEGEND FOR FIGURE 2-8. Structure of SrEnod2 5' deletion reporter gene constructs, enhancer fusions and GUS activity in transgenic L. corniculatus plants. The construct numbers are listed on the left. The base pair (bp) numbers indicate the end point of the 5' deletion, relative to the start of transcription. The interrupted open box represents the 2.4-kb 3' flanking region of the SrEnod2 gene. The open box represents the transcriptional enhancer of the CaMV35S promoter (-150 to -90 bp). For other designations, see the legend of Figure 2A. 86 GUS ACTIVITY (STAINING) Nodule 1:333: parenchyma Root Stem CONSTRUCTS tissues and vascular tissues tissues # tissues mRNA 20%—1%:22 mRNA 21 ....1"”:’ua_.7@ - - - .- mRNA 22 D—wfil + + + + mRNA aim-Ea 4+ 4+ 4+ ++ 23 87 tissues (Figure 2-6L). When the same transcriptional enhancer was fused to the -115 bp deletion fragment of the 5’ SrEnod2 gene (Figure 2-8, construct 22), an identical GUS activity staining pattern was observed as with construct 23 (Figures 2-6J, and 2-8), although the GUS activity was much lower than plants harboring construct 23 (Figure 2-8; and data not shown). These data suggest that sequences between positions -115 and -50 in the 5’ SrEnod2 region are necessary for reporter gene expression. However, they may not contain any tissue-specific cis-elements. The 3’ untranslated region (3’ UTR) of the SrEnod2 gene constitutes the minimum determinant for nodule parenchyma-specific gene expression To delineate the sequence determinants in the 3' region of the SrEnod2 gene responsible for nodule parenchyma-specific expression, a series of 3’——)5’ unidirectional deletions in the 3' region of the SrEnod2 gene were constructed (Figure 2-9A). The 3' deletion derivatives were fused to the gus coding region, flanked by the full-length 5' SrEnod2 region. The resulting constructs were introduced into L. corniculatus, and reporter gene expression was determined. As shown in Figure 2-9, deletions in the 3’ region up to position +400 bp relative to the stop codon, did not affect tissue-specific expression of the reporter gene (Figure 2-9, construct 30). The level of gene expression was greatly reduced when the 3' flanking 88 LEGEND FOR FIGURE 2-9. Structure of SrEnod2 3' deletion reporter gene constructs and determination of GUS activity in transgenic L. corniculatus plants. (A) Structure of gas fusion constructs. The construct numbers are listed to the left. The base pair (bp) numbers indicate the end point of the 3' deletion, relative to the stop codon of the SrEnod2 gene. The vertical bars attached to the boxes represent the translational start of the corresponding genes. The vertical arrow indicates the position of the polyadenylation site of the SrEnod2 gene. The interrupted lines at the 5' end represent the 1.9-kb 5' SrEnod2 region. (B) Quantification of GUS activity in transgenic L. corniculatus plants. The data represent the mean values. The error bar represents the standard deviation of the mean. The number of transgenic plants tested is indicated with an n. 89 A GUS ACTIVITY (STAINING) "l E 3 S CONSTRUCTS central Pmnchym £3; “5’1?“ and vascular # tissues tissues tissues I ~ N 0.. I Hire 5 i3} 24 25 26 27 —1 28 29 may - l: ~ ‘ I 11001.. .. 30 I ++++++++ I I - m I map - ‘1: 31 1 I nodule I root [:1 ste m 25 27 28 29 0 100 200 300 400 500 600 GUS activity (pmol MU/min/mg protein) 90 sequences were further limited to position +241 bp, although the nodule parenchyma-specific expression pattern appeared to be maintained (Figure 2-9, construct 31; data not shown). Therefore, the 241-bp fragment immediately downstream of the stop codon of the SrEnod2 gene, containing most of the SrEnod2 3’ untranslated sequence (Dehio and de Bruijn, 1992; R. Chen and F. J. de Bruijn, unpublished data), appears to be sufficient to direct gene expression to the nodule parenchyma and vascular tissues. DISCUSSION The 3' UTR of SrEnod2 confers nodule parenchyma-specific expression The Enod2 genes of several legumes have been shown to be expressed in the nodule parenchyma cell layer (van de Wiel et al., 1990a, 1990b; Allen et al., 1991; Figure 2-1). Here we show that in the case of the tropical legume S. rostrata, the (primary) determinant of this cell-specific expression pattern is located in the 3' UTR of the SrEnod2 gene. The following lines of evidence support this conclusion: a) Chimeric gus reporter constructs flanked by both 5' and 3' regions of the SrEnod2 gene are expressed in the parenchyma of transgenic L. corniculatus nodules, while constructs containing a heterologous 3' region (e.g. 3' nos) are not (Figures 2-2, and 2- .3); b) The presence of the SrEnod2 3' region is necessary for the 91 accumulation of the gus mRNA in transgenic nodules (Figure 2-4); and c) The addition of the SrEnod2 3' region to chimeric reporter gene constructs containing heterologous promoters (e. g., Srglb3 promoter, CaMV35S minimal promoter, nopaline synthase minimal promoter) extends the expression pattern normally observed with the promoters to the nodule parenchyma (Figure 2-5). One interesting observation during our analysis of gus reporter gene expression was that all the gus constructs containing the SrEnod2 3' region directed gus expression also to the vascular bundles of transgenic L. corniculatus nodules. This observation contrasts with our in-situ hybridization results (Figure 2-1), which did not reveal SrEnod2 mRNA accumulation in the vascular bundles. It also contrasts with previous results on soybean Enod2 gene/promoter expression, which did not reveal reporter gene expression in the vascular tissues either (Lauridsen et al., 1993). This observation suggests that the gus expression assay is more sensitive than in- situ hybridization, because of the stability of GUS protein (Jefferson et al., 1987), and that the expression pattern of the S. rostrata Enod2 gene differs in this respect from the soybean homologue. The latter is quite plausible, since the S. rostrata Enod2 gene is inducible by cytokinin (Dehio and de Bruijn, 1992; Silver et al., 1996), whereas the soybean Enod2 gene is not. Moreover, in the case of the soybean Enod2 gene, nodule parenchyma- specific expression has been reported to be controlled by elements far 92 upstream in the 5' region of the gene (Lauridsen et al., 1993; also see below). Another interesting observation was made with the chimeric 5'Srglb3- gus-3'SrEnod2 construct. This construct was found to direct gus expression not only in the nodule parenchyma, vascular bundles, and infected cells of the central tissue, suggesting an additive effect of cell specificity mediated by the SrEnod2 3' region and the Srglb3 promoter (Szczyglowski et al., 1994), but also in the uninfected cells of the central tissue (Figure 2-6D). This unexpected expression pattern may be due to an unusual interaction of the cis-acting elements in these 5' and 3' regions. Alternatively, since the results shown in Figure 2-6 suggest that the SrEnod2 3' region appears to confer both (a) tissue specific, as well as (an) enhancer function(s), it may reflect an enhancement of Srglb3 promoter activity in the uninfected cells to a previously undetectable level (Szczyglowski et al., 1994). Our SrEnod2 3' deletion analysis revealed that the DNA sequences downstream of position +400 bp, relative to the SrEnod2 stop codon, can be omitted without affecting the expression pattern of chimeric reporter gene constructs (Figure 2-9). In fact, even a deletion construct lacking DNA sequences 3' to position +241 directs expression to the nodule parenchyma, while expression in roots and stems/leaves is virtually non-detectable (Figure 2-9). This suggests that sequences between the SrEnod2 stop codon - and position +241, upstream of the poly A addition site (around position 93 +258, Dehio and de Bruijn, 1992; R. Chen and F. J. de Bruijn, unpublished result), are capable of directing cell-specific expression in transgenic nodules, albeit at a low level. It is possible that the enhancer-like activity of the SrEnod2 3' region is mediated by sequences located between positions +241 and +400, whereas the cell-specificity determinant(s) lie(s) in the 3' UTR Further analysis will be required to support this hypothesis. DNA elements in the 3' region of plant genes have previously been implicated in regulation of gene expression. Examples include the potato proteinase inhibitor 11 (PI-II) gene (Thomburg et al., 1987; An et al., 1989), the petunia ribulose bisphosphate carboxylase small subunit gene (Dean et al., 1989), the oilseed rape AX92 gene (Dietrich et al., 1992), the Arabidopsis GLABROUSI (GL1) (Larkin et al., 1993), and the potato sucrose synthase (Sus4; Sus3) gene (Fu et al., 1995a and b). Multiple examples have also been reported in mammalian systems. Examples include the chicken adult ,B-globin gene (Choi et al., 1986), the chicken histone H5 gene (Trainor et al., 1987), the human A7 globin gene (Bodine and Ley, 1987), the mammalian ribonucleotide reductase R1 gene (Chen et al., 1993), and the human transferrin receptor (77R) gene (Owen and Kuhn 1987; Binder et al., 1994). In these cases, the 3’ sequences act either as transcriptional enhancers or sequence determinants modulating gene expression post-transcriptionally. Cis-element(s) acting as transcriptional enhancers are generally located downstream of poly(A) addition site(s) and 94 often act as tissue-specific and/or developmental regulatory elements (e.g., Arabidopsis GL1 gene; chicken adult B-globin gene; chicken histone H5 gene; human A 7 globin gene). In those cases where post-transcriptional regulation is involved, the 3’ sequences modulate mRNA stability (petunia rbcs genes; human T fR gene), or efficacy of mRNA 3’ end formation and processing (potato PI-II gene). Additionally, different 3' end regions have also been found to influence the level of gene expression in plant cells (Ingelbrecht et al., 1989). However, our finding that (a) cis-element(s) responsible for cell- specific expression is located within the 3' UTR appears to be a unique example amongst plant nuclear genes. The mechanism by which the SrEnod2 3' UTR mediates nodule parenchyma-specific expression remains to be elucidated. In a separate but related study, we have shown that cytokinin-mediated enhancement of SrEnod2 gene expression in S. rostrata roots, in the absence of rhizobial infection, appears to occur at a post- transcriptional level (Silver et al., 1996). This aspect of SrEnod2 regulation also involves the 3' UTR of the gene (D. Silver, R. Chen, and F .J . de Brujin, unpublished observations). Whether nodule parenchyma-specific expression and cytokinin up-regulation of the SrEnod2 gene share common regulatory elements is currently unknown; however, our data favor the hypothesis that both processes involve post-transcriptional regulation. Recently, cell- - specific mRNA stability elements have been identified in the 3' UTR of the 95 a2-globin gene (Weiss and Liebhaber, 1995). It is possible that the SrEnod2 3' UTR directs nodule parenchyma-specific expression in a similar way as the 3' UTR of the a2-globin gene. Alternatively, the SrEnod2 3' UTR may act as a tissue-specific transcriptional enhancer. Experiments are in progress to address these possibilities and to further delimit the responsible cis- elements. The 5' region of SrEnod2 does not appear to harbor significant tissue- specific cis-elements Our deletion analysis of the SrEnod2 5’ region revealed that sequences upstream of position -l91, relative to the transcriptional start site, are not required for nodule parenchyma-specific expression. A deletion up to the - 50 bp position abolishes reporter gene expression in nodules, but fusing a transcriptional enhancer (CaMVBSS -150 bp to -90 bp; see Szabados et al., 1990) to the -50 bp fragment restores reporter gene expression to the nodule parenchyma/vascular tissue, to stem and root tissues, and to a limited extent to the nodule central tissue (Figure 2-6K). This pattern of reporter gene expression was found to be similar to the pattern directed by a fusion of the same transcriptional enhancer to the -115 bp fragment of the SrEnod2 5’ region (Figure 2-6J). These data suggest that sequences between positions - - 115 and -50 are required for a significant level of reporter gene expression 96 in nodules. However, these sequences are not likely to contain (a) nodule tissue-specific cis-element(s). The latter conclusion is very intriguing in light of the fact that, in the case of the soybean Enod2 (B) gene, Lauridsen et a1. (1993), have reported that the 3-kb promoter region directed reporter gene expression to the parenchyma of L. corniculatus and Trifolium repens transgenic hairy root nodules. The 3' nos terminator was used in their study. Two critical tissue- specific cis-elements were identified in regions located between positions - 1792 and -1582, and between -380 and -53, relative to the transcriptional start site. When the upstream tissue-specific cis-element was fused to the promoter region of a leghemoglobin gene (lb), the resulting gene fusion was found to direct a low level of reporter gene expression in a scattered pattern in the nodule central area. These data appear to be contradictory to the data presented in this paper, and/or suggest fisrther that there are profound differences in the mechanisms regulating SrEnod2 and the soybean Enod2(B) gene expression (also see discussion above). Further support for this idea is the finding that DNA sequences in the promoter regions of the SrEnod2 and the soybean Enod2(B) genes share remarkably little homology (de Bruijn et al., 1994), in contrast to the late nodulin leghemoglobin (lb) promoters, which are highly similar (de Bruijn et al., 1990; de Bruijn and Schell, 1992; Szabados et al., 1990; Szczyglowski et al., 1994; Stougaard et ‘ al., 1990; Ramlov et al., 1993). Perhaps this discrepancy can be explained 97 once the nucleotide sequences of the cis-element(s) responsible for the nodule parenchyma-specific gene expression are identified and directly compared. In any case, the apparent involvement of the 3' region of the SrEnod2 gene, rather than its promoter (5') region, in nodule parenchyma-specific expression is unique in a variety of ways and opens up a whole new area of investigations on the mechanisms of plant gene expression in response to rhizobial infection. 