s . £3 . 2.1;...2 . . . , . . , . 1...... .u .a . . . . inhuw. .........n.....6.. . . . . , . w: 9 , . . .. £51,...” I A In. '0. ‘ .kufmumrfidmfixzffin . préwfi. It. nah... 033328.}... . . .. Sig . it bun w. I .2- z . . . 1131.1» HW—ztil .x . :5: 1:; .113". f 3... 51.2.}... 3.1;}... x itinlbcl“ \7:"\¢"JIB‘ ." . 2x6. .1. {fili’I-I“ '- . Plt~r {r 521...... . tuna! a..vt..zl.h.a. 5.. . 13.1 A53... O~ Lit: x; A5533» Asia?! I: 3.2.. ltlgl-a.txt : !: .x 3:... .12 (95:... 1 .9. . , . : ,. . , EL. L “$9. .. _. ,: 5.4% x g... ghagr .. .. . . . m .26. .‘ L .I: iqv... megs IHIIIIIIIllllllllllllllHIIHIIHIIIIIllllllilillllillllflili 9' 1 3 1293 02048 £4200 LIBRARY Michigan State University This is to certify that the dissertation entitled GENETIC DETERMINANTS OF mRNA STABILITY IN PLANTS presented by Mark Aikens Johnson has been accepted towards fulfillment of the requirements for Ph.D. degree in Microbiology £4044 or professor Date C/éé/fld MS U is 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 11/00 cJCIRcmatooue.pes-p.14 THE GENETIC DETERMINANTS OF mRNA STABILITY IN PLANTS By Mark Aikens Johnson A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Microbiology Program in Cellular and Molecular Biology 1999 ABSTRACT GENETIC DETERMINANTS OF mRNA STABILITY IN PLANTS By Mark Aikens Johnson Regulation of mRNA stability is now known to be an important means of controlling gene expression. Despite many exciting advances in recent years, there are still many unresolved questions about the mechanisms responsible for mRNA degradation. The aim of this thesis was to address how plant cells recognize unstable messages and to understand the mechanisms the cell uses to degrade them. The work presented here provides important information as well as a foundation for mechanistic insights that will be made in the future. In Section One, Sequence-Specific mRNA Degradation, the focus is on the DST element, which targets certain plant transcripts for rapid mRNA degradation. Mutants of Arabidopsis were isolated that do not fully recognize this instability signal. At least two genes, dstl and d312, are shown to be involved in the recognition of the DST sequence. The cloning of these genes will be of tremendous interest and will likely be very informative about the recognition of unstable mRNAs. Also in this section, a dissection of the SA UR-A CI 3’ untranslated region (UTR) is described. This 3’UTR contains one DST element and several DST-like subdomains. Studies presented here indicate that redundant DST-like sequences located upstream of the DST element may play an important role in the instability fiinction of the SA UR-A C1 3’UTR. In the fiiture it will be interesting to analyze the effect of these redundant DST-like sequences on mRNA stability in the dstl and dst2 mutants. In Section Two, General Mechanisms omeNA Degradation, two sets of experiments address how the plant cell degrades mRNA. To get a glimpse into this process, poly(G) tracts were inserted into reporter transcripts and these were introduced into plant cells in an attempt to stabilize intermediates in the process of mRNA degradation. Such intermediates were not observed, which may indicate that there are some differences between the mechanisms of mRNA turnover in plants and yeast, where poly(G) tracts are known to stabilize mRNA decay intermediates. The second set of experiments addresses the role of a potential plant poly(A) ribonuclease, AtPARN, in plant mRNA decay. It has been shown that removal of the poly(A) tail leads to rapid mRNA degradation in yeast and mammals. Preliminary experiments indicate that AtPARN has poly(A) ribonuclease activity in vitro, suggesting that it may play a role in the degradation of mRNA in plants. ACKNOWLEDGMENTS I would like to thank Pam Green for being my mentor. The experiments described in this thesis would not have been possible if not for her creativity, support, guidance, encouragement, and enthusiasm. Pam has fostered a laboratory environment that helps to make our work enjoyable and intellectually rewarding. Many people have contributed to this work and while specific acknowledgements are given at the end of each chapter there are a few people who deserve special thanks. Miguel Pérez-Amador and I have worked together throughout the course of my Ph.D. He has been an extremely helpful and caring colleague without whom the selection of dst mutants would have been even more difficult. Likewise, Michael Sullivan was extremely helpful when I first joined the laboratory and laid the groundwork for the mutant selection. Linda Danhof has kept our laboratory well organized and has maintained an army of undergraduate helpers. Jonathon Vogel, who started out in Linda’s army, has made countless contributions to my work and has provided me with the pleasure of helping him to learn how to do science. Preet Lidder and I worked together during her rotation. She has and will continue to make important contributions to our study of the dst mutants and to AtPARN. Thanks also to Ellen Baker and Jim Colbert who, along with Pam, were my co-authors on the review article that became the introduction to this thesis. I am also grateful for my thesis committee: Frans deBruijn, Donna Koslowsky, and Mike Thomashow; who have taken time out of their schedules to give me excellent advice at several important stages in this process. Members of the Green Lab, past and present, have had a tremendous impact on these experiments and on the way I think and communicate about science. They have iv been a great bunch of fiiends and colleagues. My thanks go to Jim Kastenmayer, Nikki Lebrasseur, Michael Feldbriigge, Rodrigo Gutierrez, Gustavo MacIntosh, Mike Abler, Pauline Bariola, Jay De Rocher, Scott Diehn, Pedro Gil, Christy Howard, Deb Thompson, and Ambro van Hoof. My thesis is dedicated to Carol, who has made this tremendous learning experience a time of great joy for me, and to my entire family (my grand parents: Henry and Virginia Aikens, Wilhelm and Hilda Johnson; my parents: Tom and Kathy Johnson, Scott and Sherie Barnes; my brothers and sisters: Suzy, Jenny, Kate, and Andy Johnson, and Bill Barnes; and all of my aunts, uncles, and cousins) who have been with me the whole way. TABLE OF CONTENTS LIST OF TABLES ............................................................................................................. ix LIST OF FIGURES ............................................................................................................. x INTRODUCTION ............................................................................................................... l DETERMINANTS OF MESSENGER RNA STABILITY IN PLANTS mRNA INSTABILITY SEQUENCES ........................................................................... 5 Instability Sequences in SA UR genes .......................................................................... 6 AUUUA sequences ................................................................................................... 10 Premature stop codons and other translational influences on mRNA stability ......... 11 Sequences involved in regulated mRNA degradation ............................................... 14 T RANS-ACTING FACTORS INVOLVED IN mRNA STABILITY ........................... 16 RNA binding proteins ............................................................................................... 16 RNA-degrading activities .......................................................................................... 18 The role of the cap and poly(A) tail in mRNA stability ............................................ 21 Poly(A)-binding proteins ........................................................................................... 24 PATHWAYS OF mRNA DEGRADATION ................................................................ 25 POST-TRANSCRIPTIONAL GENE SILENCING ..................................................... 31 FUTURE DIRECTIONS ............................................................................................... 33 THE SCOPE OF THIS THESIS ................................................................................... 34 REFERENCES .............................................................................................................. 37 SECTION I: SEQUENCE-SPECIFIC mRNA DEGRADATION CHAPTER l-l .................................................................................................................. 49 A PROCEDURE TO SELECT MUTANTS OF ARABIDOPSIS DEFECTIVE IN SEQUENCE-SPECIFIC mRNA DEGRADATION INTRODUCTION ......................................................................................................... 50 RESULTS ...................................................................................................................... 53 Premise of the mutant selection strategy ................................................................... 53 Selection of transgenic plants to serve as the parental lines for mutagenesis ........... 58 Lines p1514-9 and p15 19-31 showed decreased hygromycin resistance ................. 62 Mutagenesis of lines p1514-9 and p15 19-31 ............................................................ 63 Selection of putative mutants .................................................................................... 64 DISCUSSION ............................................................................................................... 66 MATERIALS AND METHODS .................................................................................. 70 Plasmid construction and plant transformation ......................................................... 70 Plant material and growth conditions ........................................................................ 71 EMS mutagenesis and selection of hygromycin resistant plants .............................. 72 ACKNOWLEDGEMENTS .......................................................................................... 73 REFERENCES .............................................................................................................. 75 vi CHAPTER 1-2 .................................................................................................................. 79 MUTANTS OF ARABIDOPSIS DEFECTIVE IN A SEQUENCE-SPECIFIC mRNA DEGRADATION PATHWAY REFERENCES AND NOTES ...................................................................................... 94 CHAPTER 1-3 .................................................................................................................. 98 ANALYSIS OF THE DST ELEMENT AND DST-LIKE SUBDOMAINS WITHIN THE SA UR-A C1 3’ UNTRANSLATED REGION INTRODUCTION ......................................................................................................... 99 RESULTS .................................................................................................................... 101 Mutagenesis of all DST and DST-like subdomains ................................................ 101 Gain of function analysis of DST and DST-like subdomains ................................. 105 DISCUSSION ............................................................................................................. l 10 MATERIALS AND METHODS ................................................................................ 115 Plasmid construction and plant transformation ....................................................... 115 Plant material and grth conditions ...................................................................... 117 RNA analysis ........................................................................................................... 117 AKNOWLEDGMENTS ............................................................................................. l 18 REFERENCES ............................................................................................................ l 19 SECTION 2: GENERAL MECHANISMS OF mRNA DEGRADATION CHAPTER 2-1 ................................................................................................................ 121 USE OF POLY(G) IN SERTIONS TO GAIN INSIGHT INTO THE STEPS INVOLVED IN PLANT rnRNA DEGRADATION INTRODUCTION ....................................................................................................... 122 RESULTS .................................................................................................................... 125 DISCUSSION ............................................................................................................. 128 MATERIAL AND METHODS .................................................................................. 131 Plasmid construction and plant transformation ....................................................... 131 RNA analysis ........................................................................................................... 133 RNase H treatment .................................................................................................. 134 ACKNOWLEDGEMENTS ........................................................................................ 134 REFERENCES ............................................................................................................ 135 vii CHAPTER 2-2 ................................................................................................................ 138 CLONING AND INITIAL CHARACTERIZATION OF A POLY(A) RIBONUCLEASE FROM ARABIDOPSIS INTRODUCTION ....................................................................................................... 139 RESULTS .................................................................................................................... 141 The RNaseD family and cloning of AtPARN ......................................................... 141 AtPARN is expressed throughout the plant ............................................................ 144 Characterization of AtPARN activity in vitro ......................................................... 145 DISCUSSION ............................................................................................................. 146 MATERIALS AND METHODS ................................................................................ 150 Cloning of AtPARN ................................................................................................ 150 Analysis of AtPARN expression .............................................................................. 151 Expression of AtPARN in E. coli ............................................................................. 151 TCA PARN assay .................................................................................................... 153 ACKNOWLEDGEMENTS ........................................................................................ 154 REFERENCES ............................................................................................................ 155 viii LIST OF TABLES INTRODUCTION Table 1. Sequences that have been demonstrated to control mRNA stability in plants. .................................................................................................... 6 SECTION I: SEQUENCE-SPECIFIC mRNA DEGRADATION CHAPTER 1-1: A PROCEDURE TO SELECT MUTANTS OF ARABIDOPSIS DEFECTIVE IN SEQUENCE-SPECIFIC mRNA DEGRADATION Table l-l-l. EMS mutagenesis of transgenic Arabidopsis lines. .................................... 64 Table 1-1—2. Hygromycin selection experiments ............................................................. 65 ix LIST OF FIGURES INTRODUCTION Figure 1. Critical regions of the DST sequence ................................................................. 8 SECTION 1: SEQUENCE—SPECIFIC mRNA DEGRADATION CHAPTER l-l: A PROCEDURE TO SELECT MUTANTS OF ARABIDOPSIS DEFECTIVE IN SEQUENCE-SPECIFIC mRNA DEGRADATION Figure l-l-l. A strategy for selecting mutants defective in a sequence-specific mRNA decay pathway. .................................................................................................................. 54 Figure 1-1-2. The basic structure of transgenes used to assess various reporters and instability determinants in transgenic Arabidopsis calli. .................................................. 56 Figure 1-1-3. Accumulation of reporter transcripts in Arabidopsis ................................. 57 Figure 1-1-4. T-DNA constructs used in mutant selection strategy. ................................ 59 Figure 1-1-5. Southern blot analysis of candidate lines for mutagenesis ......................... 60 Figure 1-1-6. Northern blot analysis of candidate lines for mutagenesis ......................... 61 Figure 1-1-7. p1514-9 and p1519-3l Show increased hygromycin sensitivity ................. 62 CHAPTER 1-2: MUTANTS OF ARABIDOPSIS DEFECTIVE IN A SEQUENCE- SPECIFIC mRNA DEGRADATION PATHWAY Figure 1-2-1. Structure of transgenes and kinetics of degradation of reporter transcripts used for dst mutant selection ............................................................................................. 84 Figure 1-2-2. The dst selection strategy. .......................................................................... 86 Figure 1-2-3. mRNA abundance of selected transcripts in dst] and dst2. ....................... 87 Figure 1-2-4. Analysis of HPH-DST mRNA stability in leaves from p1519-31, dstI, and dst2 plants. ............................................................................... 91 CHAPTER 1-3: ANALYSIS OF THE DST ELEMENT AND DST-LIKE SUBDOMAINS WITHIN THE SA UR-ACI 3’ UNTRANSLATED REGION Figure 1-3-1. The SA UR-A C1 3’UTR contains one DST-element and DST-like subdomains. ..................................................................................................................... l 02 Figure 1-3-2. The structure of chimeric genes used to test mutations in the SA UR-A C1 3’UTR. ............................................................................................... 103 Figure 1-3-3. Mutations in the SA UR-A C1 3’ UTR affect polyadenylation of the Globin reporter transcript. ........................................................................................................... 104 Figure 1-3-4. Structure of chimeric genes used to test individual subdomains within the SA UR-A C1 3’UTR for instability function. .................................................................... 106 Figure 1-3-5. Analysis of the N, R, and D regions of the SA UR-A C1 3’ UTR. .............................................................................................. 107 Figure 1-3-6. Analysis of combinations of N, R, and D regions .................................... 109 SECTION 2: GENERAL MECHANISMS OF mRNA DEGRADATION CHAPTER 2-1: USE OF POLY(G) INSERTIONS TO GAIN INSIGHT INTO THE STEPS INVOLVED IN PLANT mRNA DEGRADATION Figure 2-1-1. The effect of a poly(G) insertion on the major yeast mRNA degradation pathway. .......................................................................................... 125 Figure 2-1-2. Poly(G) tracts were inserted into a variety of plant reporter transcripts .................................................................................................. 126 Figure 2-1-3. Poly(G) stabilized intermediates do not accumulate in Arabidopsis. ...... 128 CHAPTER 2-2: CLONING AND INITIAL CHARACTERIZATION OF A POLY(A) RIBONUCLEASE FROM ARABIDOPSIS Figure 2-2-1. Sequence alignment of the AtPARN and HuPARN amino acid sequences ...................................................................................................... 143 Figure 2-2-2. AtPARN mRNA is found throughout the plant. ...................................... 144 Figure 2-2-3. AtPARN has poly(A) ribonuclease activity. ............................................ 146 xi INTRODUCTION DETERMINANTS OF MESSENGER RNA STABILITY IN PLANTS In its original form, this introduction was published in “A Look Beyond Transcription: Mechanisms Determining mRNA stability and Translation in Plants”, Julia Bailey-Serres, Daniel R. Gallie, eds. Copyright 1998, American Society of Plant Physiologists (Johnson et al., 1998). Changes have been made to update the content and bring the text into the larger context of this dissertation. INTRODUCTION The abundance of an RNA molecule depends both upon the frequency with which it is transcribed and the rate at which it is degraded. This statement holds true regardless of whether one is considering mature mRNA molecules encoded by nuclear genes, plastid-encoded mRNAs, nuclear-encoded rRNAS, or the abundance of a particular species of primary transcript in the nuclear compartment. In each of these instances degradative processes play an important role in the amount of RNA that is present, and potentially, in the final level of gene expression. The focus of this introduction will be on the degradative events affecting the abundance of mature mRNAs encoded by nuclear genes in plants. It is generally assumed that these types of degradative events occur within the cytosolic portion of the cell. However, it also seems clear that nuclear degradative events may play a critical role in the establishment of cytosolic mRNA levels in particular organs, tissues, or cells (Kamalay and Goldberg, 1984; Okamuro and Goldberg, 1989). The significance of mRNA degradation to gene expression is aptly demonstrated by attempts to express the Bacillus thuringiensis toxin genes (cry genes) in transgenic plants. Wild-type versions of cry genes are typically not expressed at high levels in transgenic plants (Adang et al., 1993), even when fused to the strong cauliflower mosaic virus 358 (358) promoter. High level expression has been obtained by use of synthetic, plant-like versions of the cry genes that contain codon usage and other changes (Adang et al., 1993; Koziel et al., 1993; Stewart et al., 1996; reviewed in Diehn et al., 1996). The plant-like cry genes are generally thought to produce mRNAs that exhibit longer half-lives in plant cells (Diehn et al., 1996), leading to higher level accumulation of the toxin. There is direct evidence for this in the case of a plant-like synthetic cry] gene which encodes a transcript that is considerably more stable than the wild-type crjyIA (c) transcript (De Rocher et al., 1998). The half-lives of mRNA molecules can vary over a wide range. Typical plant mRNA molecules appear to exhibit half-lives of several hours (Siflow and Key, 1979; Sullivan and Green, 1993; Taylor and Green, 1995). Relatively unstable plant mRNAs, with half-lives of an hour or less, have also been described (McClure and Guilfoyle, 1989; Braam and Davis 1990; Seeley et al., 1992; Taylor and Green, 1995; van Hoof and Green, 1996). On the other end of the Spectrum are very stable mRNAs with half-lives of days or more, some of which may be stored or sequestered from the mRNA decay machinery. In addition, there are some well-characterized examples of mRNAs whose half-lives can vary in response to specific stimuli. In mammalian cells, the transferrin receptor mRNA is stabilized 20- to 30-fold under conditions of low intracellular iron concentration (Casey et al., 1989; Koeller et al., 1991). Mammalian histone and B- tubulin mRNAs also exhibit regulated variations in mRNA half-life (Ross, 1995). In plants, a good example is the PvPRPI mRNA which is destabilized in the presence of fungal elicitor (Zhang et al., 1993; Mehdy and Brodl, 1998). There are a number of ways to measure mRNA decay rates in plant systems. These include measuring the half-lives of endogenous or in vitro synthesized mRNAs in protoplasts (Gallie et al., 1989) or in vitro (Byrne et al., 1993; Tanzer and Meagher, 1994). Alternatively, mRNA decay rates can be monitored in transformed or nontransfonned cultured cells or intact plants, following treatment with a transcriptional inhibitor (Newman et al., 1993; Byrne et al., 1993). Recently, repressible promoters have also been used to shut off transcription of the genes of interest in cultured cells or plants so that the stability of the corresponding mRNA can be measured without the use of general transcriptional inhibitors (Weinman, et al., 1994; Gil and Green, 1996; Petracek et al., 1998). For a more detailed description of these methods and a discussion of the advantages and limitations of each, readers are referred to reviews by Abler and Green (1996) and Ross (1995). The refinement of methods for measuring mRNA stability, as well as mounting evidence that many plant genes are regulated at this level, has led to considerable progress in recent years. Information has been obtained about the contribution of the cap and poly(A) tail, and it has been Shown that plants have multiple poly(A) binding proteins that can be differentially regulated. Another important step has been the identification of Specific sequences that control mRNA decay rates. In particular, several sequences that cause rapid mRNA degradation have been characterized. These and other studies have increased our perception of the types of cellular factors that may be involved in mRNA decay and the mechanisms that may be involved. Finally, it has become apparent that mRNA decay mechanisms may play a prominent role in certain forms of gene silencing such as the phenomenon of post-transcriptional cosuppression that is sometimes observed in transgenic plants. All of these topics will be discussed in this introduction in an effort to present our current understanding of the components and mechanisms that determine the inherent stabilities of different mRNAs in plants. Emphasis will be placed on describing the most recent findings, many of which pertain to rapid mRNA decay mechanisms that allow plants to respond quickly to internal and external stimuli. The purpose of this introduction is to highlight what is known about mRNA turnover in plants and to point out how the work presented in this dissertation addresses some of the many important questions that are being addressed in this field. mRNA INSTABILITY SEQUENCES Rapid turnover of mRNA is thought to be an active process. Indeed, cis-acting sequences have been identified that target transcripts for rapid turnover in plants as well as in other systems (reviewed in Ross, 1995; Abler and Green, 1996; Caponigro and Parker, 1996; van Hoof and Green, 1997). The DST element, found in the 3' untranslated region (UTR) of the unstable small auxin up RNAS (SA URS), is one such instability sequence that so far appears to be unique to plants (McClure et al., 1989; Newman et al., 1993; Gil et al., 1994; Gil and Green, 1996; Sullivan and Green, 1996). Another element that can target transcripts for rapid turnover in plants consists of repeats of AUUUA pentamers (Ohme-Takagi et al., 1993). Multiple repeats of this pentamer are a common feature of many highly unstable mammalian transcripts (reviewed in Chen and Shyu, 1995). Premature stop codons have also been Shown to target transcripts for rapid decay in plants as well as in yeast, Caenorhabditis elegans, and mammalian cells (reviewed in Maquat, 1995; van Hoof and Green, 1997; Jacobson and Peltz, 1996). In this section, the plant instability sequences that have been most thoroughly characterized will be described. These are summarized in Table l. Table 1. Sequences that have been demonstrated to control mRNA stability in plants. Sequence Description Experimental System Ref AUUUA AUUUA repeats are a common feature of t“; measured during 1 AU-rich instability determinants found in an ActD time course labile mammalian cytokine and in stably transformed protooncogene transcripts. l l repeats of BY-Z cells" the AUUUA pentamer caused instability. DST A highly conserved 40 base sequence tm measured during 2,3 found downstream of the stop codon in an ActD time course unstable SAUR transcripts. A tandem in stably transformed dimer of DST caused instability. BY-2 cells“ SA UR-ACI Approximately 140 bases, includes one tm measured using 4 3' UTR highly conserved DST element and ToplO promoter in several repeated ATAGAT-like and stably transformed GTA-like subdomains. BY-2 cells" Premature Naturally occurring and in vitro t1 [2 measured during 5 Stop generated premature stop codons in the an ActD time course Codon coding region of the PHA transcript. in stably transformed BY-2 cells“ iLRE 5' UTR