tits}: :2:‘ .1. ‘ \. ‘ In .(5 . .31. .u. unmélfluwu: 15. I RE! i: . V ”.3. .9 "5m? m Hana. .. 4. .T ”K, i “Eighfiuu ' 15 v. .m This is to certify that the dissertation entitled GENETICS AND GENOMICS OF THE DST-MEDIATED DECAY PATHWAY IN ARABIDOPSIS THALIANA presented by PREETMONINDER LIDDER has been accepted towards fulfillment of the requirements for the PhD. degree in Cell and Molecular Biology Vail/é, // /,é[/ Mngrofessor’s Signature xl/2//0Y' MSU is an Affirmative Action/Equal Opportunity Institution u-«-—.—---------.--- -.-.- -.-.--.-.-.-.- ---—-—:-.- ._ - - W” Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 cleIRC/DateDuepes-p. 1 5 GENETICS AND GENOMICS OF THE DST-MEDIATED DECAY PATHWAY IN Arabidopsis thaliana By Preetmoninder Lidder A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Cell and Molecular Biology 2004 ABSTRACT GENETICS AND GENOMICS OF THE DST-MEDIAT ED DECAY PATHWAY IN Arabidopsis thaliana By Preetmoninder Lidder The control of mRNA stability plays a fundamental role in the regulation of gene expression in plants and other eukaryotes. The basal mRNA decay machinery, sequence- specific decay components, and regulatory factors that respond to various stimuli can influence the control of mRNA stability. Despite the importance of mRN A stability in providing the cells with a means to rapidly tailor gene expression, little is known about this process in higher eukaryotes in general, and plants in particular. The overall goal of the research described in this thesis has been to elucidate the mechanisms and cellular factors which determine sequence-specific mRN A decay in higher plants. Towards this end, Arabidopsis mutants defective in the DST-mediated decay pathway were utilized. The dst mutants were originally isolated as specifically elevating the steady-state level and increasing the half-life of DST-containing transcripts. To determine new molecular phentotypes of the dstl mutant and evaluate the biological significance of the DST-mediated decay pathway, cDNA microarray technology was employed. In addition to verifying the increase in the transgene mRNA levels, used to isolate these mutants, new genes with altered mRNA abundance in dstl were identified. RNA gel blot analysis confirmed the microarray data for all genes tested and was also used to catalog the first molecular differences in gene expression between the dstI and dst2 mutants. Clustering analysis of genes altered in dstl exposed new co-expression patterns suggesting a link between the dst] mutation and circadian rhythms. Earlier results from microarray expression data for unstable mRNAs in Arabidopsis showed that mRNA instability is associated with a group of genes controlled by the circadian clock. Experiments conducted during the course of this dissertation indicated that Ccr-like and SEN], two transcripts which are regulated at the level of mRNA stability by the clock, are also direct targets of the DST-mediated decay pathway. Not only were Cor-like and SEN] transcripts altered in their half-lives in dst], but their stabilities were also altered in the mutant relative to the parental at different times during the day. This leads to aberrant circadian oscillation of these transcripts in the mutant demonstrating that the DST 1 locus is associated with circadian control. Previous experiments (e.g. with the per gene in Drosophila) have implicated differential mRN A stability in the control of CCGs but the observations in Arabidopsis are among the first to provide direct evidence for this. Map-based cloning of the dst] mutant was initiated using RAP2.4, an endogenous target identified during the microarray experiments. The use of this marker circumvented some of the problems associated with the transgene to score homozygous mutants in the mapping population. A new mutant, dst3, was characterized in this thesis which should facilitate the investigation of the molecular components involved in DST-mediated degradation. Additionally, microarray experiments with dst2 were carried out to identify molecular markers specific for dst2 and further address the physiological role of the DST-mediated mRNA decay pathway in plants. Taken together, the results described in this dissertation should constitute significant progress toward understanding the molecular machinery responsible for sequence-specific mRNA degradation in plants. Copyright by PREETMONINDER LIDDER 2004 To my parents ACKNOWLEDGMENTS I would like to acknowledge numerous people without whose support this thesis would not be possible. But where does one begin? Indeed there have been so many who have helped and encouraged me all through that at times I have felt overwhelmed. During the period of my stay at the Plant Research Laboratory and subsequently at the Delaware Biotechnology Institute, I have been the beneficiary of the imagination, wisdom and insights of a lot of remarkable people. I owe a special debt of gratitude to my advisor, Dr Pam Green for her patient guidance at all times during my dissertation. The excellent discussions with her infiised an inquisitive attitude in my work and enabled me to be more resourceful and productive. I am also deeply indebted to my committee members, Dr David Amosti, Dr Jon Kaguni Dr Ken Keegstra and Dr Mike Thomashow, for offering invaluable advice and helpful suggestions. Thanks are also due to past and present members of the Green lab who have not only been great friends but have also helped in creating an outstanding atmosphere in the lab for great research. Thanks Linda, Nikki, Yukako, Fred, Gustavo, Jim and Cheng. This acknowledgement would not be complete without a heartfelt thanks to Mark Johnson for being a superb tutor during my rotation, Miguel Pe'rez-Amador for teaching me all about microarrays and Rodrigo Gutierrez for thought-provoking comments, timely computer- help and of course, the “drumming”. I would like to thank Monica for all her help with the mapping of DST] - it would have been extremely difficult without her and I really appreciate her assistance on this project. Thanks to the “ice-cream” club at MSU- Sam, Robin, Verna and Robert, for the refreshing breaks in between experiments. Vi My love and gratitude goes out to my parents, Jasbir and Satinder Lidder, who have stood by me when I have felt like giving up, offering encouragement and helping out in ways no one else could. A very special thank you to my brother, Nanu, for his lively remarks and never ending enthusiasm. Thanks to Dana, who has been there for the long run and I consider her to be the best friend a person could ever ask for. Last, but not the least, my all out thanks goes to my fiancee, Jashan, who has done whatever it takes to make this possible. He has been there for me through thick and thin and I want to thank him from the bottom of my heart. Thanks very much Jashan- you are number one to me and always will be. vii TABLE OF CONTENTS LIST OF TABLES XI LIST OF FIGURES XII ABBREVIATIONS XIV CHAPTER 1 Control of mRNA turnover in plants 1 Introduction 2 Determination of mRN A decay rates 4 Chemical inhibitors 5 Regulated promoters 6 RNA polymerase II 8 Microarray technology 8 Stimuli affecting mRNA stability 9 Plant hormones 9 Light 10 Sucrose 12 Nitrogen 13 Methionine 13 Biotic Stress 14 Abiotic stress 15 cis-acting determinants of mRN A stability 17 DST element 17 AUUUA sequences 20 5’ UTR 21 General mRNA decay machinery 22 Sequence-specific decay 24 Perspective 28 References 29 CHAPTER 2 New molecular phenotypes of the dst] mutant revealed by microarray analysis 38 Introduction 39 Results 43 Generation of a 600-element DNA microarray 43 The expression levels for most genes are similar in dstl and the parental plants 45 The 11,521-element microarray reveals additional genes with altered gene expression in the dst] mutant 48 RNA gel blot analysis confirms DNA microarray data 52 Identification of primary targets of the dst] mutation 56 Multiple replicates remove non-reproducible changes 58 viii Cluster analysis of genes with altered expression levels in dst] Discussion Transcripts altered in dst] have predominantly increased levels, although some decrease in abundance Molecular markers to expedite characterization of and differentiation between dst] and dst2 Circadian association of the dst] mutation Conclusions and future prospects Materials and Methods Plant material Generation of a 600-element DNA microarray 11,521 AFGC DNA Microarray Half-life measurements, total RNA Extraction, Poly(A)+ RNA Purification, and RNA Blot Hybridization Labeling of Poly(A)+ RNA DNA Microarray Hybridization and Analysis References CHAPTER 3 Circadian control of mRNA stability: Impact of the dst mutants Introduction Results Diurnal control of Ccr-likc and SEN] mRNA stability is affected in the dst] mutant DST 1 function is required for normal circadian expression of SEN] and Ccr-like mRNAs dst] affects circadian control of mRNA stability Opposite effect of dst2 on Sen] mRNA stability Impact at the whole plant level: Classical circadian phenotypes are altered in the dst mutants Discussion Materials and Methods Arabidopsis strains and grth conditions Half-life measurements, RNA preparation and analysis Leaf movement assay References CHAPTER 4 Genetic mapping of dst] Introduction Results dst] mutation does not appear to be linked to the transgene Use of RAP2.4 as an endogenous marker to score homozygous mutants Fine mapping of the dst] locus ix 61 65 65 69 7O 71 72 72 73 75 75 