. u . 461...” ”4‘33un 47'.» WW: V . t. . $79.... 3. ban a .u '0. it” I ‘ .D o I 4 .9. . . 23:35,; .2. n x 5.. up; 3.3%". ‘5.bm‘. Al mafia? r: s ..!v04 . .1... ix). . . 1...? .. l n. 3.3.9:}.usxta u... a q 533.21%]...1‘ uluv . a...) L . 7?. ' if .1. 3 .‘...li.$ 52:? .2. .. TTiE'\ib n . A ,,_CO+.. LIBRAHY Michigan State University This is to certify that the dissertation entitled Regulation and function of Arabidopsis thaliana secreted ribonucleases presented by Nicole D. LeBrasseur has been accepted towards fulfillment of the requirements for Ph . D . degree in Genetic s W% 2/ Dated 2/14 101 MS U i: an Affirmative Action/ Equal Opportunity Institution 0 12771 PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 cJCIRC/Dateouopes-nts REGULATION AND FUNCTION OF ARABIDOPSIS T HALIANA SECRETED RIBONUCLEASES By Nicole D. LeBrasseur A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Genetics Program 2001 ABSTRACT REGULATION AND FUNCTION OF ARABIDOPSIS THALIANA SECRETED RIBONUCLEASES By Nicole D. LeBrasseur While secreted ribonucleases (RNases) have been well studied at the enzymatic and structural levels, little is known regarding their biological functions. One family of secreted RNases, the RNase T2 family, is particularly widespread, with members throughout various kingdoms. In recent years, many plant members of this superfamily have been identified. Gametophytic self-incompatibility in several plant groups involves the activity of S-RNases, one subfamily of plant T2 RNases. Another subfamily, the S- like RNases, are found in self-compatible as well as self-incompatible plant species, indicating that they have different functions than have S-RNases. Expression patterns of S-like RNases in many species led to the suggestion that these enzymes are involved in remobilization and recycling of phosphate from nucleic acid sources. The project described in this thesis involves the study of secreted RNases in the model plant Arabidopsis thaliana, which has five members of the S-like subfamily. I found that several Arabidopsis RNases are induced in wounded tissues through an as-yet unidentified signaling mechanism. One S-like member, RN81, is induced both locally and systemically by wounding through a novel signal transduction pathway. The regulation of RNSI expression was studied using several reporter constructs. I found that transcriptional regulation accounts for the majority of the wound response of this gene. Several other factors were shown to affect RNSI transcript levels, including abscisic acid, salt stress, and heat shock. In addition to regulatory mechanisms controlling RN51 transcription, the function of RNSl was also studied. Tools used for these studies include plants with constitutively high RNSl activity and T-DNA mutant lines with insertions in the RNSI gene. One of these mutants, rnsI-Z, has little or no RNSl activity. Conditions that affect RNSI transcript accumulation provided a starting point for the analysis of the mutant and overexpressor. The observation that the overexpressors grew better than wild type when supplied with RNA as the sole source of Pi supports the hypothesis that RNSl is involved in P, remobilization. Analysis of the mutant and overexpressor also revealed that RNSl activity levels affect root length in Arabidopsis. This effect was not dependent on P, concentration and indicated that secreted RNases are involved in unexpected processes. Certain phenotypes of the rnsI-Z plants resembled those of the my] A mutant. Rnyl, the only RNase T2 enzyme found in Saccharomyces cerevz'seae, has been proposed to function in regulating aspects of membrane stability and permeability. Therefore, I propose that RNSl may also be involved in similar processes. Finally, the analysis of an Arabidopsis mutant with altered activity levels of several RNA-degrading enzymes is presented. Although one of the increased activities resembles RNSl, I demonstrate that this activity is not RNSl. The altered activities are specific to stems of the mutant, as are its morphological phenotypes. This mutant, arpl, provides an insight into the involvement of RNases in stem-associated processes and stem development. ACKNOWLEDGMENTS Numerous thanks are owed to many people who contributed in one way or another to this work. First and foremost, I must extend my appreciation to Dr. Pam Green for the innumerable ways in which she has helped me in the past 6 years, for taking the chance on me as an inexperienced graduate student, and for supporting me even when I chose an "alternative" scientific career. Not all mentors would have been so supportive. Thank you to members of my guidance committee, Drs Hans Kende, Mike Abler, Mike Thomashow, and Sheng Yang He, for helpful suggestions and moral support. Mike Abler was also integral in the arpl story and in my earliest exposure to benchwork; I have never forgotten his advice. Thanks to those who helped with experiments throughout this thesis, including Julie Zwiesler—Vollick for the help with the bacterial growth curves, Miguel Pérez-Amador for the microarray that got me started on wounding, Ted Farmer for great discussions and several contributions to the wounding story, Christie Howard for RNSI constructs, and Daniel Cook for blots of various stresses included in Chapter 3. For photographic and figure-making services, Marlene Cameron and Kurt Stepnitz were both patient and helpful. Thanks, Linda Danhof, for many activity gels, technical assistance, and direction of an army of undergraduate slaves (whom I would also like to thank, especially, Lindsey, Amanda, Robert, lance, and Courtney). Much of Chapter 4 would not have been included without the talents of a special undergraduate, Tracey Millard, to whom I am ever grateful for her abilities and keen interest in plant physiology. To all Green lab members past and present, I can never thank you enough. You have iv been friends, colleagues, mentors, and surrogate family all in one. To Preet Lidder and Rodrigo Gutierrez, thanks for being great roommates (the back room rules!). To Jay De Rocher and Pauline Bariola, thanks for the advice, direction, and friendship in the early years. Special thanks to Gustavo Mac lntosh, for help with experiments, for the many papers lefi on my desk, for keeping me on the right track the past three years, and for being my best friend. I wouldn't have made it without you. I also owe debt to many PRL members, who have helped scientifically and otherwise. I have never found a more friendly and productive department than the PRL. I would like to say a personal thank you to my big sister, Antje Heese—Peck, who I miss very much. Thank you to the darts crew, especially John Froelich, John Scott-Craig, Scott Peck, and Tom Newman; Mondays will never be the same. Finally, thank you to my brother, Brian, and my parents, Robert and Lauraly, for years of support, encouragement, and love across the miles. I am lucky to have you. TABLE OF CONTENTS LIST OF TABLES ............................................................................................................. ix LIST OF FIGURES ............................................................................................................. x CHAPTER 1 INTRODUCTION: PLANT RNA-DEGRADING ACTIVITIES ...................................... l Secreted ribonucleases and the RNase T2 family ........................................................ 2 Function: the S-RNases ............................................................................................... 3 The S-like RNase subfamily: general features ............................................................ 4 Plant nuclease l-type enzymes .................................................................................... 6 Regulation and putative functions of S-like and nuclease I enzymes in plants .......... 6 Dissertation topic and thesis overview ...................................................................... 14 CHAPTER 2 LOCAL AND SYSTEMIC WOUND INDUCTION OF RNASE AND NUCLEASE ACTIVITIES IN ARABIDOPSIS: RNSl AS A MARKER FOR A JASMONIC ACID- INDEPENDENT SYSTEMIC SIGNALING PATHWAY ............................................... 15 Abstract ......................................................................................................................... 16 Introduction ................................................................................................................... l 7 Results ........................................................................................................................... 19 Wounding induced several RNase activities in Arabidopsis .................................... 19 Multiple bifunctional nuclease activities are increased by wounding ....................... 22 RNSl activity is induced by wounding locally and systemically ............................. 24 The induction of RNSI and the nucleases is independent of j asmonic acid ............. 26 RNSI is not induced by the oligosaccharide pathway ............................................... 30 Discussion ..................................................................................................................... 32 Materials and Methods .................................................................................................. 38 Plant materials and treatments ................................................................................... 38 RNA extraction and RNA gel blot hybridization ...................................................... 40 Protein extraction and detection of RNase and DNase activities .............................. 41 CHAPTER 3 REGULATION OF RNSI EXPRESSION ........................................................................ 42 Abstract ......................................................................................................................... 43 Introduction ................................................................................................................... 44 Results ........................................................................................................................... 45 Run-on analysis of RNSI expression ........................................................................ 45 Regulation of RNSI expression in response to wounding ........................................ 47 Use of promoter constructs to analyze signals mediating RNSI induction ............... 52 Discussion ..................................................................................................................... 59 Sequence requirements for the wound induction of RNSI ........................................ 59 Response of RNSI to various biotic and abiotic stimuli ........................................... 61 Materials and Methods .................................................................................................. 64 Nuclei preparations and run-on analyses .................................................................. 64 vi Cloning ...................................................................................................................... 66 Plant transformation .................................................................................................. 67 Plant treatments ......................................................................................................... 67 RNA and protein analyses ......................................................................................... 68 Luciferase analysis .................................................................................................... 68 CHAPTER 4 RNSl FUNCTION: ANALYSIS OF AN RNSI MUTANT AND OVEREXPRESSOR.69 Abstract ............................................................................................ 70 Introduction ................................................................................................................... 71 Results ........................................................................................................................... 72 Isolation of ms] mutants and RNSI overexpressing plants ...................................... 72 The rnsl-Z mutant does not have increased levels of anthocyanins ......................... 77 RNSl activity appears to affect plant growth on RNA ............................................. 79 RNSl activity affects plant size ................................................................................ 81 Absence of RNSl does not increase susceptibility of the plant to a bacterial pathogen .................................................................................................................... 89 Discussion ..................................................................................................................... 92 Involvement of RNSl in phosphate remobilization .................................................. 92 Is RNSl involved in defense mechanisms? .............................................................. 94 RNSI activity negatively correlates with root length ............................................... 95 Materials and Methods .................................................................................................. 97 Plant materials ........................................................................................................... 97 Plasmid construction ................................................................................................. 98 Identification of ms] T-DNA insertional knock-out mutants ................................... 99 Protein and RNA analyses ....................................................................................... 100 DNA extraction and DNA gel blot hybridization ................................................... 100 Anthocyanin assays ................................................................................................. 101 Bacterial growth curves ........................................................................................... 101 CHAPTER 5 ANALYSIS OF AN ARABIDOPSIS T HALIANA MUTANT WITH AN ALTERED RNASE PROFILE .......................................................................................................... 102 Abstract .......................................................................................... 103 Introduction ................................................................................................................. 104 Results ......................................................................................................................... 105 An RNase-activity gel screen for mutants affecting the Arabidopsis RNase profile ................................................................................................................................. 105 The 33-kDa RNase activities are bifunctional nucleases ........................................ 109 Genetic characterization of the arp mutants ........................................................... 109 Morphology of the arpI mutant .............................................................................. 110 The arp] mutation induces an RNase activity in addition to the 33-kDa nuclease activities .................................................................................................................. l 12 The arp] RNase phenotype is stem associated ....................................................... 119 Discussion ................................................................................................................... 120 Mutants affecting single and multiple RNase activities .......................................... 120 vii The 33-kDa activities are bifunctional nucleases .................................................... 122 Possible functions for the 33-kDa nucleases ........................................................... 123 arp] is a regulatory link between the 33-kDa nucleases and an RNase activity 125 The app] phenotypes are associated with stems ..................................................... 125 Materials and Methods ................................................................................................ 127 Plant materials ......................................................................................................... 127 RNase and DNase activity gels ............................................................................... 127 Mutant growth and development ............................................................................. 127 DNA and RNA extraction and gel blot analyses ..................................................... 128 Immunoblot analysis ............................................................................................... 128 CHAPTER 6 CONCLUSIONS AND FUTURE PROSPECTS ............................................................ 129 REFERENCES ................................................................................................................ 135 viii LIST OF TABLES Table 4—1. Statistical analysis of fresh weight of WT, msI-Z, and 35S-RNSI seedlings 85 Table 4-2. Statistical analysis of root length in ms] -2, WT Ws, 35S-RNS1 and WT Col seedlings ............................................................................................................................ 86 Table 5-1. Statistical analysis of morphological differences between WT RLD and arpI plants ............................................................................................................................... 113 ix LIST OF FIGURES Figure 2-1. Multiple RNase activities are induced by wounding ................................ 20 Figure 2-2. Bifunctional nuclease activities are induced by wounding. ...................... 23 Figure 2-3. RNSl activity and transcript increase transiently in wounded and in systemic, non-wounded leaves. ............................................................................. 25 Figure 2-4. RNSl and nuclease induction by wounding is independent of jasmonic iacid........... Figure 2-5. Figure 3-1. Figure 3-2. Figure 3-3. Figure 3-4. Figure 3-5. Figure 3-6. Figure 3-7. ......................................................................................................... 28 OGA-rich fractions do not induce RNase activities. ................................. 31 Run-on analysis of RNSI in response to P; starvation ............................... 46 Constructs used to examine the regulation of RNSI . ................................ 48 Differential response of RNSI transcribed sequences to wounding .......... 50 The RNS] promoter confers wound inducibility to reporter transcripts. .. 51 5' UTR sequences are not required for the wound responsiveness of the RNSI promoter. ..................................................................................................... 53 Dehydrated plants do not induce the RNSI promoter. .............................. 54 RNSI transcript accumulates in response to salt stress. ............................ 54 ABA induces the RNSI promoter. ............................................................ 57 Figure 3-8. Figure 3-9. Figure 3-10. Figure 4-1. Figure 4-2. Figure 4-3. Figure 4-4. Figure 4-5. Figure 4-6. medium.......... Mutants in ABA signaling and biosynthesis induce RNSl by wounding. 58 Heat shock inhibits the wound-induction of luciferase activity. ............... 60 ms] mutants and overexpressor constructs ............................................... 73 RNSl activity in the rnsl-I and ms] -2 mutants. ...................................... 75 DNA and RNA gel blot analyses of rnsI-2 mutant plants. ....................... 76 RNSl activity in overexpressing plants. ................................................... 78 Anthocyanin content in mutant ms lines ................................................... 80 Growth of rns‘I-Z and 3SS-RNSI seedlings on RNA-containing . ................................................................................................... 82 Figure 4-7. Root phenotype of RNSI mutant and overexpressor ................................. 84 Figure 4-8. Absorption of dyes in wild-type Columbia (WT) and rnsI-Z mutant root cells........... ......................................................................................................... 88 Figure 4-9. Growth of Pst DC3000 in leaves of 35S-RNS1 and rnsl-Z plants. .......... 91 Figure 5-1. arp mutants are altered in RNase activity profiles .................................. 107 Figure 5-2. The 33-kDa RNase activities also degrade DNA. ................................... 108 Figure 5-3. Morphological differences in the phenotypes of wild-type and arp] mutant plants............ ................................................................................................... 111 Figure 5-4. The arpI mutant has increased levels of an activity that comigrates with RNSl....... ................................................................................................... 115 Figure 5-5. RN51 transcript and protein is not detected in arp] stems ...................... 116 Figure 5-6. A previously unidentified 23-kDa band is increased in arpI .................. 118 Figure 5-7. The cup] RNase phenotype is stem specific. .......................................... 121 Figure 6-1. Model for the control of membrane permeability by RNases. ................ 