plus the first 14 codons of t”; measured 6 F ed-I ; controls light regulation of using ToplO mRNA stability. promoter in transgenic tobacco plants ActD, actinomycin D; UTR, untranslated region; Ref, References; PHA, Phytohemagglutinin; BY-2, bright yellow-2 (NT-l, tobacco cell line); tug, half-life References: l) Ohme-Takagi et al., 1993; 2) Newman et al., 1993; 3) Sullivan and Green, 1996; 4) Gil and Green, 1996; 5) van Hoof and Green, 1996; 6) Petracek et al., 1998) *Confirmed in transgenic plants by measrmg mRNA accumulation Instability Sequences in SA UR genes mRNAs encoded by the auxin induced SA UR genes are among the most unstable plant transcripts and therefore offer an excellent model with which to study rapid mRNA turnover. The half-lives of these transcripts have been estimated to be between 10 and 50 minutes depending upon the method used in the analysis (McClure and Guilfoyle, 1989; Franco et al., 1990). Seven SA UR genes have been cloned from soybean, mungbean, and Arabidopsis thaliana all of which contain a highly conserved, 40 base pair sequence motif in their 3' UTRs (McClure et al., 1989; Yamamoto et al., 1992; Gil et al., 1994). This sequence was termed DST because ofits location downstream of the SA UR stop codon (McClure et al., 1989). The highly conserved nature of this sequence and its placement in the 3' UTR of the SA UR transcripts led to the suggestion that the DST element may control the stability of the SA UR transcripts (McClure et al., 1989; Franco et aL,1990) The destabilizing function of the DST element was tested directly by Newman et a1. (1993). They demonstrated that a synthetic dimer of the DST element, when placed in the 3' UTR of the B-glucuronidasc (GUS) or the ,BLglobin reporter transcript, is able to target these reporter transcripts for rapid turnover relative to controls lacking DST sequences. In these experiments, half-lives of reporter mRNAs were measured in stably transformed tobacco (BY-2) cell suspension cultures by inhibiting new transcription using actinomycin D and monitoring the decay of the reporter transcript over a two-hour time course. It is likely that DST sequences act as potent instability determinants in intact plants as well as in cultured cells because DST elements also cause a marked decrease in mRNA abundance relative to controls in transgenic tobacco plants. The DST element consists of three highly conserved subdomains separated by two variable regions. Site-directed mutagenesis experiments have been performed in order to determine which features of the DST element are most important for its fiinction as an instability determinant (Sullivan and Green, 1996). The results of these experiments are illustrated in Figure 1. Highly conserved subdomains are in boxes. The second and third subdomains contain residues that are invariant among all DST elements and are termed ATAGAT and GTA, respectively. Both of these subdomains have been found to be necessary for DST function. This was shown by constructing mutated DST elements in which either the ATAGAT or GTA subdomains were disrupted with five or six base substitutions, respectively. These mutant elements were tested as dimers placed in the 3' UTR of the ,B-globin reporter transcript. The effects of the mutations were analyzed by measuring mRNA accumulation and half-life in stably transformed tobacco cell cultures and mRNA accumulation in leaves of transgenic tobacco plants. These mutations resulted in mRNA half-lives that were nearly as long as a control lacking DST sequences and in the restoration of ,B-globin mRNA abundance to control levels in stably transformed tobacco cell cultures and transgenic tobacco leaves. In order to pinpoint critical bases within these two invariant subdomains, two base pair (CC) substitution mutations were analyzed. The results indicated that the first four bases of the ATAGAT subdomain are most important for DST function in tobacco cell culture. Interestingly, a CC substitution in the invariant GTA sequence led to inactivation of the DST element in transgenic tobacco leaves but not in cultured cells. This finding may indicate that the DST element is recognized differently in different cell types. Critical in all systems tested [@6- - 5 - GETAGATTIG - 7 -|CAT_I'ITIIGTA I ‘-v-’ H“ Critical in both Critical in leaves cultured cells but not and leaves cultured cells Figure 1. Critical regions of the DST sequence. DST sequences, found in unstable SA UR transcripts, consist of three highly conserved regions (boxed) separated by two variable regions. The second and third regions contain the invariant sequences ATAGAT and T-- GTA, respectively. Sequence requirements for function of the DST element were established using a series of substitution mutations introduced into a dimer of the soybean SA UR 15A DST element. All mutations were tested in both copies of the dimer. Sequences shown to be critical using five and six-base substitution mutations are highlighted above the sequence. Critical sequences identified with two-base pair substitution mutations are highlighted below the sequence. Although a function for SA UR gene products has not been demonstrated, the expression patterns of these transcripts indicate that they may have a role in early events in auxin-induced cell elongation. The SA UR-A C1 gene of A. thaliana has been carefiilly analyzed in order to determine which regions of the gene are responsible for SA UR expression characteristics, namely auxin induction and rapid mRNA turnover (Gil et al., 1994; Gil and Green, 1996). By constructing chimeric genes that consisted of the SA UR- ACI promoter, coding region, or 3' UTR fused to reporter genes, it was shown that the promoter is the Site of auxin action, the 3' UTR is largely responsible for rapid mRNA turnover and the coding region contributes to low mRNA abundance but not by decreasing message half-life (Gil and Green, 1996). It was shown that the 3' UTR of SA UR-ACI functions as a determinant of rapid mRNA turnover by fusing this region of the SA UR gene to the B-globin coding region whose expression was driven by a tetracycline-repressible promoter known as ToplO (Weinman et al., 1994). Following tetracycline treatment, mRNA stability can be measured without the use of general transcriptional inhibitors such as actinomycin D. The SA UR-A C I 3' UTR contains one canonical DST element which is located 80 bases downstream of the stop codon and 10 bases upstream of the poly(A) addition site (Gil et al., 1994). It is notable that experiments with the synthetic DST element, described above, indicated that two copies were required for instability function. It is not clear whether this apparent difference reflects an absolute DST recognition requirement or if it reflects a necessary context within the SA UR-A C l 3' UTR. Interestingly, there are several ATAGAT-like and GTA-like subdomains of the DST sequence located just upstream of the classically defined DST element within the SA UR-A CI 3' UTR. These redundancies may be serving as multiple recognition Sites for DST mediated mRNA decay within the SA UR-A C1 3' UTR. AUUUA sequences AU-rich elements are found in the 3' UTRS of several of the most unstable mammalian transcripts (Chen and Shyu, I995). The genes that encode these transcripts have functions involved in cell division and differentiation and their expression is very stringently controlled. Indeed, overexpression of some of these genes can lead to oncogenesis and it is thought that AU-rich elements serve to limit expression by targeting their transcripts for rapid decay. Repeats of the pentamer, AUUUA, are often found in these AU-rich elements and have been shown to be important for their function (Caput et al., 1986; Shaw and Kamen., 1986; Vakalopoulou et al., 1991; Shyu et al., 1991). A repeat of eleven AUUUA pentamers has been shown to target reporter transcripts for rapid degradation in plants (Ohme-Takagi et al., 1993). This was demonstrated by fusing the repeated AUUUA element downstream of the flglobin or GUS coding region and comparing the half-life of these transcripts to that of controls lacking AUUUA repeats during actinomycin D time courses. The sequence of the element rather than its A+U content appears to be most important because an AU-rich control lacking AUUUA repeats had little effect on fl-globin or GUS mRNA stability. These results indicate that the mRNA decay pathway mediated by AUUUA repeats may be conserved between animals and plants. There is no evidence that this pathway functions in yeast. The natural targets of the plant AUUUA mediated decay 10 pathway are unknown but one possible candidate is PvPRPI from Phaseolus vulgaris. The PvPRPI transcript, which appears to encode a cell wall protein, is rapidly degraded (t1/2s45') following the addition of fungal elicitor to bean cell cultures (Zhang et al., 1993). Zhang and Mehdy have reported a 50 kd protein that can be specifically UV cross-linked to the 3' UTR of the PvPRPI transcript (Zhang and Mehdy, 1994). The binding of this 50 kd polypeptide, termed PRP-BP, has been mapped to a 27 bp sequence that contains an AUUUA motif (Zhang and Mehdy, 1994). Although the 3' UTR of this transcript has not been shown to be responsible for its rapid turnover, it is interesting to consider the possibility that the PRP-BP is involved in the recognition of an AUUUA motif that is involved in regulating the stability of this transcript. This speculation does, however, beg the question of minimal sequence requirements for AUUUA repeats in plants. Although this has not yet been established, recent data in mammalian cells indicate that as few as three copies (Lagnado et al., 1994) or even one copy (Zubiaga et al., 1994) of the sequence UUAUUUAUU are sufficient to destabilize a reporter transcript. Premature stop codons and other translational influences on mRNA stability Premature stop codons function as cis-acting instability determinants in yeast, C. elegans, mammalian cells and plants, and offer compelling evidence for a link between translation and mRNA turnover that is widespread among eukaryotes (Maquat, 1995; Jacobson and Peltz, 1996; Sullivan and Green, 1993; van Hoof and Green, 1997; Marcotte, 1998). This was first recognized in plants when naturally occurring mutant ll alleles of phytohemagglutinin (PHA) and the Kunitz trypsin inhibitor were studied. These alleles contain early stop codons and result in very little mRNA accumulation (Voelker et al., 1986; Jofuku et al., 1989). Run-on transcription experiments showed that the accumulation of the Kunitz trypsin inhibitor was being limited post-transcriptionally (Jofuku et al., 1989). Recently, a study of the effects of premature stop codons in plants has been conducted using the initially isolated PHA allele and other PHA alleles that were constructed in vitro (van Hoof and Green, 1996). These experiments demonstrated that premature stop codons, when present in the first 60% of the PHA coding region, caused rapid mRNA turnover. This effect was measured by generating stably transformed tobacco cell lines with mutant and control PHA alleles and monitoring mRNA decay over an actinomycin D time course. mRNA accumulation studies carried out in transgenic tobacco leaves confirmed that premature stop codons cause mRNA instability in plants as well as in cell culture. Interestingly, a premature stop codon placed 80% of the way through the coding region did not decrease mRNA stability in stably transformed tobacco cells. One explanation for this finding is that mRNA turnover is triggered by a certain length between start and stop codons or between the stop codon and the poly(A) tail. This explanation is somewhat unsatisfying because coding regions and 3' UTRS vary greatly in length and no correlation has been found between transcript length and stability. Alternatively, nonsense mediated mRNA decay in plants may require an additional cis-acting element located downstream of the premature stop codon as has been found in yeast. In yeast and in C. elegans, genetic analysis has led to the elucidation of a pathway that is believed to have evolved to eliminate mRNAs with premature stop codons so that the cell does not produce truncated and potentially harmful polypeptides. This pathway is referred to as the nonsense mediated mRNA decay pathway. In yeast, this pathway requires an early stop codon, a downstream cis-acting element and trans-acting cellular factors (reviewed in Jacobson and Peltz, 1996). The region between 60 and 80% of the PHA transcript may contain a cis-acting element that is required for the function of the nonsense mediated decay pathway. This region does not contain sequences similar to the downstream element that has been determined in yeast (Zhang et al., 1995), however, it may contain a cis-acting element recognized by plant nonsense mediated decay machinery. The internal light regulatory element (iLRE) which comprises the 5' UTR and the first 14 codons of the F ed-I transcript appears to be another cis-acting determinant of mRNA stability that is closely linked to translation (Dickey et al., 1992). The pea Fed-I gene encodes ferredoxin I which is a major photosynthetic electron carrier. Like other nuclear encoded genes involved in photosynthesis, Fed-1 expression increases in response to light. Levels of Fed-1 mRNA are about five fold higher in light-grown versus dark-adapted plants (Elliot et al., 1989). The iLRE was identified because of its ability to confer light responsiveness to reporter genes (Dickey et al., 1992; 1994). The iLRE's location in the transcribed portion of Fed-1 and the observation that nuclear run-on transcription experiments revealed no differences in transcription rates between light-grown and dark adapted plants, provide a strong argument that the iLRE functions post-transcriptionally at the level of mRNA stability (Dickey et al., 1992). Recently, direct evidence that the iLRE modulates Fed-1 gene expression at the level of mRNA stability was obtained using the ToplO promoter system in tobacco plants. This system 13 enabled Thompson and coworkers, to measure half-lives of reporter transcripts with or without the iLRE in transgenic plants in light versus dark conditions (Petracek et al., 1998). This system was also used to analyze the effects of 3-(3,4-dichlorophenyl)-1,1- dimethylurea (DCMU), an inhibitor of photosynthetic electron transport, on iLRE- mediated mRNA degradation. The results of these experiments indicate that the photosynthetic apparatus mediates light signaling that results in degradation of iLRE- containing transcripts (Petracek et al., 1998). Previously it was shown that the iLRE requires an open reading frame in order to modulate Fed-1 mRNA levels in response to light, which indicates that control of F ed-I mRNA decay may be dependent upon translation (Dickey et al., 1994). Further experiments have shown that iLRE-containing messages are loaded on polyribosomes in illuminated plants but are on monosomes in the dark (Dickey et al., 1998). Therefore it seems clear that translation plays a major role in determining the stability of these transcripts. It would appear that active translation protects the transcript from degradation in this case. Sequences involved in regulated mRNA degradation. PvPRPI and Fed-1, discussed above, are examples of genes that are regulated at the level mRNA stability in response to a stimulus. There are many more reports in the literature of message stability that responds to various stimuli including high fluences of blue light (Anderson et al., 1999), cytokinin treatment (Downes and Crowell, 1998), and heat shock (Belanger, et al., 1986). This mode of regulation provides the plant with a 14 means to rapidly tailor gene expression, which may be particularly important for plants because they are firmly rooted in a dynamic environment. Changes in (x-amylase message stability in response to sucrose starvation is a well characterized example. The expression of several ()t-amylase genes is induced when rice suspension cells are cultured without sucrose (Yu et al., 1991). Gene-specific probes were designed that could differentiate between the eight members of this gene family and it was shown by a combination of Northern blot analysis, nuclear run-on experiments, and mRNA t( 1/2) analysis that three a-amylase genes are controlled at the level of mRNA stability (Sheu et al., 1996). The 3’UTR of aAmy3 was studied in greater detail and was shown to be largely responsible for the sugar regulation of this gene (Chan and Yu, l998a, Chan and Yu, l998b). The aAmy3 3’UTR was divided into three sections and each was introduced into the 3’UTR of the same reporter gene. Each construct was introduced stably into rice cells and the effect on message stability was assayed during an actinomycin D time course with or without the addition of sugar. Sections I and III were shown to destabilize reporter transcripts in the presence of sugar (Chan and Yu, l998b). This 3’UTR has a remarkable predicted secondary structure with several proposed stem- loops. In fact, sections I and III are predicted to form stem-loops that have the same sequence (AUAUAU) at the end of the loop (Chan and Yu, l998a). This may be just an intriguing coincidence because the contribution of these sequences to message stability has not been tested. However, there is a precedent for stem-loop structures being important for regulated mRNA stability. The transferrin receptor of mammalian cells has multiple stem-loops in its 3’UTR that serve as a binding site for the iron-responsive element binding protein (IRE-BF). The IRE-BF is thought to protect the message from 15 degradation in iron limiting conditions. These same stem-loops appear to be the recognition sequences for an endonuclease that cleaves the transcript when iron concentration is high and the IRE-BP is no longer present (Mullner and Kuhn, 1988, Casey et al., 1989, Binder et al., 1994). T RANS-ACTING FACTORS INVOLVED IN mRNA STABILITY Trans-acting factors involved in mRNA stability include both the actual ribonucleases involved in the degradative process and regulatory factors which can be supposed to act both positively (to stimulate rapid degradation) and negatively (to protect from degradation). The close relationship between mRNA degradation and translation for at least some mRNAs suggests that some factors may have dual roles in translation and turnover (Jacobson and Peltz, 1996). There are three major approaches used in the identification of new trans-acting factors: (1) cloning of genes encoding proteins with known RNA-binding motifs and subsequent assignment of function; (2) identification of proteins that bind to known cis-determinants of stability by gel mobility shift and/or cross-linking assays; and (3) biochemical purification of proteins with ribonuclease activity. RNA binding proteins The only factors with a likely role in stability to be identified by the first method are the Arabidopsis poly(A)-binding proteins (Belostotsky and Meagher, 1993; 1996; 16 Hilson et al., 1993). Poly(A)-binding proteins (PABPs) from diverse organisms have a characteristic set of four RNA-binding domains (RBDs) and a highly conserved region which can be used as a tentative first identification (Burd and Dreyfuss, 1994). One of the Arabidopsis PABPs (PAB5) has been further identified by its abilities to bind poly(A) with high affinity (Belostotsky and Meagher, 1993) and to complement certain PABP functions in yeast (Belostotsky and Meagher, 1996). The presumed function of the PABPs in mRNA turnover is discussed below. A large number of other plant RNA-binding protein cDNAs have been cloned and sequenced, using primers and probes whose design was based on other related RBD-type sequences (e. g. see Guiltinan and Niu, 1996 for a recent summary). To date, no definite functions have been assigned to any of these proteins. Many are clearly not involved in cytoplasmic mRNA stability because they localize to the chloroplast or nucleus. Since the great majority of yeast and animal RED-containing proteins whose functions are known are involved in nuclear processing and transport events (Burd and Dreyfuss, 1994), seeking RED-containing proteins may not be the most effective way to uncover trans-acting factors involved in turnover. At least ten RNP motifs have been distinguished (Mattaj, 1993), some of which may turn out to be more prevalent in stability-related RNA-binding proteins. The most direct approach to identifying trans-acting factors involved in stability is to seek proteins that bind directly to, or participate in complexes which bind to, known cis-acting stability determinants. For example, the iron response element binding protein (IRE-BP) and a number of AU-rich element binding proteins in mammalian cells have been identified in this way (e. g., Gillis and Malter, 1991; Rouault et al., 1989). Of course, the identification of such proteins is a long step away from understanding how 17 they actually participate in degradation. Although a number of plant stability determinants have been characterized (discussed above), no specific binding proteins have been identified to date. The identification of a stability determinant does not guarantee the future finding of a trans-acing factor, since not all stability determinants necessarily function by serving as binding sites. One plant mRNA-binding protein which may function in controlling stability is a 50-kD protein that binds to a sequence in the 3' UTR of the bean PvPRPI mRNA (Zhang and Mehdy, 1994) The protein binding site, a 27-nt U-rich sequence, has not yet been demonstrated to be a determinant involved in this destabilization event, but the protein binding activity in extracts increases afier elicitor treatment, consistent with this scenario. Interestingly, the binding activity of the protein is strongly influenced by its redox state. Redox regulation of RNA binding has been described for a number of other proteins, some with known functions in the control of stability and/or translation (Hentze et al., 1989; Malter and Hong, 1991; Chu et al., 1994), and could turn out to be a widespread modus operandi for such regulators (Hentze, 1994). RNA-degrading activities Ribonucleases are expected to figure prominently among the protein factors that participate in mRNA decay mechanisms. As discussed below, current models for mRNA decay pathways in plants predict the involvement of both endo- and exoribonucleases, although the relative importance of these two types of activities appears to differ depending on the mRNA in question. Unfortunately, the plant RNases that we know the most about are the least likely to be involved in mRNA degradation because these 18 enzymes enter the secretory pathway and are targeted to the extracellular space or vacuole (see Bariola and Green, 1997, for a review). It should be noted that a role for the vacuole in mRNA decay, particularly in the later stage of the process, can not be ruled out because small fragments of RNA have been shown to exist in the vacuole (Abel et al., 1990). However, it seems most likely that this RNA enters the vacuole by autophagy or some other bulk process rather by mechanisms that target specific mRNAs to the vacuole for degradation. The most common assumption is that mRNA decay mainly occurs in the cytosol. The fact that translation is often coupled to mRNA decay supports this idea. Whether some mRNAs are primarily degraded in the nucleus is, nevertheless, an open question. Clearly RNA-degrading activities must exist in the nucleus to turnover introns and 3' ends of precursor RNAS. Virtually nothing is known about these enzymes in any system but it seems possible that some may be capable of degrading fully processed transcripts prior to or during export. Recently it was shown that RNase 111, an enzyme involved in pre-rRNA processing as well as the decay of some specific mRNAs in Escherichia coli (Court, 1993), also participates in pre-rRNA processing in yeast nucleoli (Elela et al., 1996). Examining plant genes with homology to RNase III (e. g. an Arabidopsis -expressed sequence tag accession number 218464, Hofie et al., 1993) may be of particular interest because most current models for antisense RNA-mediated inhibition (N ellen and Lichtenstein, 1993) and some cosupression models argue for the involvement of a dsRNase (discussed below). Other yeast activities known to be involved in mRNA decay may also have plant homologues such the 5' to 3' exoribonuclease XRNI, which is a major participant in the decay of most yeast mRNAs. Indeed, three XRN-like genes have 19 been cloned from Arabidopsis (James Kastenmayer and Pamela J. Green, unpublished). XRNs have also been cloned from flies and mammals, indicating that these genes have been conserved throughout eukaryotic evolution (Till et al., 1998, Bashkirov et al., 1997; Shobuike et al., 1997). However. one potentially relevant observation is that insertion of a poly(G) tract into an mRNA blocks the progression of XRNI in yeast giving rise to a prominent intermediate that includes sequences 3' of the poly(G) tract (Vreken and Raué, 1992; Decker and Parker, 1993; Muhlrad et al., 1994). A 3' degradation intermediate also results from poly(G) insertion into messages in Chlamydomonas, suggesting the occurrence of a similar exonuclease activity in this organism (Gera and Baker, 1998; Drager et al., 1998). Poly(G) intermediates have been looked for but not seen in both higher plant and mammalian cell systems (Mark A. Johnson and Pamela J. Green, Chapter 2-1 of this dissertation; Goddall, G., personal communication; Maquat, L., personal communication; Shyu, A.