76 77 80 85 86 90 9O 92 96 99 99 102 106 106 106 107 109 115 116 118 118 120 120 Discussion Materials and Methods Plant material Half-life measurements, total RNA Extraction and RNA Blot Hybridization Genomic DNA extraction and markers for mapping References CHAPTER 5 Characterization of dst3, a new gene in the DST-mediated mRN A degradation pathway Introduction Results dst3 exhibits increased HPH-DS T and G US-DST mRNA levels Genetic analysis of dst3 dst3 is not allelic to dst] or dst2 Increased stability of HPH-DST mRNA in dst3 compared with 1519-31 Discussion Materials and Methods Plant material Half-life measurements, total RNA Extraction and RNA Blot Hybridization Analysis of the HPH-DSTX4 sequence element in dst3 References CHAPTER 6 Microarray analysis of dst2 Introduction Results 15k element microarray reveals genes with altered gene expression in dst2 Identification of the putative primary targets of the dst2 mutation RNA gel blot analysis of previously identified transcripts Discussion Materials and Methods Plant material 15k AFGC DNA Microarray Total RNA Extraction, Poly(A)+ RNA Purification, and RNA Blot Hybridization Labeling of Poly(A)+ RNA DNA Microarray Hybridization and Analysis References CHAPTER 7 Final remarks and future prospects 123 127 127 127 128 131 133 134 135 135 137 139 139 143 146 146 146 147 148 149 150 150 150 151 156 156 159 159 159 160 160 161 163 165 LIST OF TABLES Table 2.1. Genes included on the 600-element DNA microarray Table 2.2. Genes with increased mRNA levels in dst] vs. parental plants Table 2.3. Genes with decreased mRNA levels in dst] vs. parental plants Table 2.4. Genes with possible DST-like sequences in their 3' UTR Table 4.1. Segregation of increased RAP2.4 mRNA abundance in the progeny of crosses between dst] and 1519-31 (DS T ] ) Table 4.2. Oligonucleotides and restriction enzymes used to detect various SSLP and CAPS markers Table 5.1. Segregation of increased HPH mRNA abundance in the progeny of crosses between dst3 and 1519-31 (DST 3) Table 5.2. Segregation of increased HPH mRNA abundance in the progeny of crosses between dst3 and dst] and dst2 Table 6.1. Genes with increased mRNA levels in dst2 vs. parental plants Table 6.2. Genes with decreased mRNA levels in dst2 vs. parental plants Table 6.3. Genes with possible DST-like sequences in their 3' UTR xi 44 53 54 59 121 129 141 142 152 153 154 LIST OF FIGURES Figure 2.1 Analysis of EST 12509T7 identifies it as SA UR-like] 47 Figure 2.2. RNA gel blot analysis to confirm DNA microarray data 49 Figure 2.3. Comparison of mRN A levels in the dst] mutant and parental plants using the 11,521 element DNA microarray 51 Figure 2.4. Histogram plot comparing the gene expression values obtained using microarray analysis and RNA gel blot analysis 55 Figure 2.5. Histograms of the mutant (dst] or dst2)/parental ratios of mRNA expression for the indicated clones 57 Figure 2.6. RAP2.4, Ccr-like and SEN] mRNAs are primary targets of the DST- mediated decay pathway 60 Figure 2.7. Assessment of the reproducibility of the microarray data using fewer replicates 62 Figure 2.8. Cluster analysis of genes with altered mRNA levels in dst] 64 Figure 3.1. Regulation of Ccr-like mRNA stability is altered in the dst] mutant during the day 91 Figure 3.2. Regulation of SEN] mRNA stability is altered in the dst] mutant during the day 93 Figure 3.3. Circadian oscillation of Cor-like mRN A is altered in the dst] mutant 94 Figure 3.4. Circadian oscillation of SEN] mRNA is altered in the dst] mutant 95 Figure 3.5. Circadian oscillation of AtGRP7 mRNA is unaltered in the dst] mutant 97 Figure 3.6. Circadian regulation of Ccr-like mRNA stability is altered in the dst] mutant 98 Figure 3.7. Circadian regulation of SEN] mRNA stability is altered in the dst] mutant 100 Figure 3.8. Regulation of SEN] mRNA stability is altered in the dst] mutant during the day 101 xii Figure 3.9. Circadian rhythm of leaf movement is altered in the dst mutants 103 Figure 4.1. mRNA abundance of RAP2.4 in F2 plants from the second back-cross of dst] relative to WT (1519—31) 119 Figure 4.2. RAP2.4 mRNA is stabilized in the dst] mutant and the decay kinetics are similar in the am and pm. 122 Figure 4.3. DST 1 maps to chromosome I 124 Figure 4.4. Map-based cloning of DST] 125 Figure 5.1 HPH-DST mRNA levels are elevated in dst3 136 Figure 5.2 GUS-DST mRNA levels are elevated in dst3 138 Figure 5.3. mRNA abundance of HPH—DST in F2 plants from the first back-cross of dst3 relative to WT (1519-31) 140 Figure 5.4 Analysis of HPH-DST mRNA stability in leaves from 1519 and dst3 plants 144 Figure 6.]. RNA gel blot analysis of previously identified transcripts 157 xiii AFGC CAPS CCG CT dCAPS DNA DST EST GUS HPH MIPS mRN A ORF PCR per SAUR SMD SNP SSLP TAIR UTR ZT ABBREVIATIONS Arabidopsis Functional Genomics Consortium Adenylate/uridylate rich elements Cleaved amplified polymorphic sequence Clock controlled gene Circadian time Derived CAPS Deoxyribonucleic acid Downstream element Expression sequence tag B-glucuronidase Hygromycin phosphotransferase Munich Information Center for Protein sequences Messenger RNA Open reading frame Polymerase chain reaction Period Small auxin up RNA Stanford Microarray Database Single nucleotide polymorphism Simple sequence length polymorphism The Arabidopsis Information Resource Untranslated region Zeitgeber time xiv CHAPTER 1 CONTROL OF mRNA TURNOVER IN PLANTS INTRODUCTION Gene expression can be controlled at multiple levels in the cell and requires the integration of varied processes such as transcription, RNA processing and export, translation and post—translational events. In the past years, the significance of the plethora of mechanisms functioning at the post-transcriptional level has become exceedingly clear. The control of mRNA stability is a notable and well-known form of post- transcriptional gene regulation in eukaryotic cells. Compared to transcription, much less is known about the machinery for mRNA degradation. To provide a broad overview of the various components that contribute to the stability of mRNAs and the pathways by which mRNA decay occurs in plants, the focus of this chapter will be on the events affecting the turnover of mRN As encoded by nuclear genes in Arabidopsis. Relevant examples from other higher plants will also be highlighted. For detailed knowledge on mRN A decay pathways in other systems, the reader is referred to excellent recent reviews on the topic (Caponigro and Parker, 1996; Guhaniyogi and Brewer, 2001; McCarthy, 1998; Mitchell and Tollervey, 2001; Ross, 1995). The level of mRNA that is available in the cytoplasm for translation is a key control point in the regulation of gene expression. A useful way to consider the control of mRNA stability is in three interrelated levels (Gutierrez et a1., 1999). Recent studies indicate that the first and the most fundamental level is that of the basal mRN A decay machinery responsible for the decay of most cellular mRNAs. Superimposed on the general decay machinery is the second level of control that establishes the “inherent” degradation rates of various mRNAs, which can vary over a broad range. The average half-life of an mRNA in eukaryotic cells is on the order of several hours, but for some mRNAs, sequence-specific recognition mechanisms can facilitate half-lives as short as minutes or as long as several days (Gutierrez etal., 1999). The stability of some mRN As is differentially regulated by exogenous or endogenous stimuli and such differential regulation of mRNA stability constitutes the third level of control. There is much to be learned about the mechanisms and cellular factors that are responsible for mRNA turnover in eukaryotes in general and plants in particular. However one difficulty has been that most work on the general decay machinery has been carried out in yeast whereas work in mammals and plants has focused essentially on sequence-specific (constitutive and regulated) control of mRNA decay. Previous studies have suggested that there are distinct mechanistic differences between unicellular and multicellular eukaryotes with respect to mRNA decay. Therefore a thorough investigation of the mRNA decay machinery in multicellular systems is required to elucidate the fundamentals of how mRNA stability is controlled. Arabidopsis thaliana, a member of the mustard family, has emerged as the primary plant model organism to study mRNA stability because of the various tools available for its analysis and the technical advantages of this system that surpass those of mammals for in vivo analysis. A genetic approach in Arabidopsis may offer the best opportunity to isolate and analyze the cellular factors involved in mRNA turnover. Also the biological functions of the potential factors that govern mRNA stability can be addressed using reverse genetic approaches that are hard to reproduce in mammalian systems. Insertion mutants of Arabidopsis are available from a variety of different collections. In addition, RNAi constructs, having a “panhandle” structure, can be used to inactivate the desired gene (Waterhouse et al., 1998). Further overexpression lines are advantageous to assess the function of newly cloned proteins (Weigel etal., 2000) especially in conjunction with the inactivation lines. Finally, microarray technology (Schena et al., 1995; DeRisi et al., 1997) is a powerful approach to assay the expression of numerous genes involved in multiple aspects of mRN A stability and