132 xi CHAPTER 1 INTRODUCTION: PLANT RNA-DEGRADING ACTIVITIES The class of proteins known as ribonucleases (RNases) encompasses a diverse group of enzymes whose catalytic activity involves the degradation and catabolism of RNA. There are numerous ways in which one might categorize RNases, including primary structure homology, cellular localization, and enzymatic cleavage mechanism or other biochemical properties, to name a few. RNases are involved in many processes in cellular function. Perhaps their most obvious function is the controlled processing and degradation of mRNA as a mechanism of gene regulation, although to imagine that RNases are restricted to such processes is likely a major underestimation of their firnctions. In fact, in most systems, the genes encoding proteins involved in mRNA degradation are still unknown. In contrast, many RNA-degrading activities that are not likely to be involved in the processing or degradation of mRNA have been characterized and the genes encoding them cloned, as is the case in the field of plant RNases. Likely the most well- characterized of these activities are those that fall into the RNase T2 superfamily. More than one review has been published in recent years in the field of plant RNases and/or mRNA degradation (examples include Bariola and Green, 1997; Gutierrez et al., 1999; and Johnson et al., 1998). Therefore, this chapter will not attempt to reiterate such reviews. Instead, it will focus on recent findings in regard to the expression patterns and putative functions of plant secreted RNases and the signaling mechanisms involved in their regulation, with particular attention paid to the S-like RNases, one class of plant T2 RNases. Secreted ribonucleases and the RNase T2 family Biochemists have long used secreted RNases for structural and enzymatic studies, due to the relative ease in their purification and the availability of large quantities of proteins. IQ RNase A, for example, is a low molecular weight, stable enzyme that has been purified in large quantities from many vertebrate species. Perusing any review on pancreatic RNases, which mostly focus on RNase A family members, will enlighten the reader on several aspects of protein structure and enzymatic mechanisms (see, for example, Cuchillo et al., 1997) and is therefore quite telling regarding the focus of past studies on these particular subjects. While much is known on base-specificity, protein folding structures, and role of specific amino acids in catalysis, little is known about the physiological firnctions of these enzymes. Further, even fortuitous findings of "special" biological actions of RNases (D'Alessio et al., 1991 ), including antitumor and angiogenic activities, have not been explained biologically. The T2 superfamily of RNases has also been well-studied at the enzymatic and structural levels (Irie, 1997). First identified in the fungus Aspergillus oryzae (Sato and Egami, 1957), RNase T2 and its relatives are secreted acid endonucleases with no base- specificity, sharing two conserved stretches of amino acid residues important in the enzymes' catalytic sites (Irie, 1997). Unlike the RNase A superfamily, members of which have so far been identified only in vertebrates (Beintema et al., 1997), T2 enzymes are incredibly widespread. To date, members of this family have been found in nearly every system examined for their presence, encompassing viruses, bacteria, plants, and animals (Irie et al., 1997). The ubiquitous distribution of T2 RNases may be interpreted as the enzymes' having both an ancient origin and critical function(s) (Taylor and Green, 1991). Function: the S-RNases Despite the apparent necessity for the activity of these enzymes, very little has been demonstrated regarding their biological functions. The exception is the S-RNascs, a class of plant T2 RNases, whose activity is required for the process of self-incompatibility in several plant families, including the Solanaceae (reviewed in McCubbin and Kao, 2000). Briefly, when the S-allele of a pollen grain growing through the style of the mother plant matches either of the two S—alleles in the pistil, growth of the pollen tube is inhibited. RNase activity has been shown to be essential for this incompatibility system (Lee et al., 1994). However, the mechanism through which the S-RNase imparts incompatibility is still unknown. Available evidence indicates that the S-RNases somehow act as cytotoxins in the incompatible tubes (McClure et al., 1989). The pollen component of the interaction has not been identified, nor is it known at what site the two components interact. The two leading models in the field predict that either 1) the S-RNases are specifically imported into incompatible pollen tubes or 2) nonspecific S-RNase uptake occurs, with subsequent specific inactivation of all compatible RNases or some other similar modification (Parry et al., 1997). Recent evidence supports the latter hypothesis, in that S-RNases appear to be taken up by pollen tubes in a genotype-independent manner (Luu et al., 2000). In summary, the S-RNases are the sole members of the T2 RNase superfamily who have been assigned a function. Yet, there is much to be understood regarding the mechanism of the incompatibility reaction. The S-like RN ase subfamily: general features Soon after the identification of the S-RNases as members of the T2 family, several groups reported similar proteins in self-compatible plant species, including Arabidopsis, tomato, and Zinnia. Although these S-like RNases share certain structural features with the S-RNases, conserved regions specific to each group can also be found (for a comprehensive review of S-like RNases, see Bariola and Green, 1997). These enzymes are obviously not involved in self-incompatibility, but appear to play important roles throughout the plant kingdom, as S-like members have now been identified in both self- compatible and self-incompatible species. The Arabidopsis thaliana genome contains five genes, RNS] to RNS5, with high similarity to the S-like RNase gene family (Taylor and Green, 1991; GO Maclntosh, N.D. LeBrasseur, and P.J. Green, unpublished), and RNase activity has been demonstrated for the products of three of the RNS genes (Taylor et al., 1993; Bariola et al., 1994). Studies on S-like RNases have for the most part focused on gene expression or activity patterns, as will be discussed below. Limited biochemical and structural information is also available. As mentioned, the S-like RNases are secreted enzymes. While many are extracellular, including RNSl (Bariola et al., 1999), some have been shown to be retained intracellularly. These include RN82 (Bariola et al., 1999) and several activities in tomato that are localized to the vacuole or endoplasmic reticulum (Loffler et al., 1992; Lehmann et al., 2001). Alternate cellular localization indicates that multiple members of the S-like subfamily in the same plant may carry out separate functions or the same functions in different compartments. The crystal structure of the tomato extracellular S-like RNase LE has been determined at 1.65 A resolution (Tanaka et al., 2000). Other T2 enzymes that have been crystallized include an S-like RNase from bitter gourd (Nakagawa et al., 1999) and the fungal enzyme RNase Rh (Kurihara et al., 1996). Comparison of these and other T2 enzyme structures has revealed two structural groups: the animal/plant and the firngal subfamilies (Tanaka et al., 2000). However, the groups appear to share catalytic mechanisms despite these structural differences (Tanaka et al, 2000). Plant nuclease I-type enzymes Early classification of plant RNA-degrading activities based on biochemical properties of the enzymes identified several groups: RNase l, RNase II, nuclease I, and exonuclease I (Farkas, 1982; Wilson, 1982). While most T2 RNases appear to fall within the RNase I class, many plant activities have been identified with characteristics placing them in the nuclease I family, including activities in tobacco, barley, mung bean, zinnia, and rye (reviewed in Bariola et al, 1997). Nuclease l enzymes are also extracellular heat-stable proteins, but unlike T2 RNases, nucleases have the ability to degrade both RNA and single-stranded (ss) DNA. In general, they also have higher molecular weights, in the range of 31 to 42 kDa, are glycosylated, and have acidic pH optima. They are sensitive to EDTA and also require Zn2+ for activity and stability (Fraser and Low, 1993). Until recently, sequence information was not available for many plant nucleases. Determination of the amino acid sequences of two fungal nucleases, P1 from Penicillium citrum (Maekawa et a1, 1991) and S] from Aspergillus oryzae (lwamatsu et a1, 1991), has allowed the cloning of genes encoding similar proteins in plants. Nuclease 1 cDNA sequences are now available for daylily (Panavas et al., 1999), celery (Yang et al., 2000), barley, zinnia, and Arabidopsis (Aoyagi et al., 1998; Pérez-Amador et al., 2000). Regulation and putative functions of S-like and nuclease I enzymes in plants To date, fluctuations in RNase activity levels or gene expression have provided the most useful data for predicting RNase function. Through elucidating the biotic and abiotic cues that correlate with increases in RNase activity, we have been able to postulate functions that secreted RNases play in the plant. The following sections will briefly highlight significant examples and recent findings concerning RNase and nuclease induction under various environmental and developmental conditions and summarize conclusions drawn from these expression patterns. RNases and phosphate starvation Induction of S-like RNases and the genes encoding them is strongly associated with growth on low concentrations of inorganic phosphate (Pi). Pertinent examples include the Arabidopsis RNS] and RN32 genes and proteins (Bariola et al., 1994, 1999), and intra- and extracellular tomato activities RNases LX (Bosse and Keck, 1998) and LE (Niimberger et al., 1990), respectively. In addition to the induction of enzymatic activity, it was shown that the LE and LX mRNAs also accumulate upon P, starvation (Keck et al., 1995). The genes encoding the Nicotiana alata RNase NE (Dodds et al., 1996), and the Prunus dulcis RNase PD2 (Ma and Oliveira, 2000), both of which are highly similar to RNS], are also responsive to low Pi. ln-depth analyses of signal transduction pathways controlling Pi-sensing and response to low P, have only recently begun. Studies in the tomato cell culture system support the existence of both intracellular and extracellular mechanisms for sensing P, levels and initiating RNase induction (Glund et al., 1990; Kock et al., 1998). Maintenance of Pi homeostasis in plants is probably controlled by at least two signaling mechanisms, one at the cellular level and another regulating multiple organs and likely originating from the shoots (Raghothama, 2000). There is little direct evidence that plant hormones are involved as primary signals in the Pi-response, but ethylene and auxin may be involved in altering root architecture and root hair growth (Lynch and Brown, 1997). However, cytokinins repress the expression of Pi-starvation inducible genes in Arabidopsis, but do not affect root hair alterations that are controlled by local Pi concentrations, indicating that cytokinins are possibly involved in the repression of systemically controlled Pi-starvation responses (Martin et al., 2000). Cytosolic calcium levels and the activity of Ca2+-ATPases are thought to be involved in regulating the response and adaptation of plants to low P, (Lynch and Brown, 1997). Homeodomain leucine zipper proteins have been shown to bind to the phosphate response domain of the soybean VspB gene, placing these types of transcription factors as likely candidates for effectors of gene induction (Tang et al., 2001). Recent screens have identified mutants in the Pi-response pathway that should allow us to dissect the signal transduction pathways in detail in the future. At least one of these mutants is unable to induce RNSl under low Pi conditions (Chen et al., 2000). Future screens for constitutive Pi-starvation response mutants, such as those proposed by Ticconi et al. (2001 ), will add to the tools available to aid us in the understanding of signal pathways involved in plant molecular and developmental responses to P, limitation. The ever-growing number of RNases demonstrated to be responsive to low Pi has led to the hypothesis that S-like RNases are part of a rescue system employed by plants to recycle Pi when environmental pools are limiting (Goldstein et al., 1989). In conjunction with acid phosphatases and phosphodiesterases, RNases could be involved in the generation of Pi through the hydrolysis of RNA (Glund and Goldstein, 1993). Extracellular secretion of a fungal phytase from Arabidopsis roots has been shown to increase phosphorous nutrition in the plant (Richardson et al., 2001). Similarly, RNases secreted from the root system to the rhizosphere could breakdown organic matter in the surrounding soil into Pi. A similar function has been attributed to two secreted fungal nucleases (Fraser and Low, 1993). Alternatively, RNases could be involved in recycling cellular material from dead or dying cells, either internally (in the case of vacuolar and ER-localized RNases) or externally (in the case of secreted RNases). Experiments done in Arabidopsis have supported this theory. RNSl and RN82 activities were diminished using antisense techniques, and each line was found to have increased anthocyanin contents. The phenotype was particularly dramatic when the seedlings were starved for P. (Bariola et al., 1999). The authors suggested that this effect might be due to the plants' reduced abilities to acquire Pi, especially under low Pi-levels, although it could not be ruled out that the plants were somehow otherwise stressed, leading to high levels of anthocyanins. C ell-death pathways Programmed cell death (PCD) is an integral part of many plant developmental programs and environmental responses, including, but not limited to, senescence, the hypersensitive response (HR) during defense against pathogens, pollination, and germination (for recent reviews, see Beers and McDowell, 2001; Jones, 2001). There are numerous examples of S-like RNase and nuclease l-type activities induced during such processes. The final stages of xylogenesis have been a paradigm for studying PCD mechanisms (for a review of xylogenesis, see Roberts and McCann, 2000). This process has been well-studied in the zinnia cell culture system, in which mesophyll cells can be induced to transdifferentiate into tracheary elements (TE; reviewed in Fukuda, 2000). This process is accompanied by induction of nuclease activities, including a 43-kDa protein that is apparently involved in autolysis, the final stage of vascular differentiation in plants (Thelen and Northcote, 1989). Two cDNA clones, ZRNaseI and ZRNaseII, have been isolated from differentiating zinnia cells and encode RNases with similarity to the S-like RNase family. ZRNaseI is induced during late stages of TE differentiation (Ye and Droste, 1996). The in vivo function, as well as the substrates, of nuclease and RNase activities during TE formation is still a matter of speculation. Nuclease and RNase activities are also associated with other cell death processes, such as senescence and the hypersensitive response. Pollination-induced petal senescence in petunia causes increases in RNase and ssDNase activities (Xu and Hanson, 2000). Activities of the nuclease I family are induced in senescing tissues in Arabidopsis and zimria (Pérez-Amador et al., 2000). In Arabidopsis, RN52 and RN53 transcripts accumulate during leaf senescence (Taylor et al., 1993; Bariola et al., 1994). The gene encoding RNase LX is also induced in senescing tomato leaves (Lers et al., 1998), and the protein is expressed during xylem differentiation, germination, and senescence (Lehmann et al., 2001). In tobacco, the hypersensitive response resulting from both TMV and bacterial infection is associated with induction of extracellular nuclease activities (Mittler and Lam, 1997). Barley aleurone cells contain several nuclease activities and an S-like RNase that are induced by gibberellic acid and repressed by abscisic acid (Brown and Ho, 1987; Path et al., 1999; Rogers and Rogers, 1999). The timing of induction of these activities correlates with hydrolysis of nuclear DNA and the progression of cell death. It has been hypothesized that the barley nucleases are involved in the degradation of the endosperm in response to seed germination (F ath et al., 1999). It is possible that RNase and nuclease activities are critical components of PCD, required for the progression of the developmental pathway. It has been suggested that nucleases could be involved in the degradation of nuclear DNA during PCD (F ath et al., 1999; Mittler and Lam, 1997). However, most activities studied appear to be either 10 extracellular or associated with the vacuole or ER. Thus, they may be involved in late stages of the process, after the commitment step of vacuolar lysis, for example (Obara et al., 2001), or alter lysis of the cell itself. In such a role, they could be components of recycling pathways engaged in order to mobilize useful molecules from dying cells to growing parts of the plant. A more interesting hypothesis has been proposed by Lehmann et al. (2001). After demonstrating that RNase LX is retained in the ER, they suggested that hydrolytic enzymes, such as proteases and RNases, found in distinct ER domains or ER-derived structures such as ricinosomes play an important role in controlled degradation processes. Upon swelling and rupture of ER- or ER-derived membranes following disruption of the tonoplast (Fukuda, 1997), RNase LE and other hydrolytic enzymes would be brought in contact with its RNA substrate in the cytoplasm (Lehmann et al., 2001). Similar mechanisms may occur in other plant species during many degradative processes, such as senescence, endosperm mobilization, and xylem differentiation. RNases and defense Increases in RNase and nucleases activities during pathogen invasion have been found in several systems (for references, see Green, 1994). Two parsley pathogenesis-related (PR) proteins induced by a fungal pathogen show homology to a ginseng RNase (Moiseyev et al., 1994). These proteins are part of the large PR-lO family of intracellular plant PR proteins, and are not related to the S-like RNase or nuclease I family, but do support a role for RNases in defense responses. In Nicotiana tabacum, transcripts for the S-like RNase NE are induced by inoculation of leaves with the oomycete Phytophthora (Galiana et al., 1997). Interestingly. exogenous application of RNase A inhibited ll development of this pathogen on the leaves by as much as 90%, without eliciting general defense responses such as HR. Additionally, RNase A inhibited the formation of lesions caused by tobacco mosaic virus (TMV; Galiana et al., 1997). A 30-kDa protein sharing similarity with a tomato RNase has been isolated from leaves of Engelmanm'a pinnatifida and shown to have broad-spectrum antifungal activity (Huynh et al., 1996). An extracellular RNase activity in rust-infected wheat leaves has also been reported (Barna et al., 1989), and as many as 20 isoforrns of RNases are induced in pearl millet during SAR after challenge with downy mildew disease (Shivakumar et al., 2000). RNSl antibody (Bariola et al., 1999) appears to cross-react with a protein induced by SAR in this system (P.D. Shivakumar and V. Smedegaard-Petersen, personal communication). RNases are also induced by wounding in several systems. In tomato, for example, the transcript for RNase LE accumulates in wounded leaves (Lers et al., 1998). RNase NW, a tobacco protein with high sequence similarity to RNase NE, is also induced in wounded leaves (Kariu et al., 1998). Since both pathogen challenge and mechanical wounding have effects on RNase activity, it is possible that RNA-degrading enzymes have roles in defense responses. During wounding, RNases may function in a manner similar to their role in Pi-recycling, including recycling of nucleotides from damaged cells. They may also be involved in rebuilding of damaged vascular tissue, consistent with their induction during xylogenesis. RNases and nucleases may play a more specific role as defensive proteins to protect damaged or particularly susceptible tissues from pathogen invasion. This hypothesis is supported by the high expression levels of several RNase transcripts in pistils. Despite the fact that these structures are rich in nutrients, making them likely targets for pathogen invasion, they are rarely infected, possibly due to the presence of various defense-related proteins (Atkinson et al., 1993). These defensive proteins may include S-like RNases, as S-like genes are expressed in pistils and flower structures of Arabidopsis (RN51, RN52, and RN53; Taylor et al., 1993; Bariola et al., 1994), tobacco (NE; Dodds et al., 1996), and petunia (X2; Lee et al., 1992). New ideas regarding secreted RNase functions Recent results from our laboratory have prompted new ideas regarding functions for secreted RNases. These ideas have come from the study of the only member of the RNase T2 family found in Saccharomyees cerevisiae. Inactivation of this RNase, Rnyl, resulted in unexpected phenotypes: the mutant cells are larger than wild-type, grow as cell- aggregates, and are heat- and salt-sensitive. Complementation of the heat and salt- sensitivity phenotypes was accomplished using Rnyl , three of the Arabidopsis RNSs, and the structurally unrelated enzyme RNase A. Mutation of RNase A that caused a loss of RNase activity but retained the protein's stability and structure resulted in the inability of the protein to complement the growth phenotype (Maclntosh et al., 2001). These phenotypes, in combination with recent experiments that demonstrated that specific RNAs can bind to membranes and thereby alter membrane permeability (Khvorova et al., 1999), led to the hypothesis that secreted RNases can control membrane stability and permeability by regulating the amount of RNA in the membrane (Maclntosh et al., 2001). More recent experiments have also shown that exogenous application of RNase A to the culture medium complements the flocculation phenotype of the rnylA cells (G.C. Maclntosh and P.J. Green, unpublished). Thus, it appears that an extracellular RNase activity can affect membrane characteristics in yeast. Dissertation topic and thesis overview In summary, T2 RNases and nucleases I enzymes are widely distributed in nature and tightly regulated in response to a variety of signals. Yet in most cases, their biological roles and substrates remain elusive. The main objective of my researach was to determine the biological significance of extracellular RNases in plants. To address this question, several methods have been employed: studying the mode of regulation of the Arabidopsis S-like RNase, RNS 1; identifying signals that direct the expression of RN51 ; analyzing the effects of mutation and overexpression of RN51; and examining a mutant with altered RNase activities. Identifying signals in addition to P. starvation that affect RN51 expression should help us deduce the processes in which RNS] activity is important. Analysis of mutant and overexpressing plants during these processes may then reveal functions of RNSl. Chapter 2 will detail my findings that RNSI and other RNA- degrading activities are regulated by wounding in Arabidopsis through a novel signaling pathway. The regulation of RNSl will be firrther analyzed in Chapter 3, including other signals that control RN51 expression and their effects on the RN51 promoter. Functional questions will be addressed in Chapter 4, in which I will summarize recent data derived from the analysis of plants lacking RNSl and of plants with constitutively high RNSI activity. Finally, in Chapter 5, I will present my contribution to the analysis of an Arabidopsis mutant that has high levels of several RNase activities. 