-B., personal communication) indicating that the role of XRN—like genes may be less prominent, or different in higher eukaryotes. In any event, additional efforts to characterize RNases that may participate in mRNA decay pathways are certainly warranted. As will be discussed in the next section, the 5’ cap and poly(A) tail play an important role in determining mRNA stability. Therefore, it will be important to characterize enzymes responsible for removing these structures during the process of mRNA degradation. Recently, a poly(A) ribonuclease (PARN) was purified from calf thymus which led to the isolation of a human cDNA encoding this enzyme (Komer and Wahle, 1997; Komer et al., 1998). This is an intriguing activity because the enzyme responsible for removing the poly(A) tail, the first step in the degradation of most yeast 20 mRNAs, has not yet been identified. An Arabidopsis PARN-like gene (AtPARN) has been cloned and will be the subject of Chapter 2-2 of this dissertation. A decapping complex has been described in yeast (Dunckley and Parker, 1999; Beelman et al., 1996). There are also putative Arabidopsis homologues of these genes and the role that these play in plant mRNA tum-over will be an interesting area of future work (Gustavo MacIntosh and Pamela J. Green, unpublished). The role of the cap and poly(A) tail in mRNA stability A variety of primarily in vitro studies over many years have suggested a function for the poly(A) tail and 7-methyl guanosine cap as mRNA-stabilizing structures (reviewed in Baker, 1993; Stevens, 1993). However, the ability to definitively demonstrate a stabilizing function in cells has been limited, in part, by an inability to generate uncapped and unadenylated mRNA species in vivo. New genetic approaches in yeast have finally conclusively demonstrated stabilizing functions for these sequences in this organism. Decapping has been shown to be an early, sometimes rate-limiting, step in the major degradation pathway of many mRNAs in yeast (Hsu and Stevens, 1993; Muhlrad et al., 1994, 1995). Mutations that affect an essential component of the decapping enzyme cause a marked stabilization of both normally unstable and stable mRNAs (Beelman et al., 1996). In yeast, decapping is followed by rapid 5' to 3' exonucleolytic digestion of the rest of the mRNA (by the XRNI nuclease discussed below), and this sequence of events is likely to be the major degradative pathway for yeast mRNAs. Because decapped mRNA is extremely labile in yeast, it could be 21 detected only in mutant cells with defective 5' to 3' exonuclease activity (xrnl cells) (Hsu and Stevens, 1993; Muhlrad et al., 1994). The unavailability of similar mutants in other organisms has so far precluded a demonstration of an analogous decay pathway. The ability of an inserted tract of guanine nucleotides to trap 3' degradation intermediates by impeding exonuclease progress has allowed the demonstration that the substrate for decapping is, for many or most yeast mRNAs, a molecule which has undergone substantial poly(A) shortening (Decker and Parker, 1993). Thus, a longer poly(A) tail protects the mRNA from decapping, almost certainly via its ability to bind at least one poly(A)-binding protein (Caponigro and Parker, 1995). A similar deadenylation- dependent pathway has not yet been demonstrated to occur in other organisms. On the other hand, observation of the poly(A) status of naturally stable degradation intermediates has provided evidence against such a pathway for certain mRNAs (discussed below). Alternative experimental options to directly test for stabilizing functions for the cap and poly(A) tail in other organisms are limited. The best experimental approach to date has been the study of in vitro transcribed mRNAs introduced into cells by transfection, electroporation or injection. Only the evidence pertaining to plant cells will be discussed here. Gallie and colleagues have conducted extensive analyses of the roles of the cap and poly(A) tail in mRNA translation, using synthetic reporter mRNAs electroporated into plant protoplasts (Browning et al., 1998). Comparisons of the chemical and functional stability of capped versus uncapped and polyadenylated versus unadenylated transcripts were included in some of these reports. Chemical stability was measured simply by following the disappearance of the full length mRNA with time by northern blot. A modest stabilizing effect of a 5' cap was observed: a cap increased the 22 half-lives of either unadenylated or polyadenylated luciferase transcripts about 2-fold, from 31 and 44 minutes to 53 and 100 minutes, respectively. While all of these electroporated transcripts can be considered relatively unstable, the rather small stabilizing effect of a cap could be interpreted to mean that 5' to 3' exonuclease digestion does not represent the major pathway for degradation of electroporated mRNAs. Alternatively, it is possible that most capped transcripts are rapidly decapped following delivery, and the small stability differential reflects the decapping rate. In either case, the uncapped state does not dictate immediate degradation by a 5' to 3' exonuclease, since uncapped mRNA is detectable for several hours. Given the extremely rapid degradation that apparently follows decapping of natural mRNAs in yeast, it would be interesting to know whether uncapped transcripts electroporated into yeast spheroplasts are subject to a similar fate, or whether introduced transcripts are degraded by a different or slower mechanism (Russell et al., 1991; Everett and Gallie, 1992). The addition of a 50-nt poly(A) tail to electroporated reporter mRNAs was shown to also confer a modest 1.5 to 3-fold increase in chemical half-life (Gallie et al, 1989; Gallie, 1991). Again, it is not clear how rapidly these reporter RNAS are deadenylated following delivery, or, in this case, what fraction of delivered RNA acquires poly(A)-binding proteins (discussed below). The stabilizing effects of a cap and poly(A) tail together are consistently additive, unlike their contributions to translation, which are strongly synergistic. Another possible explanation for the failure of a cap or poly(A) tail to act as strong stabilizers of electroporated mRNAs is that the major substrates for putative mRNase activities may be messenger RNPs, not naked RNA molecules. The extent to which electroporated mRNAs associate with RNA-binding proteins is not known. It has 23 been determined, however, that only about 2% of electroporated mRNAs are polysomal 1.5 hours after delivery into carrot protoplasts (Gallie et al., 1995). It could be that only this small subset of actively translated mRNAs are substrates for the usual mRNase activities. This possibility was addressed by measuring the functional half-life of electroporated mRNAs, defined as the time required to reach 50% of the final level of protein produced. The functional half-life could be a more accurate measure of the stability of those mRNAs that assemble into an mRNP form. In fact, when functional half-lives were compared, the cap and poly(A) tail each did confer a somewhat stronger stabilizing effect than their effects on chemical half-lives (3 to 5-fold versus 1.5 to 3-fold). However, the relatively modest protection suggests that the electroporated RNAS are not substrates for potent deadenylation-dependent or decapping-dependent pathways in these cells. It is not yet clear whether this finding reflects the absence of a dominant role for the cap and poly(A) tail in plant mRNA stability or is the result of our inability to evaluate the effect of these elements on the decay of transcripts synthesized in vivo. Poly(A)-binding proteins. Poly(A)-binding proteins (PABPs) are ubiquitous and multifunctional RNA-binding proteins that have been implicated in several aspects of mRNA metabolism and translation (reviewed in Jacobson and Peltz, 1996; Baker, 1997). The yeast poly(A)-binding protein (pablp) has recently been assigned an unexpected stabilizing function: it apparently prevents decapping of polyadenylated mRNAs, presumably in cis. The evidence supporting this role is that decapping is normally triggered by shortening of 24 a poly(A) to a length too short to bind pablp (Decker and Parker, 1993; Muhlrad et al., 1994) and in pabIA cells, decapping occurs without prior poly(A) shortening (Caponigro and Parker, 1995). There are multiple PABPs in plants; in Arabidopsis, at least four different PABP genes are expressed in a tissue and organ-specific manner (Belostotsky and Meagher, 1993; 1996; Hilson et al, 1993). Whether these genes encode functionally distinct or functionally redundant proteins is not yet known. To begin to address that important question, Belostotsky and Meagher (1996) have expressed one of the Arabidopsis PABP genes (PAB5) in yeast strains depleted of endogenous PABlp to determine what functions could be complemented by the plant protein. Remarkably, although the two proteins are only 44% identical at the amino acid level, the PABS protein was able to rescue inviability and at least partially restore two functions: normal poly(A) shortening and translation initiation. However, expression of the PARS gene did not restore deadenylation-dependent decappin g of yeast mRNAs. Had it done so, this would have strongly suggested the occurrence of a similar PABP function, and thus a similar mRNA degradation pathway, in Arabidopsis. However, the inability of PABS to complement this function in yeast does not imply that it lacks this function in Arabidopsis. It will be important to see how other Arabidopsis PABPs behave in this regard. PATHWAYS OF mRNA DEGRADATION Pathways by which mRNA molecules are degraded have been proposed for both mammalian and yeast mRNAs. Mammalian mRNAs have generally been proposed to be 25 degraded in a 3' to 5' direction. Degradation of the body of message is thought to be preceded by either stepwise shortening of the poly(A) tail (e.g., MYC mRNA), or by endonucleolytic removal of the poly(A) tail (e.g., transferrin receptor mRNA) (Ross, 1996). Whether poly(A) removal is an essential rate-limiting step in the degradation of mammalian mRNAs is not known, although the biphasic degradation kinetics of c-fos and other ARE-containing mRNAs is compatible with that idea (Chen and Shyu, 1994). In yeast, 3 major pathway of degradation appears to be the deadenylation-dependent decapping pathway (Decker and Parker, 1993; Decker and Parker, 1994). After decapping, degradation of some mRNAs (e. g., PGKl mRNA) appears to proceed principally in the 5' to 3' direction. However, a more limited 3' to 5' degradation of the FOX] mRNA has also been observed following deadenylation (Muhlrad and Parker, 1994). A deadenylation-independent decapping pathway also occurs (Muhlrad and Parker, 1994). In this pathway, decapping and 5' to 3' degradation occur in the absence of ’ significant poly(A) shortening. mRNAs carrying premature nonsense codons, and perhaps other mRNAs, are targeted for rapid degradation via this route. It would be premature to conclude that yeast and mammalian cells degrade mRNAs by fundamentally distinct pathways. An equally likely possibility is that eukaryotic cells have multiple mRNA degradation pathways, with the pathway used being mRNA species specific. In some species, one or more pathways may predominate and the predominant pathways may differ among organisms. 26 Identification and analysis of in vivo produced mRNA degradation intermediates would greatly facilitate the elucidation of mRNA degradation pathways. Unfortunately, such intermediates are often not readily apparent in RNA samples from mammalian, yeast, or plant cells. However, two plant mRNAs (encoded by the soybean SRS4 and oat PH YA genes), for which putative in vivo mRNA degradation products have been reported, have provided some insight into the pathways of mRNA degradation in plant cells. It should be noted that mRNA degradation intermediates may be more prevalent than is often assumed. It is common to observe hybridizing RNA fragments smaller than the full-length mRNA in RNA blot analysis. Trimming of RNA blots prior to publication to focus attention on the full-length band or to save journal space may obscure the true frequency with which such RNA fragments occur. The problem, of course, is to determine whether the RNA fragments are degradation intermediates produced in vivo, or the result of degradation in vitro during isolation of the RNA samples. Control experiments useful to address this question have been described (Seeley et al., 1992; Thompson et al., 1992; Tanzer and Meagher, 1994). A series of discrete fragments of the soybean ribulose-l ,5-bisphosphate carboxylase small subunit (SRS4) mRNA are detected on RNA gel blots (Thompson et al., 1992). Similar mRNA fragments are present in transgenic petunia transformed with the soybean SRS4 gene under the control of the 35S promoter. These fragments appear to be bona fide in vivo degradation intermediates. This interpretation is supported by the observation that tracer RNAS are not degraded in homogenized samples during RNA isolation (Thompson et al., 1992). In addition, degradation of in vitro synthesized SRS4 RNA in cell-free systems generates the same RNA fragments observed in vivo (Tanzer 27 and Meagher, 1994; 1995). The discrete nature of the SRS4 mRNA fragments suggests the activity of an endonuclease. Sl nuclease and primer extension mapping of the SRS4 RNA fragments indicate that the fragments were indeed derived from endonucleolytic cleavage of the full-length SRS4 mRNA (Tanzer and Meagher, 1995). The endonucleolytic cleavage appears to be independent of both deadenylation and decapping. Tanzer and Meagher (1995) have proposed that the endonucleolytic cleavages are catalyzed by a stochastic endonuclease, cleaving at various sites within the SRS4 mRNA, followed by 3' to 5' or 5' to 3' exonucleolytic cleavage of the resulting fragments. The proximal (5') and distal (3') fragments would be degraded while possessing either a 5' cap or poly(A) tail, respectively. As such, this model is distinct from those proposed for mammalian and yeast mRNAs. Although, for example, a 5' capped fragment of an mRNA could be thought of as being analogous to a deadenylated mRNA, similar to deadenylation of mammalian transferrin receptor mRNA prior to degradation (Ross, 1996). RNA fragments have also been observed in RNA blot analysis of oat phytochrome A (PHYA) mRNA. These fragments also appear to be in vivo degradation products. This interpretation is based on several lines of evidence including the observation that the PH YA fragments are present in RNA samples isolated by various distinct procedures (Seeley et al., 1992). Other endogenous or exogenously-added tracer RNAS remain intact under the same isolation conditions. Unlike the discrete fiagments observed for soybean SRS4 mRNA, the PHYA fragments form a continuous distribution ranging from full-length (approximately 4 kb) down to about 200 bp (Seeley et al, 1992; Higgs and Colbert, 1994; Higgs et al, 1995). Analysis of these fragments, using distinct 28 probes corresponding to various regions of the PHYA mRNA molecule, has led to the proposal that these fragments are generated by exonucleases (Hi ggs and Colbert, 1994). RNase H mapping fails to reveal endonucleolytic cleavages near the 5' or 3' ends of the mRNA (Higgs and Colbert, 1994). In addition, analysis of the polyadenylation status of PHYA mRNA suggests that about 25% of the apparently full-length PH YA mRNA lacks a poly(A) tail. PH YA RNA fragments are present in both the polyadenylated RNA fraction and in the deadenylated RNA fraction. These observations have been incorporated into a model (Hi ggs and Colbert, 1994) proposing that about 75% of the PH YA mRNA population is degraded by a 5' to 3' exonuclease, with the poly(A) tail still attached to the RNA fragment undergoing degradation. The deadenylated portion of the PHYA mRNA population is proposed to be degraded by exonucleases acting at both the 3' and 5' ends of the molecule. If this model is correct, the bulk of the PHYA mRNA would be degraded by a pathway similar to the deadenylation-independent decapping pathway of yeast (Muhlrad and Parker, 1994). The degradation pathway of mRNA molecules is a complex process about which we have rather limited knowledge. It does seem clear that multiple pathways for mRNA degradation may be present within eukaryotic cells. Even individual mRNA species may be degraded by more than one pathway (Muhlrad and Parker, 1994; Higgs and Colbert, 1994). There are numerous unresolved questions regarding the pathway(s) of degradation of nuclear-encoded mRNAs within plant cells. Perhaps the most pressing basic question is the location of selective mRNA degradation. Two lines of evidence suggest that degradation may occur on polysomes. First, putative in vivo degradation products of both SRS4 mRNA and PHYA mRNA are present in polysomal RNA fractions 29 (Thompson et al., 1992; Byrne et al., 1993; Hi ggs and Colbert, 1994). Second, cell-free mRNA degradation systems, based on isolated polysomes, have been described for both plant (Byrne et al., 1993; Tanzer and Meagher, 1994) and animal (Brewer and Ross, 1990; Ross, 1995) cells. These cell-free systems appear to accurately reflect at least some aspects of in vivo mRNA degradation. In addition, polysome-associated ribonuclease activities have been reported (Green, 1994). However, other compartments within plant cells are also known to possess ribonuclease activity. As discussed above, plant vacuoles have high levels of ribonuclease activity (Abel and Glund, 1987; Green, 1994). The possibility that some nuclear-encoded mRNAs are targeted for decay by regulated transport into the vacuole is difficult to rule out. There is, however, only limited evidence for the movement of RNA molecules across organellar membranes (Chang and Clayton, 1987; Oda et al., 1992). Another area that warrants further development is the use of in vitro systems to dissect mRNA decay pathways. Already, the two systems described above have been shown to produce mRNA decay intermediates that mimic those observed in vivo. An important next step will be to test whether cis-acting instability sequences can be recognized in these and additional in vitro systems. Further characterization of the process by which mRNA degradation occurs may also be possible using in vitro systems. For example, recent evidence suggests that ATP is not required for the decay of PHYA mRNA in vitro (Byrne and Colbert, unpublished). Clearly, additional exploration of current in vitro systems and the development of new ones will provide further mechanistic insights. 3O POST-TRANSCRIPTIONAL GENE SILENCING Cosuppression is a form of gene silencing that in many cases, appears to involve degradation of the target mRNA. It was first identified during attempts by plant molecular biologists to overexpress various cloned plant genes. These experiments usually involved the generation of transgenic plants expressing the gene of interest driven by the strong, constitutive, 358 promoter. Much to the surprise of these early investigators, in a few of these transgenic lines, rather than over expression of the gene of interest, both the transgene and the corresponding endogenous gene were coordinately silenced (N apoli et al., 1990; van der Krol, 1990, Smith et al., 1990). This phenomenon is not limited to transgenes with endogenous counterparts because it appears that multiple copies of a transgene can suppress one another through similar mechanisms. In addition, homology dependent virus resistance, a technique which employs portions of a viral genome to generate transgenic plants that are resistant to the virus, may also function in the same manner (Lindbo et al., 1993). Many cases of cosuppression have been explained by transcriptional repression that appears to be similar to other epi genetic phenomena described in fungi, plants, and animals (Matzke and Matzke, 1996). In other cases, however, transcription rates of transgenes and endogenous genes are normal, yet very little mRNA accumulates in the cytoplasm. In these cases it has been assumed that expression is suppressed either by blocking export of mature mRNA to the cytoplasm and eventual degradation in the nucleus or by cytoplasmic mRNA degradation. Studies using cytoplasmically replicating RNA viruses as targets for homology dependent virus resistance support the idea that 31 gene silencing is occurring in the cytoplasm via rapid degradation of the suppressed mRNA (Lindbo et al., 1993). Further support for this idea comes from the finding that nuclear mRNA abundance of a silenced transgene is unaffected (De Carvalho Niebel et al,l995) Post-transcriptional gene silencing (PTGS) of this type was once thought to be limited to plants, but is now considered to be a mechanism that many organisms employ to ward off invasive nucleic acid. The phenomena known as RNA interference and quelling that have been studied in C aenorhabditis elegans and Neurospora crassa, respectively, are probably examples of PTGS (Fire et al., 1998; Romano and Mancino, 1992). Most models that describe the initiation, maintenance, and spread of PTGS in plants and other organisms call for the formation of a double-stranded RNA (dsRNA) molecule. This is an attractive idea because it accounts for the sequence-specificity that is observed in PTGS. This model implies the existence of an RNA-dependent RNA polymerase (RDRP) that would be responsible for the synthesis of RNA complementary to the target RNA and a dsRNA-specific RNase that would degrade the target RNA. RDRPs have been cloned from plants and recently an RDRP was identified in a screen for N. crassa mutants defective in quelling (Schiebel et al., 1998; Cogoni and Macino, 1999). RDRPs may generate small RNAS from specific target messages that would be able to move from one cell to another within the organism to silence the target gene. Recently evidence in support of this hypothesis has come by displaying small RNAS that are complementary to silenced viral targets in plants (Hamilton and Baulcombe, 1999). These studies represent dramatic progress in our understanding of how the PTGS signal is generated. However, how dsRNA is degraded is still an open question. Insights into 32 these questions and other lingering questions about how PTGS works may come by studying the mechanisms of rapid mRNA turnover. It is possible that the activities involved in the final steps of both of these processes, namely rapid RNA degradation, will be similar. FUTURE DIRECTIONS The development of reliable methods to measure mRNA half-lives in plants has resulted in the identification of cis-acting sequences that target transcripts for rapid turnover and in the identification of a trans-acting factor potentially involved in the recognition of one of these elements. It is also now clear that alterations in gene expression in response to extra and intra-cellular stimuli is in some cases controlled at the level of mRNA stability in plants. In addition, understanding of the pathways involved in mRNA turnover in plants has improved recently and points to possible differences between plants and animals, and yeast. The future of this work must include increased efforts aimed at the elucidation of mRNA degradation pathways and at identifying cellular factors that carry out the steps in these pathways. By characterizing the decay of a greater number of mRNAs, it will be possible to determine whether sets of mRNAs are funneled into one or more distinct pathways and to determine whether there are any common themes conserved among plants, fungi and animals. Molecular analysis of cellular factors involved in mRNA degradation will improve efforts to characterize mRNA turnover pathways and will provide insights into intriguing problems in plant molecular biology such as post-transcriptional gene silencing. 33 The progress outlined in this introduction represents a foundation upon which a rich future can be built. One exciting avenue of research will be the application of molecular genetics in Arabidopsis thaliana in order to gain insight into the mechanisms of mRNA turnover in plants and to identify cellular factors involved in this process. Knowledge of cis-acting instability sequences should allow the isolation of mutants in specific mRN A decay pathways. This may be particularly interesting in the case of AU- rich elements because genetic approaches have been lacking in mammalian systems in which the pioneering work on such elements has been done. Although most studies have thus far dealt with unstable mRNAs, it is now possible to measure long mRNA half-lives through the use of regulated promoters rather than prolonged incubation with transcriptional inhibitors. This coupled with the finding of active stabilization mechanisms in other systems opens the door for elucidating how the stabilities of long- lived plant mRNAs are controlled. Genetic studies combined with biochemical approaches aimed at purifying mRNA binding and degrading activities should lead to an era of great expansion in our knowledge of fundamental mechanisms of mRNA degradation in plants. THE SCOPE OF THIS THESIS The members of our laboratory we have begun to think of mRNA degradation as a hierarchy consisting of three tiers (Gutierez et al., 1999). At the bottom of this scheme are the basic mRNA degrading enzymes which dismantle and digest the message. Above the basal mRNA degradation machinery is sequence-specific mRNA degradation which 34 determines the degradation rate of particular mRNAs. In this scheme, instability determinants can be thought of as flags that target transcripts to the mRNA decay machinery. Above sequence-specific decay is regulated mRNA degradation which is more complex because the function of an instability sequence is determined by extra and intracellular stimuli. In my thesis, I have addressed the first two levels of this hierarchy and have therefore divided this dissertation into two sections. In Section One, I report my work aimed at understanding the mechanisms responsible for sequence-specific mRNA degradation in plants. In Chapters 1-1 and 1-2 an approach is presented that resulted in the isolation of two mutants that are defective in sequence-specific mRNA degradation mediated by the DST element. In Chapter 1-3, I report on a series of experiments that were designed to identify the critical instability sequences within the 3’UTR of SA UR- AC1, an Arabidopsis transcript that is likely to be regulated by DST-mediated degradation. Section Two is devoted to my work on understanding the basic machinery involved in degrading mRNA in plants. In a set of experiments described in Chapter 2-1, a stable secondary structure was introduced into plant reporter transcripts in an attempt to analyze intermediates in the mRNA decay process. In Chapter 2-2, preliminary work on the characterization of a potential poly(A) ribonuclease of Arabidopsis is described. The studies reported here represented steps along the way to answering some of the important questions that have been laid out in this introduction. The major contribution of this thesis is likely to be the dst mutants. These are the first mutants of 35 their kind to be isolated in any eukaryote and may be very useful in understanding the mechanisms of sequence-specific mRNA degradation. 36 REFERENCES Abel S, Blume B, Glund K (1990) Evidence for RNA-oligonucleotides in plant vacuoles isolated from cultured tomato cells. Plant Physiol. 94: 1 163-1171 Abel S, Glund K (1987) Ribonuclease in plant vacuoles: purification and molecular properties of the enzyme from cultured tomato cells. Planta 172: 71-78 Abler ML, Green PJ (1996) Control of mRNA stability in higher plants. Plant Mol. Biol. 32: 63-78 Adang MJ, Brody MS, Cardineau G, Eagan N, Roush RT, Shewrnaker CK, Jones A, Oakes JV, McBride KE (1993) The reconstruction and expression of a Bacillus thuringiensis cryIIIA gene in protoplasts and potato plants. Plant Mol. Biol. 21: 1131-1145 Anderson,M.B., Folta,K., Warpeha,K.M., Gibbons,J., Gao,J., and Kaufman,L.S. (1999). Blue light-directed destabilization of the pea Lhcbl*4 transcript depends on sequences within the 5' untranslated region. Plant Cell 11, 1579-1589. Baker EJ (1993) Control of poly(A) length. In J Belasco, G Brawerrnan, eds, Control of Messenger RNA Stability, Academic Press, inc. San Diego, pp 367-415 Baker EJ (1997) mRNA polyadenylation: functional implications. In D Morris, JB Harford, eds, mRNA Metabolism and Post-Transcriptional Gene Regulation. Wiley-Liss, Inc. New York, pp 85-104 Bariola PA and Green P] (1997) Plant Ribonucleases. In JF Riordan, G D'Alessio, eds, Ribonucleases: Structure and Function, Academic Press, Orlando, FL, pp 163-190 Bashkirov,V.I., Scherthan,H., Solinger,J.A., Buerstedde,J.-M., and Heyer,W.-D. (1997). A mouse cytoplasmic exoribonuclease (mXRNl p) with preference for G4 tetraplex substrates. Journal of Cell Biology 136 , 761-733. Baulcombe DC, English JJ (1996) Ectopic pairing of homologous DNA and post-transcriptional gene silencing in transgenic plants. Curr. Opin. Biotechnol. 7: 173-180 Beelman CA, Stevens A, Caponigro G, LaGrandeur TE, Hatfield L, Fortner DM, Parker R (1996) An essential component of the decapping enzyme required for normal rates of mRNA turnover. Nature 382: 642-646 Belanger,F.C., Brodl,M.R., and Ho,T.D. (1986). Heat shock causes destabilization of specific mRNAs and destruction of endoplasmic reticulum in barley aleurone cells. Proc. Natl. Acad. Sci. USA 83, 1354-1358. 37 Belostotsky DA, Meagher RB (1993) Differential organ-specific expression of three poly(A)-binding protein genes from Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 90: 6686-6690 Belostotsky DA, Meagher RB (1996) A pollen-, ovule-, and early embryo-specific poly(A) binding protein from Arabidopsis complements essential functions in yeast. Plant Cell 8: 1261-1275 Binder,R., Horowitz,J.A., Basilion,J.P., Koeller,D.M., Klausner,R.D., and Harford,J.B. (1994). Evidence that the pathway of transferrin receptor mRNA degradation involves an endonucleolytic cleavage within the 3' UTR and does not involve poly(A) tail shortening. EMBO J. 13, 1969-1980. Braam J, Davis RW (1990) Rain-, wind-, and touch-induced expression of calmodulin and calmodulin-related genes in Arabidopsis. Cell 60: 357-364 Brewer G, Ross J (1990) Messenger RNA turnover in cell-free extracts. Methods Enzymol. 