to monitor mRN A stability on a global scale. Arrays for Arabidopsis, containing approximately 14,000 clones, have been generated by the Arabidopsis Functional Genomics Consortium. DETERMINATION OF mRNA DECAY RATES The study of mRN A turnover in plants has been greatly aided by a number of methods for measuring mRNA decay rates. The derivation of the mRNA decay constant (kd) is based on the assumption that the degradation of mRN A obeys first-order kinetics i.e. the turnover rate of a message is proportional to the amount of mRNA present at a given time and the rate constant for decay. Usually, the stability of a message in vivo is reported as a half-life, the time required for half of the existing mRNA molecules to be degraded. The half—life of an mRNA (ti/2) is inversely proportional to its decay constant. ti/2 = In Z/kd Determination of mRNA half-life begins by blocking mRNA synthesis, isolating RNA samples at various time intervals and monitoring the loss of a particular message by analyzing equal amounts of these samples with a message-specific probe. A semi- logarithmic plot of mRNA concentration as a function of time then yields a straight line with the decay constant as the slope, as defined by: In (C/Co) = -kdt where Co is the initial mRNA concentration and C is the mRNA concentration at time t. However, there are situations in which mRN A decay might not be stochastic. For example, in some eukaryotic cells, poly(A) shortening precedes decay of the mRNA body (Decker and Parker, 1993; Chen and Shyu, 1995) and only after the poly(A) tail has been shortened to a certain extent does first-order decay occur. Thus for mRN As with biphasic kinetics, measuring half-life might not be straightforward and it is necessary to monitor the deadenylation step, or other steps which might precede decay of the body of the transcript, such as decapping. Over the past years, a number of methods to measure mRNA decay have been developed, a few of which are described in detail below. For a comprehensive review of techniques used for determining mRNA degradation rates, the reader is referred to some earlier reviews (Abler and Green, 1996; Ross, 1995). The most common method to measure the decay rates of transcripts is to first shut off transcription and subsequently determine by RNA gel blot analysis the amount of transcript present as a function of time. 1. Chemical inhibitors: The inhibition of mRNA synthesis in plants is usually accomplished by treatment with drugs such as actinomycin D, cordycepin and 0t- amanitin. Actinomycin D interferes with transcription by intercalating into the DNA, while cordycepin acts as a chain terminating adenosine analog and has been shown to be a more efficacious inhibitor than actinomycin D in leaf tissue (Holtorf et al., 1999). or- amanitin is an inhibitor of eukaryotic RNA polymerases II and III and blocks transcription by binding to the polymerase. Since these drugs inhibit the transcription of many genes, half-lives of several mRNAs can be measured simultaneously, which is especially important for genomic analysis. Such experiments are however most appropriate for measuring half-lives of short—lived mRNAs because of the potential deleterious effects of prolonged exposure to the drug. Also, it has been observed that the use of these drugs may increase the half-lives of some transcripts due to potential loss of labile factors involved in mRNA degradation (Dickey et al., 1994). In addition, the magnitude of the change in mRNA half-life in response to a particular stimulus may be dampened (Zhang et al., 1993). 2. Regulated promoters: A more specific method for measuring mRNA half-lives is to use a promoter that can be regulated to conditionally produce the transcript of interest. The rationale behind this approach is to first provide a stimulus that leads to a burst of mRNA synthesis followed by removal of the stimulus which causes rapid repression of synthesis. As a result, the mRN A of interest is synthesized for only a brief time, increases in abundance during that time and then degrades at a rate dependent on its half-life. Using this method, mRN A synthesis is repressed without resorting to toxic chemicals but with a high degree of synchrony (Ross, 1995). Also, it is unlikely that decay rates of mRNAs will be altered by depletion of labile turnover factors as they may be when global inhibitors of transcription are employed. An ideal inducible gene expression system should have low basal expression levels, high and specific inducibility, fast and efficient switch-off after the removal of the inducer and low toxicity. Inducible promoter systems that satisfy some of the aforementioned criteria have been developed in plants (Aoyama and Chua, 1997; Caddick et al., 1998; Padidam et al., 2003). In contrast to the other positively regulated promoters, the negatively regulated ToplO promoter has been used most frequently (Gossen and Bujard, 1992). The ToplO promoter sequence, which has been used in plants, contains seven copies of the tetracycline operator DNA sequence fused to a truncated version of the CaMV 358 promoter (Weinmann et al., 1994). The ToplO promoter sequence is recognized by a transactivator that acts as a synthetic transcription factor. The transactivator is a chimeric fusion protein between the operator-binding portion of the bacterial tetracycline repressor fused to the activation domain of the herpes simplex virus transcription factor, VP16 (tTA) (Weinmann et al., 1994). In the absence of tetracycline, the tetracycline repressor region of tTA has a strong affinity for the operator DNA sequences within the ToplO promoter resulting in the expression of the desired gene. In contrast, in the presence of tetracycline, tTA binds tetracycline and is rapidly inactivated, leading to the repression of transcription. This strategy has been successfully used to measure F erredoxin-l mRN A stability in response to photosynthesis in transgenic tobacco plants (Petracek et al., 1998) and also to evaluate mRNA sequences that control the rate of SA UR-A C1 transcript decay in transgenic tobacco cell lines (Gil and Green, 1996). Recently, it was shown that the ToplO promoter system is functional in Arabidopsis as well. Love et a1. (2000) generated a homozygous Arabidopsis line in which tTA was expressed from a single transgenic locus. Individual plants from this line were then crossed with transgenic plants that expressed the ER-targeted green fluorescent protein (ER-GFP) from ToplO promoter. In the resulting progeny, GF P expression was seen to be stringently controlled by teracycline and the repression of the promoter was reversible upon tetracycline removal. This system is thus an encouraging new approach for measuring mRNA half-lives in Arabidopsis. 3. RNA polymerase II: In yeast, an alternative method for inhibiting mRNA synthesis is to use a temperature-sensitive allele of the RPB] gene (rpr-I), which encodes the large subunit of RNA polymerase II (Herrick et al., 1990). Use of the rpr -1 mutation to inhibit transcription again allows the half-lives of many mRNAs to be measured simultaneously and is also attractive because it promises to have fewer nonspecific effects compared to general transcriptional inhibitors. In theory it should be possible to extend the same technology to plants. Arabidopsis has a single gene for the large subunit of Pol II (Dietrich etal., 1990) and the residue corresponding to the ts Pol 11 mutation (rpbl-I) (Scafe et al., 1990) is conserved. Recent work by Maillet et al. (1999) also indicated that a null mutant of RPB4 renders Pol II temperature sensitive in yeast and the kinetics for mRNA decay after a temperature shift are the same as in the rpr-I mutant. 4. Microarray technology With the application of genomics, monitoring mRNA stability can be achieved on a scale much larger than previously possible. Numerous groups have used DNA microarray analysis to determine mRNA half-lives on a global basis in both prokaryotic and eukaryotic systems (Bernstein et al., 2002; Holstege et al., 1998; Lam et al., 2001; Wang et al., 2002). DNA microarrays have also been used to estimate mRNA half-lives for the genes of wild type Arabidopsis, by hybridizing spotted cDNA AF GC arrays to probes corresponding to multiple time points after shutting off RNA synthesis with cordycepin (Gutierrez et al., 2002). This study indicated that at least 1% of Arabidopsis transcripts represented on that array are rapidly degraded, with estimated half-lives of less than 60 minutes. Additional transcripts with half-lives of less than 120 minutes were also identified. The microarray results were very reliable and were confirmed by statistical analysis and by performing conventional half-life measurements with multiple time points quantitated on RNA gel blots (Gutierrez et al., 2002). Ultimately, the ability to monitor mRNA decay on microarrays should help in the identification of transcripts that are differentially regulated at the level of mRNA stability in response to various stimuli. STIMULI AFFECTING mRNA STABILITY 1. Plant hormones Hormones regulate gene expression at many levels but an in depth understanding of their affect at the post-transcriptional level is lacking. The increase in the abundance of the mRNA encoding the light-harvesting chlorophyll a/b binding protein in response to cytokinin application in Lemna gibba occurs principally by a post-transcriptional mechanism (Flores and Tobin, 1988). mRNA stability has been shown to be responsible for the cytokinin-induced accumulation of SrEnodZ mRN A in Sesbania rostrata (Silver et al., 1996). Downes and Crowell (1998) demonstrated that in cytokinin-starved soybean culture cells, Cim] mRN A is stabilized upon cytokinin treatment. The Cim] protein product is similar to the B-expansin proteins which are involved in cell wall expansion. Cim] abundance increases 20-60-fold within four hours of cytokinin addition to cytokinin-starved soybean suspension cultures and analysis of Cim] stability revealed a greater than 4-fold increase in the half-life of the mRNA in response to cytokinin. In addition, Cim] accumulation is stimulated in the absence of cytokinin by the kinase inhibitor staurosporine and inhibited in the presence of cytokinin by the phosphate inhibitor okadaic acid, suggesting a role for protein dephosphorylation in cytokinin regulation of Cim] abundance (Downes and Crowell, 1998). It has also been suggested that the increase in the abundance of the tomato ER] transcript in response to ethylene (Lincoln and Fischer, 1988) and the wheat Em transcript in response to ABA (Williamson etal., 1985) is controlled at the level of transcript stability but as yet there is no direct evidence. 