14 CHAPTER 2 LOCAL AND SYSTEMIC WOUND INDUCTION OF RNASE AND NUCLEASE ACTIVITIES IN ARABIDOPSIS: RNS] AS A MARKER FOR A JASMONIC ACID-INDEPENDENT SYSTEMIC SIGNALING PATHWAY A version of this chapter is in press: LeBrasseur, N.D., Maclntosh, G.C., Pérez-Amador, M.A., Saitoh, M., and Green, P.J. (2002) Local and systemic wound-induction of RNase and nuclease activities in Arabidopsis: RN51 as a marker for a JA-independent systemic signalling pathway. Plant J. In press. 15 ABSTRACT Induction of defense-related genes is one way plants respond to mechanical injury. We investigated whether RNases are involved in the wound response in Arabidopsis thaliana. As in other plant systems, several activities are induced with various timings in damaged leaves, stems, and seedlings in Arabidopsis, including at least three bifunctional nucleases capable of degrading both RNA and DNA, as well as RNSI, a member of the ubiquitous RNase T2 family of RNases. The strong induction of RN51 expression is particularly interesting because it occurs both locally and systemically following wounding. The systemic induction of this RNase indicates that members of this family may be involved in defense mechanisms in addition to their previously hypothesized functions in nutrient recycling and remobilization. Additionally, the systemic induction appears to be controlled independently of j asmonic acid, and the local induction of RNSl and the nuclease activities is independent of both JA and oligosaccharide elicitors. Consequently, a novel systemic pathway, likely involving a third signal, appears to exist in Arabidopsis. 16 INTRODUCTION One of the mechanisms by which plants respond to wounding, either mechanical or feeding-related, is the activation of transcription of a variety of genes. The products of these genes include defensive proteins such as proteinase inhibitors (Boulter, 1993; Ryan, 1990), components of signal transduction pathways, for example, systemin in tomato and potato (Ryan and Pearce, 1998), proteins that may be involved in wound healing (Bowles, 1990), and other proteins whose functions in the wound response are as yet unknown. While the expression of some of these genes is proximal to the injury site, others are expressed at greater distances and constitute the systemic response (Green and Ryan, 1972). In tomato, multiple signals appear to be involved in activating gene expression. Local responses are thought to be controlled by carbohydrate signals released from injured plant cell walls (Farmer and Ryan, 1992). Long-distance effects are initiated by the small peptide systemin (Ryan and Pearce, 1998). Both oligosaccharides and systemin initiate the de novo production of abscisic acid (ABA) and jasmonic acid (JA), which leads to accumulation of the transcripts of wound-responsive (WR) genes (Leon et al., 2001). Ethylene also appears to be involved in the amplification of systemin-activated signaling (O’Donnell et al., 1996). In addition, electrical signals have been implicated in the systemic response (Wildon et al., 1992). In general, wound signals in tomato are thought to constitute a unified reaction that mounts both local and systemic defense and repair responses (Leén et al., 2001), although a .IA-independent WR gene has been identified (O’Donnell et al., 1998). In Arabidopsis, the emerging picture has striking differences in comparison to the 17 tomato system. In Arabidopsis, two distinct signaling pathways can be identified. Some genes whose transcripts accumulate upon wounding are independent of IA, while the expression of others requires jasmonate synthesis and perception (Titarenko et al., 1997). Recently, the JA-independent response was shown to be activated by oligosaccharides (Rojo et al., 1999). It was suggested that the two pathways work in an antagonistic manner, so that the inhibition of the JA-dependent pathway occurs in local wounded leaves through the induction of ethylene by the oligosaccharide elicitors. Conversely, only JA-dependent genes would be activated systemically, due to the limited diffusion range of the oligosaccharides (Rojo et al., 1999). However, a JA-dependent WR gene has been found that is strongly expressed both locally and systemically in Arabidopsis (Kubigsteltig et al., 1999). In other systems, RNases have been shown to accumulate in response to mechanical wounding or pathogen attack, as discussed in Chapter 1. Although several wound- responsive RNases have been identified, the signaling pathways responsible for RNase induction are rarely addressed. In order to investigate the effects of wounding on the expression of the S-like family of Arabidopsis RNases, we examined the RNase profile of wounded plants. We found not only T2 enzymes, but also multiple activities increase in damaged tissues in response to mechanical wounding, including several with both RNase and DNase activity. Within the RNase T2 family, RNSl activity increases, and the RN51 transcript accumulates in both wounded and systemic unwounded leaves of wounded plants. Unexpectedly, we found that the regulation of these activities by wounding is independent of both jasmonic acid and oligosaccharides, indicating that a novel, unidentified wound signaling pathway may be operating in Arabidopsis. Additionally, the 18 systemic induction of RN51 suggests that the enzyme may have a defensive function. RESULTS Wounding induced several RNase activities in Arabidopsis To examine the effect of wounding on RNase expression in Arabidopsis, 2-week-old seedlings, or stems or leaves of 4-week-old plants were wounded. Samples were harvested at subsequent time points, and protein extracts were analyzed on RNase activity gels (Figure 2-1). Significant alterations in RNase activities were seen in the activity gels, including the activity of one well-characterized member of the RNase T2 family, RNS], and several other activities not likely to be related to this family. In Figure 2-1A, equivalent timepoints were taken from leaves of wounded and unwounded plants to check for any possible light regulation of RNase activities. Several activities can be seen even in unwounded leaves, the most prominent of which is a strong band of activity at 25 kDa; no RNases appear to be regulated in by light (Figure 2-1A, control lanes). Similar controls were taken for stems and seedlings, and there, too, no light/dark regulation is evident (data not shown). In contrast, in leaves, stems, and seedlings (Figures 2-1A, 2-lB, and 2-1C, respectively), multiple activities were induced by wounding, including an activity that migrated at approximately 35 kDa (Figures 2-1A — 2-1C, top arrow), two at 33 kDa (middle arrow; separation of the two 33-kDa bands can be seen in Figures 2-2A and 2-2B), and a fourth of approximately 23 kDa (bottom arrow, labeled RNSl). Past studies have indicated that the activity of a well-characterized RNase T2 enzyme, RNSl, is induced by phosphate starvation in seedlings and migrates at 23 kDa on RNase activity gels (Bariola et al., 1994, 1999). In the wounded samples, a 23-kDa l9 A Control Wounded kD MW1 36122436481 36122436 48h 43— Figure 2-1. Multiple RNase activities are induced by wounding Activity gels showing the increase in RNase activity in Arabidopsis at various time points after wounding. Bands seen on the gel are areas in which RNA cast in the gel has been degraded. Each lane contains 100 pg of protein from wounded leaves (A), stems (B), and seedlings (C). Time, in hours from wounding to harvesting, is indicated above each lane of the gels. Arrows to the right of the gels indicate increased activities. Control lanes contain proteins from unwounded leaves. MW, molecular weight marker (Gibco-BRL). Sizes of the MW markers in kDa are shown to the left in (A). hpw, hours post wounding. activity was also induced (Figures 2-1A —- 2-1C; bottom arrow, labeled RNSl). This activity was seen within 3 h and remained elevated for at least 48 h after wounding. RNA gel blot analysis indicated that the RN51 transcript was induced in wounded leaves (see below). Additionally, in an insertional knock out mutant of RN51, the 23-kDa activity was not induced after wounding (see Chapter 4). Taken together, RNA gel blot analysis and mutant phenotypes confirmed that the 23-kDa activity induced by wounding was indeed RNSl. Thus, as in other plant systems, RNase T2 enzymes are also induced by wounding in Arabidopsis. However, our results also demonstrate that RNase activities not related to T2 enzymes respond to wounding in Arabidopsis. A pair of 33-kDa activities induced in all three wounding experiments has a time course of induction similar to that of RNS 1. In leaves, the doublet of activity at 33 kDa is induced within 6 h (Figure 2-1A, middle arrow). In stem tissue and in seedlings, the upper band of the doublet is evident even in unwounded samples. Still, both bands are enhanced afier wounding, reaching an apparent maximum by 24 h (Figures 2-lB and 2- 1C), and high levels of these activities can be found as late as 48 h post wounding (hpw). Interestingly, unlike the sustained induction of the 23- and 33-kDa RNases, a fourth activity, at 35 kDa, is induced transiently. In leaves, this activity appears within 6 h of wounding (Figure 2-1A). In stems and in seedlings, the 35-kDa activity appears more rapidly, with induction evident within 3 h (Figures 2-1B and 2-1C). In all three cases, the activity is no longer visible by 24 hpw. Unlike RNSl, these wound-inducible activities exhibit characteristics that distinguish them from the RNase T2 family, as discussed below. Multiple bifunctional nuclease activities are increased by wounding Previous studies using the gel assay system shown in Figure 2-1 demonstrated that a doublet of bifirnctional nucleases capable of degrading both RNA and DNA migrate at 33 kDa; these nucleases are normally expressed at low levels but are upregulated in a recently isolated mutant of Arabidopsis (M.L. Abler, N.D. LeBrasseur, L.R. Danhof, D.M. Thompson and P.J. Green, manuscript in preparation; see Chapter 5). To investigate whether the 33-kDa RNase activities increased by wounding are bifunctional nucleases, protein extracts of wounded stem tissue were run on 9% acrylamide RNase and DNase gels. While other activity gels in this study contained 11% acrylamide, the lower acrylamide percentage used in these gels results in better separation of the two activities that migrate at approximately 33 kDa, as can be seen by comparing Figures 2-1B and 2- 2A. Additionally, the increase in the intensity of the bands seen in the RNase activity gel at 33 kDa is mirrored in the DNase gel (Figure 2-2A), confirming that this doublet has bifunctional nuclease activity. In addition to the 33-kDa doublet, another activity that increases upon wounding appears to have bifunctional activity. The transient 35-kDa RNase activity seen in Figure 2-1 also degrades DNA (Figure 2-2A). Therefore, wounding causes an induction not only of T2 RNases in Arabidopsis, but also of several nuclease activities, including a previously unidentified nuclease at 35 kDa. Past studies of RNase expression profiles in Arabidopsis have shown that cultivar variation is present. Specifically, the lower band of the 33-kDa doublet can normally be seen at 10w levels in the cultivar RLD but is absent in Columbia (M.L. Abler, L.R. Danhof, and P.J. Green, unpublished). However, the lower band can be induced in A 01361224hpw B Unwounded Wounded RLD Col RLD Col Figure 2-2. Bifunctional nuclease activities are induced by wounding. (A) Protein samples (100 ug for RNase gel; 75 ug for DNase gel) from wounded stems were run on activity gels containing either RNA (upper panel) or DNA (lower panel) as a substrate. Both gels contain 9% acrylamide in order to separate the two activities at 33 kDa. Arrows to the left of the gels indicate wound-regulated bifunctional activities. hpw, hours post wounding. (B) Stem protein samples (100 ug) fi'om both Columbia (Col) and RLD (RLD) ecotypes were run on RNase activity gels. Unwounded samples are shown on the left, and stem samples taken 3 h after wounding are on the right. The lower band (lower arrow) is not found in unwounded Columbia stems but is present after wounding. Arrows indicate the upper and lower bands of the 33-kDa nuclease doublet. 23 Columbia upon wounding (Figure 2b). Thus, wounding indicates that although this activity is differentially regulated under nonwounded conditions in Columbia and RLD, it is not altogether absent from Columbia. RN 81 activity is induced by wounding locally and systemically As seen in Figure 2-1, an RNase activity that migrates at approximately 23 kDa is induced during wounding. During phosphate starvation in Arabidopsis seedlings, a 23- kDa activity is also induced (Bariola et al., 1994). The increase in this activity is due to an increase in the protein levels of RNS] (Bariola et al., 1999) and corresponds to an increase in transcript accumulation (Bariola et al., 1994). Additionally, as mentioned, wounded leaves of an RN51 T-DNA insertion mutant lack the 23-kDa activity, (see Chapter 4). Therefore, we were interested in examining whether the RN51 transcript also accumulates in response to wounding. Rosette leaves of adult plants were wounded and then harvested at subsequent timepoints. RNA gel blot analysis demonstrated that the RN51 transcript accumulated within 3 h and was still elevated as late as 48 hpw (Figure 2-3A). Immuno-blot analysis confirmed that RNS] protein accumulated with a time course similar to that of the activity increase (data not shown). We also investigated whether the Arabidopsis response to wounding includes systemic induction of RNase or nuclease activities in non-wounded tissue. Unwounded rosette leaves adjacent to wounded leaves were harvested, and RNA was examined by RNA gel blot. The RN51 transcript was induced in unwounded leaves. The increase was transient and rapidly disappeared between 6 and 8 hpw (Figure 2-3B), unlike the more sustained increase of RN51 in wounded leaves (Figure 2-3A). Wounded seedlings (9 hpw; Figure 2-3 B, lane W) were included in the blot for a comparison of transcript levels A 013 6 12243648 hpw RN81 B hpw 0 1 3 6 8 12 W C Systemic Local 013681213612hpw Figure 2-3. RNSI activity and transcript increase transiently in wounded and in systemic, non-wounded leaves. (A) Northern blot of wounded leaves. Arabidopsis leaves were wounded and harvested at subsequent timepoints. RNA was extracted, and 10 pg was separated by electrophoresis, blotted, and probed with RNSI. Blots were then stripped and reprobed with EF-Ia to control for loading. hpw, hours post wounding. (B) Systemic induction of the RN51 transcript shown by RNA gel blot analysis. Unwounded rosette leaves adjacent to the wounded leaves were harvested at the timepoints indicated above the lanes. The gel contained 10 ug total RNA per lane. RNA from seedlings harvested 9 h after wounding was included for comparison (W). The blot was stripped and reprobed with EF -1 a as a loading control. hpw, hours post wounding. (C) RNase activity gel showing systemic induction of RNSI activity. Each lane contains 100 ug protein. The arrow indicates the RNSI band. No other induced activities were detected in systemic leaves. Protein samples from wounded leaves were included for comparison of RNSI activity levels. 25 induced locally and systemically. Additionally, protein extracts were examined by RNase activity gel assay to determine whether RNSI or nuclease activities are induced in unwounded leaves of wounded plants. In the unwounded leaves, the 23-kDa RNSI band was induced (Figure 2-3C), indicating that RNases may play a role in the systemic wounding response in Arabidopsis. As in wounded leaves, the systemic RNSI activity increase could be seen slightly later than the transcript accumulation: in multiple experiments, we observed a slight increase in this activity between approximately 6 and 12 hpw. Wounded samples were included in the gel for comparison (W, Figure 2-3 C). No other bands aside from RNSI were induced systemically (not shown). Note that the activity gel in Figure 2-3C was incubated at pH 6.0, instead of pH 7.0, to enhance RNSI activity, since no other activities were seen to be induced systemically at pH 7.0 (see Materials and Methods). As a result, the relative activity of RNSI is higher compared to the constitutive 25-kDa band. It is interesting that the systemic activity increase does not appear to reflect the amount of transcript induced (compare 3 and 6 h in Figure 2-3B with 6 and 12 h in Figure 2-3C). The discrepancy may be due to a possible posttranscriptional regulation of RNS]. The induction of RN51 and the nucleases is independent of jasmonic acid Wounding responses in Arabidopsis are controlled by jasmonic acid (JA)-dependent and -independent signaling pathways (Titarenko et al., 1997). The presence of distinct wounding pathways in Arabidopsis has led to the proposal that the function of multiple signals is to control local and systemic gene expression differentially. Specifically, it has been suggested that the JA-dependent pathway controls induction of systemic responses. while an oligosaccharide-dependent pathway directs gene activation locally at the Figure 2-4. RNSI and nuclease induction by wounding is independent of jasmonic acid. (A) Top panel; activity gel showing induction of all four wound-responsive activities in both the wild type (WT) and the JA-insensitive coil mutant (coil). Arrows to the right indicate the induced activities, including the nuclease at 35 kDa (top arrow), the doublet at 33 kDa (middle arrow), and RNSI (bottom arrow). Bottom panel; systemic induction of RNSI in the coil mutant. hpw, hours post wounding. (B) Northern blot analysis of wounded wild type (WT) and coil (coil) plants. The blot contains RNA (10 pg) from unwounded leaves (U), local 6-h wounded leaves (L), and unwounded leaves of wounded plants 3 h after wounding (S). The blot was probed with RN51, stripped, and reprobed with EF -1 a as a loading control. (C) Activity gels containing 100 pg of protein from leaves of JA-treated (JA) or untreated (control) plants. Plants were treated with methyl jasmonate or buffer alone, and leaves were harvested at 0, 3, 6, and 24-h timepoints. W, protein from leaves 24 h after wounding. Arrows to the right of the gel indicate the running positions of wound- inducible activities. (D) Northern analysis of RNA (10 pg) from leaves of JA-treated (JA) or untreated (control) plants. Plants were treated as in (c). 2- and 24-h timepoints are shown. Similar results were seen at 3 and 6 h (data not shown). W, RNA from leaves 6 h after wounding. Blots were probed with RNSI, then sequentially stripped and reprobed with A05 to control for JA-treatment and with EF -1 a as a loading control. Local 0011 06240624hpw Systemic coi1 0 3 6 8 12 hpw . RN81 C Control JA 0 3 6 24 3 6 24 W h an {—33 kD <—RNS1 Control JA 036243624W h . RN81 a G... Aos EF-10t Figure 2-4. RNSI and nuclease induction by wounding is independent of jasmonic acid. 28 site of damage (Rojo et al., 1999). Under this assumption, RN51 would be expected to be regulated by the JA-dependent pathway, as its transcript is induced systemically. In contrast, the bifunctional nuclease activities, which are not systemically wound- responsive, would be regulated independently of IA. The JA-insensitive mutant coil (Feys et al., 1994) has been widely used to demonstrate dependence of the wound- induction of specific transcripts on JA (e.g., Reymond et al., 2000; Titarenko et al., 1997). We investigated whether the wound induction of RNSI or the nuclease activities are regulated by the JA-dependent pathway by wounding leaves of coil plants. As expected, like the wild type, wounded coil plants displayed increased levels of the 35- and 33-kDa activities (Figrre 2-4A). The transient 35-kDa nuclease was seen in both wild-type and mutant plants at 6 hpw, and the nuclease bands at 33 kDa were induced at 6 and 24 hpw in both wild-type and coil plants. However, wounded coil leaves unexpectedly also had increased RNSI activity. In fact, in contrast to systemic responses previously characterized in Arabidopsis, both local and systemic wound induction of RN51 is independent of IA. RNSI activity was induced in mi] local wounded tissue (Figure 2-4A, upper panel) and in unwounded leaves of the same plants (Figure 2-4A, lower panel). RNA gel blot analysis demonstrated that the RN51 transcript also accumulated in the coil mutant upon wounding (Figure 2-4B). Local 3-h and systemic 6- h timepoints are shown (Figure 2-4B, L and S, respectively). Systemic RNSI levels at 6- and 24-h timepoints were also similar in coil and wild-type plants (data not shown). To confirm that JA does not affect the induction of RNSI or the nucleases, young plants were treated with methyl jasmonate (MeJA) and examined for RNase activity (Figure 2-4C). Efficacy of the MeJA treatment was shown by the accumulation of the allene oxide synthase transcript (A05), a component of the jasmonate pathway known to be induced by JA (Laudert and Weiler, 1998; Figure 2-4D). However, MeJA treatment did not cause RN51 accumulation after 3, 6, or 24 h of treatment (Figure 2-4D). Likewise, the 35- and 33-kDa nucleases, as well as RNSI, were not induced by MeJA at the activity level (Figure 2-4C). Thus, we have demonstrated that the wound-inducible RNase and nuclease activities are not controlled by the JA-dependent signaling pathway. Recent results from an independent microarray experiment also demonstrated that the local expression of RN51 is induced in a JA-independent manner (Reymond et al., 2000), supporting our observations. RN51 is not induced by the oligosaccharide pathway It has been suggested that the JA-independent wound response in ArabidOpsis is controlled by oligosaccharide elicitors. Transcripts whose accumulation after wounding is not dependent on JA can be induced by oligogalacturonic acids (OGAs), proteinase- inhibitor inducing factor (PIIF) fractions from tomato leaves, and chitosan (Rojo et al., 1999), all of which are rich in oligosaccharides. To test if one of these known inducers of certain JA-independent wound-inducible genes would also induce RN51 transcript accumulation, an OGA-rich fraction known as 'TFA-PIIF' (generously provided by Dr. E.E. Farmer; purified according to Bishop et al., 1984) was used to treat detached rosette leaves. Leaves were floated in an MS medium either with or without the addition of TFA- PIIF and harvested at subsequent timepoints. RN51 transcript levels were analyzed by RNA gel blot analysis (Figure 2-5A). As seen in Figure 2-5A, the slight wounding involved in detaching leaves caused a 30 A +PIIF -PllF 01.5361.536Wh B +PIIF -Pl|F o 3 s 3 s h Figure 2-5. OGA-rich fractions do not induce RNase activities. Leaves were detached from 4-week-old wild type plants and floated in buffer with (+PIIF) or without (-PIIF) the addition of TEA-PIIF. Samples were harvested at timepoints indicated above the lanes (in hours). ' (A) Northern blot containing 10 pg RNA per lane. W, RNA from leaves 6 h after wounding. The blot was probed with RN51, then sequentially stripped and reprobed with Choline kinase (CK) to control for PIIF-treatment and with ch-4A as a loading control. (B) Activity gel containing 100 pg protein per lane. Arrows to the right of the gel indicate the normal running positions of the wound-inducible activities. 