181: 202-209 Burd CG, Dreyfuss G (1994) Conserved structures and diversity of functions of RNA-binding proteins. Science 265: 615-621 Byrne DH, Seeley KA, Colbert JT (1993) Half-lives of oat mRNAs in vivo and in a polysome-based in-vitro system. Planta 189: 249-256 Caponigro G, Parker R (1995) Multiple functions for poly(A)-binding protein in mRNA decapping and deadenylation in yeast. Genes Dev. 9: 2421-2432 Caponigro G, Parker R (1996) Mechanisms and control of mRNA turnover in Saccharomyces cerevisiae. Microbiol. Rev. 60: 233-249 Caput D, Beutler B, Hartog K, Thayer R, Brown-Shimer S, Cerami A (1986) Identification of a common nucleotide sequence in the 3'-untranslated region of mRNA molecules specifying inflammatory mediators. Proc. Natl. Acad. Sci. USA 83: 1670-1674 Casey J L, Koeller DM, Ramin VC, Klausner RD, Harford JB (1989) Iron regulation of transferrin receptor mRNA levels requires iron-responsive elements and a rapid turnover determinant in the 3' untranslated region of the mRNA. EMBO J. 8: 3693-3699 Chan,M.T. and Yu,S.M. (l998a). The 3' untranslated region of a rice alpha-amylase gene mediates sugar— dependent abundance of mRNA. Plant J. 15, 685-695. Chan,M.T. and Yu,S.M. (1998b). The 3' untranslated region of a rice alpha-amylase gene functions as a sugar-dependent mRNA stability determinant. Proc. Natl. Acad. Sci. U. S. A 95, 6543-6547. 38 Chang DD, Clayton DA (1987) A mammalian mitochondrial RNA processing activity contains nucleus-encoded RNA. Science 235: 1178-1 184 Chen C-YA, Shyu A-B (1994) Selective degradation of early-response-gene mRNAs: Functional analyses of sequence features of the AU-rich elements. Mol. Cell. Biol. 14: 8471-8482 Chen CYA, Shyu AB (1995) AU-rich elements: Characterization and importance in mRNA degradation. Trends Biochem. Sci. 20: 465-470 Chu S, Archer RH, Zengel J M, Lindahl L (1994) The RNA of RNase MRP is required for normal processing of ribosomal RNA. Proc. Natl. Acad. Sci. USA 91: 659-663 Cogoni,C. and Macino,G. (1999). Gene silencing in Neurospora crassa requires a protein homologous to RNA-dependent RNA polymerase. Nature 399, 166-169. Court D (1993) RNA processing and degradation by RNaseIII. In G Brawerman, JG Belasco, eds, Control of mRNA stability. Academic Press, New York, pp 70-116 De Carvalho Niebel F, Frendo P, Van Montagu M, Comelissen M (1995) Post-transcriptional cosuppression of 3-1 ,3-glucanase genes does not affect accumulation of transgene nuclear mRNA. Plant Cell 7: 347-358 De Carvalho F, Gheysen G, Kushnir S, Van Montagu M, Inzé D, Castresana C (1992) Suppression of 3- l ,3-glucanase transgene expression in homozygous plants. EMBO J. 1 1: 2595-2602 Decker CJ, Parker R (1993) A turnover pathway for both stable and unstable mRNAs in yeast: evidence for a requirement for deadenylation. Genes Dev. 7: 1632-1643 Decker CJ, Parker R (1994) Mechanisms of mRNA degradation in eukaryotes. Trends Biochem. Sci. 19: 336-340 Dehio C, Schell J (1994) Identification of plant genetic loci involved in a posttranscriptional mechanism for meiotically reversible transgene silencing. Proc. Natl. Acad. Sci. USA 91: 5538-5542 Dickey LF, Gallo-Meagher M, Thompson WF (1992) Light regulatory sequences are located within the 5' portion of the Fed-1 message sequence. EMBO J. 1 1: 2311-2317 Dickey LF, Nguyen T-T, Allen GC, Thompson WF (1994) Light modulation of ferredoxin mRNA abundance requires an open reading frame. Plant Cell 6: 1171-1176 39 Dickey,L.F., Petracek,M.E., Nguyen,T.T., Hansen,E.R., and Thompson,W.F. (1998). Light regulation of F ed—l mRNA requires an element in the 5' untranslated region and correlates with differential polyribosome association. Plant Cell 10, 475-484. Diehn SH, De Rocher EJ, Green PJ (1996) Problems that can limit the expression of foreign genes in plants: lessons to be learned from B. t. toxin genes. In JK Setlow, ed, Genetic Engineering: Principles and Methods, Vol 18. Plenum Press, New York, pp 83-99 Downes,B.P. and Crowell,D.N. (1998). Cytokinin regulates the expression of a soybean B—expansin gene by a post-transcriptional mechanism. Plant Mol. Biol. 37, 437-444. Drager,R.G., Girard-Bascou,J., Choquet,Y., Kindle,K.L., and Stern,D.B. (1998). In vivo evidence for 5'-->3' exoribonuclease degradation of an unstable chloroplast mRNA. Plant J. 13, 85-96. Dunckley,T. and Parker,R. (1999). The DCP2 protein is required for mRNA decapping in saccharomyces cerevisiae and contains a functional MutT motif [In Process Citation]. EMBO J. 18, 5411-5422. Elela SA, Igel H, Ares M (1996) RNaseIII cleaves eukaryotic preribosomal RNA at a U3 snoRNP-dependent site. Cell 85: 1 15-124 Elliott RC, Dickey LF, White MJ, Thompson WF (1989) cis-acting elements for light regulation of pea ferredoxin I gene expression are located within transcribed sequences. Plant Cell 1: 691-698 English JJ, Mueller E, Baulcombe DC (1996) Suppression of Virus Accumulation in Transgenic Plants Exhibiting Silencing of Nuclear Genes. Plant Cell 8: 179-188 Everett JG, Gallie DR (1992) RNA delivery in Saccharomyces cerevisiae using electroporation. Yeast 8: 1 007-1014 F lavell RB (1994) Inactivation of gene expression in plants as a consequence of specific sequence duplication. Proc. Natl. Acad. Sci. USA 91: 3490-3496 Franco AR, Gee MA, Guilfoyle TJ (1990) Induction and superinduction of auxin-responsive mRNAs with auxin and protein synthesis inhibitors. J. Biol. Chem. 265: 15845-15849 Gallie DR, Lucas WJ, Walbot V (1989) Visualizing mRNA expression in plant protoplasts: factors influencing efficient mRNA uptake and translation. Plant Cell 1: 301-311 Gallic DR (1991) The cap and poly(A) tail function synergistically to regulate mRNA translational efficiency. Genes Dev. 5: 2108-2116 40 Gallic DR (1993) Posttranscriptional regulation of gene expression in plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 44: 77-105 Gallie DR, Caldwell C, Pitto L (1995) Heat shock disrupts cap and poly(A) tail function during translation and increases mRNA stability of introduced reporter mRNA. Plant Physiol. 108: 1703-1713 Gera,J.F. and Baker,E.J. (1998). Deadenylation-dependent and -independent decay pathways for al-tubulin mRNA in Chlamydomonas rein/zardtii. Mol. Cell. Biol. 18, 1498-1505. Gil P, Liu Y, Orbovic V, Verkamp E, Poff KL, Green PJ (1994) Characterization of the auxin-inducible SA UR-A C1 gene for use as a molecular genetic tool in Arabidopsis. Plant Physiol. 104: 777-784 Gil P, Green PJ (1996) Multiple regions of the Arabidopsis SA UR-A CI gene control trascript abundance: the 3' untranslated region firnctions as an mRNA instability determinant. EMBO J. 15: 1678-1686 Gillis P, Malter J S (1991) The adenosine—uridine binding factor recognizes the AU-rich elements of cytokine, lymphokine, and oncogene mRNAs. J. Biol. Chem. 266: 3172-3177 Goodwin J, Chapman K, Swaney S, Parks TD, Wemsman EA, Dougherty WG (1996) Genetic and biochemical dissection of transgenic RNA-mediated virus resistance. Plant Cell 8: 95-105 Green PJ (1993) Control of mRNA stability in higher plants. Plant Physiol. 102: 1065-1070 Green PJ (1994) The ribonucleases of higher plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 45: 421-445 Guiltinan MJ, Niu XP (1996) cDNA encoding a wheat (T riticum aestivum cv Chinese Spring) glycine-rich RNA-binding protein. Plant Mol. Biol. 30: 1301-1306 Gutiérrez,R.A., MacIntosh,G.C., and Green,P.J. (1999). Current perspectives on mRNA stability in plants: multiple levels and mechanisms of control. Trends Plant Sci. 4, 429- 438. Hamilton,A.J. and Baulcombe,D.C. (1999). A species of small antisense RNA in posttranscriptional gene silencing in plants. Science 286, 950-952. Hentze MW, Rouault TA, Harford J B, Klausner RD (1989) Oxidation-reduction and the molecular mechanism of a regulatory RNA-protein interaction. Science 244: 357-359 41 Hentze MW (1994) Enzymes as RNA-binding proteins: a role for (di)nucleotide-binding domains? Trends Biochem. Sci. 18: 101-103 Herrick D, Parker R, Jacobson A (1990) Identification and comparison of stable and unstable mRNAs in Saccharomyces cerevisiae. Mol. Cell. Biol. 10: 2269-2284 Higgs DC, Barnes LJ, Colbert JT (1995) Abundance and half-life of the distinct oat phytochrome A3 and A4 mRNAs. Plant Mol. Biol. 29: 367-377 Higgs DC, Colbert JT (1994) Oat phytochrome A mRNA degradation appears to occur via two distinct pathways. Plant Cell 6: 1007-1019 Hilson P, Carroll KL, Masson PH (1993) Molecular characterization of PABZ, a member of the multigene family coding for poly(A)-binding proteins in Arabidopsis thaliana. Plant Physiol. 103: 525-533 Hobbs SLA, Kpodar P, DeLong CMO (1990) The effect of T-DNA copy number, position and methylation on reporter gene expression in tobacco transformants. Plant Mol. Biol. 15: 851-864 Hobbs SLA, Warkentin TD, DeLong CMO (1993) Transgene copy number can be positively or negatively associated with transgene expression. Plant Mol. Biol. 21: 17-26 Hofte H, Desprez T, Amselem J, Chiapello H, Caboche M, Moisan A, Jourjon M-F, Charpenteau J-L, Berthomieu P, Guerrier D, Giraudat J, Quigley F, Thomas F, Yu D-Y, Mache R, Raynal M, Cooke R, Grellet F, Delseny M, Parmentier Y, de Marcillac G, Gigot C, Fleck J, Philipps G, Axelos M, Bardet C, Tremousaygue D, Lescure B (1993) An inventory of 1 152 expressed sequence tags obtained by partial sequencing of cDNAs from Arabidopsis thaliana. Plant J. 4: 1051-1061 Hsu CL, Stevens A (1993) Yeast cells lacking 5'-->3' exoribonuclease 1 contain mRNA species that are poly(A) deficient and partially lack the 5' cap structure. Mol. Cell. Biol. 13: 4826-4835 lngelbrecht 1, Van Houdt H, Van Montagu M, Depicker A (1994) Posttranscriptional silencing of reporter transgenes in tobacco correlates with DNA methylation. Proc. Natl. Acad. Sci. USA 91: 10502-10506 Jacobson A, Peltz SW (1996a) Interrelationships of the pathways of mRNA decay and translation in eukaryotic cells. Annu. Rev. Biochem. 65: Jacobson A and Peltz SW (1996b) Interrelationships of the pathways of mRNA decay and translation in eukaryotic cells. In C Richardson, J Abelson, and C Raetz, eds, Annual Review of Biochemistry, Ed 65 pp 693-739 42 Jofuku KD, Schipper RD, Goldberg RB (1989) A frameshifi mutation prevents Kunitz trypsin inhibitor mRNA accumulation in soybean embryos. Plant Cell 1: 427-435 Kakutani T, J eddeloh JA, Richards EJ (1995) Characterization of an Arabidopsis thaliana DNA hypomethylation mutant. Nucleic Acids Res. 23: 130-137 Kamalay JC, Goldberg RB (1984) Organ-specific nuclear RNAS in tobacco. Proc. Natl. Acad. Sci. USA 81: 2801-2805 Koeller DM, Horowitz J A, Casey J L, Klausner RD, Harford JB (1991) Translation and the stability of mRNAs encoding the transferrin receptor and c-fos. Proc. Natl. Acad. Sci. USA 88: 7778-7782 K6rner,C.G. and Wahle,E. (1997). Poly(A) tail shortening by a mammalian poly(A)- specific 3'-exoribonuclease. J. Biol. Chem. 272, 10448-10456. K6rner,C.G., Wormington,M., Muckenthaler,M., Schneider,S., Dehlin,E., and Wahle,E. (1998). The deadenylating nuclease (DAN) is involved in poly(A) tail removal during the meiotic maturation of Xenopus oocytes. EMBO J. 17, 5427-5437. Koziel MG, Beland GL, Bowman C, Carozzi NB, Crenshaw R, Crossland L, Dawson J, Desai N, Hill M, Kadwell S, Launis K, Lewis K, Maddox D, McPherson K, Meghji MR, Merlin E, Rhodes R, Warren GW, Wright M, Evola SV (1993) Field performance of elite transgenic maize plants expressing an insecticidal protein derived from Bacillus thuringiensis. Bio/Technology 11: 194-200 Lagnado CA, Brown CY, Goodall GJ (1994) AUUUA is not sufficient to promote poly(A) shortening and degradation of an mRNA: The functional sequence within AU-rich elements may be UUAUUUA(U/A)(U/A). Mol. Cell. Biol. 14: 7984-7995 Lindbo JA, Silva-Rosales L, Proebsting WM, Dougherty WG (1993) Induction of a Highly Specific Antiviral State in Transgenic Plants: Implications for Regulation of Gene Expression and Virus Resistance. Plant Cell 5: 1749-1759 Malter J S, Hong Y (1991) A redox switch and phosphorylation are involved in the post-transcriptional up-regulation of the adenosine—uridine binding factors by phorbol ester and ionophore. J. Biol. Chem. 266: 3167-3171 Maquat LE (1995) When cells stop making sense: Effects of nonsense codons on RNA metabolism in vertebrate cells. RNA 1: 453-465 Mattaj IW (1993) RNA recognition: A family matter? Cell 73: 837-840 Matzke MA, Matzke AJM, Eggleston WE (1996) Paramutation and transgene silencing: a common response to invasive DNA? Trends Plant Sci. 1: 382-388 43 McClure BA, Hagen G, Brown CS, Gee MA, Guilfoyle TJ (1989) Transcription, organization, and sequence of an auxin-regulated gene cluster in soybean. Plant Cell 1: 229-239 McClure BA, Guilfoyle T (1987) Characterization of a class of small auxin-inducible soybean polyadenlyated RNAs. Plant Mol. Biol. 9: 61 1-623 McClure BA, Guilfoyle TJ (1989) Rapid redistribution of auxin-regulated RNAS during gravitropism. Science 243: 91-93 Meins F, Jr. and Kunz C (1995) Gene Silencing in Transgenic Plants: A Heuristic Autoregulation Model. In P Meyer, ed, Gene Silencing in Higher Plants and Related Phenomena in Other Eukaryotes, Vol 197. Springer-Verlag, Berlin, pp 105-120 Muhlrad D, Decker CJ, Parker R (1994) Deadenylation of the unstable mRNA encoded by the yeast MFA2 gene leads to decapping followed by 5'-->3' digestion of the transcript. Genes Dev. 8: 855-866 Muhlrad D, Decker CJ, Parker R (1995) Turnover mechanisms of the stable yeast PGKl mRNA. Mol. Cell. Biol. 15: 2145-2156 Muller,E.W. and Kuhn,L.C. ( 1988). A stem-loop in the 3' untranslated region mediates iron-dependent regulation of transferrin receptor mRNA stability in the cytoplasm. Cell 53, 815-825. Napoli C, Lemieux C, J orgensen R (1990) Introduction of a chimeric chalcone synthase gene into petunia results in reversible co-suppression of homologous genes in trans. Plant Cell 2: 279-289 Nellen W, Lichtenstein C (1993) What makes an mRNA anti-sense-itive? Trends Biochem. Sci. 18:419-423 Newman TC, Ohme-Takagi M, Taylor CB, Green PJ (1993) DST sequences, highly conserved among plant SA UR genes, target reporter transcripts for rapid decay in tobacco. Plant Cell 5: 701-714 Oda K, Yarnato K, Ohta E, Nakamura Y, Takemura M, Nozato N, Akashi K, Kanegae T, Ogura Y, Kohchi T, Ohyama K (1992) Gene organization deduced from the complete sequence of liverwort Marchantia polymorpha mitochondrial DNA. J. Mol. Biol. 223: 1-7 Ohme-Takagi M, Taylor CB, Newman TC, Green PJ (1993) The effect of sequences with high AU content on mRNA stability in tobacco. Proc. Natl. Acad. Sci. USA 90: 11811-11815 Okamuro J K and Goldberg RB (1989) Regulation of plant gene expression: General Principles. In PK Stumpf, EE Conn, eds, The Biochemistry of Plants, Vol 15. pp 1-82 44 Petracek,M.E., Dickey,L.F., Nguyen,T.T., Gatz,C., Sowinski,D.A., Allen,G.C., and Thompson,W.F. (1998). F erredoxin-l mRNA is destabilized by changes in photosynthetic electron transport. Proc. Natl. Acad. Sci. USA 95, 9009-9013. Ross J (1995) mRNA stability in mammalian cells. Microbiol. Rev. 59: 423-450 Ross J (1996) Control of messenger RNA stability in higher eukaryotes. Trends Genet. 12: 171-175 Rouault TA, Hentze MW, Haile DJ, Harford J B, Klausner RD (1989) The iron-responsive element binding protein: A method for the affinity purification of a regulatory RNA-binding protein. Proc. Natl. Acad. Sci. USA 86: 5768-5772 Russell PJ, Hambidge SJ, Kirkegaard K (1991) Direct introduction and transient expression of capped and non-capped RNA in Saccharomyces cerevisiae. Nucleic Acids Res. 19: 4949-4953 Schiebel W, Haas B, Marinkovic S, Klanner A, Sanger HL (1993a) RNA-Directed RNA Polymerase from Tomato Leaves: II Catalytic in vitro properties. J. Biol. Chem. 263: 11858-11867 Schiebel W, Haas B, Marinkovic S, Klanner A, Sanger HL (1993b) RNA-Directed RNA Polymerase from Tomato Leaves: I Purification and physical properties. J. Biol. Chem. 268: 11851-11857 Schiebel,W., Pelissier,T., Riedel,L., Thalmeir,S., Schiebel,R., Kempe,D., Lottspeich,F., Sanger,H.L., and Wassenegger,M. (1998). Isolation of an RNA-directed RNA polymerase-specific cDNA clone from tomato. Plant Cell 10, 2087-2101. Seeley KA, Byrne DH, Colbert JT (1992) Red light-independent instability of oat mRNA in-vivo. Plant Cell 4: 29-38 Shaw G, Kamen R (1986) A conserved AU sequence from the 3' untranslated region of GM-CSF mRNA mediates selective mRNA degradation. Cell 46: 659-667 Sheu,J.J., Yu,T.S., Tong,W.F., and Yu,S.M. (1996). Carbohydrate starvation stimulates differential expression of rice alpha-amylase genes that is modulated through complicated transcriptional and posttranscriptional processes. J. Biol. Chem. 271, 26998-27004. Shobuike,T., Sugano,S., Yamashita,T., and Ikeda,H. (1995). Characterization of cDNA encoding a mouse homolog of fission yeast dhpl gene: structural and functional characterization. Nucleic Acids Res. 23, 357-361. Shyu A-B, Belasco JG, Greenberg ME (1991) Two distinct destabilizing elements in the c-fos message trigger deadenylation as a first step in rapid mRNA decay. Genes Dev. 5: 221-231 45 Siflow CD, Key J L (1979) Stability of polysome-associated polyadenylated RNA from soybean suspension culture cells. Biochemistry 18: 1013-1018 Smith CJS, Watson CF, Bird CR, Ray J, Schuch W, Grierson D (1990) Expression of a truncated tomato polygalacturonase gene inhibits expression of the endogenous gene in transgenic plants. Mol. Gen. Genet. 224: 477-481 Stevens A (1993) Eukaryotic nucleases and mRNA turnover. In JG Belasco, G Brawerrnan, eds, Control of Messenger RNA Stability, Academic Press, Inc. San Diego, CA, pp 449-471 Stewart CN, Adang MJ, All JN, Boerrna HR, Cardineau G, Tucker D, Parrott WA (1996) Genetic transformation, recovery, and characterization of fertile soybean transgenic for a synthetic Bacillus thuringiensis cryIAc gene. Plant Physiol. 112: 121-129 Sullivan ML, Green PJ (1993) Post-transcriptional regulation of nuclear-encoded genes in higher plants: the roles of mRNA stability and translation. Plant Mol. Biol. 23: 1091-1 104 Sullivan ML, Green P] (1996) Mutational analysis of the DST element in tobacco cells and transgenic plants: Identification of residues critical for mRN A instability. RNA 2: 308-315 Tanzer MM, Meagher RB (1994) Faithful degradation of soybean rch mRNA in vitro. Mol. Cell. Biol. 14: 2640-2650 Tanzer MM, Meagher RB (1995) Degradation of the soybean ribulose-l,5-bisphosphate carboxylase small-subunit mRNA, SRS4, initiates with endonucleolytic cleavage. Mol. Cell. Biol. 15: 6641-6652 Taylor CB, Green PJ (1995) Identification and characterization of genes with unstable transcripts (GUT s) in tobacco. Plant Mol. Biol. 28: 27-38 Thompson DM, Tanzer MM, Meagher RB (1992) Degradation products of the mRNA encoding the small subunit of ribulose- l ,5-bisphosphate carboxylase in soybean and transgenic petunia. Plant Cell 4: 47-58 Till,D.D., Linz,B., Seago,J.E., Elgar,S.J., Marujo,P.E., Elias,M.D., Arraiano,C.M., McClellan,J.A., McCarthy,J.E.G., and Newbury,S.F. (1998). Identification and developmental expression of a 5'-3' exoribonuclease fiom Drosophila melanogaster . Mech. Dev. 79, 51-55. Vakalopoulou E, Schaack J, Shenk T (1991) A 32-kilodalton protein binds to AU-rich domains in the 3' untranslated regions of rapidly degraded mRNAs. Mol. Cell. Biol. 1 1: 3355-3364 46 Van Blokland R, Van der Geest N, Mol JNM, Kooter J M (1994) Transgene-mediated suppression of chalcone synthase expression in Petunia hybrida results from an increase in RNA turnover. Plant J. 6: 861-877 van der Krol AR, Mur LA, Beld M, Mol JNM, Stuitje AR (1990) Flavonoid genes in petunia: Addition of a limited number of gene c0pies may lead to a suppression of gene expression. Plant Cell 2: 291-299 van Hoof A, Green PJ (1996) Premature nonsense codons decrease the stability of phytohemagglutinin mRNA in a position-dependent manner. Plant J. 10: 415-424 van Hoof A and Green PJ (1997) Control of mRNA decay in plants. In DR Orris, JB Harford, eds, Post-transcriptional Gene Regulation, Wiley and Sons, New York, Voelker TA, Staswick P, Chrispeels MJ (1986) Molecular analysis of two phytohemagglutinin genes and their expression in Phaseolus vulgaris cv. Pinto, a lectin-deficient cultivar of the bean. EMBO J. 5: 3075-3082 Vongs A, Kakutani T, Martienssen RA, Richards EJ (1993) Arabidopsis thaliana DNA methylation mutants. Science 260: 1926-1928 Vreken P, Raue’ HA ( 1992) The rate-limiting step in yeast PGKI mRNA degradation is an endonucleolytic cleavage in the 3'-tennina1 part of the coding region. Mol. Cell. Biol. 12: 2986-2996 Weinman P, Gossen M, Hillen W, Bujard H, Gatz C (1994) A chimeric transactivator allows tetracycline-responsive gene expression in whole plants. Plant J. 5: 559-569 Wessenegger M, Heimes S, Riedel L, Sanger HL (1994) RNA-Directed De Novo Methylation of Genomic Sequences in Plants. Cell 76: 567-576 Yamamoto KT, Mori H, Imaseki H (1992) cDNA cloning of indole-3-acetic acid-regulated genes: Aux22 and SAUR from mung bean (Vigna radiata) hypocotyl tissue. Plant Cell Physiol. 33: 93-97 Yu,S.-M., Kuo,Y.-H., Sheu,G., Sheu,J.-J., and Liu,L.F. (1991). Metabolic derepression of or-amylase gene expression in suspension-cultured cells of rice. J. Biol. Chem. 266, 21131-21137. Zhang S, Sheng J, Liu Y, Mehdy MC (1993) Fungal elicitor-induced bean proline-rich protein mRNA down-regulation is due to destabilization that is transcription and translation dependent. Plant Cell 5: 1089-1099 Zhang S, Ruiz-Echevarria MJ, Quan Y, Peltz SW (1995) Identification and characterization of a sequence motif involved in nonsense-mediated mRNA decay. Mol. Cell. Biol. 15: 2231-2244 47 Zhang S, Mehdy MC (1994) Binding of a 50-kD protein to a U-rich sequence in an mRNA encoding a proline-rich protein that is destabilized by fungal elicitor. Plant Cell 6: 135-145 Zubiaga AM, Belasco JG, Greenberg ME (1995) The nonamer UUAUUUAUU is the key AU-rich sequence motif that mediates mRNA degradation. Mol. Cell. Biol. 15: 2219-2230 48 CHAPTER 1-1 A PROCEDURE TO SELECT MUTANTS OF ARABIDOPSIS DEFECTIVE IN SEQUENCE-SPECIFIC mRNA DEGRADATION 49 INTRODUCTION Rates of messenger RNA (mRNA) degradation are highly variable in eukaryotic cells (Reviewed in Ross, 1995; Caponigro and Parker, 1996; Johnson et al., 1998, also see the introduction to this thesis). Some messages are degraded rapidly (minutes), while others persist in the cell for much longer (days). These differences allow for precise control of gene expression. It is important to know how the cell discriminates between long and short-lived mRNAs if we are to fully understand how gene expression is regulated. Cis-regulatory elements have been identified that act as either instability or stability determinants (for example, Holcik and Liebhaber, 1997; Chen and Shyu, 1995). These elements have been defined by inserting them into messages of average stability (half-life on the order of hours) and then measuring the half-life of the resulting chimeric transcripts (Ross, 1995). Instability sequences target these otherwise stable reporters for rapid degradation while stability elements stabilize the reporter mRNA. Our studies have focused on messages that are very unstable because in many cases these messages encode important regulatory proteins. This is logical because short-lived messages allow the cell to achieve optimum control over gene expression in time and space, characteristics that are important for the expression of key regulatory genes. Several sequences that target transcripts for rapid turnover in plants have been identified. Among these are the DST element (Newman et al., 1993), the AUUUA repeat (Ohme-Takagi, 1993), and premature stop-codons (van Hoof and Green, 1996). We are interested in understanding the 50 mechanisms by which the plant cell recognizes these sequences and targets the transcripts that contain them for rapid degradation. Biochemical approaches have resulted in the isolation of sequence-specific RNA binding proteins that are involved in modulating mRNA stability. The best example is the iron-responsive element (IRE) which is bound by the IRE-binding protein (IRE-BP) under iron-limiting conditions. Binding by the IRE-BP protects the transferrin mRNA from degradation. The lRE-BP was purified from human liver by preparing biotinylated- IRE RNA that was transcribed in vitro. A specific RNA binding protein was eluted from biotin-agarose at high salt (Roualt et al., 1989). This led to the cloning of a cDNA for the IRE-BP (Roualt et al., 1990). In addition, many RNA binding proteins have been isolated that interact with AU-rich elements (ARE) in vitro. ARES are potent instability determinants that result in rapid turnover of several mammalian proto-oncogenes and often contain AUUUA repeats (Chen and Shyu, 1995). These experiments have relied on UV-crosslinking and electrophoretic mobility shift assays to show that incubation of the protein with a radiolabeled RNA probe results in a specific interaction. The challenge has been to understand the contribution of these ARE-binding proteins to ARE-mediated mRNA degradation in vivo. Isolation of sequence-specific RNA-binding proteins from plant cells has not been as successfiil perhaps because it is difficult to prepare cytoplasmic protein extracts that are free of non-specific ribonucleases (F eldbrugge et al., unpublished data). This may be due to the prominence of the plant vacuole, which is known to contain many non-specific ribonucleases and other hydrolytic enzymes (Boller and Kende, 1979; Abel and Glund, 1987). Despite these difficulties, there is at least one example of a sequence-specific 51 RNA-binding—protein that may be involved in mRNA turnover in plants. The PRP-BP binds specifically to a sequence in the 3’ untranslated region (UTR) of the Phaseolus vulgaris proline rich protein (PvPRP) mRNA which is destabilized in response to fungal ellicitor treatment of bean cell cultures (Zhang and Mehdy, 1994; Zhang et al., 1993). The gene that encodes this protein and how it affects PvPRP mRNA stability are still unknown. An alternative to purification of sequence-specific RNA binding proteins would be to isolate mutant plants that have lost the ability to recognize mRNA instability determinants. This approach allows several important advantages. First, genes identified in a mutant selection would be very likely to play a role in mRNA degradation in vivo. Second, basic information about the mechanisms of mRNA degradation may be obtained by studying mutants. Third, other cellular factors that might be involved in sequence- specific interactions, such as RNA molecules, may be identified as long as they are encoded by the nuclear genome. Finally, complications such as the presence of non- specific RNA degrading activities in protein extracts are eliminated. Threfore, parallel methods were developed to isolate mutants defective in DST- and AUUUA-mediated mRNA degradation. There were several reasons to focus on these instability determinants. The DST element has been well characterized and is likely to play an important role in regulating the expression of many plant genes (Newman et al., 1993; Sullivan and Green, 1996; Gil and Green, 1996). The AUUUA repeat contains 1 1 copies of the core sequence of the ARES, the most widely studied instability determinant (Chen and Shyu, 1995, Lagnado et al., 1994; Zubinga et al., 1995). The capacity to apply genetic analysis to ARE-mediated mRNA decay is limited in 52 mammalian cell cultures and this element is not known to function in model genetic systems such as yeast. Therefore, studies in Arabidopsis may provide a unique opportunity to study ARE-mediated mRNA degradation using mutants. The selection strategy presented here is the first one designed to select mutants that are defective in rapid mRNA degradation mediated by a specific sequence element. This approach has resulted in the isolation of mutants defective in rapid mRNA degradation mediated by the DST element (see Chapter 1-2). This procedure was based on a selectable phenotype that we engineered using transgenic Arabidopsis plants and is theoretically adaptable to any cis-regulatory element that results in rapid mRNA degradation. Utilization of this method should lead to an increase in our understanding of the basic mechanisms that are responsible for rapid degradation of mRNA in higher plants. RESULTS Premise of the mutant selection strategy In order to isolate mutants of Arabidopsis defective in a sequence-specific mRNA degradation pathway we first needed to engineer plants that would display a mRNA abundance phenotype that could be readily screened for and/or selected. This mutant selection strategy relied on two reporter genes that each contained the DST or AUUUA instability determinants in their 3’ UTR (Figure 1-1-1). In wild-type (wt) plants, the instability determinant functions effectively and the abundance of the reporter mRNA is 53 maintained at a low level. Mutants that are defective in rapid mRNA degradation mediated by the instability determinant would be expected to have an increased abundance of both reporter mRNAs. If transgenes are chosen that encode readily selectable and/or screenable enzymatic activities, it should be possible to isolate mutants by virtue of an increase in reporter gene expression. T-DNA H mm Parental Plants: Low HPH mRNA Hygromycin sensitive Low GUS mRNA Low GUS Activity I; collect seed 4} mutagenlze I} collect M2 seed Mutants: _ High HPH mRNA 3; plate on hygromycin Hygromycin resistant High GUS mRNA High GUS Activity Figure 1-1-1. A strategy for selecting mutants defective in a sequence-specific mRNA decay pathway. Transgenic plants are generated using a construct such as the one represented above. Arrows represent promotor sequences. Hygromycin phosphotransferase (HPH) and B-glucuronidase (GUS) are examples of reporter genes that confer selectable and screenable phenotypes, respectively. I.D. represents the instability determinant being studied, in this case either DST or AUUUA. 