98 LIST OF REFERENCES Allen, T., Raja, S., and Dunn, K., (1991). Cells expressing Enod2 show differential spatial organization during the development of alfalfa root nodules. Mol. Plant-Microbe Interact. 2, 139-146. An, G., Mitra, A., Choi, H.K., Costa, M.A., An, K., Thornburg, R.W., and Ryan, GA. (1989). Functional analysis of the 3’ control region of the potato wound-inducible proteinase inhibitor 11 gene. Plant Cell 1, l 1 5-122. Arora, N., Skoog, F., and Allen, O.N. (1959). Kinetin-induced pseudonodules on tobacco roots. Am. J. Bot. 46, 610-613. Averyhart-Fullard, V., Datta, K., and Marcus, A. (1988). A hydroxyproline-rich protein in the soybean cell wall. Proc. Natl. Acad. Sci. USA 85, 1082-1085. Benfey, P.N., Ren, L., and Chua, N.-H. (1989). The CaMV35S enhancer contains at least two domains which can confer developmental and tissue specific expression patterns. EMBO J. 8, 2195-2202. Binder, R., Horowitz, J.A., Basilion, J.P., Koeller, D.M., Klausner, R.D., and Harford, J.B. (1994). Evidence that the pathway of transferrin receptor mRNA degradation involves an endonucleolytic cleavage within the 3' UTR and does not involve poly(A) tail shortening. EMBO J. 13, 1969-1980. Bodine, D.M., and Ley, T.J. (1987). An enhancer element lies 3 ’ to the humane A7 globin gene. EMBO J. 6, 2997-3004. Bradford, M.M. (1976). A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye- binding. Anal. Biochem. 72, 248-254. Chen, F.Y., Amara F.M., and Wright J.A. (1993). Mammalian ribonucleotide reductase R1 mRNA stability under normal and phorbol 99 ester stimulating conditions: involvement of a cis-trans interaction at the 3' untranslated region. EMBO J. 12, 3977-3986. Choi, O.-R., and Engel, J.D. (1986). A 3' enhancer is required for temporal and tissue-specific transcriptional activation of the chicken adult B-globin gene. Nature 323, 731-734. Cooper, J.B., and Long, S.R. (1994). Morphogenetic rescue of Rhizobium meliloti nodulation mutants by trans-zeatin secretion. Plant Cell 6, 215- 225. Cox, K.H., and Goldberg, RB. (1988). Analysis of plant gene expression, In Plant Molecular Biology: A Practical Approach, Show, C.H. ed. (Oxford: IRL press), pp. 1-34. Dean, C., Favreau, M., Bond-Nutter, D., Bedhrook, J., and Dunsmuir, P. (1989). Sequences are required to activate a gene expressed in the root cortex of embryos and seedlings. Plant Cell 1, 201-208. De Block, M., and DeBrouwer, D. (1992). In-situ enzyme histochemistry on plastic embedded plant material: The development of an artifact-free B-glucuronidase assay. Plant J. 2, 261-266. De Block, M., and DeBrouwer, D. (1993). RNA-RNA in-situ hybridization using digoxigenin labeled probes: the use of high molecular weight polyvinylalcohol in the alkaline phosphatase indoxyl-nitroblue- tetrazolium reaction. Anal. Biochem. 215, 86-89. de Bruijn, F.J. (1989). The unusual symbiosis between the diazotrophic stem-nodulating bacterium Azorhizobium caulinodans ORS571 and its host, the tropical legume Sesbania rostrata. In Plant-Microbe Interactions, Molecular and Genetic Perspectives. Volume 111, T. Kosuge and E. Nester eds. (New York: McGraw Hill), pp. 457-504. de Bruijn, F.J., Szabados, L., and Schell, J. (1990). Chimeric genes and transgenic plants are used to study the regulation of genes involved in symbiotic plant-microbe interactions. Dev. Genet. 11, 182-196. 100 de Bruijn, F.J., and Schell, J. (1992). Regulation of plant genes specifically induced in developing and mature nitrogen-fixing nodules: cis-Acting elements and trans-acting factors. In Control of Plant Gene Expression, D.P.S. Verma, ed. (Boca Raton, FL: CRC Press), pp. 241- 258. de Bruijn, F.J., Chen, R.C., Fujimoto, S.Y., Pinaev, A., Silver, D., and Szczyglowski, K. (1994). Regulation of nodulin gene expression. Plant and Soil 161, 59-68. Dehio, C., and de Bruijn, F.J. (1992). The early nodulin gene SrEnod2 fi'om Sesbania rostrata is inducible by cytokinin. Plant J. 2, 117-128. Depicker, A., Stachel, S., Dhaese, P., Zambryski, P., and Goodman, H.M. (1982). Nopaline synthase: transcript mapping and DNA sequence. J. Mol. Appl. Genet. 1, 561-573. Dickstein, R.T., Bisseling, T., Reinhold, V.N. and Ausubel, EM. (1988). Expression of nodule-specific genes in alfalfa root nodules blocked at an early stage of development. Genes Dev. 2, 677-678. Dietrich, R.A., Radke, S.E., and Harada, JJ. (1992). Downstream DNA sequences are required to activate a gene expressed in the root cortex of embryos and seedlings. Plant Cell 4, 1371-1382. Dreyfus, B.L., and Dommergues, Y.R. (1981). Nitrogen-fixing nodules induced by Rhizobium on the stem of the tropical legume Sesbania rostrata. FEMS Microbiol. Lett. 10, 3 13-3 17. Ebert, P., Ha, S.B., and An, G. (1987). Identification of an essential upstream element in the nopaline synthase promoter by stable and transient assays. Proc. Natl. Acad. Sci. USA 84, 5745-5749. Farrell, L.B., and Beachy, RN. (1990). Manipulation of beta- glucuronidase for use as a reporter in vacuolar targeting studies. Plant Mol. Biol. 15, 821-825. lOl Finan, T.M., Hirsch, A.M., Leigh, J.A., Johansen, E., Kuldau, G.A., Deegan, S., Walker, G.C., and Signer, ER. (1985). Symbiotic mutants of Rhizobium meliloti that uncouple plant from bacterial differentiation. Cell 40, 869—877. Fisher, RE, and Long, S.R. (1992). Rhizobium-plant signal exchange. Nature 357, 655- 660. Franck, A., Guilley, H., Jonard, G., Richards, K., and Birth, L. (1980). Nucleotide sequence of cauliflower mosaic virus DNA. Cell 21, 285-294. Franssen, H.J., Nap, J.-P., Gloudemans, T., Stickema, W., van Dam, H., Govers, F., Louwerse, J., van Kammen, A., and Bisseling, T. (1987). Characterization of cDNA for nodulin-75 of soybean: a gene product involved in early stages of root nodule development. Proc. Natl. Acad. Sci. USA 84, 4495-4499. Fu, H., Kim, S.Y., and Park, W.D. (1995a). High-level tuber expression and sucrose inducibility of a potato Sus4 sucrose synthase gene requires 5' and 3' flanking sequences and the leader intron. Plant Cell 7, 1387- 1 394. Fu, H., Kim, S.Y., and Park, W.D. (1995b). A potato Sus3 sucrose synthase gene contains a context-dependent 3' element and a leader intron with both positive and negative tissue-specific effects. Plant Cell 7, 1395- 1403. Gloudemans, T., and Bisseling, T. (1989). Plant gene expression in early stages of Rhizobium-legume symbiosis. Plant Sci. 65, 1-4. Govers, F., Franssen, H.J., Pieterse, C., Wilmer, J., and Bisseling, T. (1990). Function and regulation of the early nodulin gene ENOD2. In Genetic Engineering of Crop Plants, G.W. Lycett and D.W. Grierson, eds (London: Butterworths), pp. 259-269. Hall, T.C., Ma, Y., Buchbinder, B.U., Pyne, J.W., Sun, S.M., and Bliss, F.A. (1978). Messenger RNA for G1 protein of French bean seeds: cell- 102 free translation and product characterization. Proc. Natl. Acad. Sci. USA 75, 3196-3200. Hirsch, A.M., Bhuvaneswari, T.V., Torrey, J.G., and Bisseling, T. (1989). Early nodulin genes are induced in alfalfa root outgrowths elicited by auxin transport inhibitors. Proc. Natl. Acad. Sci. USA 86, 1244-1248. Hirsch, A.M. (1992). Developmental biology of legume nodulation. New Phytol. 122, 211-237. Hirsch, A.M., Assad, S., Fang, Y., Wycoff, K., and Loebler, M. (1993). Molecular interactions during nodule development. In New Horizons in Nitrogen Fixation, R. Palacios, J. Moira and WE. Newton, eds (Dordrecht: Kluwer Academic Publishers). pp. 291-296. Ingelbrecht, I.L.W., Herman, L.M.F., Dekeyser, R.A., Van Montagu, M.C., and Depicker, A.G. (1989). Different 3’ end regions strongly influence the level of gene expression in plant cells. Plant Cell 1, 671- 680. Jefferson, RA. (1987). Assaying chimeric genes in plants: The GUS gene fusion system. Plant Mol. Biol. Rep. 5, 387-405. Jefferson, R.A., Kavanagh, T.A., and Bevan, M.W. (1987). GUS fusions: B-glucuronidase as sensitive and versatile gene fusion marker in higher plants. EMBO J. 6, 3901-3907. Kijne, J.W. (1992). The Rhizobium infection process. In Biological Nitrogen Fixation, G. Stacey, R.H. Burris, and H]. Evans, eds (New York: Chapman and Hall), pp. 349-398. Kunkel, T.A. (1985). Rapid and efficient site-specific mutagenesis without phenotypic selection. Proc. Natl. Acad. Sci. USA 82, 488-498. Larkin, J.C., Oppenheimer, D.G., Pollock, S., and Marks, M.D. (1993). Arabidopsis GLABROUSI gene requires downstream sequences for function. Plant Cell 5, 1739-1748. 103 Lauridsen, P., Franssen, H., Stougaard, J., Bisseling, T., and Marker, K.A. (1993). Conserved regulation of the soybean early nodulin ENOD2 gene promoter in determinate and indeterminate transgenic root nodules. Plant J. 3, 483-492. Lerouge, P., Roche, P., Faucher, C., Maillet, F., Truchet, G., Prome, J.C. and Denarie, J. (1990). Symbiotic host-specificity of Rhizobium meliloti is determined by a sulfated and acylated glucosamine oligosaccharide signal. Nature 344, 781-784. Libbenga, K.R., Van Iren, F., Bogers, R.J., and Schraag-Lamers, M.F. (1973). The role of hormones and gradients in the initiation of cortex proliferation and nodule formation in Pisum sativum L. Planta 114, 29- 39. Long, S.R. (1989). Rhizobium-legume nodulation. Life together in the underground. Cell 56, 203-214. Maniatis, T., Fritsch, E.F., and Sambrook, J. (1989). Molecular Cloning: A Laboratory Manual, 2nd ed. (Cold Spring Harbor, NY, USA: Cold Spring Harbor Laboratory Press). Mylona, P., Pawlowski, K., and Bisseling, T. (1995). Symbiotic nitrogen fixation. Plant Cell 7, 869-885. Nap, J.-P., and Bisseling, T. (1990). Developmental biology of a plant- prokaryote symbiosis: the legume root nodule. Science 250, 948-954. Ndoye, 1., de Billy, F., Vasse, J., Dreyfus, B., and Truchet, G. (1994). Root nodulation of Sesbania rostrata. J. Bacteriol. 176, 1060-1068. Odell, J.T., Nagy, F., and Chua, N.-H. (1985). Identification of DNA sequences required for activity of the cauliflower mosaic virus 35S promoter. Nature 313, 810-812. 104 Owen, D., and Kuhn, LC. (1987). Noncoding 3’ sequences of the transferrin receptor gene are required for mRNA regulation by iron. EMBO J. 6, 1287-1293. Pankhurst, C.E., Broughton, W.J., and Wieneke, U. (1983). Transfer of an indigenous plasmid of Rhizobium loti to other rhizobia and Agrobacterium tumefaciens. J. Gen. Microbiol. 129, 2535-2543. Pawlowski, K., Klosse, U., and de Bruijn, F.J. (1991). Characterization of a novel Azorhizobium caulinodans ORS571 two-component regulatory system, NtrY/NtrX, involved in nitrogen fixation and metabolism. Mol. ' Gen. Genet. 231, 124-138. Pietrzak, M., Shillito, R.D., Hohn, T., and Potrykus, I. (1986). Expression in plants of two bacterial antibiotic resistance genes after protoplast transformation with a new plant expression vector. Nucleic Acids Res. 14, 5857-5868 Ramlov, K.B., Laursen, N.B., Stougaard, J., and Marcker, K.A. (1993). Site-directed mutagenesis of the organ-specific element in the soybean leghemoglobin lbc3 gene promoter. Plant J. 4, 577-5 80. Sanders, P.R., Winter, J.A., Barnason, A.R., Rogers, S.G., and Fraley, RT. (1987). Comparison of cauliflower mosaic virus 35S and nopaline synthase promoters in transgenic plants. Nucleic Acids Res. 15, 1543- 1558. Silver, D., Pinaev, A., Chen, R, and de Bruijn, F.J. (1996). Post- transcriptional regulation of the Sesbania rostrata early nodulin gene SrEnod2 by cytokinin. Plant Physiol. 112, 559-567. Stougaard, J., Sandal, N.N., Gron, A., Kiihle, A., and Marcker, K.A. (1987). 5’-Analysis of the soybean leghemoglobin lbc3 gene: Regulatory elements required for promoter activity and organ specificity. EMBO J. 6, 3565-3569. Szabados, L., Ratet, P., Grunenberg, B., and de Bruijn, F.J. (1990). Functional analysis of the Sesbania rostrata leghemoglobin glb3 gene 105 5'-upstream region in transgenic Lotus corniculatus and Nicotiana tabacum plants. Plant Cell 2, 973-986. Szczyglowski, K., and Legocki, A.B. (1990). Isolation and nucleotide sequence of a cDNA clone encoding nodule-specific (hydroxy) proline- rich protein LEnodZ from yellow lupin. Plant Mol. Biol. 15, 361-363. Szczyglowski, K., Szabados, L., Fujimoto, S.F., Silver, D., and de Bruijn, F.J. (1994). Site-specific mutagenesis of the nodule-infected cell expression (NICE) element and the AT-rich element ATRE-BS2* of the Sesbania rostrata leghemoglobin glb3 promoter. Plant Cell 6, 317-332. Taylor, C.B., Bariola P.A., Del Cardayre S.B., Raines R.T., and Green P.J. (1993). RN82: A senescence-associated RNase of Arabidopsis that diverged fiom the S-RNases before speciation. Proc. Natl. Acad. Sci. USA 90: 5118-5122. Tempe, J., and Casse-Delbart, F. (1989). Plant gene vectors and genetic transformation: Agrobacterium Ri plasmids. In Cell Culture and Somatic Cell Genetics of Plants, Vol. 6, J. Schell and I.K. Vasil, eds. (San Diego, CA: Academic Press), pp. 25-49. Thimann, K.V. (193 6). On the physiology of the formation of nodules on legume roots. Proc. Natl. Acad. Sci. USA 22, 511-515. Thornburg, R.W., An, G., Cleveland, T.E., Johnson, R., and Ryan, CA. (1987). Wound-inducible expression of a potato inhibitor 11- chloramphenicol acetyltransferase gene fusion in transgenic tobacco plants. Proc. Natl. Acad. Sci. USA 84, 744-748. Tjepkema, J.D., and Yocum, CS. (1974). Measurement of oxygen partial pressure within soybean nodules by oxygen microelectrodes. Planta 119, 3 5 1-3 60. Trainor, C.D., Stamler, SJ., and Engel, J.D. (1987). Erythroid-specific transcription of the chicken histone H5 gene is directed by a 3 ’ enhancer. Nature 328, 827-830. 