31 higher than normal basal level of transcript in the samples after 3 h. However, in contrast to most known Arabidopsis JA-independent wound-induced transcripts, the presence of TFA-PIIF in the medium did not induce RN51 compared to the control (Figure 2-5A). Blots were also probed with choline kinase (CK), an OGA-induced transcript (Rojo et al., 1999), to demonstrate efficacy of the OGA treatment. The wounding involved in the detachment technique induced CK expression at the 90 min timepoint, regardless of the presence or absence of OGAs and consistent with the timing of CK wound-induction seen previously (Rojo et al., 1999). However, as expected, the TFA-PllF-treated samples had increased levels of CK compared to the control samples at both 3- and 6-h timepoints (Figure 2-5A). Thus, while CK levels increased due to the treatment, RN51 levels were not affected by the PIIF fraction. Interestingly, the wound-inducible nuclease activities, as well as RNSI activity, were also not induced by the treatment, as seen by RNase activity gel assay (Figure 2-SB). Throughout the time course, the only activity detected was a major band at 25 kDa, which was not regulated by wounding. This activity was also not affected by the TFA-PIIFS, and no other activities were induced, including the wound-responsive 35- and 33-kDa nucleases and RNSI. Thus, although the treatment was effective, induction of RNase activity by wounding is not mediated by oligosaccharides. These results were unexpected, considering that all the activities were also independent of J A. Therefore, we hypothesize that in addition to oligosaccharides and JA, an additional signal exists that directs expression of nuclease and RNase activities in wounded tissue. DISCUSSION In this chapter, we have demonstrated that several Arabidopsis RNase and nuclease 32 activities are coordinately regulated by wounding. Unlike other wound-inducible proteins previously studied in Arabidopsis, these activities are not controlled by jasmonic acid signaling or by oligosaccharide elicitors. The systemic induction of RNSI makes this ribonuclease especially intriguing. The systemic induction of RNSI not only supports the existence of a novel pathway for the regulation of systemic wound responses, but is also suggestive of a defensive role for this RNase. The increases in all the activities occur in a coordinated manner and provide us with a unique perspective into Arabidopsis wound signaling mechanisms. Our understanding of the wound response in Arabidopsis is currently highlighted by the presence of two distinct, antagonistic pathways: JA-dependent and -independent. The JA-independent pathway controls local induction of transcript accumulation and has been shown to be regulated by OGA elicitors probably released from injured plant cell walls (Rojo et al., 1999). The three nuclease activities and RNSI are strongly induced locally by wounding (Figures 2-1 and 2-2). However, they are not induced by treatments with OGA-rich fractions (Figure 2-5). In fact, in several repetitions of the OGA treatments, the slight wound-induction of RNSI caused by detaching the leaves appeared to be inhibited by the presence of the OGAs (for example, compare +PIIF and -PIIF at 6 h in Figure 2-5A). Given that the RNase and nuclease activities are not induced by the PIIF treatments, we conclude that another signal besides OGAs is likely operating to direct the local induction of these activities upon wounding. The local response of RNSI and the nucleases to wounding was also not controlled by the JA-dependent signaling pathway, as shown by the strong wound-induction of these activities in the coil mutant (Figure 2-4A). Further, RNSI transcript accumulated to high levels in unwounded leaves of wounded plants. IA is an important systemic signaling molecule during wound-induced gene induction in Arabidopsis (Titarenko et al., 1997). However, the systemic induction of RN51 did not depend on JA (Figure 2-4B). To our knowledge, RN51 is the first gene in Arabidopsis shown to be induced systemically by wounding in a JA-independent manner and therefore indicates the existence of an alternate long-distance signaling pathway. It has been proposed that dehydration itself might be the causal induction factor of some JA-independent wound-inducible genes, including RN51, which was shown by microarray analysis to be induced by dehydration (Reymond et al., 2000). However, the dehydration experiments performed involved detaching the rosette leaves from the roots, which itself induces RN51 (see Figure 2-5). My results indicate that dehydration itself does not induce RN51 expression but may potentiate the wound induction (see Chapter 3). The induction of certain wound-inducible genes in potato and tomato are affected by ABA levels (Pena-Cortes et al., 1989). Therefore, ABA could be considered a possible signal for the wound induction of RN51. The involvement of ABA in the wound response will be discussed further in Chapter 3. Given its involvement in the wound response in solanaceous plant species and in the regulation of various programmed cell death processes (Jones, 2001), ethylene is another candidate for the signal involved in the induction of RN51 both locally and systemically. However, the wound induction of RN51 also occurs in the ein2 mutant, suggesting that ethylene perception is not required (Reymond et al., 2000). Salicylic acid (SA) is required for many plant-pathogen defense responses (reviewed in Alvarez, 2000), and interaction between SA, J A, and ethylene pathways is emerging as an important regulatory method 34 for activating multiple resistance mechanisms (reviewed in Pieterse and Van Loon, 1999). However, SA does not appear to be involved in the regulation of RN51 by wounding, as the transcript is still induced in SA-deficient, NahG-expressing plants (E.E. Farmer, personal communication). Additionally, analysis of AF GC microarray data in the Stanford Microarray Database available on the web (http://genome-www4.stanford.edu/ MicroArray/SMD/; Wisman and Ohlrogge, 2000) indicates that RN51 is not induced by treatment with the SA-analog BTH or by various pathogens that induce systemic acquired resistance, which is known to cause the accumulation of SA within the plant (Alvarez, 2000) Other possible signals for the induction of RN51 include reactive oxygen species (ROS). ROSs are commonly produced in plants in response to both pathogen and herbivore attacks (Grant and Loake, 2000; Orozco-Cardenas and Ryan, 1999). In tomato, H202 acts as a second messenger for the induction of defense—related genes induced systemically at later timepoints than the earlier, signaling-related genes (Orozco- Cardenas et al., 2001). The timing of RNase and nuclease responses coincides with these later responses that are dependent on H202 signaling. The involvement of ROSS, as well as other possibilities, for example, electrical signals, in RNase induction during the Arabidopsis wound response will prove an important avenue of study for defining alternate pathways regulating defense responses. The systemic increase in RNSI appears to be regulated at multiple levels. Studies with reporter constructs controlled by the RN51 promoter region indicate that the local increase in RN51 transcript caused by wounding is likely due to transcriptional regulation (see Chapter 3). However, while the transcript increases to significant levels systemically, 35 the activity increase is not as great (compare Figure 2-3B, 3 and 6 h, with Figure 2-3C, 6 and 8 h). It is possible that a post-transcriptional mechanism may also exist that regulates the amount of RNSI protein translated or the activity of the translated protein in the unwounded tissue. In addition to its interesting regulatory properties, the systemic induction of RNSI implies that the RNase may have an important function during the wound response. It is striking that the RN51 transcript was the most highly induced in local wounded tissue of all transcripts examined in two independent microarray experiments: one examined 150 genes enriched for those implicated in defense responses (Reymond et al., 2000), and the second examined 600 genes, approximately half of which were hypothesized to be involved in RNA metabolism and RNA turnover (M.A. Pérez-Amador and P.J. Green, unpublished). The high level of transcript accumulation of RN51 may be indicative of an important role for the protein product of this gene. Additionally, the transcript and the activity are induced in non-damaged tissue, where recycling of nutrients and degradation of bulk cellular nucleic acid, two of the proposed functions of secreted RNases, should not be necessary. Instead, it is possible that RNSI has a defensive function. Normally, RN51 is expressed solely in flowers (Bariola et al., 1994). The presence of RNases in the pistil may contribute to protection of the structure from pathogens (Bariola et al., 1994). In fact, it has been demonstrated that application of an extracellular RNase to tobacco leaves inhibits growth of both TMV and a fungal pathogen (Galiana et al., 1997). Likewise, local and systemic induction of RNSI could contribute to protection of the plant against invasion of pathogens following wounding. Although it is not known how RNases could achieve such defensive roles, wound signal molecules are known to 36 produce membrane-associated changes within the plant cells, including membrane depolarization (Thain et al., 1995). Recently, it has been proposed that T2 RNases can affect membrane stability or permeability (Maclntosh et al., 2001), suggesting a possible link. The role of RNSI in defense mechanisms will be readily testable in a recently- isolated RN51 T-DNA insertional mutant and in plants that overexpress RNSI activity (see Chapter 4). In contrast to the systemically-induced RNSI, the bifunctional activities were induced only in local, damaged tissue. These activities may also have defensive roles in the local tissue, or they may be functioning in other aspects necessary during the wound response, such as recycling of nucleotides from damaged cells or rebuilding of damaged vascular tissue. Cells damaged by wounding would contain phosphate and other molecules that could be recycled for use in active, growing parts of the plant. Similar functions have been hypothesized for T2 RNases that are induced by phosphate-starvation and senescence in several species, including tomato and tobacco (reviewed in Howard et al., 1998). Activities at 33 kDa, as well as RNSI, are induced in P,-starved seedlings (see Figure 10 in Bariola et al., 1994), and bifunctional nucleases are induced by senescence in Arabidopsis (Pérez-Amador et al., 2000). Possibly, RNSI and the nucleases are induced during wounding as part of this wide-spread recycling mechanism. It is possible that the systemic induction of RN51 could be explained as a sort of "priming" of the plant for nucleotide recycling in as-of-yet undamaged tissue. However, the nucleases, as well as RN52 and RN53, which encode two additional RNases shown to be induced during conditions requiring nutrient recycling, such as P, starvation and senescence (Bariola et al., 1994; Taylor et al., 1993), are not induced locally or 37 systemically by wounding, at least indicating that other hydrolytic enzymes thought to be involved in P, recycling and remobilization are not generally activated by wounding. In fact, RN51 is the only one of the five Arabidopsis genes related to the RNase T2 family that is induced in either local or systemic tissue by wounding (not shown). Such specificity indicates that the enzyme may have an important role in the wound response. Although local induction of hydrolytic activities by wounding has been found in several other plant species (e.g., Galiana et al., 1997; Lers et al., 1998), little is known regarding the signal pathways involved in their regulation. We have now identified a wide array of activities that are regulated by wounding in Arabidopsis, including RNase and nuclease activities that show both sustained and transient patterns of induction. Given its unique responses to wounding, independent of all known wound regulators, RN51 will be a valuable marker for identifying novel signals that operate in the Arabidopsis wound response. More importantly, to our knowledge, RNSI is the first RNase shown to be induced systemically by wounding. This result points us toward new ideas regarding the function of the widespread T2 family of RNases. Specifically, in addition to having a role in P, recycling as has been previously hypothesized, RNSI may have a more direct defense-related function, possibly involved in protection of the plant from firrther attack after a wound-stimulus has been detected. MATERIALS AND METHODS Plant materials and treatments Unless otherwise stated, the Columbia ecotype of Arabidopsis thaliana was used throughout this study. Soil-grown plants were grown in chambers under 16 h of light in 50% relative humidity at 20°C. For seedling experiments, seeds were surface-sterilized and germinated on Arabidopsis growth medium as described (Taylor et al., 1993). The coil seeds were kindly provided by Dr. J.G. Turner (University of East Anglia, Norwich, UK). Mutant coil plants were selected by germinating on MS medium supplemented with 50 pM methyl jasmonate as described (Feys et al., 1994). The plants were then transferred to soil and grown for an additional 4 weeks before wounding treatments were performed. Stems or leaves of 4- to 6-week-old plants or leaves of 14-day-old seedlings were wounded using ridged flat-tipped tweezers and harvested at subsequent timepoints. For non-wounded material, leaves on either side of the wounded leaf were harvested. All samples were frozen in liquid N2 immediately after harvesting and stored at —80°C until used for RNA or protein extractions. Wounding experiments were performed a minimum of three times. Representative blots or gels are shown. Jasmonic acid treatments were conducted on 4-week-old plants that were placed in enclosed boxes. Methyl Jasmonate (MeJ A; Bedoukian Research, Inc., Danbury, CT) was diluted 1:25 in ethanol to a final concentration of 190 mM. A cotton-tipped applicator was soaked with 50 pL of the MeJA solution and placed in the container with the plants. An applicator soaked with ethanol alone was placed in a separate container with control plants. Four boxes were used in each of two experiments, for 2- and 24-h timepoints of both MeJA-treated and control plants. Two additional experiments were conducted for 3-, 6-, and 24-h timepoints, for a total of four replications. Representative blots and gels are shown in Figures 1-4c and l-4d. For OGA-treatments, rosette leaves of 4-week-old plants were removed by slicing the petiole with a razor blade. Approximately 20.1eaves were floated on 40 mL of MS medium (Life Technologies, Rockville, MD) supplemented with 0.5% (w/v) sucrose in each of two 250-mL flasks. One flask also contained 250 pg/mL of the OGA-rich TFA- PIIF fraction (provided by Dr. E. Farmer, Universite' de Lausanne, Switzerland). Flasks were shaken on a rotary platform at ~50 R.P.M. Six or seven leaves were removed from the flasks at 1.5-, 3-, and 6-h timepoints and immediately frozen in liquid nitrogen. OGA treatments were performed twice using this method, and the northern blot and activity gel in Figure 2-5 are representative of this method. Additionally, two replicates of treatments of seedlings grown in liquid culture, according to Rojo et al. (1999), were performed using a PIIF fraction kindly provided by Dr. G. Howe (Michigan State University). Similar results were obtained in all four repetitions; however, the greatest induction of the positive control, choline kinase, was seen using the TFA-PIIF treatments of detached leaves. RNA extraction and RNA gel blot hybridization Total RNA from Arabidopsis samples was extracted as previously described (Newman et al., 1993). RNA (10 pg per lane) was separated by electrophoresis in 3% (w/v) formaldehyde/1.2% (w/v) agarose gels and blotted to Nytran Plus nylon membrane (Schleicher and Schuell, Keene, NH). The RNA blots were hybridized as described in Taylor and Green (1991) using a 32P-labeled RN51 probe. To control for loading, the same RNA blots were stripped in distilled water at 90-95°C for at least 20 minutes and then hybridized with a 32P-labeled probe for the Arabidopsis translation elongation factor EF -1 a (EST accession number R29806) or translation initiation factor eIF-4A (Taylor et al., 1993). A choline kinase probe, kindly provided by Dr. J. Sétnchez-Seranno (Universidad Autonoma de Madrid, Spain), was used as a positive control for OGA 40 treatments (Rojo et al., 1999). A probe for allene oxide synthase (Laudert et al., 1996) was used as a positive control for MeJA treatments. Quantitation of hybridization was performed using Phosphorimager (Molecular Dynamics, Sunnyvale, CA) analysis to confirm increased CK and A05 levels in response to PIIF- and JA-treatments, respectively, as well as increased RN51 levels in local wounded and systemic leaves (data not shown). Protein extraction and detection of RNase and DNase activities Total protein was extracted as described (Maclntosh et al., 1996). Homogenates were clarified by centrifugation, and soluble protein was quantified by the method of Bradford (197 7). 