54 HPH and GUS are chosen as reporter genes We initially considered hygromycin phosphotransferase (HPH), dihydrofolate reductase (DHF R), and B-glucuronidase (GUS) as potential transgenes for this strategy. HPH and DHF R confer resistance to hygromycin and methotrexate, respectively, and there are many commercially available substrates that facilitate detection of GUS gene expression (Nunberg et al., 1980; Gritz and Davies, 1983; Jefferson et al., 1987). The impact of mRNA destabilization upon each of these selectable marker and reporter genes was an important consideration. A large difference in mRNA abundance between plants expressing a destabilized reporter versus a non-destabilized reporter would be optimal for the isolation of mutants with a wider range of mRNA abundance phenotypes. Previous experiments with transgenic tobacco plants showed that abundance of the human ,B-Globin (Globin) transcript was reduced by ~14 fold when the AUUUA repeat was inserted into its 3 ’UTR or by ~7.5 fold when a dimer of the DST element (DSTx2) was inserted (Ohme—Takagi et al., 1993; Newman et al., 1993). In order to assess the effect of AUUUA repeats and the DST element on the abundance of GUS, HPH, and DHF R transcripts in Arabidopsis, T-DNA constructs were generated that placed either the AUUUA repeat, DSTx2 or a tetramer of the DST element (DSTx4) into the 3’UTR of each selectable marker or reporter gene. Similar constructs using the Globin coding region were also analyzed so that the results in Arabidopsis could be compared with those already obtained in tobacco. These constructs were designed to have a reference gene within the same T-DNA to serve as a loading control for Northern blot analysis. A representative construct is depicted in Figure 1-1-2. 55 Test Reference Figure 1-1-2. The basic structure of transgenes used to assess various reporters and instability determinants in transgenic Arabidopsis calli. The cauliflower mosaic virus 35S promoter (35S) was used to control the expression of the DHFR, HPH, GUS, and Globin reporter genes. Polyadenlytion signals derived from the pea ribulose 1,5- bisphosphate carboxylase small subunit E9 (rch—E9, E9) gene were used as mRNA processing signals for DHFR, HPH, and Globin. The 3’ UTR of the of the rch-3C (3C) gene was used as a polyadenylation signal for GUS. The reporters were paired in test: reference combinations as follows: DHFR with GUS, HPH with GUS (shown), GUS- AUUUA with Globin, GUS-DST x2 with DHFR, GUS-DSTx4 with HPH, and Globin with GUS. The right border (RB) of the transfer-DNA (T-DNA) has been indicated by a black rectangle. The T-DNA also includes the neomycin phosphotranserase cassette so that kanarnycin resistant transgenic tissue could be selected. These constructs were used to generate transgenic Arabidopsis calli by co- cultivating Agrobacterium strains harboring the relevant constructs with root explants of Arabidopsis seedlings (Valveken eta1., 198 8). Pools of at least 100 independent transgenic calli were gathered for each construct and the mRNA abundance of each construct was analyzed by Northern blot analysis. Radiolabelled probes complementary to the DHFR, HPH, GUS, and Globin coding regions were used to determine the abundance of these reporter genes on blots that were loaded with equal amounts of total RNA from pooled calli. Signals corresponding to reference and test transcripts were quantitated by Phosphorlmager. The test signal for each construct was normalized using the reference signal and the average normalized value was plotted relative to non- destabilized control constructs (Figure 1-1-3). 56 DHFR HPH GUS Globin 100 ' " t r- P F I r- 75 I I n p 50 I I I p 25 I I I :Control - AUU UA repeat rm DSTx2 :3 DSTx4 Figure 1-1-3. Accumulation of reporter transcripts in Arabidopsis. This histogram depicts the normalized abundance of reporter transcripts in transgenic Arabidopsis calli. Data are presented as a percentage of the non-destabilized controls; where multiple experiments were done error bars have been plotted (standard deviation). One important observation from this experiment was that neither a dimer of the DST element (DSTx2) nor the AUUUA repeat destabilized the DHFR transcript to the same extent as they did HPH, GUS, or Globin. It should be noted that the absolute abundance of the DHFR transcript was lower than that of HPH, GUS, or Globin, indicating that perhaps the DHF R transcript was relatively unstable to begin with (data not shown). This might explain why insertion of instability sequences into this transcript did not result in the same reduction in abundance that was observed for HPH, GUS, and Globin. These results indicated that DHF R would not be an appropriate selectable marker for our mutant selection strategy because an increase in mRNA abundance in a putative mutant may not have been detectable above the already high relative abundance of the transcript in wt plants. On the other hand, the observed effect of the AUUUA repeat and the DST element on HPH and GUS mRNA abundance was consistent with 57 previously obtained Globin data and indicated that using these reporters would provide a wider range in which to select new mRNA abundance phenotypes in mutants. Selection of transgenic plants to serve as the parental lines for mutagenesis HPH and GUS were chosen as the reporter genes for our selection strategy (Figure 1-1-1). DSTx4 was used instead of DSTx2 because of the larger decrease in mRNA abundance it caused. New constructs were made and then used to generate transgenic Arabidopsis plants harboring both the selectable marker and reporter transcripts destabilized by AUUUA (p1514) or DSTx4 (p1519). In addition, a positive control construct was made that carried the same genes without instability determinants (Figure 1-1-4). Since it was important to isolate plants with single inserts of the T-DNA and that expressed the reporter genes to the appropriate level, several independent transgenic lines were generated with each construct and these were examined to choose the best lines for mutagenesis. Transgenic lines in which the reporter gene expression reflected the average mRNA abundance found in the experiments shown in Figure 1-1-3, became candidates to serve as the parental line for mutagenesis. 58 p1 14 HPH IAWUE gas-Mus imam Figure 1-1-4. T-DNA constructs used in mutant selection strategy. The T-DNAs also include the neomycin phosphotranserase cassette so that kanamycin resistant transgenic tissue could be selected. Also not shown is the nopaline synthase (nos) gene which contains a BamHI restriction enzyme site and is located between the GUS-3 C gene and the right border (RB). In order to identify lines with one insert of the T-DNA, T2 (progeny of the self- fertilization of the primary transforrnant) seed were plated on primary seed selection medium containing 50 ug/ml kanamycin. Three p1519 (DSTx4) and three p1514 (AUUUA) lines that showed a ratio of three resistant to one sensitive seedling, indicating inheritance of a single, dominant locus were chosen. Single inserts were confirmed by Southern blot analysis (Figure 1-1-5). 59 p1514-9 p1514-22 p1514-21 - p1519-31 ' p1519-26 p1519-13 -10Kb ‘ -fi'zu4Kb In» in! -- «In- an '1.2Kb 'AUUUA DSTx4 Figure 1-1-5. Southern blot analysis of candidate lines for mutagenesis. Genomic DNA was extracted from transgenic plants, digested with BamHI, separated on a 1% agarose gel, and transferred to a nylon membrane. The membrane was hybridized with a radiolabeled probe complementary to the nopaline synthase (nos) coding region; adjacent to the right border of the T-DNA. A 1.2 kb fragment that is part of the T-DNA is present in all of the lanes. Additional bands represent the distance to BamHI sites in the genome. The number of additional bands is indicative of the number of copies of the T-DNA. p1514-9 and p1519-31 had simple banding patterns indicating that these lines had single inserts. Northern blot analysis confirmed that each of these transgenic lines expressed the HPH and GUS reporter transcripts to levels that were expected based on previous experiments (Figure 1-1-6). Insertion of the AUUUA repeat into the 3’UTR of the HPH or GUS transcripts resulted in an increase in mRNA size of 60 nucleotides. Insertion of DSTx4 resulted in an increase of 180 nucleotides. Based on the results of this experiment and those previously described, p1514-9 and pl 519-31 were chosen as the parental lines for mutagenesis. The higher molecular weight GUS transcript (indicated by (0) in Figure 1-1-6) that is present in the p1519-31 lane is a consistent characteristic of 60 this line. We chose p1519-31 for mutagenesis despite this because this line had the simplest T-DNA insertion and because the additional band did not seem to alter GUS activity (data not shown). We measured the abundance of the lower molecular weight band, which is the expected size, throughout these experiments. p1514—9 p1514-22 p1514-21 p1519-31 p1519-26 p1519-13 i. :: 500 + AUU UA DSTx4 Figure 1-1-7. p1514-9 and p1519-31 show increased hygromycin sensitivity. Approximately 500 seeds were surface sterilized and plated on primary seed selection medium containing 50 ug/ml kanamycin plus the indicated concentration of hygromycin. The photograph was taken after four weeks. This experiment indicated that decreases in mRNA abundance due to insertion of instability sequences into the HPH transcript caused increased hygromycin sensitivity in 62 p1514-9 and p1519-31 seedlings. This was especially clear at high concentrations of hygromycin (1000 rig/ml) where very few seedlings escape this selection. Mutagenesis oflines p1514-9 and p1519-31 Plants homozygous at the T-DNA locus present in lines p1514-9 and p1519-31 were obtained by plating T3 families on kanamycin and selecting those that segregated 100% resistance. Seed from multiple T3 plants were pooled so that adequate T4 seed would be available for ethylmethane sulfonate (EMS) mutagenesis (Redei and Koncz, 1992, Li ghtner and Caspar, 1998). EMS is a standard mutagen for genetic analysis of Arabidopsis. It is thought to cause point mutations primarily due to GC to AT transitions that result from ethylation of DNA (Feldman et al., 1994). Standard EMS treatments result in multiple, hemizygous mutations in each plant derived from mutagenized seed (M.) such that only about 2000 M2 plants (self-fertilized progeny of M.) need be analyzed to find a homozygous mutation in any one of the ~ 20,000 Arabidopsis genes (Haughn and Somerville, 1986). We followed standard previously described EMS mutagenesis protocols (Redei and Koncz, 1992). Table 1-1-1 describes the mutagenesis ofp1514-9 and p1519-31. 63 Table 1-1-1. EMS mutagenesis of transgenic Arabidopsis lines. p1514-9 AUUUA EXJ). # of MI Seed Group # of Ml/Group Pilot 5000 Pilot 5000 1 100,000 1-20 5000 2 100,000 2 1-40 5000 3 50,000 41-52 4000 Total 255,000 4800 p1519-31 DSTx4 Exp. # of M1 Seed Group # of Ml/Group 1 100,000 1-20 5000 2 100,000 21-40 5000 3 25,000 41 -46 4000 Total 225,000 4900 Ml, first generation of mutagenized seed. Following mutagenesis, M. seeds were planted in groups of ~5000 plants and M2 seed were collected from each group. M2 seeds were collected in groups in order to obtain an initial approximation of the number of independent mutations, as mutants selected from different groups were likely to represent independent mutations. The effectiveness of EMS mutagenesis was evaluated by scoring the occurrence of albinism and embryo lethality in MI plants, as previously described (Redei and Koncz, 1992). Selection of putative mutants Putative mutants were selected by plating ~4000 M2 seeds on primary seed selection plates containing 1000 11ng hygromycin and 50 ug/ml kanamycin. After 3-4 weeks putative mutants, defined as seedlings that had developed at least two true leaves, were chosen from the primary plates and moved to plates lacking hygromycin. 64 Development of two true leaves was consistently absent in non-mutagenized p1519-31 and p1514-9 seedlings. After one to two weeks on plates lacking hygromycin, putative mutants were transferred to soil. Table 1-1-2 summarizes the results of the selection experiments and the number of putative mutants that were obtained. Table 1-1-2. Hygromycin selection experiments p1514-9 AUUUA M2 group # M2 plated # putative mutants Pilot 4000 4 1-20 40,000 2 21-40 127,000 126 41-52 668,000 71 Total 839,000 203 p1519-31 DSTx4 M2 group # M2 plated # putative mutants 1-20 90,000 14 2 l -40 444,000 263 41—52 200,000 46 Total 730,000 323 #, number; M2, second generation after mutagenisis The abundance of the HPH and GUS reporter mRNAs was determined by Northern blot analysis of leaf RNA extracted from M2 plants selected on hygromycin. These experiments allowed rapid scoring of the key mutant phenotype. The abundance of the HPH and GUS transcripts was normalized using the translation initiation factor 4- A (eIF4-A) transcript abundance; these were then compared to the normalized HPH and GUS mRN A abundance of parental samples that had been grown in parallel. Where this was not possible, for example if the plant was too small to obtain an adequate RNA sample, mRNA abundance was assayed in M3 plants. In addition, M3 seeds were plated 65 on hygromycin to determine whether the resistance phenotype was inherited. Hygromycin resistance was found to be heritable in approximately 20% of the M3 families tested, suggesting a high rate of false positives. 323 putative mutants derived from mutagenesis of line p1519-31 (DSTx4) and 203 derived from p1514-9 (AUUUA) were analyzed by these criteria. mRNA abundance was analyzed by Northern blot analysis for all putative mutants. Only three putative mutants met the criteria of heritable increases in HPH and GUS mRNA abundance. These were DST-M2#64 (group 21), DST-M2#l 14 (group 31), and DST-M2#290 (group 30). No mutants were isolated from the AUUUA population even though over 800,000 p1514-9 M2 seeds were plated. DST-M2#114 and DST-M2#64 were isolated relatively early in the selection process and were characterized in detail. These mutations were found to be in independent genes and have been renamed dstI (DST-M2#1 14) and dst2 (DST-M2#64) to reflect their defects in DST-mediated mRNA degradation. A detailed characterization of these mutants is the subject of the next chapter of this thesis. DST- M2#290 is currently being analyzed. The immediate priorities are to establish the mode of inheritance of the mutation and to determine whether this mutant complements either dstI or dst2. DISCUSSION Genetic selection strategies such as the one presented here have been called second generation screens or ‘targeted genetics’ because they rely on engineered phenotypes rather than those intrinsic to the plant (for example, see Hooley, 1998). 66 Targeted genetics can provide a novel class of mutants to an area where classical screens have already been saturated. The recently isolated age (guxin-responsive gene expression) mutants provide an excellent example (Oono et al., 1998). Using a fusion of an auxin-responsive promoter element to the GUS reporter gene, Oono et al., were able to isolate auxin response mutants that misexpressed the reporter. These mutants provide a useful complement to the many auxin-response mutants that have been identified by traditional means. Perhaps more importantly, targeted genetics can provide an entry into areas that have been inaccessible to genetic analyses because obvious, intrinsic phenotypes are lacking. This was the case for sequence-specific mRNA degradation mediated by AUUUA repeats or the DST element. No targets of the AUUUA-mediated mRNA degradation pathway have been definitively identified in plants. Therefore, it is difficult to imagine an intrinsic phenotype that would result from the disruption of this pathway. The DST element is likely to be involved in the constitutive instability of several Arabidopsis SA UR transcripts. Therefore, it may be possible to devise a direct screen for mutants based on the abundance of these transcripts. However, this would be extremely labor-intensive and many of the mutants would be defective in the transcriptional regulation of these genes by auxin. In fact, auxin-response mutants have been isolated that result in misexpression from the SA UR-A C1 promoter (Leyser et a1. 1996). It is for these reasons that a targeted genetic approach was necessary in order to explore the genetic basis for sequence-specific mRNA degradation in plants. The mutants isolated in this study were very rare. Three were isolated from a total of approximately 1.5 x106 mutagenized seeds that were plated during the course of 67 the DST and AUUUA mutant selections. One of the possible explanations for this observation is that genes involved in AUUUA- and DST-mediated mRNA degradation are essential. Since it is likely that mRNA degradation pathways limit the expression of key regulatory genes, it is possible that if transcripts targeted by these pathways are overexpressed or misexpressed during early development of the plant due to a mutation, the result would be embryo lethality. A relevant and interesting observation is that overexpression of AUUUA-containing oncogene messages due to mutations that delete AREs have resulted in neoplastic transformation in mammalian cells (Piechaczyk et al., 1985). This finding highlights the potential consequences that can result when mRNA degradation pathways are rendered ineffective by mutations. If these pathways are essential for development of plants, one might also expect redundancy to have evolved in these systems. This is another reason why mutants in the AUUUA pathway may not have been isolated. As is described in the next chapter of this thesis, dstI and dst2 were partially dominant mutations. If the DST-mediated mRNA degradation pathway is redundant, then one might expect to only uncover dominant alleles that interact with and negate redundant members of the pathway. Mutations in the ethylene receptor illustrate how dominant negative alleles can provide access to redundant genes (reviewed by Theologis, 1998). The original ethylene receptor mutations were all dominant. After a receptor gene was isolated by map based cloning, it was found that ethylene receptors constitute a gene family in Arabidopsis. Loss-of- function mutant phenotypes were not observed until triple and quadruple mutants were constructed (Hua and Meyerowitcz, 1998). 68 There are also technical explanations for the observed rarity of sequence-specific mRNA degradation mutants. Approximately 80% of the putative mutants that were chosen for further study were found to be false positives. This may have been due to overzealous picking of putative mutants. However, this may have been important because the mutants that were recovered, dstl and dst2, showed only slight increases in HPH and GUS mRNA abundance and may have been missed if only the healthiest plants were rescued from selection plates. In any case, this rate of false-positives, decreased the efficiency of this selection. Targeted genetics approaches have been used previously to isolate mutations that affect the function of specific promoter sequences (Susek et al., 1993; Bowling et al., 1994; Jackson etal., 1995; Martin et al., 1997; Oono et al., 1998). Here, the utility of such a procedure has been extended to cis-regulatory elements located in the 3’UTR. Therefore, the mutants isolated using this procedure have the potential to provide insight into rapid sequence-specific mRNA degradation in plants and eukaryotes in general and the cloning of dst genes may aid in the ellucidation of the molecular machinery that catalyzes the degradation of DST-containing transcripts. These studies have the potential to answer many fundamental questions about mRNA turnover in plants. For example, do DST-recognition components recruit the basal mRNA decay machinery more rapidly to DST-containing transcripts? One may find that dst gene products interact with poly(A) binding-proteins and/or poly(A) ribonucleases so that the poly(A) tails of DST-containing messages are removed more rapidly. The basic mechanisms of mRNA degradation in plants are still largely unknown. Any insight into factors that interact with dst gene 69 products may reveal clues that will lead to the basic components responsible for this process. MATERIALS AND METHODS Plasmid construction and plant transformation Standard molecular cloning procedures were used (Sambrook et al., 1989). Construction of the chimeric genes described in Figure 1-1-2 and Figure 1-1-4 was as described by Newman et al., 1993. Chimeric genes were constructed in pBluescript SKII+ (Stratagene) or pUC (New England Biolabs) and were then transferred to derivatives of pMON505 which is a binary vector that facilitates transformation of plants via Agrobacterium tumefaciens (Agrobacterium)(Rogers et al., 1987, Fang et al., 1989). Chimeric genes had the basic structure: SacI-35S-BglII-XbaI-Coding region-BamHI- poly(A) sequence-Clal. DST and AUUUA sequences were inserted at the BamHI site between the coding sequence and the poly(A) signal. The source of the Globin and GUS coding regions has been described (Newman et al., 1993). The HPH coding region was obtained as a BamHI fragment of pLG90 (Gritz and Davies, 1983). The DHFR coding sequence was described by Nunberg et al., (1980). The DST sequence used in these experiments is the same as that found in the soybean SA UR-I5A gene and was synthesized as described by Newman et a1, 1993. The construction of the AUUUA repeat was described by Ohme-Takagi et al., 1993. The use of the 35S promoter and 70 polyadenylation signals from the 3’ ends of the E9 and 3C genes was as described by Newman et al., 1993. Transgenic Arabidopsis calli were generated by co-cultivation of Agrobacterium with root explants (Valveken et al., 1988). A grobacterium strain LBA4404 was transformed with pMON505 derivatives by electroporation using the Gene-Pulser (BioRad) according to the manufacturer’s guidelines. Generation and selection of transgenic Arabidopsis plants using the vacuum infiltration method of Agrobacterium- mediated transformation has been described (N. Bechtold et al.,1993 and Web site: http://www.bch.msu.edu/pamgreen/vac.htm). Agrobacterium strain GV3101 C58C1 Rif (pMP90) (Koncz and Schell, 1986) was transformed with pMON505 derivatives by electroporation using the Gene-Pulser (BioRad) according to the manufacturer’s guidelines. Plant material and growth conditions The Columbia (Col-O) accession of Arabidopsis thaliana was used in all transformations except for p1514 and p1519. To avoid any hygromycin resistant contamination ofp1514-9 and p1519-31 during large-scale selection experiments, these constructs were introduced into a Columbia mutant that lacks trichomes (glI, Col-PRL). Arabidopsis calli were maintained on Gamborg’s BS (An, 1985) plates in an incubator set at 16 hours light (125 pE/mz) /8 hours dark and 21°C. Arabidopsis plants were grown in standard Arabidopsis soil in controlled environment growth chambers set at 16 hours light (125 uE/mz) /8 hours dark, 21°C. 71 EMS mutagenesis and selection of hygromycin resistant plants Seeds homozygous at the p1514-9 and p1519-31 loci (fourth generation of self- fertilized progeny following transformation) were soaked for 16 hours in a 0.3% (v/v, water) solution of EMS. Following mutagenesis, seeds were washed extensively with water over a 12-hour period and then sown on soil. A thorough protocol is given in Li ghtner and Caspar, (1998). The effectiveness of EMS mutagenesis was evaluated by scoring the occurrence of albinism and embryo lethality in M1 plants as described by Redei and C. Koncz (1992); and Li ghtner and Caspar, (1998). Seed selection media consisted of 4.3 g/L Murashige and Skoog salt mixture (GibcoBRL), BS vitamin mixture (100 mg/l myo-inositol, 10 mg/l thiamine hydrochloride, 1 mg/l nicotinic acid, 1 mg/l pyriodoxine), 1% sucrose, 0.5 g/L MES at pH 5.7, 0.8% phytagar, 1000ug/ml hygromycin and 50 ug/ml kanamycin. Hygromycin was omitted when selecting transgenic plants or scoring inheritance of the T-DNA. RNA analysis RNA was isolated from transgenic calli (Figure 1-1-3) and pooled seedlings (Figure 1-1-6) using the method described by Puissant and Houdebine (1990) with modifications described by Newman et al., (1993). For analysis of mRNA abundance in individual plants, two rosette leaves were harvested from mature plants and RNA was extracted using the method described by Verwoerd et a1. (1989). Northern blot analysis was as described by Newman et a1. (1993). Hybridization probes corresponding to the 72 DHFR, eIF4-A, Globin, GUS, and HPH coding regions were labeled using 32P dCTP by the random-primed method of F einberg and Vogelstein ( 1983). The translation initiation factor, eIF4-A, was used as normalization standard for northern blots as described by Taylor et al., (1993). mRNA abundance was quantitated by measuring the intensity of radioactive bands on Northern blots using a Molecular Dynamics Phosphorlmager. DNA analysis Southern blotting was performed as described in Sambrook et al., 1989. DNA was extracted from dark-grown seedlings using the method described by Saghai-Maroof et a1. (1984). ISug of DNA was digested with BamHI and loaded on a 1% agarose gel which was transferred to a nylon membrane. Prehybridization, hybridization, and washing conditions were the same as those for Northern blots (Newman et al., 1993). Hybridization probes corresponding to the nos coding region were labeled using 32P dCTP by the random-primed method of Feinberg and Vogelstein (1983). ACKNOWLEDGEMENTS I am grateful to Dr. Michael L. Sullivan who constructed the chimeric genes used in this study and got me started on this project. Dr. Sullivan and Debrah M. Thompson generated the data presented in Figure 1-1-3. The selection of putative mutants in the AUUUA pathway was mainly the work of Miguel A. Pérez-Amador. I would also like to thank Linda Danhof and Jonathan Vogel for tireless technical assistance. This work was 73 fiinded by grants from the United States Department of Energy, the United States Department of Agriculture, and the McKnight Foundation to Pamela J. Green. 74 REFERENCES Abel,S. and Glund,K. (1987). Ribonuclease in plant vacuoles: purification and molecular properties of the enzyme from cultured tomato cells. Planta I 72, 71-78. An,G. (1985). High-efficiency transformation of cultured tobacco cells. Plant Physiol. 79, 568-570. Bechtold,N., Ellis,J., and Pelletier,G. (1993). In planta Agrobacterium mediated gene transfer by infiltration of adult Arabidopsis thaliana plants. C. R. Acad. Sci. Paris 316, 1 194-1 199. Boller,T. and Kende,H. (1979). Hydrolytic enzymes in the central vacuole of plant cells. Plant Physiol. 63, 1123-1132. Bowling,S.A., Guo,A., Cao,H., Gordon,A.S., K1essig,D.F., and Dong,X. (1994). A mutation in Arabidopsis that leads to constitutive expression of systemic acquired resistance. Plant Cell 6, 1845-1857. Caponigro,G. and Parker,R. (1996). Mechanisms and control of mRNA turnover in Saccharomyces cerevisiae. Microbiol. Rev. 60, 233-249. Chen,C.