2. Light Light affects plant gene expression at many levels, including mRNA stability. The ferredoxin-encoding genes Fed-1 from pea (Dickey et al., 1998) and FedA from Arabidopsis (Vorst et al., 1993) show increased mRNA accumulation in light-grown leaves. The F edA transcript is 20-fold higher in the light relative to the dark, while the transcriptional activity is only 2-fold higher. This discrepancy suggests that the transcript is either stabilized in the light or destabilized in the dark. The pea Fed-1 mRNA is 5-fold higher in the light than in the darkness. Using the ToplO promoter in transgenic tobacco plants, it was determined that Fed-1 mRNA is post-transcriptionally regulated by light at the level of mRNA stability as its half-life is significantly reduced in the dark (Petracek et al., 1998). The region mediating the rapid degradation of the Fed-1 mRNA includes the 5’ UTR and a portion of the coding region 10 and is called the iLRE (internal light response element; Dickey et al., 1992). In order for the Fed-1 mRNA to be light responsive, an open reading frame is required (Dickey et al., 1994), suggesting that the light effect on Fed-1 mRNA is dependent on its synchronous translation. In addition, the insertion of nonsense codons within the Fed-1 coding sequence disrupts the light regulation of Fed-1 mRNA abundance (Petracek et al., 2000). When the Fed-1 gene from pea is expressed in tobacco plants grown in light, Fed-1 mRNAs are found in high molecular weight polysomes. Following a shift to the dark, the transcript rapidly dissociates from polysomes (Petracek et al., 1998), indicating that translation of the Fed-1 mRNA is repressed by dark and this may trigger rapid degradation of the Fed-1 mRNA. Treatment of plants with the electron transport inhibitor DCMU also results in destabilization of the Fed-1 mRNA (Petracek et al., 1998). Therefore, it may be the cessation of photosynthesis when plants are shifted to dark that triggers Fed-1 decay, rather than the mere absence of light. In pea, a single pulse of high-fluence blue light results in an increased rate of Lhcb transcription, with no change in the steady-state level of Lhcb mRN A (Kaufman, 1993), suggesting that the excitation of the high-fluence blue light system results in destabilization of the Lhcb mRN A. Anderson et al. (1999) demonstrated that Lhcb RNA levels in etiolated Arabidopsis are also regulated in a similar manner by the high-fluence blue light system and the 65 bp 5’ UTR of AtLhcb is necessary and sufficient for RNA destabilization. The blue light-induced destabilization response is not dependent on the phytochrome or the cryptochome receptors (Anderson et al., 1999; F olta and Kaufman, 1999) but the phototropin 1 photoreceptor seems to required for the blue-light mediated destabilization of the Lhcb transcript (Folta and Kaufman, 2003). ll 3. Sucrose a-amylases are major amylolytic enzymes and play an important role in the degradation of starch. In rice germinating embryos and cultured suspension cells, expression of a-amylase genes is activated by sugar depletion and suppressed by sugar provision (Yu et al., 1996). The half-lives of orAmy3, aAmy7 and aAmy8 have been shown to be prolonged by sucrose starvation, with the mRNA half-life of aAmy3 increasing from 1.5 h to 6 h in sucrose-starved cells (Sheu et al., 1996); however, the stability of these three mRNAs appears to be controlled by different mechanisms. The translation inhibitors cylcohexamide and anisomycin enhanced the accumulation of aAmy3 mRNA regardless of whether or not the cells were provided with sucrose, while the accumulation of orAmy7 and orAmy8 mRNA was suppressed by the inhibitors, even in cells starved for sucrose. Moreover, cyclohexamide did not significantly alter the transcription rates of a-amylase genes, suggesting that labile proteins are involved in the stabilization of the OLAmy7 and orAmy8 mRNAs but destabilization of the aAmy3 mRNA. Examination of chimeric gene expression in stably transformed rice cells has shown that the regulatory sequences in the orAmy3 3' UTR act as potent determinants of mRNA stability in response to sugar availability (Chan and Yu, 1998). A recent study conducted by Ho et al. (2001) showed that sugar activated or repressed gene expression in rice suspension cell cultures and this regulation Operates at the levels of both transcription rate and mRNA stability. Amongst the sucrose- upregulated genes that were tested, the half-lives of actin, glyceraldehyde 3-phosphate dehydrogenase, alcohol dehydrogenase and sucrose synthase P-2 mRNAs were approximately 1.6-2.6-fold longer in sucrose-provided cells than in sucrose-starved cells. 12 For all the tested sucrose-downregulated genes, the half-lives of mRNAs were 25-74- fold higher in sucrose-starved cells than in sucrose-provided cells (Ho et al., 2001). 4. Nitrogen Nitrogen is an essential plant macronutrient and plants obtain nitrogen either from the soil or through symbiotic nitrogen fixation. Nitrate and nitrogen are reduced to ammonia, which is assimilated using two enzymes Gln synthase (GS) and Glu synthase (GOGAT). In higher plants, GS occurs as two isoforms, GS; in the cytoplasm and G82 in plastids. To determine if GS, genes in alfalfa were regulated at the post-transcriptional level, Ortega et al. (2001) analyzed transgenic alfalfa plants containing the soybean GS. gene under the control of the CaMV 358 promoter. A 3-4-fold drop in the level of the soybean GS] transcript in transgenic alfalfa plants, that were fed nitrate over their non- nitrate-fed counterpart, was observed. This difference in transcript levels could not be attributed to differential promoter activity and furthermore the transcript for the alfalfa endogenous GS] gene also showed a 3-fold drop in levels in the leaves of nitrate-fed plants. These results indicate that the decrease in GS. transcript levels in nitrate-fed plants is most likely due to increased turnover of the GSI transcript. 5. Methionine Methionine biosynthesis is tightly regulated in plants and the first committed step in methionine biosynthesis is catalyzed by cystathione y-synthase (CGS), suggested to be a major regulatory site of the pathway (Ravanel et al., 1998). Analysis of Arabidopsis mto mutants that overaccumulate soluble methionine demonstrated that the gene for CGS l3 is regulated at the level of mRNA stability (Chiba et al., 1999). Treatment of Arabidopsis calli with actinomycin D showed that turnover of CGS mRNA, in the absence of methionine, was faster in wild type than the mtoI-I mutant. Methionine treatment accelerated the turnover in wild type but not in the mutant. It was also demonstrated that a 11-13 amino acid region encoded by the first exon of the CGS gene is sufficient and essential for down-regulating its own mRN A stability in response to methionine or one of its metabolites (Suzuki et al., 2001; Ominato et al., 2002) and ongoing translation is probably required (Lambein et al., 2003). 