100 pg total protein was loaded in each lane. RNase and DNase activities were assayed using activity gels as described (Yen and Green, 1991), with minor modifications. After the isopropanol wash, before incubation, gels were washed in 100 mM Tris-HCl containing 2 pM ZnCl2 for 20 min in order to restore Zn2+ required for certain RNase and DNase activities. Gels were washed and incubated at pH 7.0, except for the RNase gels of systemic activity in Figures 2-3C and 2- 4A (lower panel), which were incubated at pH 6.0 to enhance resolution of RNSI activity, which is most active between pH 5.0 and 6.0 (data not shown). Similar results were seen at pH 7.0 (data not shown). Separating gels contained 11.3% (w/v) acrylamide, except the gels in Figure 1-2, which contained 9% acrylamide for increased separation of the doublet at 33 kDa. 41 CHAPTER 3 REGULATION OF RNSI EXPRESSION ABSTRACT RNSI activity and transcript have been shown to accumulate in response to wounding (Chapter 2) and P, starvation (Bariola et al., 1994, 1999). I examined whether the induction was a transcriptional or posttranscriptional process using nuclear run-on analysis (in the case of P, starvation) and transgenic plants expressing promoter fusion constructs (wounding). Whereas the run-on analyses were inconclusive, 1 demonstrated that the RN51 promoter was responsive to wound stimuli, as shown by accumulation of the reporter transcript. I then used the promoter constructs to examine other signals that might direct transcription of RN51. Through analyses of these transgenic plants and RNA blot analyses of endogenous RN51 transcript, I found that both ABA and NaCl induced RN51 expression. Interestingly, however, ABA was not required for the wound induction of RNSI activity. Additionally, heat shock appeared to have a negative effect on RN51 induction by wounding. In contrast to previous reports, 1 determined that dehydration did not cause increases in RN51 expression, indicating that another signal directs the induction of RN51 by wounding. INTRODUCTION As seen by results presented in Chapter 2 of this thesis, the levels of RN51 transcript are responsive to external stimuli. In addition to wounding, RNSI also responds to the P, status of its environment as well as to internal cues, as shown by the increase in RN51 transcript in senescing tissues (Bariola et al., 1994). In addition to RN51, genes encoding RNases in many plant species are regulated in response to P, starvation, senescence, and wounding. Although numerous conditions that affect the levels of various RNase activities have been identified in many plant species, little is known regarding the regulatory pathways that control these responses. Previous studies addressed the regulation of RN51 in response to P, starvation; in particular, run-on analyses and transcriptional fusions were used to determine whether RN51 transcript accumulates as a result of transcriptional or posttranscriptional regulatory mechanisms (Howard, 1996). No difference could be detected by nuclear run-ons between seedlings grown on full and P,-deficient media. However, these results were inconclusive, because RN51 transcript levels were not above background. Transcriptional increases were detected by promoter fusion constructs, but the increases were not sufficient to explain the large increase in RN51 levels. Therefore, I repeated nuclear run-on analyses of P,-starved and P,-supplied seedlings to examine the level of regulation of RN51 under these conditions. Additionally, since RN51 transcript is so abundant afier wounding, I was interested in how these levels are regulated during the wound response. I was also particularly interested in other signals involved in regulating RN51 gene expression, given that known Arabidopsis wound pathways are not involved in the response of this gene during 44 wounding (see Chapter 2). RESULTS Run-on analysis of RN51 expression Run-on analyses were attempted to determine whether P, starvation leads to transcriptional induction of RN51. Comparison of run-on transcription assays and RN51 transcript accumulation in seedlings grown in the presence or absence of P, should indicate the contribution of transcriptional regulation to the increase in RN51 resulting from P, starvation. Seedlings were plated on Arabidopsis germination medium (AGM) on mesh circle inserts. On the day of germination (approximately 2.5 days after plating), the mesh inserts were used to transfer seedlings to new plates that either contained (P+) or lacked (P-) P,. After 7 d of growth, seedlings were harvested and used to isolate active nuclei and total RNA. RNA was used for RNA gel blot analysis to confirm that RN51 transcript levels were increased (Figure 3-1 B). Nuclei prepared from the two seedling samples were used for nuclear run-on analyses. Labeled nascent transcripts were purified and used to probe slot blOts containing ssDNA of plasmids of RN51, eIF4A, and empty Bluescript (BSK). The BSK plasmid was used as a control for background hybridization, and eIF4A was used as a loading control, as transcription of this gene is unaffected by the P, status of the seedlings. As seen in Figure 3-lA, hybridization of RNSI, whose expression is regulated in response to P, starvation, is below detection in samples from P,-starved and unstarved seedlings. In multiple repetitions, the hybridization signals obtained were not above background levels detected by the BSK control, although eIF 4A signals were significantly above background. The results indicate that although the nuclei were active, RN51 45 p+ p- M . eIF4A RN81 BSK B p+ p. $2.24 RN81 W W elF4A Figure 3-1. Run-on analysis of RN51 in response to P, starvation. (A) Nuclear run-on analysis was performed to compare transcriptional activity of the RN51 gene in Arabidopsis seedlings grown on medium containing (P+) or lacking (P') P,. (B) Accumulation of RN51 transcript is shown by Northern blot analysis of RNA from the same samples used to isolate nuclei in (A). 46 transcription was not at sufficient levels to measure by run-on analysis. As a result, I could not conclude anything concerning the transcriptional regulation of RN51 in response to P, starvation. Regulation of RN51 expression in response to wounding Since RN51 is so strongly induced by wounding (see Chapter 2), I were interested in whether this regulation is the result of transcriptional control of the RN51 promoter or of a posttranscriptional mechanism. Several constructs were used to examine the regulation of RN51 (Figure 3-2). The first, p1975 (referred to as nos-RNSchNA), included the RN51 cDNA fused between the constitutive, low expressing nos promoter and the pea E9 terminator. This cDNA contains the open reading frame (ORF) plus approximately 60 bases of the 5' UTR (Bariola et al., 1994). The full 5' UTR, as determined by primer extension analysis, contains an additional 40 bases (Howard, 1996). Therefore, I made a second construct, p2019 (referred to as nos-preRNSl), in which intron sequences and the missing bases of the 5' UTR not found in the cDNA were included. Transgenic plants were made using Agrobacterium tumefaciens-mediated transformation via the method of vacuum infiltration (Bechtold et al., 1993). Additionally, transgenic control plants were made by transformation with a control plasmid, p1995, which consists of the nos promoter driving the globin reporter coding region (nos-Globin). To examine the regulation of the transgenes by wounding, leaves of the transformed plants were wounded and harvested 3 h later. Unwounded leaves were also harvested to provide control samples. RNA was prepared from the samples and used for RNA gel blot analysis. Blots were probed with an oligo complementary to the transcribed nos sequences to distinguish between RNSI transcribed from the transgene and the endogenous copy. The RNA gel 47 —_-- ‘- p84811402 f ‘ 353 " ,’ Murals“ paw/1432 2mm; 59 p2081l2082 p1995 1 nos p1975 1 nos p2019 I n05 ‘3‘ preRNS1 Figure 3-2. Constructs used to examine the regulation of RNS l . Several constructs were used to transform wild-type Arabidopsis plants. Transgenic lines were then used to analyze the expression of the reporters under various conditions. Clone identification numbers are shown to the left. LUC, Luciferase coding region; GUS, ,8- glucuronidase coding region; 355, CaMV 355 promoter; nos, nopaline synthase promoter; RN51p, RN51 promoter; E9, 3' end of the pea E9 gene; preRNSI, transcribed region ofRNSl ; The RN51 cDNA was isolated by Bariola et al. (1994). 48 blots indicated no difference between nos signal in wounded and unwounded leaves of plants expressing the nos-RNSchNA construct, although endogenous levels of RN51 were clearly elevated by wounding, as shown by hybridization with the RN51 cDNA (Figure 3-3; center panels). However, in the nos-preRNSl plants, a slight but reproducible increase in nos signal was seen (Figure 3-3; left panels). It is therefore possible that a certain amount of posttranscriptional regulation of RNSI mRNA exists, which requires either the entire 5' UTR region or intron sequences, or both, whereas the cDNA sequence alone is not sufficient. However, this slight increase of approximately 2.5-fold does not account for the high levels of RN51 transcript found in wounded tissues. Since posttranscriptional regulation could not account for the increase of RNSI by wounding, I also examined the ability of the RN51 promoter to confer wound-inducibility to a reporter gene. Constructs were available (Figure 3-1) that contained 2.7 kb of RN51 promoter plus ~60 bases of RN51 5’ UTR sequence fused to either fi-glucuronidase (GUS; p1401) or luciferase (LUC; p1432) reporter genes, as well as control constructs with the same reporters driven by the constitutive CaMV 355 promoter (p848 and pl 402, respectively) (Howard, 1996). These constructs were used to transform Arabidopsis to analyze their regulation by wounding. Regulation of the GUS constructs was analyzed by RNA gel blot. In unwounded leaves of plants transformed with the RNSlp-GUS construct, little or no GUS transcript could be detected. In contrast, 3 h after wounding, wounded leaves contained high levels of GUS (Figure 3-4A). LUC lines revealed that luciferase activity also increased following wounding. Transformed RNSlp-LUC lines were wounded and examined for LUC activity 5 h alter wounding. Unwounded RNSlp-LUC and wounded and 49 nos—preRNSl—E9 nos-RNSchNA-E9 nos-Globin—E9 0h 3h 0h 31! I" HOS RNSI EF-I a Figure 3—3. Differential response of RNSI transcribed sequences to wounding. Pools of T2 Arabidopsis seedlings expressing either the entire RN51 transcribed region (left panels), the RNSI cDNA (center), or the globin transcript (right) under the constitutive nos promoter were wounded and harvested 3 h later. An oligonucleotide corresponding to a transcribed portion of nos was used as a probe in order to distinguish the transgene from endogenous RN51. Blots were then stripped and reprobed with RNSI (to control for wounding) and EF-l a (to control for loading). 50 A RNSI p-GUS-E 9 3 SS-GUS-E9 0h 3h on 3h 1?an Figure 3-4. The RNSI promoter confers wound inducibility to reporter transcripts. (A) Leaves of transgenic Arabidopsis plants expressing the GUS reporter under the control of either 2.7 kb of genomic sequence upstream of the RNSI transcription start site plus 60 bases of 5' UTR sequence or the constitutive 35S promoter were harvested before (0 h) or 3 h after wounding. Blots were probed with GUS, then stripped and reprobed with RN51 (to control for wounding) and EF -1 a (to control for loading). 3SS-GUS plants were used as controls to demonstrate that GUS is not stabilized by wounding. (B) Images of wild type (WT) plants and plants expressing LUC under control of the same upstream RN51 region as in (A) (RNSI-LUC). WND, wounded plants 5 h after wounding. The lower left leaves of the rosette were wounded. Left panel, plants in visible light; right panel, luciferase activity. 51 unwounded WT Col plants were used as controls, since no 3SS-LUC lines were ever found to have LUC activity, either with or without wounding (data not shown). As seen in Figure 3-4B, only wounded RNSlp-LUC plants have LUC activity. These results indicate that RN51 is transcriptionally induced by wounding. However, since the promoter constructs used contained approximately 60 bases of 5' UTR sequence, I removed this sequence to confirm that the promoter sequence alone confers wound- inducibility to the GUS and LUC reporters. These constructs, p2081 and p2082, are shown in Figure 3-2. As expected, plants transformed with these constructs also displayed wound-inducibility of the transgenes. RNA gel blot analysis revealed increased GUS levels 3 h following wounding of leaves of transformed plants (Figure 3-5A), while LUC activity was again seen only following wounding of p2082 lines (Figure 3-5B). My results confirm that wounding causes a transcriptional induction of RN51. Use of promoter constructs to analyze signals mediating RNSI induction One unique aspect of the wound-induced expression of RN51 is the fact that JA and OGAs are not required (Chapter 2 and LeBrasseur et al., in press). The LUC reporters offered us the opportunity to screen for various signals that might induce RN51 expression. I used the lines to examine several aspects of RN51 expression, including response to hormone treatments, interaction of the wound-induction with these hormones. and effects of other external stimuli. Since dehydration has been proposed to be the signal for the wound induction of J A- independent genes (Reymond et al., 2000), I first assessed whether dehydration caused induction of RN51. Figure 3-6 shows that dehydration did not cause induction of LUC activity in adult plants left unwatered for over two weeks, until leaves of the plants A 2081.2 2081.3 0h 3h 0h 3h RN51-LUC Figure 3-5. 5' UTR sequences are not required for the wound responsiveness of the RN51 promoter. (A) Northern blot analysis of RNA from unwounded (0 h) and wounded (3 h) leaves of T2 plants (2081.2 and 2081.3) expressing the GUS reporter controlled by 2.7 kb of the RN51 promoter with no 5' UTR sequence. (B) Luciferase activity (bottom panel) in control and wounded (WND) wild-type (WT) and transgenic plants expressing the LUC reporter under control of the RN51 promoter (RNSI-LUC) as in (A). 53 Control Wounded Figure 3-6. Dehydrated plants do not induce the RN51 promoter. Plants were dehydrated as described in the text. Luciferase activity (right panel) is seen only after wounding the dehydrated plants. No RNS] transcript was detected by Northern blot analysis of dehydrated WT seedlings (not shown). Figure 3-7. RN51 transcript accumulates in response to salt stress. Ten-day-old wild-type Arabidopsis seedlings were transferred to plates containing 250 mM NaCl and subsequently harvested at the time points indicated. RNA (15 pg) was analyzed by Northern blot using an RN51 probe. This blot was kindly provided by Daniel Cook. 54 appeared purple from high levels of anthocyanins (not shown). Wounded dehydrated plants were used as a control and had high LUC activity. The activity in the wounded dehydrated plants was the strongest seen in any of the LUC experiments performed in this work. Therefore, I speculate that although dehydration does not cause RN51 induction, it may potentiate the wound-induction. To confirm that RN51 does not respond to dehydration, I probed a blot containing RNA from dehydrated seedlings. No RN51 signal was detected on the blot (not shown), which contained RNA from time points 0-8 h of dehydration, although typical dehydration-responsive transcripts were induced (D. Cook, personal communication). Since both the LUC reporter and endogenous RN51 transcript did not respond to the dehydration treatment, I conclude that dehydration is not the signal that causes induction of RN51 by wounding. As discussed briefly in Chapter 1, 5. cereviseae has only one member of the RNase T2 family, Rnyl. Since the RNYI transcript accumulates in response to salt stress (Maclntosh et al., 2001), I were interested in examining the response of RNSI to a similar treatment. RNA blots containing RNA samples from Arabidopsis seedlings transferred to plates containing 250 mM NaCl were probed with RN51. As seen in Figure 3-7, like RNYl, RN51 also accumulated in response to high NaCl concentrations. Transcript was detected by 2 h of treatment and remained high at 8 h. Recently, a mutant deficient in ABA signaling has been isolated that affects an mRNA cap-binding protein and is suggested to affect RNSI levels through a posttranscriptional mechanism (Hugouvieux et al., 2001). In an attempt to identify any effects of ABA on RN51 expression, I treated the LUC lines with ABA. LUC activity was seen 5 h after addition of ABA to seedlings growing in liquid‘medium (Figure 3-8A). I 55 also probed a blot containing RNA from ABA-treated seedlings and found that endogenous RN51 transcript accumulated beginning 2 h after transfer of seedlings to plates containing 100 pM ABA (Figure 3-8B). RN51 levels remained high after 8 h of treatment. Thus, ABA caused induction of RN51. Most likely, this induction is transcriptional, since I detected LUC activity upon ABA-treatment. These results point toward a possible signal for the wound induction of this gene. Since ABA induced RN51, I examined whether ABA signaling is required for the wound induction of this gene. Several different mutants in the ABA pathway were wounded and analyzed for RNase activity. Two of the mutants, abiI and abi2, are blocked in ABA signal transduction, and a third, abaI-I, is deficient in ABA biosynthesis, (for a review on ABA signal transduction, see Leung and Giraudat, 1998). Twenty-four hours following wounding, all three mutants had elevated levels of RNSI activity (Figure 3-9). Thus, although ABA is able to induce RN51 transcription, it is not required for the wound induction. The 33-kDa activities were also induced in the mutants, indicating that the bifunctional nucleases are also upregulated independently of ABA. Previous studies in this laboratory indicated that wounding and heat shock may interact to regulate RN51 expression. I used the LUC and GUS reporters to examine the effects of these two stimuli on RN51 transcription. In multiple repetitions, LUC activity was not seen after wounding when combined with treatment at high temperatures (Figure 3-10A). High temperature alone had no effect on LUC activity. Since it has been reported that high temperature can affect RNase activity posttranslationally (Chang and Gallie, 1997), I tested whether the RN51 transcript was affected by heat shock. 56 0 1 2 4 6.8,.-.“ Figure 3-8. ABA induces the RNSI promoter. (A) Luciferase activity (bottom panel) in seedlings grown in liquid culture 5 h after addition of 5 pM ABA (+ABA) or buffer alone (-ABA). (B) Northern analysis of RNA isolated from seedlings treated with ABA for the indicated times. RNA gel blot in (B) provided by Daniel Cook. 57 WT abi1 abi2 aba1-1 0 24 0 24 0 24 0 24 h m <— 33 kD Figure 3-9. Mutants in ABA signaling and biosynthesis induce RNSI by wounding. Wild type (WT) and mutants abi1, abi2, and abal-l were examined for RNase activities in unwounded (0) and wounded (24) 2-week-old seedlings. Induction of RNSI and the 33-kDa activities is seen in all four lines. 58 For this experiment, stems of wild-type plants were wounded, then cut and floated in a buffer at either 22 or 42°C. The wound induction of RN51 is diminished by the heat treatment (Figure 3-IOB). These results indicate that heat shock inhibits the wound response of RN51 . DISCUSSION The findings in Chapter 2 of this thesis revealed that RNSI responds to wounding through a pathway not previously identified in Arabidopsis. This has led to the obvious question as to what signal transduces the wound stimulus to cause increased levels of RN51? l addressed this question by examining the level of regulation of this gene and found that RN51 is transcriptionally activated by wounding. I next examined several other factors that might affect RN51 transcription and found that, although ABA induces RNSI expression, it is not required for the wound induction. Sequence requirements for the wound induction of RN51 The accumulation of RN51 following wounding occurs within 2 h in Arabidopsis leaves (refer to Figure 3-IOB). In fact, previous microarray analyses of wounded leaves found that the fold-induction of RNSI at 2 h was higher than any other of the 600 spots examined (M.A. Pérez-Amador and P.J. Green, unpublished). In the unwounded state, this transcript normally can not be detected or is at very low levels in leaves of adult Arabidopsis plants (Bariola et al., 1994). Thus, wounding apparently causes a strong response of RN51. The accumulation of RN51 could be due to either a transcriptional induction or a posttranscriptional mechanism, such as an increase in stability of the transcript. Posttranscriptional regulation of this transcript has been suggested by previous experiments that showed that although P, starvation caused increased transcription of WND 22°C WND 42°C .. Control 22°C - 42°C '_ h 7, ,1;- RNS1 Figure 3-10. Heat shock inhibits the wound-induction of luciferase activity. (A) Transgenic plants expressing the RNS lp-LUC construct were wounded and placed at either 22 or 42°C for 5 h. Luciferase activity (right panel) was seen only in the wounded plants left at 22°C. Similar results were seen in seedlings (not shown). (B) RNSI accumulation in unwounded (control) and wounded stems incubated at 22 or 42°C for 2 h after wounding. Experiment in (B) was performed by Gustavo Maclntosh 60 RN51 promoter fusion reporters, transcriptional regulation could not account fully for the levels of RNSI under these conditions (Howard, 1996). In contrast to these results, I found that increased transcription of RN51, as directed by 2.7 kb of promoter sequence, causes the majority of the increased levels of this transcript by wounding. However, I did detect small effects of the transcript alone upon wounding. This effect was only seen when the introns and full 5' UTR sequence were included in the reporter; the cDNA sequence alone did not confer wound-inducibility. On the other hand, the full 3' UTR is not required for the wound induction (see Chapter 4). In conclusion, the RN51 promoter contains the sequences most critical for the regulation of the induction of the RNSI enzyme by wounding. Response of RNSI to various biotic and abiotic stimuli Previous work had shown that RN51 transcript accumulates in P ,-starved seedlings and in senescent leaves (Bariola et al., 1994). Work in this thesis has shown that multiple other signals direct expression of this gene. Wounding, as seen here and in Chapter 2, strongly influences RN51 transcription. This effect is not directed by the known wound signals jasmonic acid and OGA elicitors. Thus, I was interested in what other internal or external cues alter RN51 levels. To examine this question, I used the LUC and GUS lines to screen many different stimuli. Using this screen, I found that ABA causes increased transcription of RN51 and the LUC reporter. ABA has been suggested to affect RN51 posttranscriptionally (Hugouvieux et al., 2001). A posttranscriptional mode of regulation was suggested because the authors isolated a mutant in a gene that encodes a cap-binding protein. The mutant shows ABA- hypersensitivity and down-regulates several transcripts implicated in ABA signaling. 