-Y.A. and Shyu,A.-B. (1995). AU-rich elements: characterization and importance in mRNA degradation. Trends Biochem. Sci. 20, 465-470. Fang,R.-X., Nagy,F., Sivasubramaniam,S., and Chua,N.-H. (1989). Multiple cis regulatory elements for maximal expression of the cauliflower mosaic virus 358 promoter in transgenic plants. Plant Cell 1, 141-150. F einberg,A.P. and Vogelstein,B. (1983). A technique for radiolabeling DNA restriction endonuclease fragments to high specific activity. Anal. Biochem. 132, 6-13. Feldmann,K.A., Malmgren,C., and Dean,C. (1999). Mutagenesis in Arabidopsis. In ARABIDOPSIS, E.M.Meyerowitz and C.R.Somerville, eds. (Plainview, New York: Cold Spring Harbor Laboratory Press), pp. 137-172. Gil,P. and Green,P.J. (1996). Multiple regions of the Arabidopsis SA UR-ACI gene control trascript abundance: the 3' untranslated region functions as an mRNA instability determinant. EMBO J. 15, 1678-1686. Gritz,L. and Davies,J. (1983). Plasmid-encoded hygromycin B resistance: the sequence of hygromycin B phosphotransferase gene and its expression in Escherichia coli and Saccharomyces cerevisiae. Gene 25, 179-188. Haughn,G. and Somerville,C.R. (1987). Selection for herbicide resistance at the whole- plant level. In Biotechnology in Agricultural Chemistry, H.M.LeBaron, R.O.Mumma, 75 R.C.Honeycutt, and J .H.Duesing, eds. (Washington, DC: American Chemical Society), pp. 98-107. Holcik,M. and Liebhaber,S.A. (1997). Four highly stable eukaryotic mRNAs assemble 3' untranslated region RNA-protein complexes sharing cis and trans components. Proc. Natl. Acad. Sci. USA 94, 2410-2414. Hooley,R. (1998). Auxin signaling: Homing in with targeted genetics. Plant Cell 10, 1581-1583. Hua,J. and Meyerowitz,E.M. (1998). Ethylene responses are negatively regulated by a receptor gene family in Arabidopsis thaliana. Cell 94, 261-271. Jackson,J.A., Fuglevand,G., Brown,B.A., Shaw,M.J., and Jenkins,G.I. (1995). Isolation of Arabidopsis mutants altered in the li ght-regulation of chalcone synthase gene expression using a transgenic screening approach. Plant J. 8, 369-380. Jefferson,R.A., Kavanagh,T.A., and Bevan,M.A. (1987). GUS fusions: Beta- glucuronidase as a sensitive and versatile gene fusion marker in higher plants. EMBO J. 6, 3901-3907. Johnson,M.A., Baker,E.J., Colbert,J.T., and Green,P.J. (1998). Determinants of mRNA stability in plants. In A look beyond transcription: Mechanisms determining mRNA stability and translation in plants, J .Bailey-Serres and D.R.Gallie, eds. (Rockville: American Society of Plant Physiologists Press), pp. 40-53. Koncz,C. and Schell,J. (1986). The promoter of TL-DNA gene 5 controls the tissue- specific expression of chimaeric genes carried by a novel type of Agrobacterium binary vector. Mol. Gen. Genet. 204, 383-396. Lagnado,C.A., Brown,C.Y., and Goodall,G.J. (1994). AUUUA is not sufficient to promote poly(A) shortening and degradation of an mRNA: The functional sequence within AU-rich elements may be UUAUUUA(U/A)(U/A). Mol. Cell. Biol. 14, 7984- 7995. Leyser,H.M.O., Pickett,F.B., Dharmasiri,S., and Estelle,M. (1996). Mutations in the AXR3 gene of Arabidopsis result in altered auxin response including ectopic expression from the SA UR-A C1 promoter. Plant J. 10, 403-413. Lightner,J. and Caspar,T. (1998). Seed mutagenesis of Arabidopsis. Methods Mol. Biol. 82, 91-103. Martin,T., Hellmann,H., Schmidt,R., Willmitzer,L., and Frommer,W.B. (1997). Identification of mutants in metabolically regulated gene expression. Plant J. I 1, 53-62. Newman,T.C., Ohme-Takagi,M., Taylor,C.B., and Green,P.J. (1993). DST sequences, highly conserved among plant SA UR genes, target reporter transcripts for rapid decay in tobacco. Plant Cell 5, 701-714. 76 Nunberg,J.H., Kaufman,R.J., Chang,A.C.Y., Cohen,S.N., and Schimke,R.T. (1980). Structure and genomic organization of the mouse dihydrofolate reductase gene. Cell 19, 355-364. Ohme-Takagi,M., Taylor,C.B., Newman,T.C., and Green,P.J. (1993). The effect of sequences with high AU content on mRNA stability in tobacco. Proc. Natl. Acad. Sci. USA 90,11811-11815. Oono,Y., Chen,Q.G., Overvoorde,P.J., Kohler,C., and Theologis,A. (1998). age Mutants of Arabidopsis exhibit altered auxin-regulated gene expression. Plant Cell 10, 1649-1662. Piechaczyk,M., Yang,J.Q., Blanchard,J.M., Jeanteur,P., and Marcu,K.B. (1985). Posttranscriptional mechanisms are responsible for accumulation of truncated c-myc RNAS in murine plasma cell tumors. Cell 42, 589-597. Puissant,C. and Houdebine,L.-M. (1990). An improvement of the single-step method of RNA isolation by acid guanidinium thiocyanate-phenol-chlorophorm extraction. BioTechniques 8, 148-149. Redei,G.P. and Koncz,C. (1992). Classical mutagenesis. In Methods in Arabidopsis Research, C.Koncz, N.-H.Chua, and J .Schell, eds. (Singapore: World Scientific), pp. 16- 82. Rogers,S.G., Klee,H.J., Horsch,R.B., and Fraley,R.T. (1987). Improved vectors for plant transformation: expression cassette vectors and new selectable markers. Methods Enzymol. 153, 253-277. Ross,J. (1995). mRNA stability in mammalian cells. Microbiol. Rev. 59, 423-450. Rouault,T.A., Hentze,M.W., Haile,D.J., Harford,J.B., and Klausner,R.D. (1989). The iron-responsive element binding protein: A method for the affinity purification of a regulatory RNA-binding protein. Proc. Natl. Acad. Sci. USA 86, 5768-5772. Rouault,T.A., Tang,C.K., Kaptain,S., Burgess,W.H., Haile,D.J., Samaniego,F., McBride,O.W., Harford,J.B., and Klausner,R.D. (1990). Cloning of the cDNA encoding an RNA regulatory protein--the human iron-responsive element-binding protein. Proc. Natl. Acad. Sci. USA 87, 7958-7962. Saghai Maroof,M.A., Biyashev,R.M., Yang,G.P., Zhang,Q., and Allard,R.W. (1994). Extraordinarily polymorphic microsatellite DNA in barley: species diversity, chromosomal locations, and population dynamics. Proc. Natl. Acad. Sci. U. S. A 91, 5466-5470. Sambrook,J., Fritsch,E.F., and Maniatis,T. (1989). Molecular Cloning: A Laboratory Manual. (Cold Spring Harbor, New York: Cold Spring Harbor Laboratory Press). 77 Sullivan,M.L. and Green,P.J. (1996). Mutational analysis of the DST element in tobacco cells and transgenic plants: Identification of residues critical for mRNA instability. RNA 2, 308-315. Susek,R.E., Ausubel,F.M., and Chory,J. (1993). Signal transduction mutants of Arabidopsis uncouple nuclear CAB and RBCS gene expression from chloroplast development. Cell 74, 787-799. Taylor,C.B., Bariola,P.A., DelCardayré,S.B., Raines,R.T., and Green,P.J. (1993). RN82: A senescence-associated RNase of Arabidopsis that diverged from the S-RNases before speciation. Proc. Natl. Acad. Sci. USA 90, 5118-5122. Theologis,A. (1998). Ethylene signalling: redundant receptors all have their say. Curr. Biol. 8, R875-R878. Valveken,D., Van Montagu,M., and Van Lijsebettens,M. (1988). Agrobacterium tumefaciens-mediated transformation of Arabidopsis thaliana root explants by using kanamycin selection. Proc. Natl. Acad. Sci. USA 85, 5536-5540. van Hoof,A. and Green,P.J. (1996). Premature nonsense codons decrease the stability of phytohemagglutinin mRNA in a position-dependent manner. Plant J. 10, 415-424. Verwoerd,T.C., Dekker,B.M., and Hoekema,A. (1989). A small-scale procedure for the rapid isolation of plant RNAS. Nucleic Acids Res 1 7, 2362. Zhang,S., Sheng,J., Liu,Y., and Mehdy,M.C. (1993). Fungal elicitor-induced bean proline-rich protein mRNA down- regulation is due to destabilization that is transcription - and translation dependent. Plant Cell 5, 1089-1099. Zhang,S. and Mehdy,M.C. (1994). Binding of a 50-kD protein to a U-rich sequence in an mRNA encoding a proline-rich protein that is destabilized by fungal elicitor. Plant Cell 6, 135-145. Zubiaga,A.M., Belasco,J.G., and Greenberg,M.E. (1995). The nonamer UUAUUUAUU is the key AU-rich sequence motif that mediates mRNA degradation. Mol. Cell. Biol. 15, 2219-2230. 78 CHAPTER 1-2 MUTANTS OF ARABIDOPSIS DEFECTIVE IN A SEQUENCE-SPECIFIC mRNA DEGRADATION PATHWAY This manuscript has been submitted for publication (Mark A. Johnson, Miguel A. Pérez- Amador, and Pamela J. Green). 79 ABSTRACT One of the ways a cell can rapidly and tightly regulate gene expression is to target specific mRNAs for rapid decay. A number of mRNA instability sequences that mediate rapid mRNA decay have been identified in various eukaryotes, but the nature of the cellular components that play critical roles in sequence-specific decay in vivo has been more difficult to establish. Here, we used a novel genetic strategy to isolate rare mutants of Arabidopsis that selectively elevate the abundance of mRNAs that contain the plant instability sequence DST. Analysis of these mutants identified two loci that are important for DST-mediated mRNA decay and provided a powerful inroad into the mRNA decay machinery. Similar strategies should be applicable to the in vivo analysis of sequence-specific mRNA decay in other eukaryotes. 80 The small-auxin-up-RNA (SA UR) transcripts are among the most unstable mRNAs in plants, having half-lives of 10—50 minutes (1). The SA UR genes have been implicated in auxin-induced cell elongation and their transcripts appear to be constitutively unstable so that their abundance can be rapidly altered in response to transcriptional control by the plant hormone auxin. All unstable SA UR transcripts contain conserved DST sequences, so called because of their location (_lowns_tream of the coding region (1). A dimer of the DST element (2) or the 3’ end of the Arabidopsis SA UR-A C1 gene, which contains a DST element (3), are sufficient to destabilize reporter transcripts in plant cells. The DST sequence is approximately 45 nucleotides in length and is comprised of three highly conserved domains separated by two variable regions. Mutagenesis studies have demonstrated that residues within two of the conserved domains are necessary for instability function (4). Our detailed understanding of the DST element, its robust effect on mRNA levels, and its potential biological importance make it an attractive model for understanding how mRNA instability sequences function in plants. Although genetic strategies have proven critical for the identification of components of the general decay machinery in yeast (5), they have rarely been applied to the study of rapid mRNA decay mediated by specific sequences. Most in vivo studies of the latter have been limited to overexpression of RNA binding proteins (6) or correlations between target mRNA levels and the presence of various RNA binding proteins (7). However, it was recently demonstrated that mice made deficient for tristetraprolin by gene disruption were defective in rapid degradation of tumor necrosis factor-0t mRNA mediated by an AU-rich element (ARE). It was later shown that tristetraprolin binds ARES in vitro (8). ARES are perhaps the most well studied class of mRNA instability 81 sequence in multi-cellular eukaryotes (9). Therefore, these results highlight how a single mutant can provide important insights about sequence-specific mRNA decay mechanisms in eukaryotes. We have devised an approach to isolate mutants of Arabidopsis thaliana that are defective in the sequence-specific mRNA decay pathway mediated by the DST element. Because this strategy facilitated the isolation of rare mutations, it provided us with a unique opportunity to study a sequence-specific mRNA degradation pathway in vivo. Our approach was based on a phenotype that we engineered using two DST-containing reporter genes that were introduced into Arabidopsis via Agrobacterium tumefaciens mediated transformation (10). These reporter genes were designed to allow selection of mutants that elevate expression of these genes due to defects in DST-mediated mRNA degradation. Plasmid 1519 (p1519), used to generate the parental line for mutagenesis, contained a tetramer of the DST element inserted into the 3’ UTR of the hygromycin phosphotransferase (HPH) and B—glucuronidase (GUS) genes (Figure 1-2-l-A). HPH confers resistance to hygromycin and GUS expression is readily detected by incubation of plant tissue with substrates that are cleaved to colorimetric or fluorescent products (1 1). We used a tetramer of the DST element because previous experiments had indicated that a greater decrease in reporter mRNA abundance was achieved relative to a dimer (12). This was beneficial because it provided a greater range of potential elevated mRNA abundance mutant phenotypes. p1493 lacks the DST sequence and p1493-2 plants served as a non- destabilized control throughout our experiments (Figure 1-2-1-A). In addition to the HPH and GUS reporter genes, the T-DNA contained the nptII (neomycin phosphotransferase) cassette so that kanamycin resistant transgenic plants could be selected. Transgenic line 82 p1519-31 was chosen for mutagenesis because segregation of kanamycin resistance and Southern blot analysis indicated that it harbored a Single insert of the T-DNA (13). In addition, the abundance of the HPH-DST and G US-DST messages in this line reflected the average abundance of these transcripts in a population of p1519 transforrnants (13). The location of the T-DNA insertion in p1519-31 was mapped to chromosome 11 near position 51 (14). The DST tetramer resulted in an approximately 10 fold decrease in steady-state HPH mRNA abundance in plants harboring p1519 compared to those transformed with p1493 (1 3). The half-life of the HPH transcript was reduced by approximately 3 fold in p1519-31 plants relative to p1493-2 as measured in an actinomycinD time-course conducted on 12-day-old seedlings grown in liquid culture (Figure l-2-l-B and l-C). The disparity between the effect of the DST sequence on the steady-state abundance and the half-life of the HPH transcript (10 fold vs. 3 fold) is most likely due to dampening that is commonly observed when general inhibitors of transcription are used to measure mRN A decay rates (15). 83 A p1519 p1493 exam-Ia R. €91 13cm 0' 30’ 60’ 90’ 120150: p1519-31(4) hph mRNA 0 Remaining .‘i‘fifi 91493‘21') 0.1 ; ; . ; T O 30 60 90 120150 Time (min) Figure 1-2-1. Structure of transgenes and kinetics of degradation of reporter transcripts used for dst mutant selection. (A) Diagrams representing the plant transformation vectors, p1519 and p1493 are Shown. p1519 was used to generate transgenic Arabidopsis plants that served as the parental line for the dst mutant selection. The cauliflower mosaic virus 358 promoter was used to control the expression of the HPH and GUS genes (28). Polyadenylation Signals were derived from the 3’ ends of the pea ribulose 1,5-bisphosphate carboxylase small subunit E9 (rch-E9, E9) and rch-3C (3C) genes. The HPH and GUS transcripts encoded by p1519 have been destabilized by a tetramer of the DST sequence (DSTx4). p1493 was used to generate transgenic plants that served as a non-destabilized control during the mutant selection and in subsequent experiments. (B) Comparison of the stability of HPH transcripts in parental seedlings (p1519-31) and non-destabilized, control seedlings (p1493-2). Seedlings were grown in liquid culture for 12 days. ActinomycinD (75 ug/ml) was added, samples were removed from the cultures, and frozen in liquid nitrogen at 30 minute intervals. Northern blot analysis was performed with 10 ug of total RNA extracted from each sample (29). The blot was hybridized with a radiolabeled probe complementary to the HPH transcript. The blot displaying the p1519-31 time course has been overexposed relative to the p1493-2 blot so that the Signal at all time points would be visible. (C) Signals obtained from Northern blots were quantitated by Phosphorlmager (Molecular Dynamics) and plotted on a semi- log scale. Linear regression analysis is Shown for each set of points. 84 Destabilization of the HPH transcript by the DST element resulted in reduced resistance to hygromycin in p1519-31 seedlings relative to p1493-2 seedlings, particularly at higher concentrations of the antibiotic (Figure 1-2-2-A). We hypothesized that mutants defective in DST-mediated mRNA degradation would have increased abundance of the HPH-DST transcript and would therefore have increased resistance to hygromycin relative to the parental line. A preconstruction of the mutant selection revealed that seedlings expressing the non-destabilized HPH message could be easily identified among a lawn of seedlings expressing the destabilized HPH transcript (Figure 1-2-2-B). Because we were interested in mutations affecting the DST pathway in trans, our goal was to identify mutants with increases in both HPH-DST and GUS-DST mRNA abundance. p1519-31 seeds were subjected to ethylmethane sulfonate (EMS) mutagenesis, sown on soil and the M1 plants were allowed to self-fertilize (Figure 1-2-2-C)(16). M2 seeds were collected from groups of MI plants and were plated on seed selection media containing 1000 rig/ml hygromycin (17). Out of ~730,000 M2 seeds that were plated, 323 seedlings that appeared to exhibit increased resistance to hygromycin were transferred to soil for propagation. By plating M3 progeny from these plants on selective media, we found that approximately 20% of the original putative mutants displayed heritable hygromycin resistance. In most of these lines, resistance was not attributed to an increase in HPH-DST mRNA abundance. Three mutants, selected from independent M1 groups, Showed a heritable increase in HPH-DST mRNA abundance; these three also Showed an increase in G US-DS T mRNA abundance. Two of these mutants, dstl and dst2 that were isolated first, have been characterized in greater detail. 85 A .-. ‘- C EMS mutagenesis of p1519-31 1000 r. ' * to generaie M1 seed 730,000 M2 seeds plated on liygomycin 323 M2 seedlings rescued 500 hygromycin (pg/ml) 00 O O 20% heritable hygromycin B resistance 1 Northern Blot Analysis 2 lines (dst1 and dsQ) with heritable increases in HPH-DST and GUS-DSTmRNA Figure 1-2-2. The dst selection strategy. (A) A titration of hygromycin resistance was performed on p1493-2 and p1519-31 seedlings. Approximately 500 seeds were surface sterilized and plated on primary seed selection media containing 500, 800, or 1000 rig/ml hygromycin (17). (B) In order to preconstruct the mutant selection, five p1493-2 seeds were mixed with 500 p1519-31 seeds and plated on primary seed selection media containing 1000 11ng hygromycin. All five resistant seedlings were recovered, one of which is shown. (C) Steps in the dst selection. The abundance of the HPH-DST and GUS-DST transcripts was 3 to 5 fold higher in dstl and dst2 compared to p1519-31 plants (Figure 1-2-3-A). In contrast, the abundance of the eIF4-A transcript, used as a non DST-containing control, was equivalent in mutants and p1519-31 (Figure 1-2-3-B). In addition, two other mRNAs, encoding potential messenger ribonucleases, that do not contain DST elements were equally abundant in p1519-31 and the mutants (13). Increased HPH-DST and GUS-DST mRNA abundance is a consistent phenotype that has been observed in the progeny of all backcrosses to p1519-31 (13). 86 Because both the HPH-DST and GUS-DS T transcripts were elevated in these mutants, it was highly unlikely that the phenotype was due to a mutation in one or more of the DST elements present in the 3’UTR of the HPH transcript. To examine this possibility directly, the DST tetramer was amplified by polymerase chain reaction from genomic DNA extracted from dst] and dst2 and was sequenced (1 8). The DST tetramers in both mutants were found to be unaltered. p1519-31 v- N S ’4? 2 ”-dHPH-DST T‘I -‘ 4 GUS-DST ... ......- 4 eIF4-A { p1519-31 dst 1 dst 2 4 SAUR-AC1 any , .11]. ' .' ”1' ' m from" tows ‘ 9’F4'A Figure 1-2-3. mRNA abundance of selected transcripts in dst] and dst2. Total RNA was prepared from leaves of 60-day-old plants and analyzed by Northern blot hybridization (29). dst] and dst2 plants derived from the first backcross to p1519-31 were used as the source of material in this experiment. Radiolabeled probes were prepared against the indicated transcripts and were used in sequential hybridizations of the same Northern blot. (A) The abundance of DST-containing reporter transcripts, HPH-DST and GUS-DST, are elevated in dst] and dstZ. The higher molecular weight GUS transcript (0) is a consistent characteristic of p15 1 9- 31. It is most likely the result of a partial recognition of a cryptic downstream polyadenylation signal that is activated by incomplete integration of the T-DNA. We have measured the abundance of the lower molecular weight band (the expected size) throughout these experiments, although the two bands are coordinately regulated in the mutants. (B) The abundance of an endogenous DST-containing message, SA UR-A C1, is elevated in dst] and dst2, whereas the abundance of the eIF 4A (27) transcript, which does not contain a DST element, is equivalent in the mutants and p1519-31. Genetic analysis indicated that dst] and dst2 are partially dominant mutations in independent single genes. To arrive at this conclusion, we examined the inheritance of the dst mutations by measuring HPH-DS T mRNA abundance in the rosette leaves of mature 87 plants derived fi'om crosses of dst] and dst2 to p1519-31 by Northern blot hybridization. We used Northern blots rather than relying on reporter protein activity because direct measurements were much more sensitive to the relatively small differences in mRNA abundance observed in the mutants. Progeny of the second backcross to p1519-31 (F.) and the progeny of self-fertilizations of these plants (F2) were used in these studies. Ten F. plants were examined for each mutant. HPH-DST mRN A abundance was on average 2.1 i 0.22 fold higher in dst] F. and 2.8 i 0.23 fold higher in dst2 F. plants compared to p1519- 31 (19). These levels of HPH-DST mRNA abundance are intermediate between p1519-31 and mutant levels as would be expected if heterozygotes Showed a partial mutant phenotype. Segregation of increased HPH-DST mRNA abundance in the F2 populations was also consistent with partial dominance. Of 46 F2 plants segregating dst], nine were observed with high HPH-DST mRNA abundance (3.1 i 0.17 fold p1519-31), 23 had intermediate HPH-DST mRNA abundance (1.7 i 0.07) and 14 showed HPH-DST mRNA abundance that was similar to pl 519-31 (0.9 i 0.04). Segregation of the dstZ mutant phenotype was similar to that of dst]. Of 53 F2 plants analyzed, 13 fell into the mutant class (4.0 i 0.16), 28 fell into the intermediate class (2.2 i 0.1 1), and 12 were similar to p1519-31 (1.0 i 0.04). Segregation in dst] and dstZ F2 populations was most easily explained by a ratio of 1:2:1 (wild-type(wt):intermediatezmutant) as would be expected for inheritance of a partially dominant Single gene (dstl x2: 1.1; dst2 x2: 0.2; p > 0.5). Crossing dst] with dst2 yielded F. plants with intermediate HPH-DST mRNA abundance, indicating that dst] and dst2 are not allelic. When the F2 plants from this cross were analyzed, 10/51 showed HPH-DST mRNA abundance similar to p1519-31. Had the dst mutations been in the same gene, or if they were tightly linked, we would not have 88 expected to observe any F2 plants with HPH-DST mRNA abundance similar to p1519-31. No class of F2 plants was observed with an additive effect on HPH-DST mRN A abundance, as might be expected if dst] and dst2 function independently to destabilize DST-containing messages (20). In addition, the finding that F. plants from the dst] x dst2 cross showed an intermediate mRNA abundance phenotype indicated a lack of additive interaction between these loci because the individual mutants crossed to p1519-31 showed the same phenotype in the F.. In order to begin the characterization of the molecular basis of these mutant phenotypes, the dst] locus was mapped using molecular markers. This analysis showed that dst] is located near position 68 on chromosome 5 (14). If dst] and dst2 are indeed mutations that disrupt the DST-mediated mRN A degradation pathway in trans, one would expect the abundance of endogenous DST- containing messages to be elevated as well. We analyzed the abundance of the SA UR-A C1 transcript, the only DST-containing message in Arabidopsis that has been exarrrined at the level of mRNA stability (3). The abundance of this message is elevated in dst] and dst2, indicating that these mutations may elevate the abundance of any message destabilized by DST sequences (Figure l-2-3-B). The abundance of messages that lack DST sequences was unchanged in these mutants; for example, the abundance of the eIF4-A transcript is equivalent in the mutants and p1519-31 (Figure 1-2-3-B). In addition, initial DNA microarray experiments indicated, as expected, that most transcripts were not elevated in dstI compared to p1519-31; however, a new SA UR mRNA, containing a potential DST element, was among those that were more abundant in dst] (21). Taken together, these 89 observations suggest that the dst mutations Specifically affect the abundance of transcripts containing DST instability sequences. The Simplest explanation for the elevation of DST-containing transcripts in dst] and dst2 is a defect that results in a decreased rate of DST-mediated mRNA degradation. This is likely because the DST sequence is the only regulatory element Shared by the four transcripts that were elevated in the mutants. To test this hypothesis, we measured the kinetics of HPH-DST mRNA degradation in p1519-31, dst], and dst2 plants. We anticipated that the measurable differences in HPH-DST mRNA decay rates would be minimal because the difference in HPH-DST abundance between dst] or dst2 and p1519- 31 was only about three fold (Figure 1-2-3—A) and such differences are ofien dampened by time course analysis using general transcription inhibitors, as discussed above (15). Figure 1-2-4-A shows representative Northern blots from these experiments with the average, normalized HPH-DST mRNA abundance at each time point plotted on a semi-log scale in Figure 1-2-4-B. The average half-life of the HPH-DST transcript was found to be 24 minutes for dst], 23 minutes for dst2, and 18 minutes for p1519-31 (Figure 1-2-4-B). The differences between mutant and p1519-31 mRN A decay curves are particularly evident at later time points (Figure 1-2—4-B). We conclude that dst] and dst2 cause an increase in message stability that results in increased abundance of DST- containing mRNAs. 90 0’ 15’ 30’ 45’ 60’ A B ” “ p1519-31 g’ -: . .E '2 g 0 p1519-31 (t1,2 = 18’) “ m ’87" 03.21 at $2! dsa E U d3t1 (“,2 = 24’) I t A dst2(t1,2 = 23’) 0.01 - . . . 