5. Biotic stress In plants, an important process activated by infection or wounding is the remodeling of the cell wall (Mehdy and Brodl, 1998). A key constituent of the response to pathogens appears to be the degradation of a subset of preexisting mRNAs. An excellent example of the regulation of mRNA stability in response to biotic stress has been characterized in Phaseolus vulgaris. In bean cells treated with fungal elicitor, the transcripts of PvPRP] , a gene encoding a proline-rich protein believed to be a component of the cell wall, decrease to ~6% of the original level within 4 hr (Zhang et al., 1993). After actinomycin D treatment, the PvPRP] transcript has a half-life of 60 hr in the absence of fimgal elicitors and 18 hr in the presence of fungal elicitors. Furthermore, transcriptional rates remain constant regardless of the presence or absence of the elicitor. A subsequent study by Zhang and Mehdy (1994) identified a 50 kDa protein, PRP-BP, which specifically binds to a U-rich sequence in the 3' UTR of PvPRP] mRNA (see later for details). The RNA-binding activity of PRP-BP is redox regulated in vitro and its 14 activity is induced in response to fungal elicitor treatment prior to the onset of PvPRP] mRNA degradation. On the basis of PRP-BP activation upon elicitor treatment and its specificity of RNA binding, it has been postulated that the interaction of PRP-BP with PvPRP] mRNA may be one component in the process of elicitor-induced PvPRP] mRNA degradation (Mehdy and Brodl, 1998). In soybean cells, glucan elicitors induce a decrease of the tubBl (B-tubulin 1 isoform) transcript level, most likely by enhancing its degradation, while the transcript level of another tubulin isoform tubBZ is not affected by this elicitor (Ebel et al., 2001). Direct evidence that the major control of this down-regulation is at the level of mRNA stability was provided by the observation that in the presence of cordycepin, the half-life of tubBI mRNA decreased upon addition of the elicitor. Pre-incubation with Ca2+ modulators blocked the decrease of tubB] mRN A levels, suggesting that calcium might be involved in the regulation of tubB] message levels. The down-regulation of tubBI mRN A levels induced by these elicitors could result from a general redirection of the available cellular resources to defense-related metabolism (Ebel etal., 2001). In mammalian cells, B-tubulin mRNAs are destabilized in the presence of free tubulin heterodimers (Theodorakis and Cleveland, 1992) and a similar mechanism could be occurring in plant cells where tubBImRNA is degraded due to depolymerization of specific microtubules. 6. Abiotic stress Abiotic stress may result from water deficit or excess, extreme heat or cold, toxic substances, salinity or drought and ultraviolet light. Like biotic stress, the regulation of 15 mRNA stability is an important step for the regulation of some genes in the abiotic stress responses. The steady-state level of transcripts encoding the pyrroline-S-carboxylate reductase of Arabidopsis (At-P5R) is upregulated under salt and heat stress and this induction is mainly due to an enhanced mRNA stability (Hua et al., 2001) mediated by its 5' UTR. In carrots, heat shock disrupts the function of the 5' cap and the poly(A) tail structures, resulting in the loss of translational competence but increases in mRNA stability (Gallie et al., 1995). In wheat, it has been reported that the activity of RNases decreases following heat shock and is in correlation with an increase in mRNA half-life (Chang and Gallie, 1997). In gibberellic acid-stimulated barley aleurone layers, heat shock reduces the half-life of a-amylase mRNA (Belanger et al., 1986) and other secreted hydrolases but does not affect the mRNA levels for non-secretory proteins (Brodl and Ho, 1991). Heat shock also decreases the levels of some wound-inducible mRNAs encoding extra-cellular proteins in carrot root disks and using cordycepin as a transcriptional inhibitor, it was shown that this decrease in mRNA levels is due to accelerated mRN A turnover during heat shock (Brodl and Ho, 1992). Plants vary widely in response to cold temperature and there is evidence that altered gene expression occurs during cold acclimation (Thomashow, 1998). The increase in transcript levels for some cor (cold-regulated) genes from Arabidopsis, in response to cold stress, has been shown to be mainly at the posttranscriptional level, possibly due to increased transcript stability (Hajela et al., 1990). In alfala, transcripts of a cold acclimation-specific gene c0518 are destabilized upon return to warm temperature (Wolfraim et al., 1993). 16 cis- ACTING DETERMINANTS OF mRNA STABILITY The majority of mRNAs fall in the stable range of mRNA half-lives for a given organism (Taylor and Green, 1995; Ross, 1996). This observation led to the hypothesis that mRNAs might have sequence elements that can either act constitutively to establish the inherent instability of a particular transcript or modulate the stability of an mRNA in response to certain physiological, developmental, or environmental cues. Many cis-acting elements have been identified that target transcripts for rapid turnover in plants as well as other systems (reviewed in Abler and Green, 1996; Ross, 1995). A number of studies conducted over several years have suggested a function for the 7-methyl-G cap at the 5' end and the poly(A) tail at the 3' end, two cis-acting determinants that are common to all plant mRN As, as mRN A stabilizing structures. It is known that in the major mRNA degradation pathway of yeast these stabilizing structures are removed, or at least in the case of the poly(A) tail, greatly shortened, prior to degradation catalyzed by exoribonucleases (Decker and Parker, 1993). l. DST element In plants, an instability determinant called DST (Newman etal., 1993) has been studied in detail. The DST (gowngtream) element was first identified as a highly conserved sequence in the 3' UTRs of the soybean SA UR genes (McClure et al., 1989). The SA UR (small-auxin-up—RNAS) genes encode unstable transcripts whose half-lives have been estimated to be on the order of 10-50 minutes (McClure and Guilfoyle, 1987; Franco et al., 1990). Although the function of the SAUR proteins is unknown, the temporal and spatial expression of SA UR genes correlates with auxin-induced cell 17 elongation (McClure and Guilfoyle, 1989). Detailed studies of SA UR gene expression in Arabidopsis have been carried out on the SA UR-A C] gene, which contains the DST element in its 3' UTR. Examination of chimeric gene expression has shown that the promoter region is responsible for auxin induction, the 3' UTR is largely responsible for rapid mRN A turnover and the coding region contributes to low mRN A abundance but not by decreasing mRNA half-life (Gil and Green, 1996). A synthetic dimer of the DST element has been demonstrated to be sufficient to destabilize normally stable reporter transcripts in stably transformed tobacco cell suspension cultures (Newman et al., 1993). 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 prototype DST, from SA UR15A of soybean, is about 45 base pairs in length and consists of three highly conserved subdomains separated by two variable regions (Newman et al., 1993). Site-directed mutagenesis studies have been performed in order to determine which features of the DST element are critical for its instability function (Sullivan and Green, 1996). Residues within the conserved second and third subdomains, the ATAGAT and the GTA regions respectively, have been found to be necessary for the DST element to function as an instability determinant. As mentioned above, the DST element has 3 conserved subdomains, of which the ATAGAT and the GTA subdomains are critical for its instability function. Five- and six- base substitutions in the ATAGAT and the GTA regions resulted in inactivation of the instability function of the DST element in suspension cell cultures as well as in transgenic plants, while smaller two-base substitution mutations resulted in inactivation of DST in 18 transgenic tobacco leaves but had varying effects on DST function in tobacco cell culture (Sullivan and Green, 1996). These results suggest that the DST element might be recognized differently in different cell types. Interestingly, the SA UR-A C] 3' UTR contains one canonical DST element located 80 bp downstream of the stop codon and 10 bp upstream of the poly(A) addition site (Gil et al., 1994). Earlier experiments have indicated that two copies of the synthetic DST element are required for instability function (Newman et al., 1993). 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 C] 3' UTR. These sites may be serving as multiple recognition sites for DST-mediated decay within the SA UR-AC] 3' UTR. Recently additional genes have been identified that have DST-like subdomains rather than a classical DST sequence like those in soybean SA UR genes (Pérez-Amador et al., 2001). Furthermore there is accumulating data in favor of multiple subdomains being sufficient for instability function (Feldbriigge et al., 2002). Thus far, the DST element appears to be unique to plants. Visual inspection of known eukaryotic instability determinants has not led to the identification of any elements that contain all the characteristic features of a DST element. However, similarities have been noticed in sequence between the GTA region and animal mRNA elements which are bound by proteins containing a Pumilio-like RNA-binding domain (Feldbrfigge et al., 2002; Zamore et al., 1999). To investigate whether the plant DST mRNA instability determinant is also recognized in mammalian cells, DST element variants were tested in mouse NIH3T3 fibroblasts, a well-defined model system to monitor mRNA decay in mammalian cells (Zubaiga et al., 1995). From the 19 aforementioned experiments, it seems that the plant DST element is not recognized in animal cells with the same sequence requirements as in plant cells. Therefore, its mode of recognition may be plant-specific (Feldbriigge et al., 2002). Interestingly, the GTA region interacts with the human Pumilio-like protein HsPUM in in vitro binding studies (Feldbriigge et al., 2002). This raises the intriguing possibility that this region is recognized by Arabidopsis proteins containing the same type of RNA-binding domain i.e. AtPUMs. 