106 Truchet, G., Roche, P., Lerouge, P., Vasse, J., Camut, S., de Billy, F., Prome, J.C. and Denarie, J. (1991). Sulphated lipooligosaccharide signals from Rhizobium meliloti elicit root nodule organogenesis in alfalfa. Nature 351, 670-673. Tsien, H.C., Dreyfus, B.L., and Schmidt, EL. (1983). Initial stages in the morphogenesis of nitrogen-fixing stem nodules of Sesbania rostrata. J. Bacteriol. 156, 888-897. van de Wiel, C., Scheres, B., Franssen, H., van Lierop, M.J., van Lammeren, A., van Kammen, A., and Bisseling, T. (1990a). The early nodulin transcript ENOD2 is located in the nodule parenchyma (inner cortex) of pea and soybean root nodules. EMBO J. 9, 1-7. van de Wiel, C., Norris, H.J., Bochenek, B., Dickstein, R., Bisseling, T., and Hirsch, M.A. (1990b). Nodulin gene expression and ENODZ localization in effective, nitrogen-fixing and ineffective, bacteria-free nodules of alfalfa. Plant Cell 2, 1009- 1017. van Kammen, A. (1984). Suggested nomenclature for plant genes involved in nodulation and symbiosis. Plant Mol. Biol. Rep. 2, 43-45. Verma, D.P.S. and Delauney, AJ. (1988). Root nodule symbiosis: nodulins and nodulin genes. In Temporal and Spatial Regulation of Plant Genes, R.B. Goldberg, and D.P.S. Verma, eds (New York: Springer- Verlag), pp. 169-199. Verma, D.P.S. (1992). Signals in root nodule organogenesis and endocytosis of Rhizobium. Plant Cell 4, 373-3 82. Verwoerd, T.C., Dekker, B.M.M., and Hoekema, A. (1989). A small- scale procedure for the rapid isolation of plant RNAs. Nucleic Acids Res. 17, 2362. Vijn, I., das Neves, L., van Kammen, A., Franssen, H., and Bisseling, T. (1993). Nod factors and nodulation in plants. Science 260, 1764-1765. 107 Weiss, I.M., and Liebhaber, S.A. (1995). Erythroid cell-specific mRNA stability elements in the a2-globin 3' nontranslated region. Mol. Cell. Biol. 15, 2457-2465. Witty, J.F., Minchin, F.R., Skot, L., and Sheehy, J.E. (1986). Nitrogen fixation and oxygen in legume nodules. Oxford Surv. Plant Mol. Cell Biol. 3, 275-314. Yang, W.-C., de Blank, C., Meskiene, 1., Hirt, H., Bakker, J., van Kammen, A., Franssen, H., and Bisseling, T. (1994). Rhizobium Nod factors reactivate the cell cycle during infection and nodule primordium formation, but the cycle is only completed in primordium formation. Plant Cell 6, 1415-1426. CHAPTER 3 IDENTIFICATION OF CIS-ELEMENTS WITHIN THE 3' UNTRANSLATED REGION OF THE SrENODZ GENE AND INTERACTING PROTEINS 108 109 ABSTRACT Unlike some other plant nuclear genes, whose transcripts are polyadenylated at multiple sites yielding highly heterogeneous 3'-ends, SrEnod2 transcripts terminate at three consecutive nucleotide positions. Previously, we have shown that the 3' UTR of the SrEnod2 gene contains cis-elements responsible for nodule parenchyma-specific expression. Using deletions and truncations within the 3' UTR that are fused to the gus reporter gene in transgenic plants, we have delineated sequence motifs within the SrEnod2 3' UTR, deletion of which affects tissue-specific expression. Using a protein-RNA UV-cross linking approach, we have identified several nodule-specific proteins that interact with the RNA transcribed from the SrEnod2 3' UTR. Competition experiments have revealed that sequences between positions +1 and +163 nt downstream of the stop codon are efficient to compete out the observed protein-RNA interactions, suggesting that the cis-acting elements responsible for nodule parenchyma-specific expression co-localize with specific protein binding sequences. 110 INTRODUCTION The S. rostrata SrEnod2 gene, encoding a putative cell wall protein (Dehio and de Bruijn, 1992), and is expressed in the nodule parenchyma during early stages of both stem and root nodule development (Chen et al., 1996; Chapter 2), as well as in roots after cytokinin treatment (Dehio and de Bruijn, 1992; Silver et al., 1996). By examining gus reporter gene expression in transgenic L. corniculatus plants, cis-elements responsible for the nodule parenchyma-specific expression have been delimited to the 3' untranslated region (3' UTR) of the SrEnod2 gene (Chen et al., 1996). The same 3' UTR appears to confer cytokinin enhancement of the gus reporter gene expression in transgenic Arabidopsis (D. Silver, R. Chen, and F.J. de Bruijn, unpublished data). The presence of the SrEnod2 3' UTR is required for accumulation of the gus reporter transcript in nodules of transgenic plants, indicating that a post-transcriptional mechanism may be responsible for the observed nodule parenchyma-specific expression. Cytokinin up-regulation of SrEnod2 gene expression has been further characterized using nuclear run-off experiments and application of transcriptional and translational inhibitors. It has been shown that cytokinin up-regulation takes place in the cytoplasm, and occurs at a post- transcriptional level (Silver et al., 1996). This process requires concurrent protein synthesis, and appears to involve protein phosphorylation as well 111 (Silver et al., 1996). These data suggest that trans-acting protein factor(s) may be involved in both tissue-specific and cytokinin induced expression of the SrEnod2 gene by interacting with RNA sequences encoded by the SrEnod2 3' UTR. To test this hypothesis, we carried out RNA-protein UV-cross linking experiments which identified nodule-specific proteins that interact with RNA sequences encoded by the SrEnod2 3' UTR. Additionally, the effect of various deletions in and truncations of the 3' UTR of the SrEnod2 gene was examined in RNA-protein competition experiments, as well as in transgenic L. corniculatus plants. These analyses revealed that the cis-acting elements required for nodule parenchyma-specific expression co-localize with protein binding sites in the RNA sequences derived from the 3’ UTR of the SrEnod2 gene. METHODS Plant growth and nodulation Sesbania rostrata seeds were surface-sterilized and germinated on petri dishes in the dark at 28°C for one day, as described by Pawlowski et al., (1991). Seedlings were grown for three weeks in autoclaved soil mix (MetroMixzsand; 2: 1) in grth chambers with an 18-hr light (28°C) and 112 6-hr dark (22°C) cycle. Plants were inoculated with a two-day-old culture of Azorhizobium caulinodans ORSS71 (Dreyfus and Dommergues, 1981) on the stems for stem nodulation. Stem tissues and ten-day-old stem nodules were collected and processed (see below). Lotus corniculatus transgenic plants were grown for one week in autoclaved soil mix (VenniculitezsandzMetroMix; 10:10:1) in growth chambers with a 16-hr light (24°C) and 8-hr dark (18°C) cycle. Plants were inoculated with a three-day-old culture of Rhizobium loti NZP2037 (Pankhurst, 1983) and grown for 21 days. Plant tissues (nodules, infected roots, and leaves plus stems) were collected and assayed for GUS activity (see below). End-labeling and S1 nuclease protection assay The 2.4 Kb SacI-EcoRI fragment of the SrEnod2 3’ region was previously generated using a site-directed mutagenesis approach (see Methods of Chapter 2). The resulting plasmid was linearized by SacI digestion and the resulting fragment was end-labeled by T4 DNA polymerase, and Klenow in the presence of dCTP, dGTP, dTTP and a32P-dATP. The reaction was controlled in such a way that about 30 nucleotides at each end were labeled (Maniatis et al., 1989). The labeled fiagment was digested with HindIII, and the 2 kb SacI end-labeled fragment was purified and used as the S1 probe. 113 S] nuclease protection analysis was performed according to Binder et a1. (1994). Total RNA was isolated from ten-day-old stem nodules and ethanol precipitated together with 1x105 cpm (2x106 cpm/pmol) of the prepared S1 probe. Samples were resuspended in 40 ul of 80 % deionized forrnamide, 0.4 M NaCl, 0.04 M PIPES (pH 6.4) and 0.005 M EDTA, heated at 85°C for 10 min and annealed at 30°C overnight. Samples were diluted with 9 volumes of 0.4 M NaCl, 0.03 M sodium acetate (pH 4.8), 0.001 M ZnClz, 10 ug/ml of sonicated and denatured salmon sperm DNA and S1 nuclease (1000 units/ml; BRL), and incubated at 370C for 1 hr. The reaction was terminated by adding of EDTA to a final concentration of 0.015 M, extracted with phenol-chloroform and ethanol precipitated. Alkaline hydrolysis was performed by dissolving dried samples in 100 111 of 0.2 N NaOH and incubation at 42°C for 1 hr. Samples were neutralized with 20 ul of 1 M Tris-HCl (pH 7.4) and 0.2 N HCl, and ethanol precipitated with 10 u g of carrier yeast tRNA. Samples were washed in 70% ethanol and air dried. Samples were resuspended in 4 ul of loading dye, loaded on 6% polyacrylamide sequencing gels and autoradiographed. Mapping of 3’-ends of SrEnod2 mRNA by 3' RACE Mapping of the 3’-ends was carried out by the 3’ RACE (random amplification of cDNA ends) method described by Frohman et a1. (1988) 114 and Frohman (1990). Total RNA was reverse transcribed in 20 [.11 solution containing 50 mM Tris-HCl (pH 8.15), 6 mM MgC12, 40 mM KCl, 1 mM DTT, 1 mM dNTP, 0.5 ug of dT17-adapter primer (5'GACTCGAGTCGAC ATCGATI73'), 10 units of RNasin, 10 units of AMV reverse transcriptase, at 42°C for 1 hr. An aliquot of this reaction was used to PCR amplify the 3' ends of the SrEnod2 transcripts. The PCR reaction was carried out in a solution, containing 70 mM Tris-HCl (pH 8), 7 mM MgC12, 20 mM (NH4)2SO4, 10% DMSO, 1.5 mM dNTPs, 25 pmol adapter primer (5'GACT CGAGTCGACATCG3'), 25 pmol gene-specific primer (5'GAGCTCCAAC CACTACC3'), 1 ul of reverse transcription reaction mixture and 2.5 units of Taq polymerase. Construction of chimeric reporter gene fusion The 1.9-kb EcoRI-BamHI fragment, which contains the entire 5' untranslated upstream region of the SrEnod2 gene (see Methods of Chapter 2), was inserted in fi'ont of the gus coding region carried on plasmid pBIlOl.1 (Clontech Laboratories, Palo Alto, CA) to generate the pBIlOl.1- S’SrEnodZ vector. The EcoRI end of the fragment was filled in by Klenow before cloning. The 400 bp and the 241 bp deletions in the SrEnod2 3’ region were generated previously (see Methods of Chapter 2). To generate construct 32, the 400 bp deletion fragment (SacI-XhoI) was digested with 115 NheI. The resulting 163 bp-long SacI-Nhel fragment was inserted into the SacI-XbaI sites of pBluescript KS- vector (Stratagene, La Jolla, CA). A SacI-EcoRI fragment was excised from the resulting plasmid and inserted into the SacI-EcoRI sites of the pBIlOl.1-5’SrEnod2 vector. To generate construct 33, the 400 bp deletion fragment was digested with MboI. The resulting 360 bp MboI-Xhol fragment was inserted into the BamI-II-Xhol sites of pBS KS- vector. The resulting plasmid was digested with BstXI and SpeI, Klenow filled-in and religated to remove the polylinker sequence between the Sacl and BamHI sites. A SacI-Xhol fragment was then excised and inserted into the SacI-XhoI sites of pBIlOl.1-5’SrEnod2 vector. To generate construct 34, the 400 bp fragment was digested with Nhel, and the resulting 237 bp NheI-XhoI fragment was inserted into the XbaI-Xhol sites of pBS KS- vector. A SacI-Xhol fiagrnent was excised from the resulting plasmid and inserted into the SacI-Xhol sites of the pBIlOl.1-5’SrEnod2 vector. To generate construct 35, a 40 bp SacI-MboI subfragment of the 400 bp fragment was first inserted into the SacI-BamI-II sites of pBS KS- vector, a 237 bp NheI (blunt-ended by Klenow fill-in)-Xhol subfragment was subsequently inserted at the SmaI-Xhol sites of the resulting plasmid. A SacI-Xhol fragment was then excised and inserted into the SacI-XhoI sites of the pBI 1 01 . 1-5 ’SrEnodZ vector. 116 Regeneration of Transgenic Lotus corniculatus Plants Transgenic Lotus corniculatus cv Rodeo plants were generated according Szabados et a1. (1990). Binary vectors were conjugally transferred into the Agrobacterium rhizogenes strain A4 (Tempe and Casse-Delbart, 1989). Re generated plants were transferred to growth chambers (Conviron, Asheville, NC) and nodulated as described (Szabados et al., 1990). Histochemical Staining of GUS Enzymatic Activity GUS enzymatic activity was histochemically analyzed according to the procedure described by Jefferson (1987) with the following modifications. Hand-cut nodule sections were incubated at 37°C in a solution containing 50 mM NazPO4 and NaHPO4, pH 7, 1 mM ferri- ferro-cyanide and 20% methanol, and 1 mM X-Gluc (5-bromo-4-chloro-3-indolyl B-D—glucuronide; Gold Biotechnology Inc., St. Louis, MO). Stained sections were cleared through an ethanol and ethylene series and analyzed under a dissection microscope (Wild MPS52, Heerburgg, Switzerland). 