0 15 30 45 60 time (min) Figure 1-2-4. Analysis of HPH-DST mRNA stability in leaves from p1519-31, dst], and dst2 plants. Rosette leaves from 12 F3 plants derived from the second backcross of dst] and dst2 to p1519-31 were harvested. p1519-31 rosette leaves were harvested from 12 plants. Leaves were incubated in buffer for 30 minutes before addition of 150 rig/ml cordycepin (30). Four leaves were harvested every 15 minutes for 60 minutes and frozen in liquid nitrogen. Total RNA was extracted from each sample and analyzed by Northern blot hybridization (29). Blots were hybridized with a radiolabeled probe complementary to the HPH-DS T transcript and radioactive signals were quantititated by Phosphorlmager (Molecular Dynamics). HPH-DST Si gnals were normalized by rehybridizing the blots with a probe complementary to the eIF 4A transcript (27). (A) Representative Northern blots Showing the HPH-DST signal over the 60-minute time-course in p1519-31, dst], and dst2. (B) Average normalized HPH-DST signals (i standard error of the mean) from the three time-courses are plotted on a semi-log scale. The average half-life of the HPH- DST transcript is given on the graph. Results are expressed as the relative amount of HPH-DST mRNA remaining at each time point after normalization with the eIF 4A Si gnal compared to the first time point. Neither of the dst mutants had any obvious defects in development or morphology. Because expression of an auxin responsive gene, SA UR-A C], was altered in these mutants, we analyzed phenotypes associated with this hormone. dst] and dst2 displayed normal root gravitropism as indicated by the wavy root assay or by rotating vertical plates following germination of seedlings (22). Additionally, there was no loss or enhancement of apical 91 dominance in these plants, nor were there alterations in root length or morphology, other phenotypes that have been associated with auxin response mutants (23). The SA UR gene family is large in Arabidopsis and, interestingly, there are several family members that are highly similar in the coding region to SA UR-A C 1 yet lack DST elements (24). Therefore, it is not surprising that a modest increase in mRN A abundance of one or a few of these genes does not cause any obvious altered phenotypes. Although the number of endogenous targets of the DST-mediated decay pathway is unknown, the dst mutants should be powerful tools to address this question. For example, DNA microarray analysis Should allow additional targets of the pathway to be identified by revealing mRNAs that accumulate preferentially in either dst] or dst2. The dst mutant selection had two interesting characteristics. First, these mutants were extremely rare having been isolated at a frequency of less than l/200,000. In Arabidopsis, null mutations in nonessential genes are typically expected at frequencies of ~1/2000 in EMS populations (25). Second, the dst mutations did not fully restore the abundance of the HPH or GUS transcripts to the levels found in plants with non-destabilized transcripts. These observations indicate that dst] and dst2 may be relatively weak alleles that allow partial function of the DST-mediated mRNA decay pathway. If proper regulation of DST-containing mRNAs is an essential function for Arabidopsis then this may have precluded the isolation of stronger alleles. Altematively, functional redundancy among components that recognize instability determinants may also contribute to the low frequency of sequence-specific decay mutants. Perhaps it is by necessity rather than coincidence that both dst mutants are partially dominant. Whether the corresponding genes encode DST-binding proteins, DST-specific ribonucleases or other effectors, 92 dominant negative interactions between mutant derivatives and redundant proteins or other components of the decay machinery are easy to imagine. Similar obstacles may explain why successful selections for sequence-specific mRNA decay mutants have yet to be reported in other eukaryotic systems. Based on our work, selections for mutants defective in rapid mRNA degradation mediated by Specific sequence elements should be feasible in other eukaryotic model organisms if selectable marker genes that facilitate detection of small expression differences can be engineered. The potential impact of this approach is demonstrated by a series of very informative mutants of Saccharomyces cerevisiae and Caenorhabditis elegans that are defective in a general mRNA decay pathway responsible for the degradation of transcripts with premature stop codons (26). Finally, by providing the means to clone dst genes, this work opens a new avenue to elucidate the molecular machinery responsible for rapid sequence-Specific mRNA degradation in plants and will likely provide novel information applicable to other eukaryotes. 93 REFERENCES AND NOTES 1. B. A. McClure and T. J. Guilfoyle, Science 243, 91(1989); C. S. Brown, M. A. Gee, T. J. Guilfoyle, Plant Cell 1, 229 (1989); A. R. Franco, M. A. Gee, T. J. Guilfoyle, J.Biol. Chem. 265, 15845 (1990); B. A. McClure, G. Hagen,; K. T. Yamamoto, H. Mori, H. Imaseki, Plant Cell Physiol. 33, 93 (1992); P. Gil et al., Plant Physiol. 104, 777 (1994) 2. T. C. Newman, M. Ohme—Takagi, C. B. Taylor, P. J. Green, Plant Cell 5, 701 (1993). 3. P. Gil and P. J. Green, EMBOJ. 15, 1678 (1996). 4. M. L. Sullivan and P. J. Green, RNA 2, 308 (1996). 5. G. Caponigro and R. Parker, MicrobiolRev. 60, 233 (1996). 6. R. G. Jain, L. G. Andrews, K. M. McGowan, P. H. Pekala, J. D. Keene, M01. Cell.Biol. 17, 954 (1997); X. C. Fan and J. A. Steitz, EMBOJ. 17, 3448 (1998); S. S. Peng, C. Y. Chen, N. Xu, A. B. Shyu, EMBOJ. 17, 3461 (1998). 7. P. Loflin, C. Y. A. Chen, A. B. Shyu, Genes and Development 13, 1884 (1999); P. R. Bohjanen, B. Petryniak, C. H. June, C. B. Thompson, T. Lindsten, Mol. Cell Biol. 11, 3288 (1991). 8. W. S. Lai et al., Mol. Cell. Biol. 19, 4311 (1999); E. Carballo, W. S. Lai, P. J. Blackshear, Science, 281, 1001 (1998); G. A. Taylor et al., Immunity, 4, 445 (1996). 9. C.-Y. A. Chen and A.-B. Shyu, Trends BiochemSci. 20, 465 (1995). 10. Constructs were introduced into Arabidopsis using the vacuum infiltration method of Agrobacterium tumefaciens mediated transformation [N. Bechtold, J. Ellis, G. Pelletier, C.R.Acad.Sci.Paris 316, l 194 (1993) and Web site: http://www.bch.msu.edu/pamgreen/vac.htm]. To avoid any hygromycin resistant contamination of pl 5 19-31 during large-scale selection experiments, this construct was introduced into a mutant that lacks trichomes (Col-PRL, glI). p1493 and other constructs that confer hygromycin resistance were introduced into wt Columbia plants (Col-0). 11. L. Gritz and J. Davies, Gene 25, 179 (1983); R. A. Jefferson, T. A. Kavanagh, M. A. Bevan, EMBOJ. 6, 3901 (1987). 12. M. L. Sullivan, D. M. Thompson, P. J. Green, unpublished data. 13. M. A. Johnson and P. J. Green, unpublished data. 14. Mapping of the p1519-31 transgene locus and the dst] locus were carried out using SSLP markers [C. J. Bell and J. R. Ecker, Genomics 19, 137 (1994) and at the Web Site 94 http://genome.bio.upenn.edu/SSLP_info/SSLP.html]. F3 families derived from a cross between p1519-31 and Landsberg erecta (glI-I) [M. Koomeef, L.W. Dellaert, J .H. van der Veen, Mutation Research/, 93, 109 (1982)] were scored for kanamycin resistance. 15 families were identified that showed 100% kanamycin resistance. These were used to map the transgene locus and were scored for 7 markers Spread across the genome. Linkage was detected only to markers on chromosome 11, the closest marker being ngal 126 (position 5 l , 0 recombinants out of 30 chromosomes analyzed). To map the dst] locus, leaves of F2 plants from a cross of a'stI with Landsberg erecta (glI-I) were scored for 3 fold or higher HPH-DST mRNA abundance compared to p1519-31 by Northern blot analysis (19). 30 F2 plants were identified and scored for 18 markers spread across the genome. Linkage was again detected with nga1126 (0 recombinants out of 16 chromosomes analyzed) due to the necessity of the p1519 transgene for scoring the mutant phenotype. In addition, linkage was detected to nga139 (map position 51, 2 recombinants out of 56 chromosomes analyzed) and nga76 (map position 68, 0 recombinants out of 56 chromosomes analyzed), defining the location of dst] on chromosome 5. 15. A.-B. Shyu, M. E. Greenberg, J. G. Belasco, Genes Dev. 3, 60 (1989); J. G. Belasco and G. Brawerman, in Control of messenger RNA stability, J. Belasco and G. Brawerman, Eds. (Academic Press, San Diego,1993), ch. 18; S. Zhang, J. Sheng, Y. Liu, M. C. Mehdy, Plant Cell 5, 1089 (1993); G. Caponigro and R. Parker, Microbiol.Rev. 60, 233 (1996); H. Holtorf, H. Sch6b, C. Kunz, R. Waldvogel, F. Meins, Jr., Plant Cell 11, 471 (1999) 16. Seeds homozygous at the p1519-31 locus (fourth generation of self-fertilized progeny following transformation) were soaked for 16 hours in a 0.3% (v/v, water) solution of EMS. Following mutagenesis, seeds were washed extensively with water over a 12 hour period and then sown on soil. The effectiveness of EMS mutagenesis was evaluated by scoring the occurrence of albinism and embryo lethality in M. plants as described in G. P. Redei and C. Koncz, in Methods in Arabidopsis Research, C. Koncz, N.-H. Chua, J. Shell, Eds. (World Scientific Publishing, Singapore, 1992) pp.16-82. l7. Seed selection media consisted of 4.3 g/ L Murashige and Skoog salt mixture (GibcoBRL), BS vitamin mixture (100 mg/l myo-inositol, 10 mg/l thiamine hydrochloride, 1 mg/l nicotinic acid, 1 mg/l pyriodoxine), 1% sucrose, 0.5 g/L MES at pH 5.7, 0.8% phytagar, 1000pg/m1 hygromycin and 50 ug/ml kanamycin. l8. Genomic DNA was prepared from dst] and dst2 F2 plants from the first backcross to p1519-31 as described in M. A. Saghai-Maroof, K. M. Soliman, R. A. Jorgensen, R. W. Allard, Proc. Natl. Acad. Sci. 81, 8014 (1984). PCR primers complementary to the 3’ end of the HPH coding region and the 5’ end of the E9 3’UTR were used to amplify the DST elements with a polymerase with proof-reading capability (Pfu, Stratagene) under standard PCR cycling conditions. Products were ligated into a derivative of pBluescript SKII(-) and multiple individual clones were sequenced for each reaction. 95 19. Signals from Northern blots were quantitated with a Phosphorlmager (Molecular Dynamics). The HPH-DS T Signal was normalized using the elF 4A (27 ) signal. p1519- 31 samples were analyzed in parallel and the average fold increase of segregants compared to p1519-31 HPH-DST mRNA abundance is reported i the standard error of the mean. RNA was extracted from two rosette leaves of individual plants as described by T. C. Verwoerd, B. M. Dekker, A. Hoekema, Nucleic Acids Res 17, 2362 (1989). Northern blots were processed as described previously (2). 20. In the F2 of a cross between two partially dominant mutations with no additive effect on one another, such as is proposed here, one would expect a ratio of 1:8:7 (wt:Intermediate:Mutant) or 12826 if double mutants are not viable. We observed 10 wt, 26 intermediate, and 15 mutant plants out of 51 F2 plants that were tested. This gives a ratio of ~3:8:5, which does not precisely fit simple models of inheritance. Because it is clear that dst] and dst2 are partially dominant mutations in single genes, this may indicate subtle interactions between these loci. For example, dst! and dst2 may partially suppress one another, explaining the higher than expected proportion of wt segregants. 21. M. A. Perez-Amador, M. A. Johnson, J. Landgraf, E. Wisman, P. J. Green, unpublished data. 22. C. J. Bell and E. P. Maher, MGG 220, 289 (1990); K. Okada and Y. Shimura, Aust.J.Plant Physiol. 19, 439 (1992). 23. W. Boerjan et al., Plant Cell 7, 1405 (1995); L. Hobbie and M. Estelle, Plant Cell Environ. 17, 525 (1994); H. M. O. Leyser, F. B. Pickett, S. Dharrnasiri, M. Estelle, Plant J. 10, 403 (1996). 24. T. J. Guilfoyle, personal communication; M. A. Johnson and P. J. Green, unpublished sequence database search results. 25. G. Haughn and C. R. Somerville, in Agricultural Chemistry, H. M. LeBaron, R. O. Mumma, R. C. Honeycutt, J. H. Duesing, Eds. (American Chemical Society, Washington, DC, 1987) p. 98. 26. P. Leeds, S. W. Peltz, A. Jacobson, M. R. Culbertson, Genes Dev. 5, 2303 (1991); R. Pulak and P. Anderson, Genes Dev. 7, 1885 (1993); Y. Cui, K. W. Hagan, S. Zhang, S. W. Peltz, Genes Dev. 9, 423 (1995); F. He and A. Jacobson, Genes Dev. 9, 437 (1995); B. S. Lee and M. R. Culbertson, Proc.Natl.Acad.Sci. USA 92, 10354 (1995); B. M. Cali, S. L. Kuchma, J. Latham, P. Anderson, Genetics 151, 605 (1999). 27. The eukaryotic translation initiation factor 4A (eIF4-A) gene was cloned from Arabidopsis by C. B. Taylor, P. A. Bariola, S. B. DelCardayre', R. T. Raines, P. J. Green, Proc.Natl.Acad.Sci. USA 90, 51 18 (1993). 28. p1519 and p1493 are derivatives of pMON505-70 [R.-X. Fang, F. Nagy, S. Sivasubramaniam, N.-H. Chua, Plant Cell 1, 141 (1989)]. The HPH coding region was obtained as a BamHI fragment of pLG90 [L. Gritz and J. Davies, Gene 25, 179 (1983)]. 96 Use of the 35S promoter, polyadenylation Signals from the 3’ ends of the E9 and 3C genes, and synthesis and insertion of the DST tetramer were as described previously (2). 29. RNA extraction and Northern blot analysis was performed essentially as described by T. C. Newman, M. Ohme-Takagi, C. B. Taylor, P. J. Green, Plant Cell 5, 701 (1993). 30. Cordycepin (150 ug/ml) was used to inhibit transcription in mature leaves because actinomycinD was less effective in this tissue (13). This protocol, including the leaf incubation medium (lmM Pipes, pH 6.26; lmM sodium citrate; lmM KCl; 15 mM sucrose) was adapted from previous reports [K. A. Seeley, D. H. Byrne, J. T. Colbert, Plant Cell 4, 29 (1992); H. Holtorf, H. Schéb, C. Kunz, R. Waldvogel, F. Meins, Jr., Plant Cell 11, 471 (1999)]. 31. We thank Dr. Ambro van Hoof, Dr. Michael Thomashow and James Kastenmayer for reading the manuscript and Dr. Joanne Chory for providing pLG90. We also thank Dr. Michael Sullivan and Debrah Thompson for construction of pl 519 and p1493; and Jonathan Vogel and Linda Danhof for technical assistance. This work was funded by grants from the United States Department of Energy, the United States Department of Agriculture, and the McKnight Foundation to P.J.G. M.A. P-A. received postdoctoral fellowships from NATO-Spain and from the Ministerio de Educacion y Ciencia, Spain. 97 CHAPTER 1-3 ANALYSIS OF THE DST ELEMENT AND DST-LIKE SUBDOMAIN S WITHIN THE SA UR-A C1 3’ UNTRANSLATED REGION 98 INTRODUCTION The DST element is perhaps the best-studied instability determinant that is plant- specific. The demonstration that this sequence is sufficient to destabilize messenger RNA (mRNA) came when Newman et al. showed that a dimer of the element, when inserted into the 3’ untranslated region (UTR), targeted otherwise stable reporter RNAS for rapid turnover (1993). The DST element was originally found in the 3’ UTR of the unstable small-auxin-up-RNA (SA UR) genes and has been found in the 3’UTR of several SA URs from diverse plant Species (McClure et al., 1989; Yamamoto et al., 1992; Gil et al., 1994). The element consists of three conserved subdomains separated by two variable regions. These subdomains are referred to as GGA, ATAGAT, and GTA after the DNA sequence that comprises the core of each conserved region. Sullivan and Green showed that substitution mutations in either the ATAGAT or GTA subdomains abolished the effect of the DST dimer (1996), demonstrating the necessity of these sequences for DST function. The experiments described above explored the firnction of a synthetic DST sequence derived from the soybean SA UR-15A gene (McClure et al., 1989). In order to study a DST element in its native context, Gil and Green cloned the SA UR-A C1 gene from Arabidopsis (Gil et al, 1994). The 3’UTR of this gene is sufficient to target reporter transcripts for rapid turnover, reducing the half-life of the human ,B-globin (Globin) transcript to a Similar extent as did a dimer of the DST element (Newman et al., 1993; Gil and Green, 1996). While these two experiments are not directly comparable because the half-life measurements were done using different methods, this result raises some 99 interesting questions because the SA UR-A C1 3 ’ UTR contains only one canonical DST element. Previous experiments indicated that a monomer of the DST element was not sufficient to cause rapid turnover of reporter transcripts (Newman et al., 1993) and experiments reported in Chapter 1-1 of this thesis indicate that a DST tetramer is more effective as an instability determinant than a dimer. These results would support the hypothesis that DST elements serve as recognition sites for the mRNA degradation machinery and become more and more effective as additional copies are added, with a minimum of two copies required for function. The function of the SA UR-A CI 3’UTR as an instability determinant, which contains one DST element, would appear to contradict this assertion. To test the contribution of the DST element to SA UR-A C1 3’UTR instability function, Gil and Green made substitution mutations in the ATAGAT and GTA subdomains (unpublished data). These mutations, tested individually, reduced the abundance of reporter transcripts to the same extent as did the wild type SA UR-A C1 3’UTR, indicating that individually, they are not required for SA UR-A C1 3’UTR instability function. Interestingly, upon analysis of the sequence of the SA UR-A C1 3’UTR upstream of the DST element, several potentially redundant ATAGAT- and GTA- like subdomains were noted. This raised the possibility that perhaps these redundant elements were contributing to the instability fiinction of the SA UR-A C1 3’UTR and were providing the additional DST sequences that experiments with synthetic elements had Shown to be necessary. The purpose of the experiments outlined in this chapter was to analyze the contribution of these ATAGAT and GTA-like subdomains and to enhance our understanding of DST function by studying the element in its native context. Two 100 approaches were taken. First, all DST and DST-like subdomains were replaced with the same substitutions that had been shown to abolish the function of the DST dimer (Sullivan and Green, 1996). In the second approach, the SA UR-ACI 3’UTR was divided into three regions and each was tested for instability function. Both of these approaches were somewhat risky because the DST element overlaps polyadenylation Signals within the SA UR-A C I 3 ’UTR. We expected that chimeric transcripts might be aberrantly polyadenylated, however, because so little is known about polyadenylation signals in plants, there was no way to know for certain until we analyzed the constructs in transgenic plants. In some cases, this limitation has made the results of these experiments difficult to interpret. However, this study suggests a possible role for upstream redundant ATAGAT and GTA-like subdomains in the instability function of the SA UR-ACI 3’ UTR and provides some interesting tools for future analysis of DST-mediated mRNA degradation. RESULTS Mutagenesis of all DST and DST-like subdomains The 3’UTR of the SA UR-A C1 transcript contains one DST element along with other DST-like subdomains that are iterated upstream. In Figure 1-3-1, the DST element, ATAGAT-, and GTA-like subdomains have been highlighted. The ATAGAT subdomain consists of the sequence ATAGAT and the GTA subdomain consists of the sequence 101 CAATGCQE, ATAGAT-like subdomains were defined as having one mismatch and GTA-like subdomains were defined as having one or two mismatches. In order to assess the contribution of the DST element and ATAGAT and GTA- like subdomains to SA UR-A C1 instability function, all of the residues that are highlighted in Figure 1-3-1 were mutated. ATAGAT and ATAGAT-like subdomains were replaced with GCATGC, changing 5 of the Six conserved residues. GTA subdomains (CAATGCGTA) were replaced with CATAGGCCT which changes the GTA residues, conserved in all SA UR DST elements, and three residues upstream. Either of these mutations iS known to abolish the function of the synthetic DST dimer (Sullivan and Green, 1996). I! AGTAC TATA CTACAACATTTC CA TA'ITT'ITI'ITAG AéI'I'G TTA GCTAATTTC cc crc. GAGATAA'lTG TAAATTGTITCAATG AG I :54»: '45:? . DST element TTGCATGTTAAAAA '~' ~ - r" ...m..- GTA-like sequence mew-«3.4:.» ATAGAT-like sequence Figure 1-3-1. The SA UR-A C1 3’UTR contains one DST-element and DST-like subdomains. The sequence of the SA UR-A C1 3’UTR immediately 3’ of the stop codon to the beginning of the poly(A) tail is shown (Gil et al., 1994). The DST subdomains (“GGA”, “ATAGAT”, and “GTA”, are highlighted by gray boxes. DST-like subdomains are highlighted by bars and arrows as described in the figure. The fully mutated SA UR-A C1 3’ UTR was constructed by polymerase chain reaction using four overlapping oligonucleotides. These mutations altered 29% of the nucleotides in the SA UR-A C1 3’UTR, changing the GC content from 29% to 45%. The fully mutated 3 ’UTR was inserted 3 ’ of the Globin coding sequence and the resulting 102 chimeric gene was introduced into a binary vector containing a GUS reference gene so that the effect on message stability could be assessed in transgenic plants (Figure 1-3-2). Control constructs containing the wild type SA UR-A C I 3’UTR and the E9 3’UTR, were also generated. -[Es>| Globin I 59] , T t @I Globin [SAUR] es Reference Figure 1-3-2. The structure of chimeric genes used to test mutations in the SA UR-ACI 3’UTR. The cauliflower mosaic virus 35$ promoter (35S) was used to control the transcription of test and reference genes. The reference gene, B-glucuronidase (GUS), was included in the same binary vector as the test gene to facilitate normalization of mRNA abundance data. The 3 ’UTRS of the pea ribulose 1,5-bisphosphate carboxylase small subunit E9 (E9) and 3C (3C) genes were used as a negative control (lacking instability activity) and as a 3’UTR for the reference gene, respectively. SA UR represents the wild type SA UR-A C1 3’UTR, while saur (shaded) represents the fully mutated version. The right border (RB) of the transfer-DNA (T-DNA) has been indicated by a black rectangle. The T-DNA also includes the neomycin phosphotransferase cassette so that kanamycin resistant transgenic plants could be selected. These constructs were introduced into Arabidopsis via the vacuum infiltration method of Agrobacterium tumefaciens mediated transformation (Bechtold et al., 1994). Approximately 100 independent, 12-day-old primary transforrnants (T.) from each construct were pooled and harvested for Northern blot analysis (Figure 1-3-3). 103 Figure 1-3-3. Mutations in the SA UR-ACI 3’ UTR affect polyadenylation of the Globin reporter transcript. A Northern blot containing 10 pg of total RNA from pools of T. seedlings was he < Globin hybridized sequentially with radiolabelled Globin ' A i and GUS probes. The 3’UTR of each Globin 1 2 3 reporter is given above each lane (E9, wild type SA UR-A C1, fully mutated SA UR-A C1). MMM I Globin I a has I es -Es>rclobinl o I ab E355>I GUS I acJ-m Reference Figure 1-3-4. Structure of chimeric genes used to test individual subdomains within the SA UR-A C1 3’UTR for instability function. A. The SA UR-A C1 gene structure is depicted with the 3’UTR divided into sections called N, R, and D as described in the text. B. Constructs used in this study were derivatives of those described in Figure 1-3-2. Transgenic plants expressing these constructs were tested for instability function using the same method described above. Figure 1-3-5 Shows the results of Northern blot analysis of RNA isolated from pools of T. seedlings. Pools of seedlings expressing the Globin reporter with the E9 3’UTR (Globin-E9) alone were included in the analysis as a negative control. Insertion of the N, R, and D regions resulted in an increase of the Globin transcript size that was expected for each (Figure 1-3-5-A). By comparing the ratios of Globin/ GUS mRNA accumulation presented in Figure 1-3-5-B, it was found that the N region, had almost no effect on message accumulation in transgenic seedlings. In contrast, the D region had a mild effect (71% of E9) and the R region had a large effect on the accumulation of the Globin reporter mRNA, reducing the abundance to 37% compared to plants expressing Globin with the E9 3’UTR alone. 106 A B 1.2- 3 3 3 (0 921, 37% 71% IBM“ I I I a z 0: o 3 g 0.9- EIBlot2 I M “U 3' exoribonuclease degradation of an unstable chloroplast mRNA. Plant J. 13, 85-96. Drager,R.G., Higgs,D.C., Kindle,K.L., and Stern,D.B. (1999). 5' to 3' exoribonucleolytic activity is a normal component of chloroplast mRNA decay pathways. Plant Journal 19, 521-531. Dunckley,T. and Parker,R. (1999). The DCP2 protein is required for mRNA decapping in saccharomyces cerevisiae and contains a functional MutT motif. EMBO J. 18, 5411- 5422. Fang,R.-X., Nagy,F., Sivasubramaniam,S., and Chua,N.-H. (1989). Multiple cis regulatory elements for maximal expression of the cauliflower mosaic virus 358 promoter in transgenic plants. Plant Cell I, 141-150. Gera,J.F. and Baker,E.J. (1998). Deadenylation-dependent and -independent decay pathways for al-tubulin mRNA in Chlamydomonas reinhardtii. Mol. Cell. Biol. 18, 1498-1505. 135 Gil,P. and Green,P.J. (1996). Multiple regions of the Arabidopsis SA UR-ACI gene control trascript abundance: the 3' untranslated region functions as an mRNA instability determinant. EMBO J. 15, 1678-1686. Higgs,D.C. and Colbert,J.T. (1994). Oat phytochrome A mRNA degradation appears to occur via two distinct pathways. Plant Cell 6, 1007-1019. Koncz,C. and Schell,J. (1986). The promoter of TL-DNA gene 5 controls the tissue- specific expression of chimaeric genes carried by a novel type of Agrobacterium binary vector. Mol. Gen. Genet. 204, 383-396. LaGrandeur,T.E. and Parker,R. (1998). Isolation and characterization of Dcplp, the yeast mRNA decapping enzyme. EMBO J. 17, 1487-1496. Muhlrad,D., Decker,C.J., and Parker,R. (1994). Deadenylation of the unstable mRNA encoded by the yeast MFA2 gene leads to decapping followed by 5'-->3' digestion of the transcript. Genes Dev. 8, 855-866. Newman,T.C., Ohme-Takagi,M., Taylor,C.B., and Green,P.J. (1993). DST sequences, highly conserved among plant SA UR genes, target reporter transcripts for rapid decay in tobacco. Plant Cell 5, 701-714. Ohme—Takagi,M., Taylor,C.B., Newman,T.C., and Green,P.J. (1993). The effect of sequences with high AU content on mRNA stability in tobacco. Proc. Natl. Acad. Sci. USA 90,11811-11815. Puissant,C. and Houdebine,L.-M. (1990). An improvement of the Single-step method of RNA isolation by acid guanidinium thiocyanate-phenoI-chlorophorm extraction. BioTechniques 8, 148-149. Rogers,S.G., Klee,H.J., Horsch,R.B., and Fraley,R.T. (1987). Improved vectors for plant transformation: expression cassette vectors and new selectable markers. Methods Enzymol. 153, 253-277. Sambrook,J., Fritsch,E.F., and Maniatis,T. (1989). Molecular Cloning: A Laboratory Manual. (Cold Spring Harbor, New York: Cold Spring Harbor Laboratory Press). Stevens,A. (1978). An exoribonuclease from Saccharomyces cerevisiae: effect of modifications of 5' end groups on the hydrolysis of substrates to 5' mononucleotides. Biochem. Biophys. Res Commun. 81, 656-661. Stoeckle,M.Y. (1992). Removal of a 3' non-coding sequence is an initial step in degradation of groa mRNA and is regulated by interleukin- 1. Nucleic Acids Res. 20, 1123-1127. Thompson,D.M., Tanzer,M.M., and Meagher,R.B. (1992). Degradation products of the mRNA encoding the small subunit of ribulose-1,5-bisphosphate carboxylase in soybean and transgenic petunia. Plant Cell 4, 47-58. 136 Vreken,P. and Raue’,H.A. (1992). The rate—limiting step in yeast PGKI mRNA degradation is an endonucleolytic cleavage in the 3'-terminal part of the coding region. Mol. Cell. Biol. 12, 2986-2996. 