2. AUUUA sequences The most widely studied instability determinants in mammalian cells are the AU- rich elements (ARES). ARES are found in the 3' UTRs of several of the most unstable mammalian transcripts, such as lymphokine, cytokine and proto-oncogene mRNAs (Chen and Shyu, 1995). Repeats of the pentamer, AUUUA, are often found in these AU-rich elements and have been shown to be important for their instability function (Shyu et al., 1991; Vakalopoulou et al., 1991). Many functional ARES mediate deadenylation as the first step in mRNA decay, although different classes of ARES exhibit different reaction kinetics (Chen and Shyu, 1995). AUUUA repeats are likely to be of broad significance in higher eukaryotes, since they can target transcripts for rapid decay in plants (Ohme-Takagi et al., 1993), and recent evidence suggests that the ARE-mediated decay pathway is functional in yeast as well (Vasudevan and Peltz, 2001). Reporter transcripts containing 11 repeats of the AUUUA motif were degraded more rapidly in plants as compared to an AU-rich control lacking AUUUA repeats (Ohme-Takagi et al., 1993). AUUUA motifs present in the 3' 20 UTR of PvPRPI have been proposed to trigger mRNA degradation (Zhang and Mehdy, 1994) in common bean. Three AUUUA motifs are present in the 3' UTR of tubB] but none have been found in the 3' UTR of tubBZ, which suggests differential regulation via degradation or stabilization of the message for the two different tubulin isoforms, as discussed earlier (Ebel et al., 2001). In rice, the entire aAmy3 3' UTR and two of its subdomains can independently mediate sugar-dependent repression of reporter mRN A accumulation (Chan and Yu, 1998). Examination of the nucleotide sequences has revealed that domains I and 111 each contain a stretch of a 9-bp AU-rich conserved sequence. Moreover, RNA structure prediction of the 3' UTR identified extensive regions of putative duplex formation, and regions encompassing domains I, II and 111 each contained a putative stem-loop structure. Also the 9-bp conserved AU-rich sequence is located in the loop regions of both domains I and III (Chan and Yu, 1998). 3. 5' UTR Theoretically, the half-life of all mRNAs can be affected by how its 5' UTR influences its translational efficiency. Light-mediated changes in transcript stability are known to occur for the Fed-1 mRNA in pea (Dickey et al., 1992). A major light response element, iLRE, in the pea F ed—I gene is located within the transcription unit and spans a portion of the 5' UTR and the first 20 codons of the coding region. A CATT repeat element, located near the 5' UTR, has been identified as being essential for light regulation (Dickey et al., 1998). This element is important for mRNA stability since two different mutations in the CATT repeat element altered dark-induced F ed—I mRNA disappearance. Recently it was demonstrated that mRNA containing the F ed-I iLRE 21 ceases translation and dissociates from polyribosomes soon after plants are transferred from light to darkness, providing support for the model that Fed-1 mRNA is protected from degradation in light by association with ribosomes and/or the act of translation (Hansen et al., 2001). The 5' UTR of the pea Lhcb] *4 transcript contains a sequence involved in the regulation of transcript stability (Anderson et al., 1999). No known specific mechanisms by which the destabilization mediated through the 5' UTR occurs have been reported. However, it is plausible to postulate that the element somehow stalls translation, exposes the RNA which is then subject to digestion by RN ase activity (Abler and Green, 1996). The first 92 bp region of the At-P5R 5' UTR is sufficient to mediate transcript stabilization during salt and heat stress (Hua et al., 2001). In silica analysis has predicted that extensive secondary structures in the 5' UTR and high GC content further stabilize the secondary structures. The At-P5R 5'UTR also contains a 26 bp sequence, repeated seven times, which has sequence identity with a part of the 3' non-coding region of a potential retrovirus (Hua et al., 2001). It is possible that the 26 bp region binds protein factors stabilizing the secondary structures during stress. GENERAL mRNA DECAY MACHINERY In contrast to the factors that are involved in the recognition of specific sequences within mRNAs, proteins which are responsible for general mRNA degradation can be considered to constitute the basal mRNA degradation machinery. The basal mRNA decay machinery catalyzes the degradation of mRNAs from many different genes, while 22 sequence specific mRNA binding proteins probably recruit the basal mRNA decay machinery to specific target molecules, or modulate the degradative activity of the basal machinery. In yeast, mRNA turnover is mediated by mRNA decay pathways that use shared, as well as distinct components of the basal mRNA decay machinery (reviewed in Tourriere et al., 2002). It is likely that mRNA decay in Arabidopsis, as well as in other higher plants also occurs through multiple degradation pathways. There are several Arabidopsis homologs of the yeast basal mRNA decay machinery suggesting that mRN A turnover in Arabidopsis may resemble the mRN A decay pathways of yeast (Gutierrez et al., 1999). However, plant-specific mRNA decay pathways and mechanisms are also likely to be functioning since apparent differences have been observed between the two systems (Kastenmayer et al., 2001). The mechanisms responsible for mRNA degradation are most well understood in yeast. In yeast, the deadenylation-dependent-decapping pathway appears to be the main pathway for mRNA degradation. In this pathway, mRNAs are deadenylated by a complex of proteins including Caflp and Ccr4p (Daugeron et al., 2001; Tucker and Parker, 2000). This deadenylation reaction shortens the poly(A) tail to a length of 10—12 adenylates (Decker and Parker 1993). Shortening of the poly(A) tail then triggers decapping catalyzed by Dcplp (LaGrandeur and Parker, 1998) and associated factor Dcp2p (Dunckley and Parker, 1999), after which the decapped transcripts are degraded from the 5’ end by the 5’-3’ exoribonuclease Xmlp (Muhlrad et al., 1994). In addition to the major mRNA decay pathway, several minor mRNA decay pathways operate in yeast. The nonsense mediated decay pathway, responsible for the degradation 23 of mRNAs containing premature nonsense codons, is similar to the major decay pathway; however, 5’-3 degradation catalyzed by Xmlp is not dependent on prior deadenylation (Muhlrad and Parker, 1994). Degradation of mRNAs is also catalyzed from the 3’ end following deadenylation by a multi-protein complex, the exosome. The exosome and its associated co-factors in yeast consists of greater than 15 proteins, at least 10 of which are homologous to 3’-5’ exoribonucleases from E. coli (van Hoof and Parker, 1999). SEQUENCE-SPECIFIC DECAY Sequence-specific mRN A degradation determines the degradation rate of particular mRNAs. The instability determinants can be thought to target certain transcripts to the mRNA decay machinery. In mammals, sequence-specific RNA binding proteins have been identified that interact in vitro with sequences that regulate mRN A stability. A number of proteins have been hypothesized to fimction in ARE-mediated decay based on in vitro activity, the best characterized of which are the human HuR/HuA protein and AUFl/hnRNP D. HuR belongs to a family of RNA-binding proteins related to the Drosophila embryonic lethal abnormal visual (ELAV) proteins. Overexpression of HuR in vivo (Fan and Steitz, 1998) and increased levels of HuR or other ELAV-like proteins in vitro (Ford et al., 1999) lead to stabilization of ARE-containing mRNAs. In contrast, in vivo depletion of AUFl/hnRNP D leads to a strong stabilization of diverse ARE-containing mRNAs whereas ectopic expression restores destabilization (Loflin et al., 1999). These data suggest that HuR and hnRNP D have antagonistic effects on the stability of ARE-containing mRNAs. 24 For another mammalian protein, tristetraprolin, a mouse insertional mutant was available, and demonstrating a role in ARE-mediated decay was more definitive. Half- life measurements showed that TNF-a mRN A, which contains AUUUA elements in its 3'UTR, was stabilized in cells from the tristetraprolin deficient mutant relative to wild- type (Carballo et al., 1998). Subsequently, the protein was shown to bind AUUUA sequences in vitro indicating that it functions to target TNF-a for ARE-mediated decay (Lai et al., 1999). More recently, it has been shown that tristetraprolin deficient mice are also defective in the deadenylation of ARE-containing GM-CSF mRNA (Carballo et al., 2000) To date, no mRNA-binding proteins have been identified in plants that are known to play a role in the control of mRN A stability. One plant mRNA-binding protein which may function in controlling stability is a 50 kDa protein that binds to a sequence in the 3' UTR of the bean PvPRPI mRN A (Zhang and Mehdy, 1994). The PvPRP] transcript, which appears to encode a cell wall protein, is rapidly degraded following the addition of fungal elicitor. 