61 RN51 is one of the transcripts with lower basal levels in this mutant. These results may be slightly misleading, since RN51 transcript is usually just at or below the level of detection, suggesting that the wild-type plants used for Northern and microarray analyses in their study may have been stressed. Nonetheless, if RN51 is affected in this mutant, it is logical to infer that this transcript is normally posttranscriptionally regulated by ABA. I found that ABA causes transcriptional induction of a reporter controlled by the RN51 promoter (Figure 3-8A); however, I cannot rule out the possibility that certain ABA effects are carried out through a posttranscriptional mechanism. Treatment of the nos- RNSchNA or nos-preRNSl lines with ABA should address this possibility. Although ABA does affect transcription of RNSI, it is not required for the wound induction of RNSI activity. Similarly, the wound-induction of the 33-kDa activities does not require ABA signaling or biosynthesis (Figure 3-9). It is possible that local RNSI induction is independent of ABA, while systemic induction requires ABA. The systemic regulation should be addressed in future experiments. One interesting result derived from the dehydration experiments was a particularly strong induction of the LUC activity in wounded dehydrated plants. Dehydration is known to induce ABA production in plants (Thomashow, 1999). Since ABA caused increased LUC activity in the transgenic lines, it is possible that the combined effect of wounding and increased ABA levels may have an amplification effect on the induction of the reporter. This effect may explain the increased RN51 transcript levels identified in the dehydration experiment by Reymond et al. (2000), since dehydration was achieved by detaching rosette leaves from the roots, possibly causing a weak wound response that was amplified by the dehydration. It appears, however, that the dehydration-elicited ABA increase alone is not sufficient for RN51 induction, since both luciferase activity in the transgenic lines (Figure 3-6) and RN51 transcript in WT plants (not shown) did not accumulate in response to dehydration. The differential control of RNSI by ABA and dehydration is particularly interesting since overexpression of CBF! or CBF2 results in accumulation of RN51 (M.F. Thomashow, personal communication). CBF] and CBF2 are transcription factors rapidly induced by cold temperature in Arabidopsis (Gilmour et al., 1998), and overexpression of CBF] results in increased cold tolerance (J aglo-Ottosen et al., 1998). Interestingly, RNSI is not induced by cold treatment (not shown), nor is it induced by overexpression of CBF3 (M.F. Thomashow, personal communication), another in the cold-induced CBF family (Gilmour et al., 1998). Therefore, cross-talk between pathways that control dehydration, cold, and ABA responses appears to regulate RNSI accumulation in a highly specific manner. Alternatively, overexpression of the transcription factors may cause secondary effects that result in increased RNSI accumulation. ’ While dehydration may potentiate the wound-induction of RN51, we found that heat shock inhibited this response. Posttranslational effects on RNase activity by heat shOck have been suggested to be one mechanism by which plants reduce mRNA turnover until recovery has occurred (Chang and Gallic, 1997). RNSI is an extracellular enzyme and therefore is not expected to be involved in mRNA decay. Additionally, this effect is apparently not posttranslational, according to the reduced RNSI transcript levels by heat shock (Figure 3-IOB). The wound-induced 35-kDa nuclease activity was also inhibited by the heat treatment (G.C. Maclntosh and P.J. Green, unpublished). I am currently using the GUS reporter lines to determine whether RNSI is transcriptionally affected by heat 63 shock. Further, effects on the 33-kDa nuclease activities will be examined. During the screen, I also examined several other factors that did not cause increased LUC activity (not shown). These include auxin and cytokinin. Additionally, phosphatase and kinase inhibitors did not affect the induction of the LUC reporter by wounding. It is possible that the treatments were not effective and that further studies will uncover regulation by these factors; however, it currently appears that these two hormones and inhibitors of other Arabidopsis wounding signal transduction pathways (Rojo et al., 1998) are not involved in the wound induction of RN51. In conclusion, transcription of the RN51 gene is regulated by several factors. Wounding, senescence, and P,-starvation were already known to control RNSI levels (LeBrasseur et al., in press; Bariola et al., 1994). I have now shown by promoter fusion analysis that the wound induction is largely a transcriptional response. ABA and NaCl are other factors that regulate RNSI transcription, while stresses such as cold and dehydration do not. Although dehydration does not cause RN51 induction, it may amplify the induction by wounding. In contrast, heat shock has an inhibitory effect on RN51 wound- induction. There appear to be multiple networks controlling the expression of RN51. These may be important for tightly regulating RNSI activity levels under various stress conditions. Understanding of the conditions regulating RNSI activity is important in order to elucidate the function of this and related enzymes in plants. This question will be addressed in the following chapter. MATERIALS AND METHODS Nuclei preparations and run-on analyses Seedlings for nuclei purification were grown on Pi and PT media as described (Bariola et 64 al., 1994). Nuclei were purified as follows. Frozen 7-day-after-transfer seedlings were ground in Nuclear Isolation Buffer (NIB; 1.0 M sucrose, 5 mM MgCl2, 10 mM Tris pH 7.2) to a powder. The ground tissue was filtered through two layers of miracloth, Triton X-100 was added to 0.5% (v/v) to lyse the chloroplasts, and the filtrate was spun at 1,000 X g at 4°C for 10 min. The pellets were resuspended in 0.5 ml NIB plus 0.5% Triton and spun as above, for 15 min. Pellets were then resuspended in Nuclei Storage Buffer (50% glycerol, 5 mM MgCl2, 2 mM DTT, 20 mM HEPES pH 7.2) and frozen at ~80°C until used for run-on analysis. An aliquot of nuclei from each preparation was DAPI-stained, and nuclei were counted to estimate the number of nuclei purified. Nuclear run-on reactions were begun by adding each of the following to a 50-pl aliquot of nuclei for a final volume of 100 pl: 1 x transcription salts (10 mM MgCl2, 100 mM ammonium sulfate, 20 mM HEPES pH 7.9), 10 pl glycerol, 1 pl 0.1 M DTT, 0.5 mM ATP, GTP, and CTP, 70 units of RNase inhibitor (Promega), and 100 pCi 32P-UTP (3000 Ci/mmol). After 20 min, 5 pl DNase RQI (Promega) was added to each tube, and reactions were placed at 30°C for 5 min. Reactions were stopped with 300 pl phenol/chloroform (50:50 v/v) and 200 pl stop solution (1% SDS, 20 mM EDTA, 100 mM LiCl, 10 mM aurintricarboxylic acid). The tubes were spun in a microcentrifuge for 10 min at 4°C. The supernatant was transferred, and 75 pl of 10 M ammonium acetate and 940 pl EtOH was added. After precipitation, the transcripts were resuspended and used to probe slot blots containing 2.5 pg ssDNA of the appropriate plasmids. ssDNA was purified using a procedure from Stratagene. Briefly, a single colony of bacteria (E. coli strain DH5a F ') carrying the appropriate clone was inoculated into 5 ml of 2 x YT medium containing 50 pg/ml ampicillin and l x 108pfu/ml VCM13 helper 65 phage and grown overnight at 37°C. Cells were pelleted, and 1 ml of supernatant plus 150 pg 20% PEG and 2.5 M NaCl was precipitated on ice for 15 min. After centrifugation, the pellet was resuspended in 400 pl 0.3 M sodium acetate (pH 6.0) and 1 mM EDTA. The ssDNA was extracted with phenol/chloroform and precipitated. Cloning The RN51 cDNA was digested with Smal and blunt-end ligated into a nos-globin gene construct, p2031, with the globin gene insert removed. This construct, p1966, was then digested with Pstl and ligated into the Pstl site of the plant transformation vector pCambia 1301, which has the hygromycin resistance plant selection marker. This clone was named p1975. The nos-globin control in the plant vector pCambia 2301 was named pl 995 and confers kanamycin resistance to transformed plants. To clone the entire RN51 transcribed region, including the full 5' UTR and introns, primers (PG749 and PG750) were designed for PCR amplification using wild-type genomic DNA. Each primer contained a BamHI recognition sequence. The PCR product was isolated from an agarose gel (Qiagen) and ligated into pGEM-T Easy (Promega). The insert was sequenced to confirm the sequence of the PCR product and ligated with BamHl/BglII-digested p203]. The orientation of the insert was confirmed, the nos-RN51 insert was retrieved by digestion with Pstl and ligated into the Pstl site of pCambia 2301. Constructs p848, pl401, p1402, and p1432 were made previously (Howard, 1996). To remove 5' UTR sequences from these constructs, primers corresponding to the transcription start site (3' primer PG866) and to a site in the promoter approximately 420 bp upstream of the transcription start site (5' primer PG853) were synthesized. The 3' primer inclUded a BglII restriction site. The PCR product was isolated and ligated into 66 l pGEM-T Easy (p2028), sequenced to confirm the PCR product sequence, and digested with BgIII and SpeI to isolate the correct 3' end of the promoter fragment. This 383-bp fragment, which lacks the 5' UTR sequences present in constructs pl401 and p1432, was used to replace the 475-bp BgIII/Spel fragment in the original vectors. These pMON- derivative plasmids carry a kanamycin resistance gene. Plant transformation All plasmids were transformed into Agrobacterium tumefaciens strain GV3101 C58C1 Rif (pMP90) (Koncz and Schell, 1986) by electroporation, using a Gene-Pulser apparatus (Bio-Rad) as per the manufacturer's instructions. Arabidopsis thaliana ecotype Columbia plants were transformed by vacuum infiltration according to the method of Bechtold et a1. (1993). Seeds from transformed plants were plated on selection medium containing either kanamycin at 50 pg/ml or hygromycin at 30 pg/ml. Resistant seedlings were transferred to soil after 2 weeks. Expression of the transgene was tested in adult plants by Northern analysis (not shown). Lines expressing the transgenes were used for further studies. Plant treatments Plants used for various treatments were selected on medium with appropriate drugs and transplanted to soil. Seedlings grown in liquid culture were sterilized for 7 min. in 50% bleach/0.02% Triton X-100, washed three times in ddH2O, and resuspended in 0.1% agarose. Approximately 25 — 30 seeds were added to each well in 12-well ELISA plates. Each well contained 4 mL of MS medium (Life Technologies, Rockville, MD) supplemented with 0.5% (w/v) sucrose. Plates were shaken on a rotary platform at ~50 R.P.M. in a growth chamber at 16 h light and 22°C. Treatments were performed on 10- or 67 14-day-old seedlings. ABA was added to the liquid medium from a 5-mM stock solution dissolved in IN NaOH. Final ABA concentration was 5 pM. Controls were treated with equal amounts of 1N NaOH. Wounding was done using ridged flat-tipped tweezers. Approximately 30% of the leaf surface was wounded. RNA and protein analyses RNA and protein extraction, RNase activity assays, and Northern hybridization were performed as described in Chapter 2. Luciferase analysis After each set of treatments, LUC-expressing and control plants were sprayed with the substrate luciferin (Promega; 5 pg/ml in 0.011% triton). In some cases, luciferin was added directly to the liquid medium instead of being sprayed onto the plants. Luciferase activity was examined using a photonic camera 40 min after addition of the luciferin substrate. 68 CHAPTER 4 RNSI FUNCTION: ANALYSIS OF AN RNS] MUTANT AND OVEREXPRESSOR 69 ABSTRACT I used a T-DNA insertional mutant, rnsl-2, and plants overexpressing the RN51 gene to analyze the function of the RNSI ribonuclease. The rnsl-2 mutant lacked detectable RNSI activity, even in wounded leaves. Plant morphology was examined. Unlike antisense RNSI lines, rnsI-2 plants did not have elevated anthocyanins. However, in support of a previous hypothesis regarding the function of S-like RNases in P,- remobilization (Goldstein et al., 1989), the overexpressors were better able to utilize RNA as a source of P,. I also found an unexpected effect of RNSI activity on root length in the plants. While the mutant had longer roots, the overexpressing plants had shorter roots than the wild type. These results indicate that RNSI is involved in processes other than recycling P,. since the phenotypes were seen independently of P, concentrations in the medium. 70 INTRODUCTION For decades, RNases have been used as models for structural and enzymatic studies. RNase A in particular has been extensively analyzed, and crystallized structures of wild type as well as enzymatic mutants have been determined (Gilliland, 1997). Despite the knowledge gained on enzymatic activity, little is known concerning physiological functions of secreted ribonucleases. In the T2 family, the exception is the S-RNases involved in self-incompatibility in the Solanaceae (Parry et al., 1997). Previously, antisense methods were used to reduce Arabidopsis S-like RNase activity to address the function of members of this family. These experiments revealed that seedlings with low RNSI or RNS2 activity had high anthocyanin content, an effect that was amplified by P, starvation (Bariola et al., 1999). These results are consistent with the hypothesis that S- like RNases are involved in the mobilization and recycling of P, during P, starvation and senescence. However, in the antisense experiments, RN51 transcript levels were never reduced below 10% of that of the wild type (Bariola et al., 1999). Also, the segregation of kanamycin resistance in the antisense lines indicated that they may have had silencing or T-DNA rearrangement complications (P.A. Bariola, G.C. Maclntosh, and P.J. Green, unpublished). Recent results, including the systemic induction of RNSI during wounding (LeBrasseur et al., in press) and phenotypes of the insertional mutant of the yeast T2 RNase (Maclntosh et al., 2001), have indicated new directions regarding the functions of secreted RNases. This chapter will focus on the use of insertional mutagenesis (Krysan et al., 1999) to examine the function of the RNSI enzyme. Along with this project, I also overexpressed RN51 in plants to examine the effect of increased RNSI activity. Previous attempts to overexpress RN51 in roots did not lead to increased levels of the transcript 71 beyond that induced by P, starvation (Bariola, 1996). In this study, I expressed the RN51 cDNA throughout the plant under the control of the strong cauliflower mosaic virus (CaMV) 355 promoter. RESULTS Isolation of ms] mutants and RNSI overexpressing plants To get a more accurate assessment of the function of RNSI, I screened the University of Wisconsin T-DNA-tagged lines for insertions within the RN51 gene. PCR primers complementary to sequences in the promoter region of RNSI (650 bp upstream of the transcription start site) and 450 bp 3' of the end. of the cDNA were synthesized Two putative insertion lines identified in the primary screen were pursued further. The locations of the insertions are shown in Figure 4-1. The alleles were named rnsl-I and rnsl-2. The rnsl -1 allele lies within the 3' UTR, while the rnsI-2 allele is 85 bp 5' of the translation start site in the 5' UTR. PCR analysis revealed that the left border (LB) of the T-DNA flanked both ends of the insertion in both alleles (not shown). Plants homozygous for each allele were identified by PCR analysis. To analyze RNSI function firrther, I transformed plants with the RNSI gene under the control of the strong constitutive CaMV 355 promoter (Figure 4-1). Once plants homozygous for the T-DNA insertions were identified, I examined RNSI activity in the two mutants. Protein extracts from unwounded and 6-h wounded leaves were run on RNase activity gels. The rnsl-I plants still had significant levels of wound- induced RNSI activity (Figure 4-2A), as might be expected since the T-DNA insertion lies in the 3' UTR. Not only was there still activity, but also regulation of the gene appeared to be unaffected by the insertion: there was no activity in the absence of 72 RN51 1" l Figure 4-1. rnsl mutants and overexpressor constructs. Two RN51 T-DNA insertion alleles were identified. The first insertion, rnsI-l, is downstream of the RNSI polyadenylation site (asterisk). The second, rns1-2, lies in the 5' UTR, 85 bases 3' of the translation start site (arrow). The bent arrow indicates the transcription start site. Boxes represent exons, and lines represent introns. To overexpress RN51, the cDNA was cloned between the CaMV 355 promoter and the pea E9 3' end. This construct was named p2020. 73 wounding, and strong activity after wounding. The regulation in the mutant is consistent with sequences required for the wound-induction, as demonstrated in Chapter 3. In contrast, rnsI-2 mutants lacked RNSI activity, even after wounding (Figure 4- 28). I therefore studied this allele further. To confirm that the rnsI-Z line was homozygous for the T-DNA insertion, DNA gel blot analysis was performed. Genomic DNA digested with EcoRI was subjected to DNA gel blot analysis using an RNSI probe. While the probe hybridized to a 5-kb band in wild-type DNA, only a 2-kb band was seen in the rnsl-2 lane (Figure 4-3 B). The absence of the S-kb band in this lane indicates that the rnsl-Z allele is in the homozygous state in this line. The blot was also hybridized with a T-DNA probe to determine the number of T-DNA insertions in the genome. Sizes of the expected bands, based on restriction maps of the T-DNA vector and genomic region surrounding RN51, are 1.0, 1.7, 4.6, and 5.1 kb. Bands of these sizes were observed (Figure 4-3A), and no other bands were found on the blot, indicating that the T-DNA had inserted at only one site in the rnsI-Z line. However, at least two copies of the T-DNA were present at this allele, since PCR fragments were obtained using primers corresponding to the LB of the T-DNA and to either the 5' or 3' ends of the RN51 genomic region (not shown). The exact number of T-DNA copies inserted was not determined. Since this insertion lies in the 5' UTR of the gene and the gene might therefore produce transcripts including the full open reading frame, I performed RNA gel blot analysis of wounded and unwounded leaves from the mutant. Several aberrantly sized bands could be detected using an RN51 probe (Figure 4-3C). These were found only upon wounding, confirming previous results that the 5' UTR is not required for wound 74 WT rns1-1 WT rns1-2 03 62403 624hpw w‘m“ “YT“ . . If <~“35 kD «RN51 Figure 4-2. RNSI activity in the rnsI-l and rnsl-2 mutants. (A) Wounded (7 hours post wounding, hpw) and unwounded (0) leaves of 4-week-old rnsl-l and wild-type (WT) Ws Arabidopsis plants were analyzed by RNase activity gel assay. (B) Leaves of WT Ws and rns1-2 mutants were wounded, harvested at timepoints indicated, and analyzed by RNase activity gel assay. RNSI activity (bottom arrow) can be found only in WT wounded samples. Induction of the 35-kDa RNase activity (top arrow) provides a positive control for the wound response of the rnsl-Z leaves. hpw, hours post wounding. 75 A [WT rnst-Z I“ — 5.1 —— 1.7 . ——1.0 C WT rns1-2 0 3 624 03 624hpw Figure 4-3. DNA and RNA gel blot analyses of rns1-2 mutant plants. Genomic Southern analysis of wild-type (WT) and rnsl-2 plants using the T-DNA (A) or RNSI (B) probe. Estimated sizes of hybridizing bands are shown to the right. (C) Northern analysis of wounded WT and rnsl-2 leaves harvested at the indicated timepoints using an RNSI probe. hpw, hours post wounding. 76 induction of RN51 (see Chapter 3). The various sizes could be due to irregular splicing within the T-DNA and RN51. One of the smaller bands was approximately the same size as the wild-type transcript; however, RT—PCR analysis of RNA from wounded rns1-2 leaves did not detect any polyadenylated transcript that included the start of the open reading frame (not shown). Additionally, RN51 transcript induced in the wild type was significantly more abundant than the aberrant bands produced. Thus, although it is possible that a small amount of wild-type transcript is made and properly translated in this mutant, I predict the amount to be minimal. Plants with high RNSI activity in the absence of wounding were identified from a T, population of the 3SS-RNS1 lines. Leaves were harvested and analyzed by activity gel assay. Multiple T. plants had high RNSI activity; two examples are shown in Figure 4-4. Seeds from a homozygous T2 progeny plant of line 2020.2 were used in further studies. The rnsI-2 mutant does not have increased levels of anthocyanins Since RNSI antisense lines have high levels of anthocyanins, particularly when grown on P,‘ medium, I looked for a similar effect in the rns1-2 mutant, which had lower RNSI activity induced by wounding than found in antisense plants subjected to P,-starvation (compare Figure 6 in Bariola et al., 1999 and Figure 4-2b of this thesis). Thus, I expected a similar or more severe anthocyanin phenotype to be associated with the rns1-2 mutant. The anthocyanin content of rns1-2 seedlings was measured after 7 d of growth on medium either containing (P+) or lacking ( P‘) P,. Antisense RN52 lines, which have high anthocyanin levels similar to antisense RN51 plants (Bariola et al., 1999), and wild-type Ws seedlings were used as positive and negative controls, respectively. As expected, the 77 M 3SS-RNS1 1 2 <— RNS1 Figure 4-4. RNSI activity in overexpressing plants. Protein extracts of unwounded leaves of wild-type (WT) and two T, 35S-RNS1 plants were analyzed by RNase activity gel assay. High RNSI activity (arrow) is seen only in the overexpressing plants. 