137 CHAPTER 2-2 CLONING AND INITIAL CHARACTERIZATION OF A POLY(A) RIBONUCLEASE FROM ARABIDOPSIS 138 INTRODUCTION With few exceptions, the poly(A) tail is common to all eukaryotic messenger RNA (mRNA) molecules. There is also recent evidence that this structure is present on many bacterial RNAS (reviewed in Carpousis et al., 1999). The poly(A) tail appears to be involved in controlling mRNA stability in prokaryotes and eukaryotes; in prokaryotes it is thought to be a signal for rapid mRNA degradation while in eukaryotes it protects mRNA from degradation (reviewed in Jacobson, 1996). The poly(A) tail is also an enhancer of translation in eukaryotes (reviewed in Gallie, 1998). The role of the poly(A) tail in controlling mRNA function has made it a target for developmental regulation. For example, in Xenopus, poly(A) tails are added, or removed to promote or inhibit expression of certain genes during oocyte maturation (Fox and Wickens, 1990; Vamum et al., 1992; Sheets et al., 1995). The poly(A) tail is also a point of post-transcriptional control of gene expression in general, and in yeast it has been found that deadenylation triggers the degradation of the majority of cytoplasmic messages (reviewed in Caponigro and Parker, 1996). The major yeast mRNA degradation pathway is thought to begin with deadenylation which is followed by cleavage of the of the 7-methyl guanosine cap (decapping) and eventual degradation of the body of the mRNA by a 5’ to 3 ’ exoribonuclease. These three steps are dependent upon one another such that decapping cannot occur without deadenylation and 5’ to 3’ degradation will not occur without prior decapping (reviewed in Caponigro and Parker, 1996). Nucleases that are responsible for decapping and degradation of the body of the transcript have been cloned and their role in 139 the major yeast mRNA degradation pathway has been demonstrated genetically and biochemically (Muhlrad et al., 1994; Beelman et al., 1996; Legrandeur and Parker, 1998; Dunckley and Parker, 1999). In contrast, the enzyme(s) responsible for complete degradation of the poly(A) tail has not been identified. Two candidates for this activity have been studied in yeast. First, a poly(A) binding-protein (PAB)-dependent poly(A) nuclease was cloned, however, recently it was Shown that this activity is most likely responsible for trimming poly(A) tails to their mature length (Brown and Sachs, 1998). The exosome, an additional candidate, is a recently described complex of ~10 3’ to 5’ exoribonucleases involved in degradation of yeast mRNA, but the substrate for this complex is thought to be deadenylated mRNA (reviewed in van Hoof and Parker, 1999). Therefore the deadenylating nuclease remains unknown in yeast. A 3’ to 5’ exoribonuclease that preferentially degrades poly(A) was purified from calf thymus (Komer and Wahle, 1997). This enzyme, was Shown to have characteristics that identify is a candidate poly(A) riboguclease and was called PARN. In vitro, PARN is Mg2+-dependent, and at low salt concentrations was dependent on Spermidine or PAB. At physiological salt concentrations, sperrnidine activation was not as strong and PAB was inhibitory (Komer and Wahle, 1997). Incubation of PARN plus PAB with radiolabeled capped and polyadenylated mRNA Species results in phased degradation products that are consistent with protection of the poly(A) tail by PAB (Komer and Wahle, 1997). A cDNA clone corresponding the human homologue was subsequently obtained from the human genome sequencing project and is referred to as HuPARN (Komer et al., 1998). The protein encoded by the HuPARN cDNA was expressed in Escherichia coli and purified and was found to have similar properties to the enzyme 140 originally purified from calf thymus. Experiments conducted in Xenopus oocytes implicate HuPARN as a potential deadenylating nuclease in vivo (Komer et al., 1998). A sequence similar to HuPARN was identified in the Arabidopsis genomic sequence by Komer et al. (1998). To determine if AtPARN (for Arabidopsis thaliana PARN-like gene) is involved in mRNA degradation in plants, a cDNA corresponding to this gene was cloned and we have begun to characterize its function. These studies have the potential to answer many interesting questions about PARN in particular and mRNA degradation in general. For example, it is not known what role if any this type of protein may play in general mRNA degradation mechanisms. Is PARN responsible for triggering general mRNA degradation by removal of poly(A) tails? Is AtPARN important for post-transcriptional control of gene expression in plants? Does conservation of PARN indicate deadenylation plays an important role in plant mRNA decay pathways as it does in yeast? By applying the molecular genetic techniques available in Arabidopsis to the study of AtPARN, we may answer some of these important questions. RESULTS The RNaseD family and cloning of AtPARN Sequence analysis revealed that HuPARN was a member of an ancient gene family of RNases named after it’s founding member, RNaseD (Komer et al., 1998; Mian, 1997; Moser et al., 1997). RNaseD is involved in processing of tRNAs in E. coli (reviewed in Deutscher, 1993). This family includes PAN2 from yeast which is a subunit I41 of the PAB-dependent poly(A) trimming activity described above. The feature that unites this gene family are three 3’ to 5’ exoribonuclease (Exo) domains which contain acidic residues involved in the coordination of metal ions necessary for nuclease activity (Bamad et al., 1989; Joyce and Steitz, 1994). This catalytic domain iS also common to the proofreading domain of E. coli DNA polymerase I. Mian used a hidden Markof model to find RNaseD family members and to characterize important amino acids in these proteins (1997). This algorithm allows one to find specific domains within unrelated polypeptides; outside of the Exo domains many of the RNaseD family members are not similar. For example, there are several open reading frames within the yeast genome that lack Exo domains, but are more similar to HuPARN than is PAN2, an RNaseD family member (Komer et al., 1998). A region of the Arabidopsis genome was identified that bore Significant similarity to the HuPARN sequence. Among the conserved residues were the three Exo domains, the hallmark of the RNaseD family. The genomic sequence was used to design primers for rapid amplification of cDNA ends (RACE) by polymerase chain reaction (PCR). The predicted amino acid sequence of the resulting clone is Shown in an alignment with the HuPARN amino acid sequence in Figure 2-2-1. AtPARN and HuPARN are 21% identical at the amino acid level. Interestingly, AtPARN has a 38 amino acid N-terrninal extension. Eleven out of 38 residues in this N-terminal sequence are serine or threonine, amino acids that are often over-represented in N-terrninal chloroplast targeting signals. However, the AtPARN polypeptide failed to enter chloroplasts in a targeting assay (John Froehlich, data not Shown). 142 WPLR LVCSP AAETVTTSTAASAEAAFP TTLNDLRSL :0 O ::::...:O O O O FVSGRHNFFVFPRQELTFDPPAHEFLCQTTSMDFLAKYQFDFNTCIHEGISYLBRREEEB ... ::.:::. .. . : .:.::..:.::::. 33:: ...:: :z...:: YITKSFNFYVFPKP - FNRSSPDVKFVCQSSSIDFLASQGFDFNKVPRNGI PMQII- - ASKRLKMLHGEDGIDSSGETEBLKLVRLADVLPAARMEKLLNIWRSGLLHGGNLSSIFPR 00:. O : ...: :0 : CO. C : C. O. O. . --RQLREQYDEKRSQANGAGA-LSYVSPN----TSXCPVTIPBDQKKPI ----- DQVVER ISNGSNOSMETVFHHMRPALSLKGFTSHQLRVLNSVLRKHPGDLVYIHSNDKSSSSRDIV : O 3: :0 30:. :0 : ... 0.: O. .0... C C. 3 :: IED-LLQSEEN ------ KNLDLEPCTGFQRKLIYQTLSWKYPKGIHVITLITEKKIRYIV I---SKVDEEE--RKRREQQKHAKBQEELNDKVGPSRVIHAIANS Exmll vrtsmcpnps'rmrvnsrumpxrmxnmomssrsnssussmp O C: III. .8 0 :g ..z::.. 0.0:. O .0: :0 O 8 TVHQFYCPLPADLSIFREMTTCVPPRLLDGKLMASTQPF---KDIINNTSLA8LIKRL-R QIE?SSRSSDSPLQQRVNIDVEIDNVRC‘ ”L;' ;_"w 6 IPAQACNHLGPD . :0 O O: C O I. .0. “I I memems -AEGI'PSYD‘1‘ASEQL --------- " ' - 1’ encrrsnmm-s Exolll PKQHSQLDDPAQNIKLEKYINRLYLSWT -------- RGDIIDLRTGH----SN1DNWRV8 PLSPPKIHVSARSKLIEPFPNKLIEMRVNDIPYLNLEGPDLQPKRDHVLHVTFPXIWKI8 Kr ----- KYENIVLIW-NFPRKLKARGIKBCICKAFGSASVTSVYHVDDSAVPVLFKNS! 3: o : o o o o I o :0... o :00. 3 o. :0 I DLYQLPSAPGNIQISW1DDTSAFVSLSQPEQVKIAVNTSKYAESYRIQTYARYMGRKQIB KQIKRKWTEDSWKEADSKRLNPQCIPYTLQNHYYRNNSFTAPSTVGKRNLSPSQEEAGLE DQAITVGVKSRTRPNAQCETBTRE--ENTVTVTHKASDLIDAFLANR----VEVETATSN DGVSGEISDTELEOTDSCAEPLSEGRKKAKKLKRMKKELSPIGSISKNSPATLPEVPDTW Figure 2-2-1. Sequence alignment of the AtPARN and HuPARN amino acid sequences. The AtPARN polypeptide, listed as the first line, is predicted to have 689 amino acids; HuPARN consists of 639 amino acids. The N-terminal extension found in AtPARN is boxed. Three Exo domains are also boxed and key residues are shaded. Sequence identity is indicated by a :, chemically similar residues (.) are also indicated. 143 RnaseD family members have diverse functions (Deutscher, 1993, Moser et al., 1997). Therefore it may be important to analyze the relative sequence similarity between different family members in order to make functional approximations. AtPARN and HuPARN share Similarity outside of the Exo domains and are more similar to each other than they are two other members of the RNaseD family. BLAST (basic local alignment search tool) searches using the predicted AtPARN amino acid sequence find only HuPARN and a gene from Caenorhabditis elegans (K10C81, also described in Komer et al., 1998). AtPARN is not Significantly Similar with other members of the RNaseD family, including PAN2, in these searches. The close relationship between HuPARN and AtPARN relative to other members of the RNaseD family may indicate functional similarity between these proteins. AtPARN is expressed throughout the plant To determine the expression pattern of A tPARN in Arabidopsis plants, Northern blot analysis was performed on total RNA isolated from roots, stems, leaves, and flowers. AS Figure 2-2-2 indicates, the AtPARN mRNA was found in all of these plant organs. or {'3 a g g 9 8 a 8 3 C U) ...I ll. " Q ‘ AtPARN Figure 2-2-2. AtPARN mRNA is found throughout the plant. A Northern blot containing 20 pg of total RNA from roots, leaves, stems, and flowers of 4,; . w 4 elF4-A Arabidopsis plants was hybridized sequentially with AtPARN and eIF4-A probes. A photograph of the ethidium-bromide stained gel is included to highlight differences in abundance of chloroplast 4 rRNA , ,2; RNA in these samples. 144 The Size of the AtPARN transcript was approximately 2 Kb, about that of the cDNA clone which was 2193 base-pairs excluding the poly(A) tail. Northern blot analysis indicated that AtPARN mRNA was more abundant in roots and flowers than in leaves and stems (Figure 2-2-2). These data must be interpreted with caution, however, because the differences in AtPARN mRNA abundance may reflect differences in loading due to the smaller amount of chloroplast rRNA in roots and flowers. The eIF4-A transcript abundance was to be used as loading control, however, the abundance of this message also varied in these samples. Characterization of AtPARN activity in vitro To analyze the enzymatic activity of the AtPARN polypeptide, the coding sequence was expressed in E. coli as an N-terminal six-histidine fusion. This construct facilitated the partial purification of the AtPARN polypeptide using a Ni2+-NTA column. Denaturing polyacrylamide gel electrophoresis showed that two major proteins were eluted from the Ni2+-NTA column, only one of which cross-reacted with anti— pentahistidine antibodies (data not Shown). These proteins were approximately 70 kd, matching the predicted size of AtPARN, the smaller one may be the result of an N- terminal degradation that occurred after purification (Figure 2-2-3-A). Poly(A) ribonuclease activity was analyzed using a trichloroacetic acid (TCA) precipitation assay that monitors the release of mononucleotides from a uniformly radiolabeled poly(A) substrate following incubation with AtPARN. Preliminary assays indicate that AtPARN has poly(A) ribonuclease activity (Figure 2-2-3-B). Thus far, only conditions optimal for HuPARN have been used in these assays. 145 *3 3 B |.I.l E coll Substrate only HuPARN AtPARN Figure 2-2-3. AtPARN has poly(A) ribonuclease activity. A. A denaturing polyacrylamide gel was loaded with approximately equal amounts of protein from the flow through (FT), the first wash (W1), the second wash (W2), and the eluate of a Ni2+- NTA column. A photograph of the Coomassie stained gel is shown. B. The AtPARN eluate was tested using the TCA PARN assay (described in materials and methods). The average counts per minute (CPM) detected in two experiments is plotted i the standard deviation. The eluate from BL21(DE3)pLysS cells (E. coli) that were not expressing AtPARN and the assay mixture with no protein added (Substrate only) served as negative controls. HuPARN was used as a positive control. DISCUSSION Deadenylation is the first step in the degradation of many yeast and mammalian mRNAs and the rate of deadenylation is important in determining mRNA stability in these systems (Shyu et al., 1990; Muhlrad et al., 1994, Ross, 1995; Caponigro and Parker, 1996). There is evidence that deadenylation may be involved in the degradation of at least one plant message, the oat phytochrome A transcript (Higgs and Colbert, 1994). A thorough study of poly(A) degrading activities in plants is necessary to determine the role 146 of this process in plant mRNA stability. In addition, studying a poly(A) ribonuclease in plants may provide important information about the mechanism of deadenylation in other eukaryotes. Eukaryotic mRNA turnover pathways have been studied almost exclusively in yeast. In Saccharomyces cerevisiae, genes have been identified that encode every enzyme in the major mRNA decay pathway except the deadenylating nuclease. Thus far, the best candidate in any eukaryote for such an activity is HuPARN, a 3’ to 5’ exoribonuclease that degrades poly(A) preferentially (Komer and Wahle, 1997, Komer et al., 1998). The characteristics of this protein in vitro are consistent with a role in deadenylation (discussed above, Komer et al., 1997). Perhaps the best evidence that HuPARN may play a role in deadenylation of mRNA comes from studies in Xenopus oocytes. A deadenylating nuclease activity that catalyzes the default removal of poly(A) tails in mature oocytes was identified almost ten years ago (Varnum et al., 1990). Purification of a protein responsible for this activity and cloning of the gene that encodes it revealed strong Similarity to HuPARN (Komer et al., 1998). Antibodies directed against HuPARN blocked the default deadenylation of transcripts in mature oocytes and ectopic expression of HuPARN promoted deadenylation in enucleated oocytes lacking Xenopus PARN. These experiments indicated that HuPARN is a deadenylase in vivo (Komer et al., 1998). We have cloned a cDNA from Arabidopsis that bears significant Similarity to HuPARN. Initial experiments indicate that this gene, called AtPARN, encodes a protein with poly(A) nuclease activity in vitro. Thus far only assay conditions that were optimal for HuPARN activity have been tested. The AtPARN mRNA is expressed throughout the 147 plant as might be expected for a gene involved in a ubiquitous process such as deadenylation of mRNA. Future experiments will address the role of this gene in mRN A degradation in Arabidopsis. We have initiated the search for a transfer-DNA (T-DNA) insertion in AtPARN. A collection of ~68,000 plants that collectively have ~100,000 T-DNA inserts Spread throughout the genome has been screened by PCR using primers corresponding to the borders of the T-DNA and the AtPARN genomic sequence (Hirsch et al., 1998). Initially, 30 large pools containing DNA from the entire collection were screened to identify those with inserts in AtPARN. These pools must be dcconvoluted through a series of sub-pools in order to find single plants with the T-DNAS of interest. The PCR products corresponding to five T-DNA inserts in AtPARN have been sequenced. One T- DNA is inserted just after the first Exo domain within the third cxon of the AtPARN genomic sequence. A pool of nine plants has been identified that contains plants with this T-DNA insertion. The T-DNA is ~8 Kb long, so insertions will likely result in a null mutation. Based on database searches and Southern blot analysis (data not Shown), A tPARN appears to be the only PARN-like sequence in the Arabidopsis genome. Therefore, a mutant in this gene Should be informative. Of course if AtPARN is important for the control of mRNA stability in plants, loss of function may be lethal. One of the advantages of isolating mutants by PCR screening is that heterozygotes can be identified if the mutation is recessive. Analysis of the progeny of these heterozygous plants allows one to determine if a mutation is lethal, an opportunity not available to classical geneticists. The chief goal of these studies will be to study the effect on poly(A) tail length in these mutants and analyze any phenotypic changes that become apparent. 148 A fusion of HuPARN to the green fluorescent protein (GFP) was localized to the nucleus and cytoplasm of Cos-1 cells (Komer et al., 1998). It is unclear why HuPARN would be found in the nucleus in COS-1 cells, but one would expect a deadenylase involved in mRNA degradation to be found in the cytoplasm. In addition, Xenopus PARN isoforms are found in the nucleus and in the cytoplasm of mature oocytes (Komer et al., 1998). In Xenopus, PARN activity is thought to be regulated by sequestration in the nucleus. One interesting question raised by these studies is whether PARN is a specialized activity involved in the default deadenylation of mRNAs developing zygotes or whether it is involved in general mRNA degradation. The initial purification of mammalian PARN from calf thymus and the expression of HuPARN mRNA in many different types of human cells would suggest that this is not true of HuPARN, and whether this is true of Xenopus PARN remains to be tested (Komer et al., 1998). AtPARN is expressed throughout the plant, but we do not yet know where it is localized within the plant cell or whether it is expressed preferentially in certain cell types. We have initiated these localization studies by introducing an N-terrninal GFP fusion into Arabidopsis plants. Genetic strategies are limited in mammalian cells and in Xenopus, therefore, the ability to test the function of AtPARN directly by isolating a T-DNA insertion mutant may provide us with the unique opportunity to address the role of this type of enzyme in cytoplasmic mRNA degradation in somatic cells. 149 MATERIALS AND METHODS Cloning of AtPARN Standard cloning procedures were used as described in Sambrook et al. (1989). Two products of RACE experiments corresponding to the 5’ and 3’ ends of the AtPARN cDNA were generated using primers complementary to the AtPARN genomic sequence and to the adapter sequence that was 1i gated to the ends of cDNAs from Arabidopsis seedling mRNA. The details of this protocol are described in the package insert of the Marathon cDNA kit (Clonetcch). The 5’ product consisted of the first 1.5 kb of the AtPARN cDNA and was amplified using primer pg-444 (5’- TGAAGTCAAAACCGAGATGATTGC -3’). This product was inserted into the EcoRV site of a derivative of pBlueskript SKII(-) called p948, the resulting clone was called p1843. The 3’ product consisted of the latter 0.9 kb of the AtPARN cDNA and was amplified using primer pg-609 (5’-AGGTTGCATCTTTGCGCAGGC -3’). This product was inserted into the EcoRV site of p948, the resulting clone was called p1920. The overlap Shared by these two PCR products contained a unique Sphl Site that allowed the two products to be joined, creating p1932. The final cDNA clone was sequenced and checked against two available genomic sequences. No errors in the cDNA were found. The AtPARN Genomic region consists of ~3000 Kb on bacterial artificial chromosome f14j16 which has been mapped to chromosome 1 near position 80. Comparison of the cDNA with the genomic sequence revealed that AtPARN consists of 7 cxons. 150 Analysis of A {PARN expression Total RNA was isolated from the leaves, stems, and flowers of mature (6 week old) Arabidopsis (Col-0) plants. Root RNA was prepared from the roots of seedlings grown in liquid culture for 3 weeks (2.3% Gamborg’s medium [GibcoBRL, Inc.], pH 5.0). RNA was prepared using the method described by Puissant and Houdebine (1990) with modifications described by Newman et al., (1993). Northern blot analysis was as described in Newman et al. (1993). The use of the translation initiation factor, eIF4-A, as a loading control for Northern blots was as described by Taylor et al., (1993). Hybridization probes corresponding to the AtPARN and eIF4-A coding regions were labeled using 0t-32P dCTP by the random-primed method of Feinberg and Vogelstein (1983). Expression of AtPARN in E. coli The vector used to express HuPARN in E. coli (pGMMCS) was obtained from Wahle and coworkers (Komer et al., 1998). This is a derivative of the pET vectors (Novagen). To fuse Six histidine residues to the N-terrninus of the AtPARN coding sequence and to facilitate cloning into pGMMCS, the primers pg-718 (5’- T GCC ATG GCT CAC CAT CAC CAT CAC CATGTC GAC A_TG CGC CGG CAC AAG CGA TG — 3’) and pg-727 (5’ — CG TGC TCG AGT TAA TTA CTC GTA GCA GIT TCG - 3’) 151 were used to amplify the AtPARN coding sequence by PCR. Stop and start codons are underlined, the codons encoding the six histidine residues are italicized. The resulting PCR product was inserted at the EcoRV site of p948 creating pl938. The AtPARN coding region was sequenced and no errors were found. An Ncol, Xbal fragment of this clone was then introduced into pGMMCS creating p1949, the AtPARN E. coli expression vector. BL21(DE3)pLysS (Novagen) cells were transformed with p1949 by electroporation using the Gene-Pulser (BioRad) according to the manufacturer’s guidelines. Induction of AtPARN expression was carried out in LB medium supplemented with chloramphenicol (33 pg/ml), ampicillin (100 pg/ml) and IPTG (1 mM) as described on pages 46 and 47 of the QIAexprcssionist (1997, Qiagen Inc.). 10 ml cultures from three clones that expressed AtPARN to high levels were combined and a soluble lyste was prepared as described on page 63 of the QIAexprcssionist (1997, Qiagen Inc.). AtPARN was purified using a Ni2+-NTA (nitrlotriacctic acid) resin that was incubated with p1949 soluble lysates in a batch culture at 4°C as described on pages 66 and 67 of the QIAexprcssionist (1997, Qiagen incorporated). The denaturing polyacrylamide gel shown in Figure 2-2-3-A consisted of 10% polyacrylamide and 5% SDS. Lysates from BL21(DE3)pLysS (Novagen) cultures were prepared in parallel and were resolved using NiZ+-NTA resin so that eluates from these columns could be used as a negative control in PARN assays. 152 TCA PARN assay This assay monitors the release of TCA-soluble products from homogeneously labeled poly(A). The substrate was prepared as described in Brown et al.(l996). The reaction mixture consisted of 500 units of yeast poly(A) polymerase (U .S. Biochemicals), 50 pCi (1:32P rATP, 0.5 pM l2 nucleotide oligo(A) (obtained from Dharmacon Research, Inc.), 167 mM rATP, 1X poly(A) polymerase buffer (U.S. Biochemical). The reaction was incubated at 30°C for 60 minutes. Polyacrylamide gel electrophoresis showed that labeled poly(A) was greater than 1 Kb in length. Unincorporated nucleotides were removed using a spin column (S-200, Phannacia). 100,000 CPM of the poly(A) substrate was incubated with 10 pls of BL21 eluate, 10 pls of AtPARN eluate, or 1 pl of HuPARN or with the PARN assay mixture alone. The concentrations of AtPARN and HuPARN were not determined. The assay mixture and conditions were as described in Komer and Wahle (1997) and modified based on Brown et al. (1996). Briefly, eluates were added to a reaction mixture consisting of 2mM sperrnidine plus dilution buffer (5mM Hepes, pH 7.4; 2mM MgC12; 14.4mM B- mercaptoethanol) in a total volume of 50 pls and were incubated at 37°C for 30 minutes. 200 pls of cold 20% TCA was then added and polynucleotides were precipitated at -20°C for 10 minutes. Pellets were spun down and 100 pls of the supernatant was counted by liquid scintillation. Two experiments that were conducted one after the other are reported in Figure 2-2-3-B. 153 ACKNOWLEDGEMENTS I would like to thank Drs. Ambro van Hoof and Roy Parker of The University of Arizona who made us aware of the AtPARN genomic sequence before publication of Komer ct a1. (1998); and Dr. Eva Devlin in Dr. Elmar Wahlc’s group at the Universitiit Hallc Institut fiir Biochemie, who provided pGMMCS and an aliquot of HuPARN. I thank Jonathan Vo gel for his great interest in this project and for conducting the TCA PARN assays reported in Figure 2-2-3-B. I would also like to thank Prectrnoninder Lidder who constructed a GFP-AtPARN fusion protein and iS currently analyzing its localization in transgenic Arabidopsis. Thanks also to James Kastenmayer who generated the cDNA library that was used to clone AtPARN and Miguel A. Pérez-Amador who generated the Northern blot used for Figure 2-2-2. 154 REFERENCES Beelman,C.A., Stevens,A., Caponigro,G., LaGrandeur,T.E., Hatfield,L., Fortner,D.M., and Parker,R. (1996). An essential component of the decapping enzyme required for normal rates of mRNA turnover. Nature 382, 642-646. Bernad,A., Blanco,L., Lazaro,J.M., Martin,G., and Salas,M. (1989). A conserved 3'----5' exonuclease active site in prokaryotic and eukaryotic DNA polymerases. Cell 59, 219- 228. Brown,C.E., Tarun,S.Z., Jr., Boeck,R., and Sachs,A.B. (1996). PAN3 Encodes a Subunit of the Pabl p-Dependent Poly(A) Nuclease in Saccharomyces cerevisiae. Mol. Cell. Biol. 16, 5744-5753. Caponigro,G. and Parker,R. (1996). Mechanisms and control of mRNA turnover in Saccharomyces cerevisiae. Microbiol. Rev. 60, 233-249. CarpousiS,A.J., Vanzo,N.F., and Raynal,L.C. (1999). mRNA degradation - a tale of poly(A) and multiprotein machines. Trends Genet. 15, 24-28. Deutscher,M.P. (1993). Promiscuous exoribonucleases of Escherichia coli. J. Bacteriol. I 75, 4577-4583. Dunckley,T. and Parker,R. (1999). The DCP2 protein is required for mRNA decapping in saccharomyces cerevisiae and contains a functional MutT motif [In Process Citation]. EMBO J. 18, 5411-5422. Feinberg,A.P. and Vogelstein,B. (1983). A technique for radiolabeling DNA restriction endonuclease fragments to high specific activity. Anal. Biochem. 132, 6-13. Fox,C.A. and Wickens,M. (1990). Poly(A) removal during oocyte maturation: A default reaction selectively prevented by specific sequences in the 3' UTR of certain maternal mRNAs. Genes Dev. 4, 2287-2298. Gallie,D.R. (1998). A tale of two termini: A functional interaction between the termini of an mRNA is a prerequisite for efficient translation initiation. Gene 216, 1-1 1. Higgs,D.C. and Colbert,J.T. (1994). Oat phytochrome A mRNA degradation appears to occur via two distinct pathways. Plant Cell 6, 1007-1019. Hirsch,R.E., Lewis,B.D., Spalding,E.P., and Sussman,M.R. (1998). A role for the AKTl potassium channel in plant nutrition. Science 280, 918-921. See also www.biotech.wisc.edu/Arabidopsis/ J acobson,A. and Peltz,S.W. (1996). Interrelationships of the pathways of mRNA decay and translation in eukaryotic cells. Annu. Rev. Biochem. 65, 693-739. 155 Joyce,C.M. and Steitz,T.A. (1995). Polymerase structures and function: variations on a theme? J. Bacteriol. I 7 7, 6321-6329. K6rncr,C.G. and Wahle,E. (1997). Poly(A) tail Shortening by a mammalian poly(A)- Specific 3'-exoribonuclease. J. Biol. Chem. 272, 10448-10456. K6mcr,C.G., Wormington,M., Muckenthaler,M., Schneider,S., Dehlin,E., and Wahle,E. (1998). The deadenylating nuclease (DAN) is involved in poly(A) tail removal during the meiotic maturation of Xenopus oocytes. EMBO J. 1 7, 5427-5437. LaGrandeur,T.E. and Parker,R. (1998). Isolation and characterization of Dcplp, the yeast mRNA decapping enzyme. EMBO J. 17, 1487-1496. Mian,I.S. (1997). Comparative sequence analysis of ribonucleases H11, 111, II PH and D. Nucleic Acids Res. 25, 3187-3195. Moser,M.J., Holley,W.R., Chatterjee,A., and Mian,I.S. (1997). The proofreading domain of Escherichia coli DNA polymerase I and other DNA and/or RNA exonuclease domains. Nucleic Acids Res. 25, 51 10-5118. Muhlrad,D., Decker,C.J., and Parker,R. (1994). Deadenylation of the unstable mRNA encoded by the yeast MFA2 gene leads to decapping followed by 5'-->3' digestion of the transcript. Genes Dev. 8, 855-866. Newman,T.C., Ohme-Takagi,M., Taylor,C.B., and Green,P.J. (1993). DST sequences, highly conserved among plant SA UR genes, target reporter transcripts for rapid decay in tobacco. Plant Cell 5, 701-714. Puissant,C. and Houdebine,L.-M. (1990). An improvement of the Single-step method of RNA isolation by acid guanidinium thiocyanate-phenol-chlorophorm extraction. BioTechniques 8, 148-149. Qiagen Incorporated. The QIAexprcssionist. March 1997. Qiagen Inc. See also www.qiagen.com for protocols. Ross,J. (1996). Control of messenger RNA stability in higher eukaryotes. Trends Genet. 12, 171-175. Sambrook,J., Fritsch,E.F., and Maniatis,T. (1989). Molecular Cloning: A Laboratory Manual. (Cold Spring Harbor, New York: Cold Spring Harbor Laboratory Press). Sheets,M.D., Wu,M., and Wickens,M. (1995). Polyadenylation of c-mos mRNA as a control point in Xenopus meiotic maturation. Nature 3 74, 51 1-516. Shyu,A.-B., Belasco,J.G., and Greenberg,M.E. (1991). Two distinct destabilizing elements in the c-fos message trigger deadenylation as a first step in rapid mRNA decay. Genes Dev. 5, 221-231. 156 Taylor,C.B., Bariola,P.A., DelCardayré,S.B., Raines,R.T., and Green,P.J. (1993). RNS2: A senescence-associated RNase of Arabidopsis that diverged from the S-RNaseS before speciation. Proc. Natl. Acad. Sci. USA 90, 51 18-5122. van Hoof,A. and Parker,R. (1999). The cxosome: a proteasome for RNA? Cell 99, 347- 350. Varnum,S.M. and Worrnington,W.M. (1990). Deadenylation of maternal mRNAs during Xenopus oocyte maturation does not require specific cis-sequences: a default mechanism for translational control. Genes Dev. 4, 2278-2286. Varnum,S.M., Hurney,C.A., and Worrnington,W.M. (1992). Maturation-specific deadenylation in Xenopus oocytes requires nuclear and cytoplasmic factors. Dev. Biol. 153, 283-290. 157 IV. nrcnrcan sran UN LIB 11111111111111111111111111 11111111 31293020 8 RRRIES 4 8759