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 in response to fungal elicitor treatment, consistent with this scenario. Genetic approaches have rarely been applied to mRN A stability problems in multicellular organisms. The majority of the studies aimed at understanding the cellular factors in mRNA decay pathways have involved characterization of proteins that bind the instability sequences in vivo. Many RNA-binding proteins have been isolated that interact with ARES in vitro (discussed above). Isolation of sequence-specific RNA- 25 binding proteins, involved in mRNA stability, from plant cells has not been as successful, perhaps because it is difficult to prepare cytoplasmic protein extracts that are free of non- specific ribonucleases. This may be due to the presence of the plant vacuole, which is known to contain many non-specific ribonucleases and other hydrolytic enzymes that are released upon cell rupture during cell extract praparation. One exception iS the recent isolation of mutant mammalian cell lines that stabilize green fluorescent protein (GFP)-IL-3 reporter transcripts (Stoecklin et al., 2000). In this screen, two loci involved in rapid mRNA turnover mediated by the interleukin-3 (IL-3) 3'UTR were identified by screening mutagenized human HT1080 fibrosarcoma cells for elevated green flourescence. One of the mutants was rescued by a cDNA corresponding to butyrate-response factor-1 (BRFl), which is a Zn finger protein homologous to tristetraprolin (Stoecklin et al., 2002). In order to understand the molecular basis of sequence-specific recognition and degradation of unstable mRNAs, an approach was devised to isolate Arabidopsis mutants defective in DST-mediated mRNA degradation (Johnson et al., 2000). This approach allows several important advantages, such as: 1. Genes identified in a mutant selection would be very likely to play a role in mRNA degradation in vivo. Of particular interest would be the cellular factors that might be involved in sequence-specific interactions. 2. Basic information about the mechanisms of mRNA degradation may be obtained by studying mutants since a genetic approach offers an alternative to biochemical approaches that have proved difficult. 26 dst], dst2 and dst3 were isolated based on their ability to stabilize specific DST- containing transgene mRNAs (Johnson et al., 2000; Chapter 5). For this strategy, two transgenes, hygromycin phosphotransferase (HPH) and B-glucuronidase (GUS), both containing a tetramer of the consensus DST element in the 3' UTR, were introduced into Arabidopsis as selectable and screenable markers, respectively. The presence of the DST elements decreased the mRN A levels for both transgenes, HPH-DST and GUS-DS T, resulting in decreased resistance to hygromycin and low GUS activity. The dst mutants were isolated as hygromycin resistant plants with increased GUS activity due to stabilization of the corresponding mRNAs. In addition, dst mutants have elevated levels of SA UR-A C1 mRN A, the only endogenous DST-containing unstable transcript characterized thus far (Gil and Green, 1996; Johnson et al., 2000). The dst mutants that we isolated are extremely rare (3/~800,000), are all codominant and exhibit no obvious morphological or developmental phenotype (Johnson et al., 2000). The simplest explanation to account for the mutant phenotype observed in the dst mutants would be that the gene corresponds to a sequence-Specific RNA-binding protein or a ribonuclease. Alternatively, it could encode for a regulatory factor that controls one of these components. To evaluate the physiological significance of the DST-mediated mRNA degradation pathway, we sought to identify additional molecular markers of dst] using DNA microarrays. In addition to the identification of novel targets of dst], the microarray experiments provided the first indication of an association between circadian rhythms and DST-mediated decay (Pe'rez-Amador et al., 2001; Chapter 3). 27 PERSPECTIVE Recent advances have begun to shed light on the mechanisms that underlie mRNA turnover in eukaryotic systems. Exciting progress has been made with cis- and trans-acting determinants of mRN A stability, but there is much to be learned about how other factors such as exogenous or endogenous stimuli interact with and affect mRN A decay. The identity of all the players participating in the mRNA degradation process as well as the rules governing Specific RNA degradation remains to be elucidated. By combining the power of genomics with traditional genetic and biochemical approaches, it should be possible to better define the influence of transcript stability on gene expression. 28 REFERENCES Abler, M. L., and Green, P. J. (1996). Control of mRNA stability in higher plants. Plant Mol Biol 32, 63-78. 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, 1579-1590. Aoyama, T., and Chua, N. H. (1997). A glucocorticoid—mediated transcriptional induction system in transgenic plants. Plant J 11, 605-612. Belanger, F. C., Brodl, M. R., and Ho, T. H. (1986). Heat shock causes destabilization of specific mRNAs and destruction of endoplasmic reticulum in barley aleurone cells. Proc Natl Acad Sci U S A 83, 1354-1358. Bernstein, J. A., Khodursky, A. B., Lin, P. H., Lin-Chao, S., and Cohen, S. N. (2002). Global analysis of mRN A decay and abundance in Escherichia coli at single-gene resolution using two-color fluorescent DNA microarrays. Proc Natl Acad Sci U S A 99, 9697-9702. Brodl, M. R., Ho T.h-D. (1991). Heat shock causes selective destabilization of secretory protein mRNAs in barley aleurone cells. Plant Physiol 96, 1048-1052. Brodl, M. R., Ho T.h-D. (1992). Heat shock in mechanically wounded carrot root discs causes the destabilization of stable secretory protein mRNA and the dissociation of endoplasmic reticulum lamellae. Physiol Plant 86, 253-262. Caddick, M. X., Greenland, A. J., Jepson, 1., Krause, K. P., Qu, N., Riddell, K. V., Salter, M. G., Schuch, W., Sonnewald, U., and Tomsett, A. B. (1998). An ethanol inducible gene switch for plants used to manipulate carbon metabolism. Nat Biotechnol 16, 177-180. Caponigro, G., and Parker, R. (1996). Mechanisms and control of mRNA turnover in Saccharomyces cerevisiae. Microbiol Rev 60, 233-249. Carballo, R, Lai, W. S., and Blackshear, P. J. (1998). Feedback inhibition of macrophage tumor necrosis factor-alpha production by tristetraprolin. Science 281, 1001-1005. Carballo, E., Lai, W. S., and Blackshear, P. J. (2000). Evidence that tristetraprolin is a physiological regulator of granulocyte-macrophage colony-stimulating factor messenger RNA deadenylation and stability. Blood 95, 1891-1899. Chan, M. T., and Yu, S. M. (1998). The 3' untranslated region of a rice alpha-amylase gene mediates sugar-dependent abundance of mRN A. Plant J 15, 685-695. 29 Chang, S. C., and Gallie, D. R. (1997). RNase Activity Decreases following a Heat Shock in Wheat Leaves and Correlates with Its Posttranslational Modification. Plant Physiol 113, 1253-1263. Chen, C. Y., and Shyu, A. B. (1995). AU-rich elements: characterization and importance in mRNA degradation. Trends Biochem Sci 20, 465-470. Chiba, Y., Ishikawa, M., Kijima, F., Tyson, R. H., Kim, J., Yamamoto, A., Nambara, E., Leustek, T., Wallsgrove, R. M., and Naito, S. (1999). Evidence for autoregulation of cystathionine gamma-synthase mRNA stability in Arabidopsis. Science 286, 1371-1374. Daugeron, M. C., Mauxion, F ., and Seraphin, B. (2001). The yeast POP2 gene encodes a nuclease involved in mRNA deadenylation. Nucleic Acids Res 29, 2448-2455. Decker, C. J ., and 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. DeRisi, J. L., Iyer, V. R., and Brown, P. O. (1997). Exploring the metabolic and genetic control of gene expression on a genomic scale. Science 278, 680-686. Dickey, L. F ., Gallo-Meagher, M., and Thompson, W. F. (1992). Light regulatory sequences are located within the 5' portion of the Fed-1 message sequence. Embo J 11, 2311-2317. Dickey, L. F., Nguyen, T. T., Allen, G. C., and Thompson, W. F. (1994). Light modulation of ferredoxin mRN A abundance requires an open reading frame. Plant Cell 6, 1171-1176. Dickey, L. F ., Petracek, M. E., Nguyen, T. T., Hansen, E. R., and Thompson, W. F. (1998). Light regulation of Fed-1 mRNA requires an element in the 5' untranslated region and correlates with differential polyribosome association. Plant Cell 10, 475-484. Dietrich, M. A., Prenger, J. P., and Guilfoyle, T. J. (1990). Analysis of the genes encoding the largest subunit of RNA polymerase II in Arabidopsis and soybean. Plant Mol Biol 15, 207-223. Downes, B. P., and Crowell, D. N. (1998). Cytokinin regulates the expression of a soybean beta-expansin gene by a post-transcriptional mechanism. Plant Mol Biol 37, 437-444. Dunckley, T., and Parker, R. (1999). The DCP2 protein is required for mRN A decapping in Saccharomyces cerevisiae and contains a functional MutT motif. Embo J 18, 5411- 5422. 30 Ebel, C., Gomez, L. G., Schmit, A. C., NeuhauS-Url, G., and Boller, T. (2001). Differential mRN A degradation of two beta-tubulin isoforrns correlates with cytosolic Ca2+ changes in glucan-elicited soybean cells. Plant Physiol 126, 87-96. Fan, X. C., and Steitz, J. A. (1998). Overexpression of HuR, a nuclear-cytoplasmic shuttling protein, increases the in vivo stability of ARE-containing mRNAs. Embo J 17, 3448-3460. Feldbrugge, M., Arizti, P., Sullivan, M. L., Zamore, P. D., Belasco, J. G., and Green, P. J. (2002). Comparative analysis of the plant mRNA-destabilizing element, DST, in mammalian and tobacco cells. Plant Mol Biol 49, 215-223. Flores, S. and. Tobin, EM. (1988). Cytokinin modulation of LHCB mRNA levels: the involvement of post-transcriptional regulation. Plant Mol Biol 11, 409-415. F olta, K. M., and Kaufman, L. S. (1999). Regions of the pea Lhcbl *4 promoter necessary for blue-light regulation in transgenic Arabidopsis. Plant Physiol 120, 747-756. F olta, K. M., and Kaufman, L. S. (2003). Phototropin 1 is required for high-fluence blue- light-mediated mRNA destabilization. Plant Mol Biol 51, 609-618. Ford, L. P., Watson, J ., Keene, J. D., and Wilusz, J. (1999). ELAV proteins stabilize deadenylated intermediates in a novel in vitro mRNA deadenylation/degradation system. Genes Dev 13, 188-201. Franco, A. R., Gee, M. A., and Guilfoyle, T. J. (1990). Induction and superinduction of auxin-responsive mRNAs with auxin and protein synthesis inhibitors. J Biol Chem 265, 15845-15849. Gallie, D. R., Caldwell, C., and 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. Gil, P., Liu, Y., Orbovic, V., Verkamp, E., Poff, K. L., and Green, P. J. (1994). Characterization of the auxin-inducible SAUR-AC1 gene for use as a molecular genetic tool in Arabidopsis. Plant Physiol 104, 777-784. Gil, P., and Green, P. J. (1996). Multiple regions of the Arabidopsis SAUR-AC1 gene control transcript abundance: the 3' untranslated region functions as an mRNA instability determinant. Embo J 15, 1678-1686. Gossen, M., and Bujard, H. (1992). Tight control of gene expression in mammalian cells by tetracycline-responsive promoters. Proc Natl Acad Sci U S A 89, 5547-5551. 