78 antisense line had a modest increase in anthocyanins when grown on Pi medium and a stronger increase on' P" medium. However, the rnsl-2 mutant had anthocyanin contents similar to wild-type seedlings (Figure 4-5). This unexpected result was seen regardless of whether the rnsl-2 seedlings were grown on kanamycin-containing selection medium (kanamycin selection was used for the antisense lines in Bariola et al., 1999). Thus, it appears that the antisense lines have high anthocyanins for a reason other than the absence of RNSI or RNS2 activity. RNSI activity appears to affect plant growth on RNA Expression patterns of RN51 during P,-starvation indicated that the RNSI enzyme may be involved in nutrient recycling during P, stress (Bariola et al, 1999). Therefore, Iexamined the effect of low P, concentrations in the growth medium on rnsI-Z and 3SS-RNSI seedlings. No obvious morphological difference between wild-type and transgenic populations grown on concentrations of P, varying from 0 to 1250 pM was found (not shown). Further studies were conducted to examine the ability of the mutant lines to grow when provided with RNA as the sole source of phosphate. In the initial experiment, seeds were plated densely on AGM medium with either 1250 pM P, (P+) or 0.6 mg/ml RNA. On RNA-containing medium, the 35S-RNS1 seedlings were larger and bolted and flowered sooner than the WT Col seedlings, while the rnsI-2 mutants were smaller and lagged developmentally compared to WT Ws (Figure 4-6A). I then repeated the analysis of the 3SS-RNSI and WT Col seedlings with only 12 seedlings per plate to increase spacing between the'plants. Once again, the overexpressing lines grew larger on the RNA plates (Figure 4-6B). While 35S-RNS1 seedlings were larger on RNA medium, on PL 79 Anthocyanin levels of ms mutants ElP+' "Pi Relative anthocyanin levels rns1-2 AIS rns'2 WT Line ' Figure 4-5. Anthocyanin content in mutant ms lines. Anthocyanin contents of rns1-2, wild-type Ws (WT), and antisense RNS2 (A/S rnsZ) lines grown in the presence (white bars) or absence (gray bars) of P,. Like antisense RNSI seedlings, the antisense RN32 line had high anthocyanin levels, particularly in the absence of P,. The rnsl-Z line had anthocyanin levels similar to that of the wild type. 80 medium, as determined by fresh-weight analysis, the 35S-RNS1 seedlings were smaller than WT Col (Table 4-1; see below for more on this effect), indicating that high levels of RNSI activity aid the overexpressor line in growing on RNA. RN S1 activity affects plant size When I first examined the rns1-2 mutant, I expected the mutant to be less able to grow on medium containing low concentrations of P,. I did not observe such effect, at least based on morphological examination (not shown). However, during a test in which I plated the wild type and mutant on various concentrations of P,, I noticed that the roots of the mutant appeared different. I then plated the lines again and grew them vertically, to allow the roots to grow downward. I plated on two different concentrations of P,: 1250 pM (concentration of P, in P+ plates) and 62.5 pM. The latter concentration allowed the plants to grow but caused severe starvation phenotypes, including high anthocyanin levels and severely reduced growth. rnsI-2 mutants grown on both of these P, concentrations appeared to have longer roots than the wild type Ws seedlings (Figure 4-7A). To quantitate these differences, roots were measured and analyzed by t-test (Table 4-2). In each of five repetitions, the differences in root length were statistically significant (P<0.001). They were also different independent of the P, concentration in the medium, indicating that this phenotype is not related to the putative function of RNSI in P, remobilization. Since RNSI activity affected root length in the mutant, I also examined the 35S-RNS1 roots. In two repetitions, the 3SS-RNSI seedlings had significantly shOrter roots than the Columbia wild type (Figure 4-7B; P<0.001). Therefore, it appears that levels of RNSI activity negatively correlate with root length in Arabidopsis. 81 ' 3SS-RNS1 Figure 4-6. Growth of msI-Z and 35S-RNS1 seedlings on RNA-containing medium. (A) 35S-RNS1, wild-type Columbia (WT Col), ms] -2, and wild-type Ws (WT Ws) seeds were grown on P,‘ plates containing 0.6 mg/ml RNA. After 34 days, the transgenic lines exhibited differences from their wild-type counterparts. While the WT Ws plants had bolted, rns1-2 plants had not. Similarly, the WT Col seedlings, which normally bolt later than Ws, had still not yet bolted, while many of the 35S-RNS1 plants had bolted and flowered. Arrowheads point to flowers in the Ws and 35S-RNS1 plants. (B) 3SS-RNSI and WT Col seedlings grown on P,’ plates containing 0.6 mg/ml RNA. 82 Figure 4-7. Root phenotype of RN51 mutant and overexpressor. (A) Root length of 12-day-old rnsl-2 and wild-type Ws (WT Ws) seedlings grown on vertical plates containing either 1250 or 62.5 pM P,. (B) Root length of l2-day-old 3SS-RNSI and wild-type Columbia (WT Col) seedlings grown on 1250 pM P,. 83 WTWs ms 1-2 M u. 5 2 6 WT Col 35S-RNS1 Figure 4-7. Root phenotype of RNSI mutant and overexpressor. 84 Table 4-1. Statistical analysis of fresh weight of WT, rnsI-2, and 35S-RNSI seedlings' Line Ave. weight per seedling t P (mg) wr Col P+ 12.9 2.3 <0.05 35S-RNS1 P+ 8.8 WT WS P+ 9.6 4.3 <0.01 rnsI-2 P+ 12.9 WT Col RNA 10.8 4.8 <0.01 35S-RNS1 RNA 15.7 IThe t-test was used to determine statistical significance of the weight differences between 9-day-old WT C01 and 35S-RNS1 and WT Ws and rns1-2 seedlings grown on P+ medium and between 37- day-old WT C01 and 3SS-RNSI seedlings grown on P' medium containing 0.6 mg/ml RNA. Results shown are from one experiment. A P-value of 0.05 or less is considered statistically significant. 85 Table 4-2. Statistical analysis of root length 1n ms] -,2 WT Ws, 35S-RNS1 and WT Col seedlings Age Ave Root Test Line P, (pM) (days) length (mm) t P rns1-2 10.0 WT Ws 62.5 6.6 12.5 <0.001 1 rnsl-2 13 48 3 WT Ws 1250 31.0 12.6 <0.001 rnsI-2 7.1 WT Ws 62.5 3.1 8.1 <0.001 2 ms! 2 12 37 4 WT Ws 1250 29.5 8.0 <0.001 rnsl-2 4.5 WT Ws 62.5 3.6 6.0 <0.001 3 rnsI-2 7 l6 9 WT Ws 1250 13.5 5.3 <0.001 rns1-2 36.8 4 WT Ws 1250 8 29.1 5.3 <0.001 35S-RNSI 41.2 5 WT C01 1250 12 50.3 4.4 <0.001 rnsI-Z 24.6 6 WT Ws 1250 9 18.3 5.2 <0.001 35S-RNS1 ’ 19.5 WT C01 1250 24.3 5.2 <0.001 IThe t-test was used to determine statistical significance of the differences in root lengths in seedlings grown on two different P, concentrations. A P- value of 0.05 or less is considered statistically significant. 86 In addition to the roots, the size of the green tissue of the rnsl-2 and overexpressor seedlings appeared to differ from the size of the wild types. I quantitated this difference by growing the four lines on P+ medium and weighing the seedlings. I removed the roots to eliminate any effect of the differences I observed in root tissue. T-test analysis of the weights indicated that rnsl-2 seedlings were significantly larger than WT Ws seedlings (P<0.01), and 35S-RNS1 seedlings were smaller than WT Columbia (P<0.05) (Table 4-1). These results are preliminary but are consistent with the phenotype seen in root lengfli The difference in root length could be due to cell size, cell number, or a combination of the two. To address this issue, I examined the roots by laser confocal microscopy. Roots of the four lines were fixed and stained using two dyes, DAPI to stain nucleic acid and calcofluor to stain the cell wall. I did not observe a large increase in size of the rnsl-2 cells or smaller 35S-RNS1 cells (not shown), but a small difference may be difficult to visualize using this technique. A more involved experiment, including measurements of hundreds of cells, will be a more thorough method to assess any differences. During the analysis of these roots, I noticed a difference in the rate at which the rnsl-2 and wild-type roots absorbed the dyes. Five minutes after adding dyes to the WT roots, the intensity of the stain was low (Figure 4-8). After 45 min, the staining was stronger, and a larger portion of the root could be viewed. However, even at this point, the intensity of the WT root was not as high as the rnsI-2 root only 5 min after the dyes were added. These results are preliminary and require repetitions to ensure that the effect was not an artifact of the staining preparation procedure. However, as will be discussed 87 Figure 4-8. Absorption of dyes in wild-type Columbia (WT) and rns1-2 mutant root cells. The dyes calcofluor and DAPI were added to fixed intact roots and photographed at the indicated times. Staining of the mutant root cells was greater in 5 min than that of the WT cells even after 45 min. These photos were taken by Gustavo Maclntosh and Kirk Czymmek. 88 below, the differential penetration of the dyes was consistent with the phenotype seen in a yeast mutant in the RNYI gene, the only homolog of RN51 in 5. ccreviseae, and supports the hypothesis that RNase T2 enzymes are involved in regulation of membrane permeability (Maclntosh et al., 2001). Absence of RN S1 does not increase susceptibility of the plant to a bacterial pathogen As discussed in Chapter 2, the induction of RNSI activity by wounding in unwounded leaves is not easily explained as part of a nucleotide-recycling mechanism, since this part of the plant has not been injured and should not require such a system. I hypothesized that RNSI might also play a role in defense of the plant from opportunistic pathogens that invade after wounding. Thus, as a first examination of this putative function, I tested the ability of a bacterial pathogen, Pseudomonas syringae pv tomato (Pst) DC3000 to infect the rnsI-Z and 3SS-RNSI plants in comparison to their respective WT. Leaves of the rns1-2, WT Ws, WT C01, and 3SS-RNS1 plants were inoculated with Pst, and growth of the pathogen was monitored for three days. In two repetitions, growth of the pathogen on the rnsI-2 mutant was similar to that on WT Ws (Figure 4-9A). Unlike rns1-2, the overexpressor appeared to have an effect on the growth of the bacteria; in four independent experiments, I consistently detected a lower Pst grth rate in 35S- RNSl compared to that in WT Col (Figure 4-9B). However, error bars indicated that variation between growth on the two lines was too great to make a statistically valid conclusion. Additionally, analysis of variance indicated no significant difference between growth of the bacteria on the 2nd day after infection. Further repetitions Of this experiment, particularly focusing on day two, may indicate valid differences. However, I currently cannot conclude that RNSI activity affected grth of this pathogen. 89 Figure 4-9. Growth of Pst DC 3000 on leaves of 3SS-RNS1 and rnsl-2 plants. (A) Leaves of 4-week-old wild-type Ws (Ws) and rnsl-2 mutants were infiltrated with P. syringae. Starting on the day of infiltration (day 0), bacterial growth was measured until day 3. Representative results of one oftwo experiments are shown. (B) Growth of P. syringae on wild-type Columbia (Col) and 355-RNSI plants was monitored as in (A). Averaged results from four independent experiments are shown. 90 P. syringae Growth Curve 7 A 6 ‘3 t ” " -- 4 ‘ C 4+ -, S 3 1/‘. Q +WS a, / ~--» rns1-2 2 2 '. .5 1 ,X 3 ll/ ,_ . o o . '1 T T r I 0 1 2 3 4 5 Days Combined Growth Curve 6l — — 4 — — 1 5 a e a 4 . § ' / ..- .— T— .... ~.. g 3 v ‘V f N a J a f, “'" IL 2 /” 5 2 i / l o ,T J . i n.—4——m_ ,._ - ___ , o i 2 3 4 ; Days j Figure 4-9. Growth of Pst DC3000 in leaves of 35S-RNS1 and rns1-2 plants. 91 DISCUSSION The vast majority of studies done on extracellular RNases, including the RNase T2 family, have focused on structural or enzymatic studies, or have addressed regulatory questions. Gene expression or protein induction patterns have been particularly usefirl in providing insights into the function of S-Iike RNases, but many questions still remain. I have attempted to gain further information on the function of plant secreted RNases through basic plant biology techniques, namely, overexpression and gene knock out of RN51. The phenotypes of these plants have not only reinforced previous ideas concerning the role of this family of enzymes, but also indicate that they may perform unexpected functions in the plant. Involvement of RNSI in phosphate remobilization Previous data, including gene expression patterns and the phenotype of antisense RN51 plants, have been consistent with the idea that RNSI is involved in the remobilization of P, from nucleic acid sources, either in the extracellular space or possibly from an intracellular source (Bariola et al., 1999). I therefore expected similar insights to be gained by studying the rnsI-2 null mutant and RN51 overexpressor. One phenotype of the plants studied here is consistent with this previously proposed function: the overexpressing plants appear better able to grow when provided with RNA as the only source of phosphate. This phenotype is interesting because it is the first evidence that RNSI may be secreted from the roots in order to degrade organic material in the surrounding rhizosphere to provide the roots with P, for uptake. Introduction and secretion of an Aspergillus phytase activity into Arabidopsis has been shown to improve the plants' ability to utilize phosphorus from phytate, a major component of soil organic 92 phosphorus (Richardson et al., 2001 ). By extension, induction of secreted RNases may be an endogenous method plants use to acquire P, from other organic sources. It will be important to study this effect further, to quantitate the differences in plant weight and to ensure that this effect is specific to grth on RNA by using DNA-containing medium as a control. Although I did not have the opportunity to quantitate differences, the observed growth effects on RNA are in contrast to the sizes of the plants grown on high P,, in which case the 3SS-RNS1 plants appear to be smaller than the WT, indicating that the effect of the RNA overcomes the differences in grth on medium containing high P,. I did not detect any differences in the sizes of the plants grown on low concentrations of P, that are consistent with a role for RNSI in P,-recycling. One possible interpretation of this result is that the presence of RNS2, which is also induced by P, starvation, provides the rns1-2 seedlings with sufficient RNA-degrading activity. However, the ms]- 2 mutant was less able to grow on RNA plates than the wild type. The different localizations of the two enzymes may cause these effects; RNSI is extracellular, and RNS2 appears to be intracellular, most likely either vacuolar or ER-associated (Bariola et al., 1999). Therefore, RNS2 may be unable to degrade RNA in the medium because the substrate is not accessible. I was also surprised that the rns1-2 mutant did not have high levels of anthocyanins, either in the presence or absence of P,. The antisense lines had significantly higher anthocyanin contents, and this was shown to be specific for reduced RNSI or RNS2 activity; the vector controls did not have high anthocyanin content (Bariola et al., 1999). Therefore, there appears to be another reason for the accumulation of anthocyanins in the antisense lines. For example, as mentioned in the introduction to this chapter, kanamycin 93 segregation indicates that there may be an unusual silencing of the transgenes in the antisense lines. The reduced kanamycin resistance may somehow result in increased anthocyanins, even in the kanamycin resistant seedlings. Alternatively, the sequences used to silence RN51 or RN52 may have an effect on the accumulation of RN54 and/or RN55, which may then cause increased anthocyanins. Expression of RN52 was not affected in the antisense RN51 lines, and vice versa (Bariola et al., 1999), but RN54 and RN55 had not been identified at the time of the analysis of the antisense lines. The expression of these genes in the antisense lines was therefore not explored. Is RNSI involved in defense mechanisms? One interesting hypothesis borne from the observation that RNSI is induced systemically upon wounding is the idea that this enzyme may be involved in defense of the plant. My initial results do not support this hypothesis. The rns1-2 mutant did not have increased susceptibility to Pst. However, RN51 is not normally induced by infection with this pathogen (not shown). Therefore, the mutant would not be expected to have increased sensitivity. Additionally, Pst is virulent in Ws, and therefore, I may not be able to observe further susceptibility. Interestingly, though, the RN51 overexpressing plants appeared to have an increased resistance to Pst. This effect is only preliminary; further replications must be done for this to be considered a reliable result. Other pathogens may be more likely to reveal that the presence or absence of RNSI can have an impact on defense mechanisms. For instance, it was shown that application of RNase A in the extracellular space of tobacco leaves prevented the growth of the oomycete Phytophthora parasitica (Galiana et al., 1997). Inoculation of tobacco leaves with this pathogen also induced expression of the gene encoding RNase NE, another S- 94 like RNase. In the future, a pathogen more closely related to P. parasitica, such as Peronospora parasitica, may be a better tool to analyze the role of RNSI in defense. RNSI activity negatively correlates with root length Possibly the most surprising phenotype that resulted from altering the expression levels of RNSI was the difference in root length. This effect may extend to the green tissue of the plant as well, as preliminary evidence suggests that the weight of this tissue is also increased in the mutant and reduced in the 35S-RNS1 plants (not shown). These effects are particularly unexpected because RN51 transcript was not detected in roots or leaves by Northern blot (Bariola et al., 1994). Under non-inducing conditions, RNSI was detected only in flowers. Most likely, RNSI is expressed in pistils, not petals, based on microarray analysis of the apetalaZ and agamous mutants (SMD experiment ID 12302; http://genome-wwwS.stanford.edu/MicroArray/SMD/). However, RN51 can be detected in leaves by RT-PCR (not shown), indicating that some activity may also be present in tissues other than flowers. Perturbation of this small amount of activity in roots and possibly throughout the plant is evidently sufficient for the phenotypes I observed. Similarly, although the yeast enzyme Rnyl cannot be detected by activity gel assay in wild-type cells, the knock-out still shows the enlarged cell-size phenotype (Maclntosh et al., 2001). Examination of the cells in the roots did not reveal gross changes in the cell size. It is possible that there is a small difference, e.g., 10%, in the cell size between the different lines. Differences on this order would be difficult to detect by eye. As such, more in depth analysis of the size of the cells is required. I hypothesized that cells in rns1-2 would be larger than WT cells based on results 95 seen in a mutant in the RNYI gene of 5. cereviseae. In this case, the mutant cells were approximately lO-fold larger in diameter than the WT cells and had enlarged vacuoles (Maclntosh et al., 2001). Thus, I postulated that a similar effect may be causing the rnsl- 2 roots to be larger. However, the presence of the cell wall or factors regulating cell-cycle may limit the expansion of the rns1-2 cells. As such, the cells may not be enlarged, but may instead divide more rapidly. We plan to compare cell division rates in the mutant and overexpressor lines by microscopic video analysis of the root cells during growth. During the analysis of the root cells, we also noticed that rns1-2 cells seemed to absorb the dyes more rapidly than the WT cells (Figure 4-8B). This effect has also been seen in the rnylA mutant using syrol3, a dye that does not normally enter cells, but which is taken up by the rnyIA cells (G.C. Maclntosh and P.J. Green, unpublished observations). These results are also preliminary; however, two of the phenotypes in the yeast mutant are consistent with my observations of the plant knock out: the yeast rnyl A cells and the rns1-2 mutant are bigger under normal grth conditions, and both appear to be more permeable to stains. Maclntosh et a1. (2001) hypothesized that natural RNAs may exist in the membrane, possibly as an ancient mechanism to control membrane stability and permeability in the "prebiotic world". They further postulated that secreted ribonucleases may be used to control the amount of RNA in the membrane. In the my] A mutant, this control is lacking, resulting in excess RNA in the membrane and a corresponding increase in membrane permeability. The cells may compensate for this change by increasing in size and enlarging the vacuole. Similarly, the rns1-2 mutant is lacking an extracellular RNase activity in the same family as Rnyl. The cells may be more permeable, and consequently, the plant is larger. 96 My results indicate that several phenotypes result from altering the expression of RN51. It will be important in the future to confirm these results by complementing the phenotypes using a genomic fragment of RN51. The rnsl-Z mutant has already been transformed with a construct that consists of the RN51 coding region plus 1.4 kb of upstream promoter sequence and 1.8 kb of sequence downstream of the stop codon, and primary transforrnants have been selected. In the near future, seeds from these plants and from plants transformed with the vector control should be plated for analysis of root length. If lack of RNSI causes longer roots, the complemented plants should have roots of WT length. This result is expected, since the 35S-RNS1 roots are shorter than roots of Columbia wild type. Since the rnsI-2 line has a minor ap2 phenotype as a result of the T-DNA used in the mutagenesis, we are in the process of isolating additional alleles. The T-DNA insertions in these rnsl-3 and rnsI-4 alleles lie in the RN51 coding region (+602) and promoter (-80), respectively. Since the T-DNA lies between the two conserved histidine residues highly conserved in S- and S-like RNases, the rns1-3 allele may prove to be a valuable null mutant. Any phenotypes seen in rnsl-2 should also be present in new alleles identified. MATERIALS AND METHODS Plant materials Throughout this chapter, Arabidopsis thaliana ecotypes Columbia or Ws were used. The Ws ecotype was used as a WT control for the rnsl mutants, since Ws was used to generate the Wisconsin T-DNA collection. The 35S-RNS1 construct was transformed into the Columbia ecotype; therefore, Columbia was used as a control for experiments 97 involving these lines. For plating, seeds were sterilized as described (Pérez-Amador et al., 2000). Selection medium contained 50 pg/ml kanamycin. RNA-containing medium was made according to Chen et al. (2001). Root length was determined by growing seedlings vertically in a chamber (Percival) at 22°C. Lengths were measured at different ages, including 9, 12, and 14 days. Four repetitions of rnsl-2 versus WT W5 and two repetitions of 35S-RNS1 versus WT Columbia were performed. The first three sets of rnsl-Z/Ws seedlings were grown on 1250 and 62.5 pM P,. In total, more than 320 seedlings each of rnsI-2 and WT Ws and more than 40 seedlings each of 35S-RNS1 and WT Columbia lines were measured and analyzed for a statistically significant variation by t-test. Values of t for each of the nine sets exceeded the cut-off value for a probability of 0.1% (P<0.001). Fresh weight of green tissue was determined by excising this tissue from the root base. Seedlings were measured in sets based on the plate they were grown on. The total weight per line per plate was measured and a per seedling weight was calculated for each plate. A t-test was used to determine whether the differences between the weights of the four lines were significant. Using at least 37 seedlings per line, P<0.05 for the 35S-RNS1 vs. WT Columbia set and P<0.01 for the rns1-2 vs. WT Ws set. Plasmid construction A clone containing the RN51 cDNA fused between a doubly enhanced c0py of the cauliflower mosaic virus 355 promoter and the nos terminator was constructed previously (Bariola, 1996). This plasmid was digested with Xbal to remove the entire promoter/cDNA/terminator cassette, and the