117 RNA-protein UV-cross linking experiments Cellular protein extracts were prepared as following. Plant materials were grinded to a fine powder in liquid nitrogen and resuspended in an equal volume of 2x NP-40 extraction buffer (20 mM HEPES, pH 7.9; 80 mM KC]; 6 mM MgC12; 2 mM DTT; 10 % glycerol; 0.4 % NP-40; 16 ug/ml Aprotinin; 4 ug/ml Leupeptin; and 1 mM PMSF). After mixing, an equal volume of 1x NP-40 extraction buffer was added. The mixture was centrifuged at 14,000 rpm for 2 min. The supernatant was collected, and protein concentrations were determined according to Bradford (1976). Crude extracts were aliquotted and stored at -70°C until use. Plasmid pBS KS- DNA, containing the 241 bp 3’ UTR fragment was linearized at the X1101 or Sacl sites and used as template to produce sense or antisense RNA in in vitro transcription reactions using T7 or T3 RNA polymerase, respectively, according to the manufacturer’s instructions. Labeled RNA transcripts were produced by adding a32P-UTP (800 Ci/mmol) in the reaction, and the resulting transcripts had a specific activity of approximately 3x 107 cpm/11g. Cold RNAs were prepared essentially the same way except that cold UTP was used in the reactions. Poly(A), poly(C), poly(G), and poly(U) (Sigma, St. Louis, M0) were resuspended in 1x binding buffer and stored at -70°C until use. 118 Cellular protein extracts (10 ug) were incubated at room temperature for 30 min with 125,000 cpm 32P-labeled RNA in 1x buffer containing 40 mM KCl, 10 mM HEPES (pH 7.9), 3 mM MgC12, 1 mM dithiothreitol, and 5 % glycerol. The volume of each reaction was 20 ul. RNase-T1 (20 units; Boehringer Mannheim, Mannheim, Germany) was added to cleave unbound RNA, and Heperan sulfate (0.1 mg; Sigma) was added to reduce nonspecific binding. The reaction mixtures were UV-cross linked using 254 nm UV radiation (250 mJ/cmz) with a Stratagene UV cross-linking apparatus (Stratagene, La Jolla, CA) and then separated by electrophoresis on 10 % SDS-polyacrylamide gels under reducing conditions. Gels were dried and exposed to x-ray films at -70°C. RESULTS Polyadenylation sites of the SrEnod2 transcript Sequence analysis of SrEnod2 cDNA clones revealed that the SrEnod2 transcript was polyadenylated at the T residue at position +258 nt downstream of the TAA stop codon (Strittmatter, et al., 1989; Dehio and de Bruijn, 1992). Since multiple AATAAA-like sequences were present in the SrEnod2 3' UTR (Figure 3-2), we tested whether multiple polyadenylation events occurred using both 81 nuclease protection and 3' RACE (rapid 119 LEGEND FOR FIGURE 3-1. S1 nuclease protection assay. Total RNA of 10-day-old stem nodules of S. rostrata was mixed at the concentration indicated on top of each lane with 1x 105 cpm Sl probe prepared from the SrEnod2 3’ region. A single band of 255 nt was protected after S1 nuclease digestion. Bands of higher molecular weights were derived fi'om the double-stranded S1 probe. 120 0 10 20 50 150 ugRNA < 255 nt 121 amplification of cDNA ends; Frohman et al., 1988). For S1 nuclease protection assays, total RNA isolated from 10 day-old stem nodules of S. rostrata plants, hybridized with a 2 kb SacI-HindIII fragment of the SrEnod2 3' flanking region labeled at the SacI site. As shown in Figure 3-1, a single protected band of about 255 nt was detected after 81 nuclease digestion. Since poly(A) signal sequences are normally located 10 to 30 nt upstream of the poly(A) sites (Hunt, 1994), a single protected band suggests that the AATAAA-like sequences located between 20 and 30 nt upstream of the poly(A) site are likely to be the functional poly(A) signals (Figure 3-2). The polyadenylation sites were further analyzed using 3' RACE and found to correspond to the A or T at positions +256 and +257. Together with the T residue found at position +258, these three polyadenylation sites are very close to each other, indicating that they may be generated by the same polyadenylation process. Sequence motifs in the SrEnod2 3' UTR As shown previously (Chen et al., 1996; Chapter 2), the 241 bp fragment downstream of the SrEnod2 stop codon contains the minimal cis-element(s) responsible for nodule parenchyma-specific expression. To delineate sequence motifs within the SrEnod2 3' UTR that are important for nodule parenchyma-specific expression, motif deletions and truncations within the 122 CCGGGATACAATCCACCACCATATGGCCACTATCCACCGTCCAAAAAAAATTAA P G Y N P P P Y G H Y P P S K K N * CAACCACTACCACTACACGTGGCATGCATATTTTGGTGAICTAATAAATGTCAA GATTAGTTGTTTGTCATAGTAAAGTTTTTGTTTTCTCCCTCCTTCCCTTTAATI AATTACCACTTTTAAGATGCATGTGAACTAGCIAGCTACAATGTTCTCCTTCAA AATAAGGGCTCTATGCATGCCCCTCTCTCTGTAACCTTTGCTTTGCATTTCCCT tit TTATCTTGTAATATCAATCAATAAGEACTTTCTCGTTTTATTATCTATGCCTAC CATACAATTACATTCTTTATCATACTTACAAATTAATGATAAGCTAAACGGAAA ATCTAAACATACATAACACACAGTAGATACAATAATTGTAACACTGTTATGTTA GTACCAAAAAAAAAAGAGAGAT FIGURE 3-2. Sequence of the 3’ region of the SrEnod2 gene. A short 54 108 162 216 270 324 378 stretch of C-terminal amino acid sequences is shown. * indicates the stop codon. The underlined nucleotide sequences represent AATAAA-like sequences. Nucleotides in bold represent sequences recognized by the MboI, Nhel, and RsaI enzymes, from the 5’ to 3’direction, respectively. Arrows indicate the polyadenylation sites identified by using 3 ’ RACE and cDNA sequencing. 123 3’ UTR were generated. The resulting fragments were fused to the gus reporter gene and the full-length SrEnod2 5' flanking region (Figure 3-3). The chimeric reporter gene fusions were introduced into L. corniculatus plants and their GUS expression was histochemically analyzed. As summarized in Figure 3-3, a deletion from position +241 to position +163 maintains nodule parenchyma-specific expression (Construct 32), indicating that sequences between positions +1 and +163 are sufficient for tissue-specific expression. It should be noted that the potential poly(A) signal sequences between position +235 and +240 are deleted in this construct. However, other AATAAA-like sequences are present within this fragment. For examples, a canonical poly(A) signal sequence AATAAA is located at position +42, and two additional AATAAA-like sequences are present at positions +105 and +109, respectively (see Figure 3-2). It is likely that they may be recognized as alternative poly(A) signal sequences in order to form the 3'-end of the transcript. A construct with a deletion in the SrEnod2 3' region fi'om positions +1 to +38 still directed tissue-specific expression of the gus reporter gene (Figure 3-3, Construct 33). A construct with a deletion from +1 to +163 also directed nodule parenchyma-specific expression, but GUS activity could also be detected in the nodule-infected-tissues of some transgenic plants, indicating a less stringent tissue-specific control of gene expression (Figure 3-3, Construct 34). This data seems to be in contrast with the observation 124 LEGEND FOR FIGURE 3-3. Structure of gus fusion constructs and GUS expression in transgenic L. corniculatus. The gas reporter gene was flanked by the SrEnod2 5’ region and various 3’ deletion and truncation fragments. The nucleotide numbers indicate the end- points of the deletion and truncation fragments, relative to the stop codon. The arrows represent the poly(A) sites. + and - indicate positive and negative GUS activity staining, respectively. Constructs 30 and 31 were previously analyzed (Chen et al., 1996; Chapter 2). 125 5'SrEnod2 3'SrEnod2 ‘65—“— GUS ACTIVITY STAININ G CONSTRUCTS # \ Infected Nodule Root Other tissue parenchyma tissues 30 ' +400 I I. u n 31 +241 - + . . 32 +163 I + I I 33 +38 v +400 - -l- - . 34 +163 V +400 U- + . .. 35 __-_t-38 +163 V +400 * I. _ _ 126 that sequences between positions +1 and +163 are sufficient for tissue- specific expression (see above). However, it is possible that redundant cis- elements may exist in the SrEnod2 3’ UTR, and that deletion of some elements resulted in a slightly altered tissue-specific control. Support for this hypothesis came from the analysis of the expression pattern of a construct with a deletion of sequences between positions +38 and +163 (Figure 3-3, Construct 35), which was found to direct a similar reporter gene expression, namely in the nodule parenchyma, and in nodule central area of some transgenic plants. In constructs 33, 34 and 35, the potential poly(A) signals remain intact, suggesting that a normal polyadenylation process may take place. These results taken together suggest that sequences between positions +38 and +163 may be important for tissue-specific gene expression, and sequences between positions +163 and +400 may contain redundant cis-elements. Proteins from nodule extracts interact with the RNA encoded by the 3' UTR of the SrEnod2 gene The SrEnod2 gene is rapidly induced during early nodule development (Dehio and de Bruijn, 1992). To test whether proteins from nodule extracts could interact with the RNA encoded by the SrEnod2 3’ UTR, total soluble proteins prepared from stem cortical tissues and 10-day old stem nodule 127 tissues, were UV-cross linked to 32P-labeled sense and antisense RNAs generated by transcribing the 241 bp SrEnod2 3’ UTR (see Methods). As shown in Figure 3-4, several protein bands fi'om stern nodule extracts were found to be UV-cross linked to the sense RNA of the SrEnod2 3’ UTR (lane 2). Some of the protein bands could also be detected in the UV-cross linking reaction with proteins from stem cortical tissues in a prolonged exposure, however, the intensity of labeling was extremely low (lane 1). In UV-cross linking experiments using the antisense RNA probe, proteins from both stem and stem-nodule tissues were cross-linked to the RNA, and the labeling was found to be profuse and identical in stem- and stem-nodule extracts, indicating a non-specific interactions between the antisense RNA and proteins from these tissues (data not shown). Competition experiments using nonspecific competitors were carried out. As shown in Figure 3-4, UV-cross linking between stem-nodule proteins and the sense RNA probe was not affected by a 1000-fold excess of nonspecific competitors: poly(A) and poly(C) (lanes 3-5, and 6-8, respectively). The cross-linking was totally inhibited by a 1000-fold excess of poly(G) (Figure 3-4, lanes 9-11). Surprisingly, pre-incubation of the nodule protein extracts with small amounts of poly(G) shifted the UV-cross linked bands upwards, although a similar pattern was maintained (Figure 3- 4, lanes 9-11). Pre-incubation of nodule protein extracts with poly(U) 128 LEGEND FOR FIGURE 3-4. RNA-protein UV-cross linking and competition with non-specific competitors. Protein extracts (10 1.1g protein) of stem and 10-day-old stem nodule tissues were incubated with 125,000 cpm 32P-labeled RNA derived from the 241 nt- long SrEnod2 3’ UTR. Labeled proteins were visualized by autoradiography. Competition was performed by pre-incubation with 10-, 100-, and 1000-fold (w/w) non-specific competitors: poly(A), poly(C), poly(G), and poly(U), as indicated. S: stem; SN: stem nodule. 129 234567891011121314 LANE 1 S SN SN SN SN SN SN SN SN SN SN SN SN SN PROTEIN ++++++++++++++32P-3’UTR - .. I POLYIAI I I POLYEI I I POLYIGI I I POLYIUI I COMPETITOR MW(kDa) — 200 — 116 97.4 —66 —45 —31 130 inhibited the cross-linking reactions, but to a lesser extent (Figure 3-4, lanes 12-14). Competition experiments using specific competitors were also carried out. As shown in Figure 3-5, the 241-nt RNA derived from the SrEnod2 3’ UTR inhibited the cross-linking reactions (lanes 3-4; see Figure 3-3, Construct 31). The RNA transcribed from a fragment containing sequences between positions +38 and +400 of the SrEnod2 3’ region inhibited the cross-linking (lanes 5-6; see Figure 3-3, Construct 33). The RNA transcribed from the +163 to +400 fragment was found to be less efficient in the competition experiments (lanes 7-8; see Figure 3-3, Construct 34). RNA transcribed from the SrEnod2 3’ region (+1 to +400 bp) with a deletion of sequences between positions +38 and +163, or with a deletion of sequences between positions +163 to +400, inhibited the cross-linking reactions as well (lanes 7-8, and 9-10, respectively; see Figure 3-3, Constructs 35 and 32). These data suggest that the cross-linking between nodule proteins and the RNA derived from the SrEnod2 3’ UTR are mediated by sequences located between positions +1 and +163, since deletion of these sequences results in a RNA that is inefficient in competition. Furthermore, the presence of multiple labeled bands also suggests that multiple proteins may interact with the RNA transcribed from the 3’ UTR of the SrEnod2 gene. 131 LEGEND FOR FIGURE 3-5. RNA-protein UV-cross linking and competition with specific competitors. Protein extracts (10 ug protein) of stem and 10-day-old stem nodule tissues were incubated with 125,000 cpm 32P-labeled RNA derived from the 241 nt- long SrEnod2 3’ UTR. Labeled proteins were visualized by autoradiography. Competition was performed by pre-incubation with 10-, 100-fold (w/w) RNAs derived fi'om the deletion and truncation fragments of the SrEnod2 3 ’ UTR shown in Figure 3-3. S: stem; SN: stem nodule. 132 1 2 3 4 5 6 7 8 9 10 11 12 LANE S SN SN SN SN SN SN SN SN SN SN SN PROTEIN + + + + + + + + + + + + 32P-3’ UTR - - II] CE [233:] Ill 22:! COMPETITOR /l/l/I44 MW(kDa) — 200 — 116 97.4 ,—66 —45 —31 133 DISCUSSION SrEnod2 mRNA 3’-end formation S1 nuclease protection, 3’ RACE and cDNA sequencing experiments reveal that SrEnod2 transcripts appear to be polyadenylated at three closely- positioned sites. Three potential poly(A) signal sequences could be identified and located 20, 24, and 30 nt upstream of position +256, respectively. Several other AATAAA-like sequences were also found dispersed throughout the 3’ UTR region. During nodule development, these upstream AATAAA-like sequences may not be recognized as poly(A) signals, as evidenced by the lack of detection of poly(A) addition near these sites (Figure 3-1). However, a reporter gene construct with a deletion of these potential poly(A) signal sequences still directed reporter gene expression to the nodule parenchyma (Figure 3-3, Construct 32), suggesting that some upstream poly(A) signal sequences may be alternatively used in the absence of these potential signal sequences. Analysis of sequence motifs within the SrEnod2 3’ UTR Analysis of chimeric SrEnod2-gus reporter gene expression in transgenic plants indicates that sequences between positions +1 and +163 could direct 134 the gus reporter gene expression in the nodule parenchyma, suggesting that this sequence is sufficient to confer tissue-specific expression. This observation is supported by an altered reporter gene expression directed by constructs deleted of this sequence (Figure 3-3, Constructs 34 and 35). In these cases, reporter gene expression was found in nodule parenchyma and in the infected tissue of some transgenic plants. The reporter gene expression in the nodule central tissue is particularly interesting, because this suggests that the SrEnod2 5’ region may have an extremely weak nodule-specific activity. The presence of the nodule parenchyma-specific cis-elements within the 3’ UTR inhibits this weak activity in the infected tissue, and stimulates nodule parenchyma expression. Another interesting observation is that nodule parenchyma expression remains observable in transgenic plants expressing constructs deleted of sequences between positions +1 and +163. This suggests that either redundant cis-elements exist in regions 3’ to the position +163, or sequences between position +163 and +400 is a part of a large cis-element, which after truncation, is still able to direct nodule parenchyma expression. Proteins that interact with the RNA encoded by the SrEnod2 3' UTR Using in vitro protein-RNA UV-cross linking experiments, several proteins of different molecular weights interacting with the RNA transcribed from 135 the SrEnod2 3' UTR have been detected in nodule extracts. These interactions seem to be specific, since they can not be competed out by adding a 1000-fold excess of poly(A) and poly(C) homopolymers (Figure 3- 4), but can be competed out by adding a 1000-fold excess of poly(G), and to a less extent poly(U), indicating that the sequences that interact with these proteins are GU-rich. Pre-incubation of nodule protein extracts with poly(G) polymers shifted cross-linking bands upwards. The reason for that is not clear. Furthermore, presence of multiple proteins interacting with the RNA derived fiom the SrEnod2 3’ UTR suggests that they may interact with cis- elements within the SrEnod2 3’ UTR to form a complex. Competition experiments with specific competitor RNAs derived from the SrEnod2 3 ’ UTR suggests that sequences between positions +1 and +163 interact with nodule proteins. Thus, these in vitro data correlate nicely with the in vivo reporter gene expression patterns, suggesting that the sequences between positions +1 and +163, important for the nodule parenchyma expression, co-localize with the specific protein-binding activity. The significance of these interactions remains to be elucidated, but these findings open up possibilities for the isolation of tissue-specific trans-acting factors by screening cDNA expression libraries using riboprobes (For example, see Schumacher et al., 1995), or by using the yeast three—hybrid system developed by SenGupta et a1. (1996). l 36 LIST OF REFERENCES Binder, T., Horowitz, J., Basilion, J.P., Koeller, D.M., Klausner, R.D., and Harford, J.B. (1994). Evidence that the pathway of transferrin receptor mRNA degradation involves an endonucleolytic cleavage within the 3' UTR and does not involve poly(A) tail shortening. EMBO J. 13, 1969-1980. Bradford, M.M. (1976). A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye- binding. Anal. Biochem. 72, 248-254. Chen, R., Silver, D., and de Bruijn, F.J. The nodule parenchyma-specific expression of the Sesbania rostrata early nodulin gene SrEnod2 is mediated by its 3’ untranslated region (3’ UTR). (manuscript is submitted). Dreyfus, B.L., and Dommergues, Y.R. (1981). Nitrogen-fixing nodules induced by Rhizobium on the stem of the tropical legume Sesbania rostrata. FEMS Microbiol. Lett. 10, 313-317. Frohman, M.A. (1990). Race: rapid amplification of cDNA ends. In PCR Protocols: A guide to methods and applications. M.A. Innis, D.H., Gelfand, J .J ., Sninsky, and T.J. White, eds (Academic Press, Inc.), pp. 28-38. Forhman, M.A., Dush, M.K., and Martin, GR. (1988). Rapid production of full-length cDNAs from rare transcripts: amplification using a single gene-specific oligonucleotide primer. Proc. Natl. Acad. Sci. USA 85, 8998-9002. Hunt, A.G. (1994). Messenger RNA 3’ end formation in plants. Ann. Rev. Plant Physiol. Plant Mol. Biol. 45, 47-60. Jefferson, RA. (1987). Assaying chimeric genes in plants: The GUS gene fusion system. Plant Mol. Biol. Rep. 5, 387-405. 137 Pankhurst, C.E., Broughton, W.J., and Wienek, U. (1983). Transfer of an plasmid of Rhizobium loti to other rhizobia and Agrobacterium tumefaciens. J. Gen. Microbiol. 129, 2535-2543. Pawlowski, K., Klosse, U., and de Bruijn, F.J. (1991). Characterization of a novel Azorhizobium caulinodans ORS571 two-component regulatory system, NtrY/NtrX, involved in nitrogen fixation and metabolism. Mol. Gen. Genet. 231, 124-138. Schumacher, J.M., Lee, K., Edelhoff, S., and Braun, RE. (1995). Spnr, a murine RNA-binding protein that is localized to cytoplasmic microtubules. J. Cell Biol. 129, 1023-1032. SenGupta, D.J., Zhang, B., Kraemer, B., Pochart, P., Fields, S., and Wickens, M. (1996). A three-hybrid system to detect RNA-protein interactions in vivo. Proc. Natl. Acad. Sci. USA 93, 8496-8501. Silver, D., Pinaev, A., Chen, R, and de Bruijn, F.J. (1996). Post- transcriptional regulation of the Sesbania rostrata early nodulin gene SrEnod2 by cytokinin. Plant Physiol. 112, 559-567. Szabados, L., Ratet, P., Grunenberg, B., and de Bruijn, F.J. (1990). Functional analysis of the Sesbania rostrata leghemoglobin glb3 gene 5'- upstream region in transgenic Lotus corniculatus and Nicotiana tabacum plants. Plant Cell 2, 973-986. Tempe, J., and Casse-Delbart, F. (1989). Plant gene vectors and genetic transformation: Agrobacterium Ri plasmids. In Cell Culture and Somatic Cell Genetics of Plants, Vol. 6, J. Schell and I.K. Vasil, eds. (San Diego, CA: Academic Press), pp. 25-49. CHAPTER 4 THE LjENOD2 GENE IS INDUCIBLE BY ETHYLENE IN ROOTS OF LOTUS JAPONIC US 138 139 ABSTRACT The SrEnod2 gene of Sesbania rostrata is induced during early nodule development, and also in S. rostrata roots treated with the plant hormone cytokinin in the absence of rhizobial infection. However, since S. rostrata is not amenable for molecular genetic analysis, we decided to study cytokinin regulation of nodulin gene expression and its involvement in Lotus japom'cus, an emerging model legume. The LjEnodZ gene was isolated and found to be inducible in roots of L. japonicus, but not by cytokinin. Submergence of the roots in water for several hours was sufficient to induce Enod2 gene expression. Further analysis revealed that ethylene was the primary inducer. Application of ethylene inhibitors reduced the Enod2 gene expression in roots. Pretreatment with the protein synthesis inhibitor cycloheximide abolished ethylene induction of Enod2 gene expression, indicating that this process requires concurrent protein synthesis. 140 INTRODUCTION The Enod2 gene is an early nodulin gene, first isolated from soybean (Franssen et al., 1987). Its homologues have been isolated from other legumes, such as pea (van de Wiel et al., 1990a); alfalfa (van de Wiel et al., 1990b); lupine (Szczyglowski et al., 1992) and Sesbania rostrata (Dehio and de Bruijn, 1992). Expression of the Enod2 gene in “empty-nodules” induced on alfalfa roots by Rhizobium nodulation-deficient mutant strains (Nod’, van de Wiel et al., 1990b), as well as by strains deficient in exopolysaccharide synthesis (Exo'; Dickstein et al., 1988), indicates that the Enod2 gene product is not involved in the infection process per se, but rather in nodule morphogenesis. The Enod2 gene has also been found to be expressed in “pseudo- nodules” elicited by auxin transport inhibitors G-Iirsch et al., 1989), or by a Rhizobium nodulation deficient (Nod') strain carrying a cytokinin biosynthetic gene tzs (Cooper and Long, 1994) on alfalfa roots. Plant hormones, therefore, seem to be involved in early nodule development (Cooper and Long, 1994; for a review, see Hirsch, 1992), as well as in early nodulin gene expression. Indeed, Dehio and de Bruijn (1992) found that the Enod2 gene of Sesbania rostrata is induced specifically in roots after cytokinin treatment. Similar results were reported for alfalfa, although the 141 size of the cytokinin regulated transcript is larger than the corresponding transcript in nodules (Hirsch, et al., 1993). Although the cytokinin up-regulation of Enod2 gene expression in roots is not a common phenomenon in legume plants, the observation that the SrEnod2 gene is specifically induced in roots of S. rostrata by cytokinins does provide a model for studying the mechanism of cytokinin regulated gene expression, as well as cytokinin involvement in nodule development. We have recently shown that the cytokinin up-regulation of the SrEnod2 gene expression in roots of S. rostrata occurs at a post- transcriptional level, takes place mainly in the cytoplasm, requires protein synthesis, and appears to involve protein phosphorylation and dephosphorylation (Silver et al., 1996). To study the developmental and tissue-specific regulation of Enod2 gene expression, we have identified the 3' untranslated region of the SrEnod2 gene as the minimal DNA sequence determinant responsible for nodule parenchyma-specific expression (Chen et al., 1996; Chapter 2). Therefore, we were interested in determining whether the same DNA sequence was responsible for the cytokinin up-regulation of SrEnod2 gene expression. Unfortunately, we found that the Enod2 genes of L. corniculatus, the plant used for our previous reporter gene expression studies, were not induced by cytokinin treatment. Therefore, the cytokinin 142 effects on Enod2 gene expression could not be studied in L. corniculatus plants in parallel with the tissue-specific expression. In a search for other transformable legumes where the Enod2 gene was regulated in a developmental/tissue—specific manner, as well as by hormone treatment, we found L. japonicus. The LjEnodZ gene of L. japonicus was isolated and found to be induced early during nodule development, and to be induced in roots in the absence of rhizobial infection, but not by cytokinin treatment. Rather, we found that incubating of roots in water was sufficient to induce LjEnodZ gene expression. Here, we present experimental evidence in support of the notion that in L. japonicus roots, ethylene can induce the expression of the Enod2 gene. METHODS Plant growth and treatment Lotus japonicus Gifu B-129 seeds were surface sterilized and germinated according to Hansberg and Stougaard (1992). One-week-old seedlings were transferred into soil (vermiculite:sand, 6:1) and watered with B+D nutrient solution (Broughton and Dilworth, 1971) containing 0.5 mM KNO3. Plants were grown in a growth cabinet (Conviron, Asheville, NC) at a 18 hr light 22°C and 6 hr dark 18 °C cycle, with alight intensity of 250 J.J.E/sec/m2 and 143 70 % RH. Two to three weeks old seedlings were gently washed free of soil and incubated in distilled water, containing the appropriate chemicals (Sigma, St. Louis, MO). 6-benzylaminopurine, aminoethoxyvinylglycine, silver thiosulfate, and cycloheximide were used at concentrations of 2, 100, 200, and 100 nM, respectively. To prepare silver thiosulfate, equal volumes of 0.2 mM silver nitrate and 1.6 mM sodium thiosulfate were mixed before use, according to Veen (1979). For ethylene treatments, seedlings were directly removed fiom soil, placed between water-soaked Whatrnann papers (Schleicher & Schull, Keene, NH) and sealed in jars filled with either air or 20 ppm C2H4. Isolation of cDNA clones and DNA sequence analysis A nodule cDNA library of L. japonicus constructed using a oligo-dT primer in Lambda ZAPII vector (Stratagene Inc, La Jolla, CA), was kindly provided by Dr. Jens Stougaards, Aarhus, Denmark. Screening of cDNA clones was carried out according to standard procedures (Maniatis et al., 1989) and manufacturer’s instructions. DNA sequences were determined by using the Sequenase 2.0 kit (United States Biochemical Inc., Cleveland, OH). 144 Southern blot analysis Genomic DNA was isolated from leaves of L. japonicus using a CsCl ultracentrifugation protocol (Maniatis et al., 1989). 