fragment was ligated into the Xbal site of 98 the plant transformation vector pCambia 2301. WT ArabidOpsis thaliana ecotype Columbia plants were transformed with this construct, p2020, by vacuum infiltration (Bechtold et al., 1993). Identification of rns1 T-DNA insertional knock-out mutants The University ofWisconsin T-DNA insertion line collection was screened using primers PG904 (positioned at —572 from the RN51 transcription start site) and PG905 (+1645). Primers were tested on WT Ws genomic DNA and resulted in the amplification of a 2.2- kb fragment. These primers were sent to the University of Wisconsin for PCR-screening of their T-DNA insertion line collection using each of the two primers and a primer corresponding to the left border (LB) of the T-DNA. An aliquot of each reaction was electrophoresed on 0.8% (w/v) agarose gels and transferred to Nytran Plus nylon membrane (Schleicher and Schuell, Keene, NH). The RNA blots were hybridized as described in Taylor and Green (1991) using a 32P-labeled RN51 probe. The RN51 probe hybridized strongly with the PCR products from two reactions: LB/PG904 reaction 15 and LB/PG905 reaction 19. Bands corresponding to these signals were gel-isolated and sequenced to confirm the position of the insertion. The LB/PG904 insertion (rns1-1) lay 20 bases 3' of the stop codon, and the LB/PG905 insertion (rns1-2) caused a deletion of 21 bases in the 5' UTR, 65 to 86 bases upstream of the RN51 translation start site. Seeds were obtained from Wisconsin for the appropriate 25 pools of 9 insertion lines each. Seeds were grown and harvested as pools for PCR analysis to identify the pools that contained the desired insertions. Seedlings from those two pools were then transferred to soil, and individual plants were screened by PCR analysis to identify single plants harboring at least one copy of the rnsl alleles. At least one heterozygous plant was 99 identified for both alleles. These plants were allowed to self-fertilize. and their progeny were analyzed similarly to identify a homozygous line. Seeds from the homozygous lines were used for further analyses. Protein and RNA analyses Protein and RNA purification and analyses were performed as described in Chapter 2. DNA extraction and DNA gel blot hybridization For genomic Southei'n analysis. DNA was extracted from rosette leaves of mature plants using Plant DNAzol (GibcoBRL) according to the manufacturer's instructions. DNA was quantified by ethidium staining and comparison against a known amount of Lambda DNA. After digestion with EcoRI, 3 pg DNA was separated by electrophoresis in a 1.0% (w/v) agarose gel and blotted and hybridized as above using a 32P-labeled RN51 or T-DNA probe. The T-DNA probe was a PCR product of the entire transferred region of the T-DNA in the rnsl-2 allele. For genomic DNA extraction for PCR, a shorter extraction protocol was used to isolate DNA from one to two leaves of young plants. Tissue was pulverized and incubated for 30 min at 65°C in 500 pl CTAB buffer [2% (w/v) CTAB (Sigma); 1.4 M NaCl; 0.2% (v/v) B-mercaptoethanol; 20 mM EDTA; 100 mM Tris, pH 8.0]. The solution was extracted twice with chloroform:IAA [24:1 (v/v)], precipitated, and resuspended in 100 pl ddH2O containing RNase A (l pg/pl). PCR was performed according to the University of Wisconsin web site (www.biotech.wisc.edu/Arabidopsis/Guidelines.html) using 1 pl of DNA for each reaction. 100 Anthocyanin assays Anthocyanin contents of WT Ws, rns1-2, and RN52 antisense lines were measured according to Bariola et al. (1999). Levels of rnsl-2 and WT anthocyanins were determined twice, by growing the lines once on medium lacking selection and once on kanamycin selection medium (except WT). The RN52 antisense line was included in the latter as a positive control for high anthocyanin content. Representative anthocyanin content of one experiment is shown; results from the other were very similar. Error bars are not included because the RNS2 antisense line was examined only once. Bacterial growth curves Inoculations with Pseudomonas syringae pv. tomato DC3000 were performed as described by Li et a1. (2000). Samples of inoculated tissue from four different leaves of each line were taken daily by excision with a cork borer (area/leaf = 0.250 cmz). Bacteria inside the leaf discs were released by grinding the tissue in a microfuge tube in sterile water and plated on LB medium. The bacterial population was determined based on the numbers of colonies formed on LB plates, as described by Bertoni and Mills (1987). Mean results from four repetitions of the 35S-RNS1 and WT Col lines are shown. Representative results of two repetitions of the rns1-2 and WT Ws lines are shown. 101 CHAPTER 5 ANALYSIS OF AN ARABIDOPSIS THALIANA MUTANT WITH AN ALTERED RNASE PROFILE ABSTRACT Plants produce a complex series of ribonucleases (RNases), but little is known about their biological function and regulation at the molecular level. In Arabidopsis, stem tissues are particularly rich in RNases, suggesting that RNase regulation and function may be linked at the molecular level to stem growth and development. To address this possibility and to generate tools to study RNase function, we screened stem extracts of Arabidopsis to identify mutants with altered RNase profiles (mp mutants). Several mutants affecting different size classes of RNA-degrading enzymes were isolated. Of particular interest was arpl, which overproduces a doublet of 33-kDa RNases. Based on our analysis, we were able to demonstrate genetically that these activities are bifunctional nucleases that degrade both RNA and DNA. Additionally, arpl overproduces a previously unidentified small RNase of approximately 23 kDa that comigrates with the well-characterized RNase, RNSI. The elevated RNase activity caused by the arpl mutation is observed only in stem tissue and is not found in leaves or in seedlings. Other phenotypes of arpl are also stem-associated, including overproduction of the 33-kDa nucleases, short stature, and increased branching. These data indicate that the ARPI locus constitutes a novel regulator of the production of several RNase and nucleases in stems and that ARPl may be an important factor in stem growth and development. INTRODUCTION Although RNases have been used for decades as models for protein structure studies, the biological roles of these enzymes and the mechanisms by which they are regulated in plants and in other higher eukaryotes are largely unknown. In particular, functional analyses of secreted RNases that do not appear to be involved in mRNA or rRNA processing are limited. To date, fluctuations in RNase activity levels or gene expression have provided the most useful data for predicting RNase function, although the machinery involved in these regulatory pathways has not been elucidated. Additionally, organ-specific expression patterns of RNases can be found that may lend clues to their functions in vivo. For instance, the most complex pattern in mature Arabidopsis plants is found in stem extracts (Yen and Green, 1991). Our knowledge of the molecular pathways controlling stem-specific processes in plants, e.g., branching, or other prOcesses prominent in stem tissue, such as tracheary element formation (for reviews, see Schmitz and Theres, 1999, and Roberts and McCann, 2000, respectively), is limited. However, the high level of activity found in stems suggests that RNases could play important roles in such processes. The induction of multiple nuclease activities that occurs in zinnia cells undergoing differentiation into tracheary elements (Fukuda, 2000) is consistent with this argument. In contrast to the many studies on the expression patterns of RNases, very little is known concerning the factors that regulate these pathways at the molecular level. For example, while P, starvation in Arabidopsis induces RN51 and RN52, but not RN53 (Bariola et al., 1994) or BFNl (Pérez-Amador et al., 2000) expression, senescence induces all four of these activities (Taylor et al., 1993; Bariola et al., 1994; Pérez-Amador 104 et al., 2000). Evidently, there is a high degree of specificity of RNase regulation by different stimuli, but the molecular components controlling these patterns are not known, neither in plants, nor in most higher eukaryotic systems. The identification of regulators is the first step in the elucidation of the mechanisms plants use to mobilize certain sets of enzymes in response to developmental or environmental cues. To further our understanding of the regulation of RNases and thereby gain insight into their potential functions, we designed a screen to isolate mutants of Arabidopsis with altered RNase profiles. Through a gel-based screen of EMS-mutagenized M2 plants, we identified several mutants that show altered expression of at least one RNase activity. In particular, one mutant with altered levels of at least three activities provides a useful tool for the elucidation of regulatory pathways involved in RNase expression in Arabidopsis. This alteration of RNase patterns appears to be specifically associated with the stem, and physiological phenotypes of the mutant are also related to modifications in stem morphology, possibly offering new insights into processes involved in stem development and differentiation. RESULTS An RN ase-activity gel screen for mutants affecting the Arabidopsis RN ase profile To screen for mutants with altered RNase profiles, a gel-based assay for RNase activity (Yen and Green, 1991) was used to analyze protein extracts from the stems of EMS- mutagenized M2 Arabidopsis plants. Stem extracts were assayed because the stem RNase profile is more complex. containing more bands of RNase activity, than the leaf RNase profile (Yen and Green, 1991). It was previously reported that up to 13 bands of RNase activity could be distinguished on a gel containing Arabidopsis stem extracts (Yen and 105 Green, 1991). Eight bands of RNase activity were consistent and intense enough to screen for alterations in their activities. Protein extracts from 2,500 M2 plants were scored on the RN ase activity gels to identify plants whose RNase profiles differed from those of wild-type stems. Figure 5-1 depicts an example of the identification of a this type of mutant (compare the intensity of bands at approximately 33 kDa in lane 2657 with their intensities in other lanes). In the primary screen, 112 putative mutants with alterations in the Arabidopsis RNase profile were identified. The phenotypes of 60 of the 112 putative mutants could be reproduced on subsequent gels. To determine whether the phenotypes observed in such potential mutants were heritable, several progenies (M3) of each were grown, stem extracts were prepared, and the extracts were examined using the activity gel assay. Those mutants with heritable phenotypes were studied further. In total, nine gltered RNase profile (arp) mutants that showed the original phenotypes in all M3 progeny have been isolated. This chapter focuses on those mutants that affect the activity of a doublet of 33-kDa RNase activities for several reasons. Six of the nine mutants isolated display alterations in the levels of these activities. Previously, 33-kDa activities were seen on both RNase and DNase activity gels (Yen and Green, 1991). The doublet is also interesting because the RNase activities are induced in seedlings starved for inorganic phosphate (P,) (see Figure 10 in Bariola et al., 1994) and by wounding (Chapter 2). One of the six mutants affecting the 33-kDa activities was plant 2657, first identified in Figure 5-1 as showing increased RNase activity of the doublet. The phenotype of 2657 was seen more clearly in a lower percentage acrylamide gel that resolved the two bands of the doublet (Figure 5-2A). In contrast to the increased activity of the 33-kDa RNases in line 2657, relative to the wild-type activity (first and last lanes), 1 ()6 2654 2655 OFNM mmmm £0000 NNNN Figure 5-1. arp mutants are altered in RNase activity profiles Primary identification of an M2 RNase activity mutant. Stems of plant 2657 showed increased levels of two bands of RNase activity at 33 kDa. Plant numbers are shown above the gels. Approximate molecular weights (kDa) are indicated to the left. Arrows on the right point to the RNase activities affected in the mutant extract. 100 pg of protein was loaded in each lane. An extract from aerial portions of wild-type plants (leaves, stems, flowers), indicated by "A", was included as a standard. This experiment was performed by Michael Abler and Linda Danhof. 107 N 1.0 ‘0 N 3270 2048 2438 VVT 3433 VVT 2657 2972 3270 3433 2048 2438 VVT Figure 5-2. The 33-kDa RNase activities also degrade DNA. (A) RNase activity gel of stem samples from the six lines in which the 33-kDa RNase activities are affected. Plant numbers are shown above the gels and approximate molecular weights (kDa) are indicated to the left. WT, wild-type RLD stem extracts. (B) DNase activity gel of the six mutants in which the 33-kDa RNase activities are affected. Denatured single-stranded DNA was substituted for RNA as the substrate in the gel. This experiment was performed by Michael Abler. 108 plants 2972, 3270, 3433, 2048, and 2438 all greatly diminish or abolish the activity of the lower band of the doublet (Figure 5-2A). The 33-kDa RN ase activities are bifunctional nucleases Previous data from our laboratory indicated that there is also a doublet of DNase activity at 33 kDa, but it was not known whether both the RNase and DNase activities were derived from the same enzyme(s) (Yen and Green, 1991). The availability of several plants affecting the 33-kDa RNase activities allowed us to test this hypothesis by analyzing the DNase activities in these mutants. Figure 5-2A shows the RNase gel phenotypes for the six plants affected in the activity of the 33-kDa doublet. Shown in Figure S-ZB, protein extracts from the same plants were run in a gel containing denatured single-stranded DNA as the substrate for the enzyme activity. The profile of DNase activities in Arabidopsis stem extracts is far simpler than the RNase profile, and the 33-kDa doublet is the most abundant of the DNase activities in wild-type stems. In each of the plants, the alteration in the DNase gel phenotype corresponds with that in the RNase gel for the 33-kDa activities. Because all of the mutations affect the DNase and RNase activities in the same manner, these data demonstrate genetically that the 33-kDa doublet corresponds to a pair of bifunctiOnal nuclease activities, capable of degrading either RNA or DNA. Genetic characterization of the arp mutants Genetic analyses of the plants affected in the 33-kDa activities were performed. Each line was backcrossed to WT RLD to determine the inheritance patterns of the mutations. The RNase profiles of the F , plants from each of the crosses were indistinguishable from that of the wild type, thereby indicating that all nine mutations were recessive (Abler et al., 109 manuscript in preparation). Segregation of the mutant phenotype in the F2 generation showed that the mutations were inherited in a Mendelian fashion, fitting a 3:1 ratio of wild type to mutant (Abler et al., manuscript in preparation). Complementation tests of the six mutants affecting the 33-kDa nuclease activities indicated that the mutations correspond to two loci and that the 2657 mutation affects a locus distinct from that affected in the other four mutants (Abler et al., manuscript in preparation). Accordingly, the mutant locus generating the gel phenotype exhibited by plant 2657 (increased 33-kDa activities) was designated arpl. During subsequent mapping experiments, it was determined that the gel phenotype of mutant lines 2048, 2438, 2972, 3270, and 3433 resembled the pattern of stem RNase activities in the Columbia ecotype of Arabidopsis (not shown and LeBrasseur et al., in press). From this observation and subsequent RFLP analysis, we concluded that these lines were in fact Columbia contaminants in the screen, which had been performed using the ecotype RLD. Nevertheless, our results demonstrate that a single locus confers ecotype-specific control of the lower band of the 33-kDa doublet in wild-type stems. Morphology of the arpl mutant Examinations of growth and development indicated that a visible phenotype is associated with altered expression of the 33-kDa nuclease activities in arpl. As shown in the example in Figure 5-3A, arpl plants appear shorter and more branched than wild type. To quantify these visible phenotypic differences, two blind studies were performed in which approximately 40 plants each (wild type and arpl) were scored. Seeds from the arpl line and from wild-type plants were planted in pots, and the order of the pots was 110 LLLI III 000 40 -- 883%“ Plant Height (cm) 20 -» Figure 5-3. Morphological differences in the phenotypes of wild—type and arpl mutant plants. (A) Six-week-old arpl and wild-type (WT) plants are shown. The arpl plants appear shorter and more branched than wild type plants. (B) Scatter plot of final plant height of wild-type and arpl plants. Individuals in the respective populations are represented by circles. The average plant heights for the populations are represented as horizontal bars. This experiment was performed by Michael A bier. lll randomized. The time for each plant to reach certain developmental stages (e.g., germination, flowering) was measured, as were physical (e.g., plant height) characteristics of the plants. Differences between populations were analyzed by the Mann-Whitney rank-sum test. Significant growth and developmental differences between wild-type Arabidopsis and arpl plants were confirmed in the blind study. Figure 5-3B depicts the differences in height between arpl and wild type. The data in Table 5-1 demonstrate statistically that arpl plants were more heavily branched and had shorter siliques than did wild type. The differences between the wild-type and arpl plants were evident in plant height, number of branches, and silique length. Developmental processes (days to bolting, appearance of first flower and branch flower) did not differ between the populations. The altered morphological phenotypes of the arpl mutants are likely due to the arpl mutation itself, since these phenotypes have always cosegregated with the gel phenotypes through four generations of backcrosses (not shown). The arpl mutation induces an RNase activity in addition to the 33-kDa nuclease activities Another difference observed in the RNase profiles of arpl and wild-type inflorescence stems was the appearance of an approximately 23-kDa band of RNase activity in arpl plants. Like the morphological phenotypes, the 23-kDa band of RNase activity consistently cosegregated in plants showing the increased 33-kDa nuclease arpl phenotype through four generations of backcrosses. Initially, this 23-kDa band was not scored because it was not seen consistently in wild-type Arabidopsis plants. However, as shown in Figure 5-4, RNSI produced in yeast comigrates with the 23-kDa RNase activity Table 5-1. Statistical analysis of morphological differences between WT RLD and arpl plants“ arpl vs. WT Parameterb T P Plant height (final) 6.89 <0.000001 6.93 <0.000001 Number of branches 3.56 0.00037 4.26 0.000020 Silique length 9.03 <0.000001 1 1.51 <0.000001 aThe Mann-Whitney rank-sum test was used to determine the statistical significance of the morphological differences between populations. Parameters with reproducibly significant differences are listed in the first column. Ir’Other parameters (measured in days from planting) that were not significantly different include germination, 2, 4, and 6 rosette leaves, bolting, first flower, lO-cm bolt, 15-cm bolt, branch flower, and axillary stem flower. This analysis was performed by Michael A bier that is elevated along with the 33-kDa doublet in arpl plants. Both the original mutant (arpl M6) and a mutant recovered after four generations of backcrosses (arpl BC4®) have RNase activities that comigrate with yeast-produced RNSI. Neither the WT plants (WT), nor plants from the BC4® generation with a wild-type ARPI allele (ARPI BC4®), have activities that comigrate with RNSI (Figure 5-4). Since the 23-kDa band comigrates with RNSI, RNA gel blot and immunoblot analyses were performed to determine whether RN51 mRNA or protein levels were affected in arpl. For RNA blot analysis, RNA was isolated from stems of 6-week-old arpl and WT RLD plants. Poly(A)+ RNA was purified from total RNA preparations so that small effects might be detected. RN51 transcript was not detected in either total or poly(A)+ RNA in WT RLD or arpI stems by RNA gel blot analysis using an RN51 probe (Figure 5-5A). RNA from wounded seedlings was included as a positive control for RN51 hybridization [Wnd and Wnd Poly(A)+]. Since we did not detect increased RNSI mRNA levels in arpl stems, we were interested in whether the RNSI protein was elevated because of a possible posttranscriptional effect on RNSI. To examine this possibility, proteins from arpl and WT RLD stems and seedlings were separated by SDS-PAGE and subjected to immunoblot analysis (Figure 5- SB). RNSI purified from yeast cells and protein from P,-starved and P,-supplied seedlings were included as controls (RNSI, P-, and P+, respectively). Although the RNSI antibody (Bariola et al., 1999) detects the yeast-produced RNSI and RNSI in P,- starved seedlings, no RNSI was detected in arpl stems. Therefore, significant increases in RN51 transcript as well as protein could not be found in arpl stems compared to WT. 114 Figure 5-4. The arpl mutant has increased levels of an activity that comigrates with RNSI. Approximate molecular weights are shown to the left of the gel. 100 pg of protein was loaded in the arpl and wild-type RLD (WT) lanes. Lane RNS] contains an aliquot of culture medium from yeast cells expressing the Arabidopsis RN51 gene from a yeast secretion vector as described (Bariola et al., 1994). ARPl BC4®, F l progeny plant of the fourth backcross of arpl with wild-type RLD displaying wild-type levels of the 33-kDa nucleases; arpl BC4®, F, progeny plant of the fourth backcross of arpl with wild-type RLD displaying increased levels of the 33-kDa nucleases; arpl M6, sixth generation of mutant line 2657. This experiment was performed by Michael Abler. 115 A $.21, <