10 pg of DNA was digested with the BamHI, EcoRI, HindIII, SalI, XbaI, XhoI (Boehringer Mannheim, Mannheim, Germany), respectively, electrophorized in 0.8 % agarose gel. Gels were denatured and renatured according to standard methods (Maniatis et al., 1989) and blotted onto a Nylon membrane with a 0.45 um pore size (Hybond-N; Amersharn Life Science Inc., Arlington Heights, IL). Membranes were pre-hybridized and hybridized in 0.5 M phosphate buffer, pH 7.2, containing 7 % SDS and 1% BSA (Sigma) at 65°C, according to Church and Gilbert (1984). An or 32P-dATP-labeled DNA fragment containing the SrEnod2 coding region was generated by random priming (Boehringer Mannheim) and used as a probe. Filters were washed at 65°C in 2x ssc, 0.1 % SDS for 20 min; and 0.5x SSC, 0.1 % SDS for 20 min. Filters were exposed to X-ray films at -70°C overnight. Isolation of total RNA and northern blot analysis Total RNA was isolated according the hot phenol method of Verwoerd et al., (1989), with minor modifications (Chen et al., 1996; Chapter 2). Ten . micrograms of total RNA were electrophorized in 1 % (w/v) agarose gels in 145 1x MOPS buffer (20 mM MOPS, 1 mM EDTA, 5 mM sodium acetate, pH 7) containing 7% (v/v) formaldehyde. Gels were soaked in 10x SSC solution for 30 min and blotted onto 0.22 um nitrocellulose membranes (Nitro/Plus; MSI, Westborough, MA). Membranes were probed with an a 32P-dATP-labeled LjEnod2 cDNA fragment, generated by random priming (Boehringer Mannheim). Filters were re-probed with an HSP70 probe from Arabidopsis (Wu et al., 1988), an 18S rRNA DNA probe from rice (Takaiwa et al., 1984), or an ubiquitin probe from S. rostrata (kindly provided by M. Holsters, University of Gent, Belgium) as loading controls. All filters were washed at 65°C in 2x SSC, 0.1% SDS for 20 min; 0.5x SSC, 0.1% SDS for 20 min; and 0.2x SSC, 0.1% SDS for 20 min. The signals were quantified using Phosphorimager analysis (Model 400B, Molecular Dynamics, Sunnyvale, CA). RESULTS cDNA isolation and DNA sequence analysis The genomic organization of the Enod2 genes in L. japonicus was analyzed by Southern blotting, using a 32P-labeled DNA fragment containing the SrEnod2 coding region (Dehio and de Bruijn, 1992). As shown in Figure 4- ' l, a single hybridizing band was found to be present in the BamI-II, HindIII, 146 SalI and XbaI-digested genomic DNA, and two bands in the EcoRI-digested genomic DNA, suggesting that L. japonicus contains a single Enod2 gene. To isolate a LjEnodZ cDNA, a nodule cDNA library of L. japonicus was screened using the SrEnod2 probe. A plasmid with a 1 kb insert was isolated and its DNA sequence was determined. The deduced amino acid sequence from this partial cDNA was found to contain proline-rich motifs (Figure 4-2), resembling those present in the Enod2 proteins of other legumes (Franssen et al., 1987; van de Wiel et al., 1990a; Dehio and de Bruijn, 1992; Szczyglowski and Legocki, 1992). Northern blot analysis revealed that the size of the Enod2 transcript in L. japonicus nodules was about 4 kb (see below). Expression of the LjEnod2 gene occurs at an early stage of nodule development To examine the temporal expression of the LjEnodZ gene during L. japonicus nodule development, root and nodule tissues were harvested at different time after rhizobial infection. Total RNA was isolated and subjected to northern blot analysis, using the 32P-labeled partial LjEnodZ cDNA as a probe. As shown in Figure 4-3, expression of the LjEnodZ gene was induced between 7 and 10 days after infection, when nodule primordia ' were becoming visible on the roots. The expression was further enhanced to 147 LEGEND FOR FIGURE 4-1. Genomic organization of the Enod2 gene in Lotus japonicus plants. Genomic DNA of L. japonicus plants (10 ug/lane) was digested with BamI-II, EcoRI, I-IindIII, SalI, Xbal, and X1101, respectively, and separated by electrophoresis in a 0.8 % agarose gel. The blot was hybridized with a 3”'P-labeled DNA fragment containing the SrEnod2 coding region. A single hybridizing band was detected in digestions with all enzymes, except EcoRI, which gave two hybridizing bands. Molecular markers are indicated. 148 149 LEGEND FOR FIGURE 4-2. Nucleotide and the deduced amino acid sequences of a partial LjEnodZ cDNA clone. Number 1 is given to the first nucleotide at the 5’ end of the insert. A star (*) indicates the stop codon. The poly(A) tail is in bold. A single letter designation is given to each amino acid deduced. 61 121 181 241 301 361 421 481 541 601 661 721 781 841 901 150 GGCACGAGCCAACCTCCACCGTATGAGAACCCCCCACCAGTGCATCCGCCTCCTACACAT G T 8449 P P P Y E N P P P ‘V H P P P T H CAGAAGCCATCACCAGGATACCAACCGCCTCATGTGAAACCACCACATATTGCGCCACCA 9 K P 8 P G Y Q P P H V’ K P P H I .A P P CATGAAAAAACACCACCCAAATACCAACCACCTCATGAGAAGCCACCCTTTGAGAAGCCA H E K T P P K Y Q P P H E K P P F E K P CCACTTGAGAAACCACCCCATGAAAAGCCACCACCAGAGTACCAACCTCCTCATAAGAAG P L E K P P H E K P P P E Y Q P P H K K CCACCATCACAGTACCAACCACCACCAGAGTACGAACCTCCTCAGGAAAAACCACCACCT P P S 2 Y Q P P P E Y E P P Q E K P P P GTGTACCTACCCCCATATTGGAAACCACCACCTGCGTATCCACCTCCATATGAGAAACCA ‘V Y L P P Y ‘W’ K P P P A. Y P P P Y E K P CCACCAGAGTATGAACCGCCTCAAGAAAAACCACCACCAGTGTATTCGCCTCCATATGAA P P E Y E P P Q E K P P P ‘V Y 8 P P Y E AGGCCACCCACAGGGCATCCGACTCCTTACCCACCATTTTATCACCTACCACCTGTGTAT R P P T G H P T P Y P P F Y H L P P ‘V Y CACCCCCCTTACGAGAAGCCGCCACCACTATACCAACCTCCTCATGAGAAACCACCCATT H P P Y E K P P P L Y Q_ P P H E K P P E TACGAACCTCCCCATGAGAAACCACCCATTTATGAACCTCCTCACGAGAAACCACCACTT Y E P P H E K P P I Y E P P H E K P P L TATGAACCACCTCCAGGATACAACCCTCCACCATATGGTCACTATCCACCGTCTAAAAAT Y E P P H E K P P I Y E P P H E K P P L TAATGACAACTACTAGTGAAGAACGTGGCAATGTGGCATGCATGTTTTGGTTAACCAAAA * * GCAAGAGCAGTTCGTTTGTCACATAAATTTTCACTTTGTTTTTAATTTGATTCATGTTTA GGAAGCACTTTTAAGATGGGGAGTACAATACTCCACTTCTGCAATGGTTTGTAAAAATAA GGGCTCTCAATGCCTCTGTGTGTAATAATACCTTTCATTTCCGTTTCAAAAAAAAAAAAA AAAAA 60 120 180 240 300 360 420 480 540 600 660 720 780 840 900 151 LEGEND FOR FIGURE 4-3. The LjEnodZ gene is expressed early during nodule development. Total RNAs were isolated from roots and nodules of L. japonicus at different times after rhizobial infection, as indicated by the numbers on top of the lanes. Total RNA was also isolated from roots submerged in water for 8 hr (lane labeled “R”). Ten micrograms of RNA were loaded in each lane and the blot was hybridized with a 32P-labeled DNA probe prepared from the insert of the partial LjEnodZ cDNA clone. The same blot was stripped and re-hybridized with a 32P-labeled constitutively-expressed Hsp70 DNA probe to calibrate sample loading. 152 7 10 14 22 30 R Days <- LjEnod2 kw” “7.x“ Arte-ab 3!"?5553 4- Hsp 70 153 reach a plateau at day 14, when mature nodules had developed. The LjEnodZ transcript level remained the same up to 30 days after infection. Thus, the temporal expression pattern of the LjEnodZ gene in root-nodules of L. japonicus is similar to that of the S. rostrata SrEnod2 gene during stem-nodule development, and different from the SrEnod2 gene expression during root-nodule development, where it is expressed in a transient fashion (Dehio and de Bruijn, 1992). The LjEnodZ gene is induced in water-submerged roots of L. japonicus Since the Enod2 gene from S. rostrata had been reported to be induced in roots treated with cytokinin (Dehio and de Bruijn, 1992; Silver et al., 1996), we examined whether the LjEnodZ gene could also be induced in roots by cytokinin application. Two-week-old seedlings, with roots submerged in water containing 2 M of benzylaminopurine (BAP) or lacking this hormone, were grown for 8 hr. Total root RNA was isolated and subjected to northern blot analysis. The result is shown in Figure 4-4. The LjEnodZ transcript level was found to be greatly enhanced after 8 hr submergence in water (lanes 1 and 2). The LjEnodZ transcript level in roots submerged in 2 1.1M BAP solution was also enhanced, but to a lower level (Figure 4-4, lane 3). These data indicate that submergence in water alone is sufficient to ' activate LjEnodZ gene expression in roots of L. japonicus. This finding 154 LEGEND FOR FIGURE 4-4. LjEnod2 gene expression in roots of L. japonicus plants treated with cytokinin. (A) Northern blot analysis. Roots of two-weeks old seedlings were submerged in water solution containing 0 (lane 2) or 2 uM benzylaminopurine (BAP; lane 3) for 8 hr. Total RNAs were isolated and subjected to northern blot analysis with a 32P-labeled LjEnod2 probe. The primary LjEnod2 transcript is indicated with an arrowhead. The same blot was re-hybridized with a 32P-labeled DNA probe encoding for the 188 rRNA to calibrate sample loading. The position of the ribosomal RNAs is also indicated. (B) LjEnod2 transcript levels relative to the 18S rRNA level. Relative LjEnod2 transcript levels were calibrated with the 18S rRNA levels using Phosphorimager analysis. The RNA level in the root sample at time 0 was arbitrarily assigned the value 1. 155 a m m 4 4 18S rRNA 0000000 654321 SAME mwgmtomgb NESMQ 156 contrasts with the observation made with the SrEnod2 gene, which has been shown to be specifically induced by cytokinin and not by submergence of roots in water (Dehio and de Bruijn, 1992). In fact, cytokinin does not only fail to induce the LjEnod2 gene expression, but has a slightly negative effect on the level of the expression induced by water treatment (Figure 4-4, lanes 2 and 3). To test whether the Enod2 gene expression in roots was specific to submergence in water or to other side effects, such as wounding resulting from removing plants from soil, we tested the Enod2 gene expression in seedlings grown undisturbed in pots. Additionally, to test whether the induction of the Enod2 gene expression in roots required whole seedlings, we tested the Enod2 gene expression in excised roots. As shown in Figure 4-5, the LjEnod2 transcript in roots of undisturbed plants could be induced by water flooding (lanes 1, 4 and 7), indicating that water treatment alone sufficiently induced the LjEnod2 gene expression. However, the induction was found to be smaller than the corresponding induction after submerging roots of plants removed from soil (compare lanes 3 and 4; 6 and 7), suggesting that other effects, such as wounding could enhance this induction. The LjEnod2 transcript was also found to be induced in roots excised from the plants, indicating that the induction did not need whole seedlings (lanes 1, 2 and 5) 157 LEGEND FOR FIGURE 4-5. LjEnod2 gene expression in roots of L. japonicus plants. (A) Northern blot analysis. Three-weeks old seedlings were subjected to various treatments, such as excision of roots (R), removal fiom soil (S), before roots were submerged in water, or undisturbed in soil and pots being submerged (S*), for various time, as indicated at the bottom of panel B. Total RNA was isolated and subjected to northern blot analysis with a 32P- labeled LjEnod2 probe. The same blot was re-hybridized with a 32P-labeled DNA probe encoding the 18S rRNA to calibrate sample loading. The position of the ribosomal RNAs is also indicated. (B) LjEnod2 transcript levels relative to the 18S rRNA level. Relative LjEnod2 transcript levels were calibrated with the 188 rRNA levels using phosphorimager analysis. The RNA level in the root sample at time 0 was arbitrarily assigned the value 1. 158 a m m 4 25$— 18S— < 18S rRNA B 0, 5 0 5 0 5 0 3 2 2 1 1 <2”: £3888: mpeema treatment S* 159 The LjEnod2 gene is induced by ethylene in roots of L. japonicus It has been observed that over-watering of L. japonicus plants causes accumulation of anthocyanines in the stems, indicating a stress response. It is also known that ethylene production in plants is induced by a variety of environmental factors, including wounding, anaerobiosis, chilling and water stress (Yang and Hoffman 1984; Abeles et al., 1992). To test whether ethylene was involved in the response of LjEnod2 gene expression to water treatment, we determined LjEnod2 transcript levels after the application of ethylene. Three-weeks old seedlings were removed from soil and placed between two layers of wet Whatmann paper and incubated in sealed jars containing air or ethylene (20 ppm) for 8 hr. Total RNA was isolated and subjected to northern analysis. As shown in Figure 4-6, the ethylene treatment induced the LjEnod2 transcript level to approximately 3.5- and 29-fold, as compared to the control treatment and the initial level, respectively (compare lanes 3 and 4, and 1 and 4). In the control treatment, the LjEnod2 transcript level was increased from the initial level to approximately 8-fold (Figure 4-6, lanes 1 and 3). In the same experiment, the LjEnod2 transcript level was found to be induced from the initial level to approximately 36-fold by water submergence (Figure 4-6, lanes 1 and 2). Since wounding induces ethylene production (Yang and Hoffman 1984; Abeles et al., 1992), it is likely that the 8-fold induction in control samples 160 LEGEND FOR FIGURE 4-6. The LjEnod2 gene expression in roots of L. japonicus plants upon ethylene treatment. (A) Northern blot analysis. Three-weeks old seedlings were gently removed from soil and sealed in jars filled with either air or 20 ppm ethylene for 8 hr, as indicated at the bottom of panel B. Total RNAs were isolated and subjected to northern blot analysis with a 32P-labeled LjEnod2 probe. The same blot was re-hybridized with a 32P-labeled DNA probe encoding an ubiquitin mRNA to calibrate sample loading. (B) LjEnod2 transcript levels relative to the ubiquitin mRNA level. Relative LjEnod2 transcript levels were calibrated with the ubiquitin RNA levels using phosphorimager analysis. The RNA level in the root sample at time 0 was arbitrarily assigned the value 1. 161 LjEnod2 iquitin mRNA m . ... ‘ B wwmm flag gm 3%.: «#4 % 0505050 433221150 aromas Easemeeemfi hr treatment air C2H4 162 (air) is due to the action of ethylene produced in response to wounding. Taking this consideration into account, the ethylene treatment appears to induce the LjEnod2 gene expression in roots of L. japonicus to a similar level as the water submergence treatment (Figure 4-6, lanes 2 and 4). Inhibitors of ethylene biosynthesis and action reduce the induction of the LjEnod2 transcript level Since it appeared that ethylene could induce the LjEnod2 gene expression in roots of L. japonicus, we examined whether application of inhibitors for ethylene biosynthesis and action could reduce the ethylene mediated induction of LjEnod2 gene expression. Three-week old seedlings were incubated in water containing 100 pM aminoethoxyvinylglycine (AVG) and 0.2 mM silver thiosulfate (Ag2S203), respectively. AVG is an inhibitor of ethylene biosynthesis, since it inhibits specifically ACC synthase, a rate- limiting enzyme for ethylene synthesis (Yu and Yang, 1979). Silver ion is an inhibitor of ethylene action (Beyer, 1976; 1979).Total RNA was isolated and subjected to northern blot analysis. As shown in Figure 4-7, the LjEnod2 transcript level remained the same after 4 hr treatment with either AVG or Ag28203 (compare lanes 2, 3, and 4), although the ethylene production was inhibited by these treatments (data not shown). However, ' after 8 hr treatment with the inhibitors, the level of the LjEnod2 transcript 163 LEGEND FOR FIGURE 4-7. LjEnod2 gene expression in roots of L. japonicus plants treated with ethylene inhibitors. (A) Northern blot analysis. Three-weeks old seedlings were treated with 100 M AVG or 200 M silver thiosulfate for various time, as indicated at the bottom of panel B. Total RNAs were isolated and subjected to northern blot analysis with a 32P-labeled LjEnod2 probe. The same blot was re-hybridized with a 32P-labeled DNA probe encoding an ubiquitin mRNA to calibrate sample loading. The position of the ribosomal RNAs is also indicated. (B) LjEnod2 transcript levels relative to the ubiquitin mRNA level. Relative LjEnod2 transcript levels were calibrated with the ubiquitin RNA levels using phosphorimager analysis. The RNA level in the root sample at time 0 was arbitrarily assigned the value 1. transcript tin 1qu1 LjEnod2/ub 164 AVG Ag+ VG 4 LjEnod2 4 Ubiquitin mRNA hr treatment 165 was reduced to 76 % and 52 % by the AVG and Ag+ treatment, respectively (compare lanes 5, 6, and 7). Silver ion treatment was found to be more effective than the AVG treatment, consistent with their mode of actions. Cycloheximide inhibits the LjEnod2 gene expression in roots To test whether ethylene could enhance the LjEnod2 transcript accumulation in the absence of translation, we monitored the LjEnod2 transcript levels in roots after a pretreatment with a translational inhibitor cycloheximide. Three-week-old seedlings were pretreated for 1 hr with 100 M cycloheximide, a concentration known to be effective for several plants (Silver et al., 1996). The seedlings were then washed extensively and incubated in water for 3 and 7 hr, respectively. Messenger RNA levels were determined by northern blot analysis. As shown in Figure 4-8, pretreatment with cycloheximide inhibited the accumulation of the LjEnod2 transcript after subsequent incubation in water up to 7 hr (lanes 4 and 5), suggesting that de novo protein synthesis may be involved in the ethylene stimulation of the LjEnod2 transcript accumulation. 166 LEGEND FOR FIGURE 4-8. LjEnod2 gene expression in roots of L. japonicus plants treated with the protein synthesis inhibitor cycloheximide (A) Northern blot analysis. Three-weeks old seedlings were pretreated with 100 uM CHX for 1 hr, and roots were submerged in water for additional 3 and 7 hr, as indicated as 4 and 8 hr CHX, respectively, at the bottom of panel B. Total RNAs were isolated and subjected to northern blot analysis with a 32P-labeled LjEnod2 probe. The same blot was re-hybridized with a 32P—labeled DNA probe encoding an ubiquitin mRNA to calibrate sample loading. The position of the ribosomal RNAs is also indicated. (B) LjEnod2 transcript levels relative to the ubiquitin mRNA level. Relative LjEnod2 transcript levels were calibrated with the ubiquitin RNA levels using phosphorimager analysis. The RNA level in the root sample at time 0 was arbitrarily assigned the value 1. 167 = 4 LjEnod2 4 Ubiquitin mRNA 18S B 505050 2211 £888 Easemeeemb CHX + + 168 DISCUSSION The LjEnod2 gene is highly conserved and expressed early during nodule development Our analysis of the partial LjEnod2 cDNA confirms that the coding region of Enod2 genes is highly conserved among legume plants. However, the length of Enod2 transcripts varies significantly. The LjEnod2 gene encodes a transcript of about 4 kb; whereas for example, the SrEnod2 transcript is about 1.4 kb. This is consistent with the notion that the nature of the repeating amino acid motifs, encoded by Enod2 genes, are functionally more important than their length (van de Wiel et al., 1990a). Like the Enod2 genes isolated from other legume plants, the LjEnod2 gene is induced at an early stage of nodule development. Unlike the SrEnod2 gene expression in root nodules of S. rostrata, which has been shown to be expressed in a transient fashion, the LjEnod2 transcript level reaches a plateau at 14 days after nodulation, and remains at the same level up to 30 days. The LjEnod2 gene is induced by ethylene in roots That ethylene appears to induce LjEnod2 gene expression is surprising, but interesting. Ethylene has previously been implicated in nodulation, since 169 certain nodulation mutant phenotypes can be restored to wild-type by adding inhibitors of ethylene biosynthesis and action. For example, when the sym5 pea mutant which nodulates poorly at 20°C, is treated with AVG and Ag, the number of nodules formed is increased (F earn and LaRue, 1991). It is also known that nitrate inhibits nodulation. This process may involve ethylene as well, since application of AVG releases nitrate inhibition of nodulation (Ligero et al., 1991). Moreover, the “thick and short roots” (Tsr) phenotype of Vicia sativa subsp. nigra, induced by its microsymbiont Rhizobium leguminosarum bv Viciae, can be suppressed by the ethylene inhibitor AVG (Zaat et al., 1989). Unlike other hormones, such as auxin and cytokinin, which have stimulating effects on nodulation, ethylene generally appears to have negative effects on early stages of nodule development (Lee and LaRue, 1992). Even though the mechanism and significance of LjEnod2 gene induction by ethylene remains to be elucidated. It is not known at present why the expression of Enod2 genes from S. rostrata, L. japonicus, and other legumes such as soybean are regulated differently in roots, one by cytokinin and one by ethylene, while many others are not induced by these hormones. Clearly certain cytokinin effects on plant development can be coupled to ethylene responses. For examples, the observed cytokinin inhibition of root and hypocotyl elongation in dark-grown Arabidopsis seedlings mimics the plant response to ethylene (Cary et al., 1995). Application of the ethylene 170 inhibitor AVG partially releases the cytokinin inhibition (Cary et al., 1995). Furthermore, the coupling between cytokinin and ethylene mediated responses is evident in the cytokinin-resistant mutant ckrI, which is resistant to ethylene, and is allelic to the ethylene-resistant mutant ein2 (Cary et al., 1995). Specific induction of Enod2 genes by cytokinin versus ethylene may not involve a direct coupling between cytokinin and ethylene action. However, the two signal transduction pathways underlying their mode of action may overlap. The observation that the induction of LjEnod2 gene expression by ethylene is inhibited by pretreatment with the protein synthesis inhibitor cycloheximide, suggests that de novo protein synthesis may be required for the ethylene mediated induction. In this respect, LjEnod2 induction by ethylene is similar to the cytokinin up-regulation of the SrEnod2 gene expression (Silver et al., 1996). The LjEnod2 gene is the first nodulin gene that has been shown to be ethylene-inducible. This may provide a tool to study ethylene involvement in nodule development and a tool to study regulation of other nodulin genes by ethylene. 1 7 1 LIST OF REFERENCES Abeles, F.B, Morgan, P.W., and Saltveit, Jr. ME. (1992). Ethylene in plant biology. Academic Press, New York. Beyer, J r., EM. (1976). A potent inhibitor of ethylene action in plants. Plant Physiol. 58, 268-271. Beyer, J r., EM. (1979). Effect of silver ion, carbon dioxide, and oxygen on ethylene action and metabolism. Plant Physiol. 63, 169-173. Broughton, W.J., and Dilworth, M.Y. (1971). Control of leghemoglobin synthesis in snake beans. Biochem. J. 125, 1075-1080. Cary, AJ., Liu, W., and Howell, SH. (1995). Cytokinin action is coupled to ethylene in its effects on the inhibition of root and hypocotyl elongation in Arabidopsis thaliana seedlings. Plant Physiol. 107, 1075- 1082. Chen, R., Silver, D., and de Bruijn, F.J. The nodule parenchyma-specific expression of the Sesbania rostrata early nodulin gene SrEnod2 is mediated by its 3’ untranslated region (3’ UTR). (manuscript is submitted). Church, G.M., and Gilbert, W. (1984). Genomic sequencing. Proc. Natl. Acad. Sci. USA 81, 1991-1995. Cooper, J.B., and Long, S.R. (1994). Morphogenetic rescue of Rhizobium meliloti nodulation mutants by trans-zeatin secretion. Plant Cell 6, 215- 225. Dehio, C., and de Bruijn, F.J. (1992). The early nodulin gene SrEnod2 fi'om Sesbania rostrata is inducible by cytokinin. Plant J. 2, 117-128. Dickstein, R.T., Bisseling, T., Reinhold, V.N. and Ausubel, RM. (1988). Expression of nodule-specific genes in alfalfa root nodules blocked at an early stage of development. Genes Dev. 2, 677-67 8. 172 Fearn, J.C., and LaRue, T.A. (1991). Ethylene inhibitors restore nodulation to sym5 mutants of Pisum sativum L. cv Sparkle. Plant Physiol. 96, 239-244. Franssen, H.J., Nap, J.-P., Goudemans, T., Stiekema, W., van Dam, H., Govers, F., Louweres, J., van Kammen, A., and Bisseling, T. (1987). Characterization of cDNA for nodulin-75 of soybean: a gene product involved in early stages of root nodule development. Proc. Natl. Acad. Sci. USA 84, 4495-4499. ' Hansberg, K., and Stougaard, J. (1992). Lotus japonicus, an autogamous, diploid legume species for calssical and molecular genetics. Plant J. 2, 487-496. Hirsch, A.M., Bhuvaneswari, T.V., Torrey, J.G., and Bisseling, T. (1989). Early nodulin genes are induced in alfalfa root outgrowths elicited by auxin transport inhibitors. Proc. Natl. Acad. Sci. USA 86, 1244-1248. Hirsch, A.M. (1992). Developmental biology of legume nodulation. New Phytol. 122, 211-237. Hirsch, A.M., Assad, S., Fang, Y., Wycoff, K., and Loebler, M. (1993). Molecular interactions during nodule development. In New Horizons in Nitrogen Fixation, R. Palacios, J. Moira and WE. Newton, eds. (Kluwer Academic Publishers, Dordrecht). pp. 291-296. Lee, K.H., and LaRue, T.A. (1992). Exogenous ethylene inhibits nodulation of Pisum sativum L. cv Sparkle. Plant Physiol. 100, 1759- 1763. Ligero, F., Caba J.M., Lluch, C., and Olivares, J.(1991). Nitrate inhibition of nodulation can be overcome by the ethylene inhibitor aminoethoxyvinylglycine. Plant Physiol. 97, 1221-1225. 173 Maniatis, T., Fritsch, E.F., and Sambrook, J. (1989). Molecular Cloning: A Laboratory Manual, 2nd ed. (Cold Spring Harbor, NY, USA: Cold Spring Harbor Laboratory Press). Silver, D., Pinaev, A., Chen, R, and de Bruijn, F.J. (1996). Post- transcriptional regulation of the Sesbania rostrata early nodulin gene SrEnod2 by cytokinin. Plant Physiol. in press. Szczyglowski, K., and Legocki, A.B. (1990). Isolation and nucleotide sequence of cDNA clone encoding nodule-specific (hydroxy) proline-rich protein LEnodZ from yellow lupin. Plant Mol. Biol. 15, 361-363. Takaiwa, F., Oono, K., and Sugiura, M. (1984). The complete nucleotide sequence of a rice 17S rRNA gene. Nucleic Acids Res. 12, 5441-5448. van de Wiel, C., Scheres, B., Franssen, H., van Lierop, M.J., van Lammeren, A., van Kammen, A., and Bisseling, T. (1990a). The early nodulin transcript ENOD2 is located in the nodule parenchyma (inner cortex) of pea and soybean root nodules. EMBO J. 9, 1-7. van de Wiel, C., Norris, H.J., Bochenek, B., Dickstein, R., Bisseling, T., and Hirsch, M.A. (1990b). Nodulin gene expression and ENOD2 localization in effective, nitrogen-fixing and ineffective, bacteria-free nodules of alfalfa. Plant Cell 2, 1009-1017. Veen, H. (1979). Effects of silver on ethylene synthesis and action in cut carnations. Planta 145, 467-470. Verwoerd, T.C., Dekker, B.M.M., and Hoekema, A. (1989). A small- scale procedure for the rapid isolation of plant RNAs. Nucleic Acids Res. 17, 2362. Wu, C.H., Caspar, T., Browse, J., Lindquist, S., and Somerville, C. (1988). Characterization of an HSP70 cognate gene family in Arabidopsis. Plant Physiol. 88, 731-740. Yang, S.F. and Hoffman, N.E. (1984). Ethylene biosynthesis and its regulation in higher plants. Annu. Rev. Plant Physiol. 35, 155-189. 174 Yu, Y.B., and Yang, S.F. (1979). Auxin-induced ethylene production and its inhibition by aminoethoxyvinylglycine and cobalt ion. Plant Physiol. 64, 1074-1077. Zaat, S.A.J., van Brussel, A.A.N., Tak, T., Lugtenberg, B.J.J., and Kijne, J .W. (1989). The ethylene-inhibitor aminoethoxyvinylglycine restores normal nodulation by Rhizobium Ieguminosarum biovar. Vicia on Vicia sativa subsp. nigro by suppressing the ‘thick and short roots’ phenotype. Planta 177, 141-150. CHAPTER 5 FUNCTION OF THE ENOD2 PROTEIN IN NODULES 175 176 ABSTRACT The Enod2 gene encodes a putative cell wall protein, and Enod2 transcripts are localized in the nodule parenchyma, which has been hypothesized to function as a barrier for oxygen diffusion. To examine the proposed role of nodulin Enod2 in the formation of this oxygen barrier, we attempted to down-regulate Enod2 gene expression in transgenic Lotus corniculatus plants by expressing antisensc Enod2 constructs. In transgenic plants, the endogenous Enod2 transcript levels were found to be altered, but not completely abolished. Expressing the antisense Enod2 constructs did not affect plant growth and nodulation, and nodules were found to be able to reduce acetylene, indicating the presence of a functional nitrogenase enzyme. However, nodules formed on several transgenic antisense plants appeared to be smaller than those on control plants, indicating an alteration in nodule morphogenesis. Oxygen microelectrode measurements revealed that the oxygen diffusion profile of some transgenic plants may be altered. 177 INTRODUCTION Oxygen plays an important role in symbiotic nitrogen fixation. On the one hand, the nitrogenase activity requires high energy (ATP) input, which has to be generated by bacterial oxidative phosphorylation in the nodule infected cells. On the other hand, the nitrogenase complex is irreversibly and rapidly inactivated by oxygen (Robson and Postgate, 1980). This apparent oxygen paradox is reconciled by the creation of a microaerobic environment inside of nodules. Direct measurements of nodule internal oxygen concentration (Oi) have provided evidence for the presence of physical, as well as variable oxygen diffusion barriers in nodules, to allow a continuous flow of oxygen to the infected cells of the nodule, while maintaining the Oi at a low level to prevent nitrogenase from inactivation. Using oxygen microelectrodes, Tjepkema and Yocum (1974) were able to measure the free oxygen concentration in nodules of soybean (Glycine max Merr.). They found that the 02 concentration decreased sharply across the nodule cortex and was below the detection level in central tissues. These data suggested that the resistance to O2 diffusion into the nodule primarily occurred within the cortex. Subsequently, Witty et a1. (1987) confirmed this observation in nodules of pea (Pisum sativum) and French bean (Phaseolus vulgaris). Detailed microscopic examination of the nodule anatomy revealed that the inner cortex (named nodule parenchyma 178 thereafter) has compact and small cells, with fewer intercellular spaces than in other nodule tissues (van de Wiel et al., 1990). Since oxygen diffusion is approximately four orders of magnitude slower in solution than in air, the nodule parenchyma, therefore has been hypothesized to be the site of resistance to oxygen diffusion in nodules (van de Wiel et al., 1990). A nodule variable resistant barrier to oxygen has been identified in experiments, in which the nitrogenase was found to be active in nodules exposed to higher rhizosphere 02 concentration (Sheehy et al., 1983; for a review, see Hunt and Layzell, 1993) The change in resistance to 02 has been linked to inhibition of nitrogenase activity by a number of environmental and physiological factors (for a review, see Hunt and Layzell, 1993). However, the site of the variable barrier and its relationship to the physical barrier are currently not well understood. The expression of the Enod2 gene has been localized to the nodule parenchyma (Nap and Bisseling, 1990; van de Wiel et al., 1990). Since it is a putative cell wall protein, the Enod2 protein has, therefore, been postulated to contribute to the special morphology of the nodule parenchyma, and the formation of the oxygen diffusion barrier (van de Wiel et al., 1990). To test this hypothesis, we have used a genetic approach to down-regulate Enod2 gene expression in transgenic L. corniculatus plants, by expressing antisensc Enod2 transcripts. We have generated antisensc expressing constructs using both a strong constitutive promoter and a 179 nodule parenchyma-specific promoter. Here, we report the phenotypic and physiological analysis of these transgenic plants. METHODS Plant transformation and growth L. corniculatus plants were grown in soil mix (V ermiculitezsandzMetroMix, 5:3:1) in pots maintained in a controlled environment cabinet (Conviron, Asheville, NC) at 24°C during the day and 20°C at night, with a 16 hr photoperiod (250 pE/sec/m2). For nitrogenase activity measurements, six cuttings of each plant to be analyzed were generated and grown for one week to regenerate roots. These plants were inoculated with a two-day-old culture of Rhizobium loti NZP 2037 (Pankhurst, 1983) and grown for two weeks. Plants were subsequently repotted into silica sand, in sealable pots, and grown for two weeks. During this grth period, plants were watered with B+D nutrient solution (Broughton and Dilworth, 1971), containing 0.5 mM KNO3. 180 Isolation of total RNA and northern blot analysis Total RNA was isolated following the hot phenol method of Verwoerd et al., (1989), with modifications described previously (Chen et al., 1996; Chapter 2). Northern blot analysis was carried out following standard procedures (Maniatis et al., 1989). A a32P-dATP-labeled DNA fragment containing the SrEnod2 coding region was used as a probe. Antisense Enod2 constructs preparation A 1.3 kb MspI fiagment, containing the 5’ UTR and most of the SrEnod2 coding region (Dehio and de Bruijn, 1992), was cloned into the Ach site of pUC18 (GIBCO BRL, Gaitherburg, MD) to generate plasmid pCD21 (Dehio and de Bruijn, 1992). To generate the antisense Enod2 construct containing coding sequences (Figure 5-1, Construct 37), a 900-bp BamI-II- SpeI fragment was excised and inserted into the BamHI-SpeI sites of the binary vector pROK2275 (Szabados et al., 1990), harboring a modified polylinker between the CaMV3SS promoter and the gas coding region (Ach-BamHI-SalI-SmaI-Spel). To generate the antisense Enod2 construct containing primarily 5’ sequences (Figure 5-1, Construct 36), a 400-bp SpeI-PstI fragment was excised from pCD21 and inserted into the XbaI-PstI sites of pUC18. The resulting plasmid was digested with PstI and blunt- 181 ended with T4 DNA polymerase (Maniatis et al., 1989). The linearized plasmid was digested with BamHI, and the 400 bp resulting fragment was purified and inserted into the BamHI and the blunt-ended SalI sites of the modified binary vector pROK2275. To generate the antisense Enod2 construct containing primarily 3’ sequences (Figure 5-1, Construct 38), a 280-bp BamHI-Kpnl fragment was excised fiom pCD22A (Dehio and de Bruijn, 1992), and cloned into the BamHI-Kpnl sites of pROK2275. To generate the antisense Enod2 constructs without the gus gene, the 1.3 kb BamHI and blunt-ended HindIII fragment of pCD21 was cloned into the BamHI and blunt-ended SacI sites of the binary vector pB1121 (Clontech Laboratories, Palo Alto, CA) (Figure 5-1, Construct 39), or the binary vector 5 ’SrEnod2-gus-3 ’SrEnodZ (Figure 5-1, Construct 40; see Chapter 2). Oxygen microelectrode measurement of nodule oxygen tensions Oxygen microelectrode measurements were carried out essentially as described by Tjepkema and Yocum (1974) with the following modifications. Briefly, nodules were detached from roots 21-28 days after nodulation and embedded in 0.5 % agarose positioned on top of a layer of solidified 1 % agarose in a 1 cm3 chamber. Nodules were positioned in such a way that the apical region points to the side of the chamber and that the nodules were 1 to 1.5 cm under the surface. An oxygen electrode (ca. 10 pm 182 in tip width; kindly provided by Dr. David Emerson, Michigan State University, MI) was used to monitor the oxygen tension. Measurement of nitrogenase activity using an open-flow gas exchange system The open-flow gas exchange system developed by Hunt et al., (1989) was utilized to measure nitrogenase activity. Briefly, nodulated plants were grown in silica sand for two weeks and the pots were sealed one day before gas exchange. A computer-controlled mixture of gases was passed through the pots and both H2 and CO2 evolution in the efflux was measured using a 112 analyzer (Model H-lSO, Morgen Instruments, Andover, MA) and an infi'ared CO2 analyzer (Model 225, Mark III, Analytical Development Corporation, Hoddesdon, U.K.). Apparent nitrogenase activity (ANA) was defined as the level of 1-12 evolution in air. Total nitrogenase activity (TNA) was defined as the level of 112 evolution in ArzOz. Potential nitrogenase activity (PNA) was defined as the maximum rate of H2 evolution in ArzOz, during a linear increase in p02 from 20 to 100 % in a 30-min period (Diaz del Castillo et al., 1992). For dry weight measurement, nodules were first dehydrated in a 80°C baking oven for two days and then weighed. 183 RESULTS Generation of antisense Enod2 constructs Three antisensc Enod2 constructs were generated using fragments covering the 5', internal and 3' portions of the SrEnod2 genomic locus (Figure 5-1, Constructs 36, 37, and 38; Dehio and de Bruijn, 1992). To facilitate the detection of expression of these antisensc constructs, these fragments were fused, in an antisense orientation, to a bacterial gus reporter gene (Jefferson, 1987). A strong promoter from CaMVBSS (Benfey et al., 1989) was employed to generate antisensc transcripts. The constructs were introduced into L. corniculatus plants by A. rhizogenes-mediated plant transformation (Szabados, et al., 1990). The expression of the antisense transcripts was monitored in roots of transgenic plants by GUS activity staining, and found to be positive in 12 out 30, 7 out of 23, and none out of 16 plants, harboring the antisense constructs containing the 5’, 3’ and the coding sequences, respectively (data not shown). The other two antisensc constructs (Figure 5- 1, Constructs 39 and 40) were constructed such that a DNA fragment containing the 5’ and internal portions of the SrEnod2 gene was fused in an antisense orientation to either the CaMV35S promoter or the SrEnod2 5 ’ and 3’ sequences responsible for directing expression to the nodule 184 LEGEND FOR FIGURE 5-1. Antisense Enod2 constructs. (A) The SrEnod2 gene locus. The black box represents the SrEnod2 coding region. Open boxes at each end of the black box represents the 5’ and 3’ untranslated regions of the SrEnod2 gene. (B) Subfi'agments of the SrEnod2 gene. 5’-, internal-, and 3’- indicate subfragments corresponding to the 5’, internal, and 3’ portions of the SrEnod2 gene, respectively. (C) The binary vector used to generate the 5’-, Intemal-, and 3’- antisensc Enod2 constructs. The insertion site and orientation are indicated by an arrow. The construct numbers and description are provided. (D) A antisensc Enod2 construct. The fragment containing the 5’ and internal portions of the SrEnod2 gene was inserted between the CaMV35S promoter and the 3’ nos terminator. The construct number and description are provided. (E) A antisensc Enod2 construct. The fragment containing the 5’ and internal portions of the SrEnod2 gene was inserted between the SrEnod2 5’ and 3’ flanking sequences (Chen et al., 1996; see Chapter 2). The construct number and description are provided. 185 5'- Internal- ‘\ ’1 / ._ ,x/’/ Cansss _ 3'nos Construct 36: 5'-Antisense Construct 37: Intemal-Antisense Construct 38: 3'-Antisense cwzsm 3'nos Construct 39: 35S-Antisense s mm m s W ! Construct 40: SrEnod2-Antisense I 186 parenchyma (Chen et al., 1996; see Chapter 2). 40 and 80 transgenic plants ' harboring these two antisensc constructs, respectively, were generated. Phenotype of antisense Enod2-expressing transgenic plants The majority of the 69 transgenic L. corniculatus plants harboring the antisense constructs 36, 37, and 38, as well as 8 plants transformed with the control construct, 5'Srglb3-gus-3 'nos (designated as pLP32; Szabados, et al., 1990), appeared to be normal as compared to wild type plants. However, a couple of plants harboring the 5'-antisense Enod2 construct (Figure 5-1, Construct 36), showed severely altered phenotypes (reduced intemode length and early senescence of leaves). The reason for that is not clear. One possibility is that the expression of rolABC genes (White et al., 1985) from the A grobacterium rhizogenes strain used to transform L. corniculatus plants in these experiments caused the observed plant grth phenotypes, since tobacco plants expressing these genes showed similar phenotypes (Sinkar et al., 1988; Estruch et al., 1991a; b). However, the possibility that the expression of the 5’-antisense Enod2 transcript caused the phenotype can not be completely ruled out. Transgenic plants were also tested for their nodulation characteristics. All plants, including those plants showing severe phenotypes, could be effectively nodulated. Nodules were pink, indicating that they were able to 187 fix nitrogen. This was confirmed by monitoring nitrogenase activity using the acetylene reduction assay (data not shown). When transgenic plants with altered endogenous Enod2 transcripts levels (see below) were chosen to measure their nodule fresh weight and number, nodules formed on several of these plants were found to be smaller than those formed on the wild type and control plants (Figure 5-3). Endogenous Enod2 mRNA levels in transgenic antisensc plants Endogenous Enod2 mRNA levels in nodules of transgenic antisense plants were examined by Northern blot analysis. Two major Enod2 transcripts were found in wild type nodules (Figure 5-2, Panel A). Transgenic antisense plants were found to contain various levels of these two transcripts (Figure 5-3, Panel B). A total disappearance of the two endogenous Enod2 transcripts was not observed in 69 individual transgenic plants harboring the three different antisensc constructs (Figure 5-1, Constructs 36, 37, and 38; data not shown). Similar results were obtained after analyzing 120 transgenic plants transformed with the other two constructs (Figure 5-1, Constructs 39 and 40; data not shown). Accumulation of the antisense transcripts was also examined in the transgenic plants, and the antisense transcripts could not be detected. Since the activity of the GUS fusion protein had been detected in transgenic 188 LEGEND FOR FIGURE 5-2. The Enod2 gene expression in L. corniculatus plants. (A) Northern blot analysis. The Enod2 gene expression during nodule development of wild type L. corniculatus plants is shown. Two major transcripts were detected 14 days after rhizobial infection. (B) Northern blot analysis. The endogenous Enod2 gene expression in transgenic plants harboring the 5’- and Intemal- antisensc Enod2 constructs is shown. The position of ribosomal RNAs is indicated. 189 21 28 Days 14 m2..— 5.: I: 3% E. .m cudm 190 LEGEND FOR FIGURE 5-3. Nodule fresh weight of transgenic L. corniculatus plants harboring antisense Enod2 constructs. Nodules formed between 4 and 5 weeks after rhizobial infection were collected and weighed. Six replicates for each transgenic line were assayed. The solid black boxes represent average values, and the error bars represent the standard deviation. 191 < bra: < Ymmmq < E < \- Tm < cmfi < warm < m TEH < v.2: u < w-m .1. A ”.5