31 Guhaniyogi, J ., and Brewer, G. (2001). Regulation of mRNA stability in mammalian cells. Gene 265, 11-23. Gutierrez, 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. Gutierrez, R. A., Ewing, R. M., Cherry, J. M., and Green, P. J. (2002). Identification of unstable transcripts in Arabidopsis by cDNA microarray analysis: rapid decay is associated with a group of touch- and specific clock-controlled genes. Proc Natl Acad Sci U S A 99,11513-11518. Hajela, R. K., Horvath, D.P., Gilmour, S.J., Thomashow, M.F. (1990). Molecular cloning and expression of cor (cold regulated) genes in Arabidopsis thaliana. Plant Physiol 93, 1246-1252. Hansen, E. R., Petracek, M. E., Dickey, L. F ., and Thompson, W. F. (2001). The 5' end of the pea ferredoxin-1 mRNA mediates rapid and reversible light-directed changes in translation in tobacco. Plant Physiol 125, 770-778. Herrick, D., Parker, R., and Jacobson, A. (1990). Identification and comparison of stable and unstable mRNAs in Saccharomyces cerevisiae. Mol Cell Biol 10, 2269-2284. Ho, S., Chao, Y., Tong, W., and Yu, S. (2001). Sugar coordinately and differentially regulates growth- and stress-related gene expression via a complex Signal transduction network and multiple control mechanisms. Plant Physiol 125, 877-890. Holstege, F. C., Jennings, E. G., Wyrick, J. J., Lee, T. 1., Hengartner, C. J., Green, M. R., Golub, T. R., Lander, E. S., and Young, R. A. (1998). Dissecting the regulatory circuitry of a eukaryotic genome. Cell 95, 717-728. Holtorf, H., Schob, H., Kunz, C., Waldvogel, R., and Meins, F ., Jr. (1999). Stochastic and nonstochastic post-transcriptional silencing of chitinase and beta-1,3-glucanase genes involves increased RNA tumover-possible role for ribosome-independent RNA degradation. Plant Cell 11, 471-484. Hua, X. J ., Van de Cotte, B., Van Montagu, M., and Verbruggen, N. (2001). The 5' untranslated region of the At-P5R gene is involved in both transcriptional and post- transcriptional regulation. Plant J 26, 157-169. Johnson, M. A., Perez-Amador, M. A., Lidder, P., and Green, P. J. (2000). Mutants of Arabidopsis defective in a sequence-Specific mRNA degradation pathway. Proc Natl Acad Sci U S A 97, 13991-13996. 32 Kastenmayer, J. P., Johnson, M. A., and Green, P. J. (2001). Analysis of XRN orthologs by complementation of yeast mutants and localization of XRN-GFP fusion proteins. Methods Enzymol 342, 269-282. Kaufman, L. S. (1993). Transduction of Blue-Light Signals. Plant Physiol 102, 333-337. LaGrandeur, T. E., and Parker, R. (1998). Isolation and characterization of Dcplp, the yeast mRNA decapping enzyme. Embo J 17, 1487-1496. Lai, W. S., Carballo, E., Strum, J. R., Kennington, E. A., Phillips, R. S., and Blackshear, P. J. (1999). Evidence that tristetraprolin binds to AU-rich elements and promotes the deadenylation and destabilization of tumor necrosis factor alpha mRNA. Mol Cell Biol 19, 4311-4323. Lam, L. T., Pickeral, O. K., Peng, A. C., Rosenwald, A., Hurt, E. M., Giltnane, J. M., Averett, L. M., Zhao, H., Davis, R. E., Sathyamoorthy, M., et al. (2001). Genomic-scale measurement of mRNA turnover and the mechanisms of action of the anti-cancer drug flavopiridol. Genome Biol 2, 1-11. Lambein, 1., Chiba, Y., Onouchi, H., and Naito, S. (2003). Decay kinetics of autogenously regulated CGSl mRN A that codes for cystathionine gamma-synthase in Arabidopsis thaliana. Plant Cell Physiol 44, 893-900. Lincoln, J. E., and Fischer, R. L. (1988). Diverse mechanisms for the regulation of ethylene-inducible gene expression. Mol Gen Genet 212, 71-75. Loflin, P., Chen, C. Y., and Shyu, A. B. (1999). Unraveling a cytoplasmic role for hnRN P D in the in vivo mRNA destabilization directed by the AU-rich element. Genes Dev 13, 1884-1897. Love, J., Scott, A. C., and Thompson, W. F. (2000). Technical advance: stringent control of transgene expression in Arabidopsis thaliana using the ToplO promoter system. Plant J 21, 579-588. Maillet, I., Buhler, J. M., Sentenac, A., and Labarre, J. (1999). pr4p is necessary for RNA polymerase II activity at high temperature. J Biol Chem 274, 22586-22590. McCarthy, J. E. (1998). Posttranscriptional control of gene expression in yeast. Microbiol Mol Biol Rev 62, 1492-1553. McClure, B. A. and. Guilfoyle, T. (1987). Characterization of a class of small auxin- inducible soybean polyadenlyated RNAS. Plant MolBiol 9, 611-623. McClure, B. A., Hagen, G., Brown, C. S., Gee, M. A., and Guilfoyle, T. J. (1989). Transcription, organization, and sequence of an auxin-regulated gene cluster in soybean. Plant Cell 1, 229-239. 33 McClure, B. A., and Guilfoyle, T. (1989). Rapid redistribution of auxin-regulated RNAS during gravitropism. Science 243, 91-93. Mehdy, M. C. and. Brodl, MR. (1998). The role of stress in regulating mRNA stability. In A look beyond transcription: Mechanisms determining mRNA stability and translation in plants (J Bailey-Serres and DR Gallie, eds American Society of Plant Physiologists), 64-67. Mitchell, P., and T ollervey, D. (2001). mRNA turnover. Curr Opin Cell Biol 13, 320-325. Muhlrad, D., Decker, C. J ., and Parker, R. (1994). Deadenylation of the unstable mRNA encoded by the yeast MF A2 gene leads to decapping followed by 5'-->3' digestion of the transcript. Genes Dev 8, 855-866. Muhlrad, D., and Parker, R. (1994). Premature translational termination triggers mRNA decapping. Nature 370, 578-581. Newman, T. C., Ohme-Takagi, M., Taylor, C. B., and Green, P. J. (1993). DST sequences, highly conserved among plant SAUR 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 mRN A stability in tobacco. Proc Natl Acad Sci U S A 90,11811-11815. Ominato, K., Akita, H., Suzuki, A., Kijima, F ., Yoshino, T., Yoshino, M., Chiba, Y., Onouchi, H., and Naito, S. (2002). Identification of a short highly conserved amino acid sequence as the functional region required for posttranscriptional autoregulation of the cystathionine gamma-synthase gene in Arabidopsis. J Biol Chem 277, 36380-36386. Ortega, J. L., Temple, S. J ., and Sengupta-Gopalan, C. (2001). Constitutive overexpression of cytosolic glutamine synthetase (GSl) gene in transgenic alfalfa demonstrates that GSl may be regulated at the level of RNA stability and protein turnover. Plant Physiol 126, 109-121. Padidam, M., Gore, M., Lu, D. L., and Smimova, O. (2003). Chemical-inducible, ecdysone receptor-based gene expression system for plants. Transgenic Res 12, 101-109. Perez-Amador, M. A., Lidder, P., Johnson, M. A., Landgraf, J ., Wisman, E., and Green, P. J. (2001). New molecular phenotypes in the dst mutants of Arabidopsis revealed by DNA microarray analysis. Plant Cell 13, 2703-2717. Petracek, M. E., Dickey, L. F., Nguyen, T. T., Gatz, C., Sowinski, D. A., Allen, G. C., and Thompson, W. F. (1998). Ferredoxin-l mRNA is destabilized by changes in photosynthetic electron transport. Proc Natl Acad Sci U S A 95, 9009-9013. 34 Petracek, M. E., Nuygen, T., Thompson, W. F., and Dickey, L. F. (2000). Premature termination codons destabilize ferredoxin-l mRNA when ferredoxin-1 is translated. Plant J 21, 563-569. Ravanel, S., Gakiere, 8., Job, D., and Douce, R. (1998). Cystathionine gamma-synthase from Arabidopsis thaliana: purification and biochemical characterization of the recombinant enzyme overexpressed in Escherichia coli. Biochem J 331 ( Pt 2), 639-648. 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. Scafe, C., Martin, C., Nonet, M., Podos, S., Okamura, S., and Young, R. A. (1990). Conditional mutations occur predominantly in highly conserved residues of RNA polymerase II subunits. Mol Cell Biol 10, 1270-1275. Schena, M., Shalon, D., Davis, R. W., and Brown, P. O. (1995). Quantitative monitoring of gene expression patterns with a complementary DNA microarray. Science 270, 467- 470. 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. 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 mRN A decay. Genes Dev 5, 221-231. Silver, D. L., Pinaev, A., Chen, R., and De Bruijn, F. J. (1996). Posttranscriptional Regulation of the Sesbania rostrata Early Nodulin Gene SrEnod2 by Cytokinin. Plant Physiol 112, 559-567. Stoecklin, G., Ming, X. F ., Looser, R., and Moroni, C. (2000). Somatic mRNA turnover mutants implicate tristetraprolin in the interleukin-3 mRNA degradation pathway. Mol Cell Biol 20, 3753-3763. Stoecklin, G., Colombi, M., Raineri, I., Leuenberger, S., Mallaun, M., Schmidlin, M., Gross, B., Lu, M., Kitamura, T., and Moroni, C. (2002). Functional cloning of BRFl, a regulator of ARE-dependent mRNA turnover. Embo J 21, 4709-4718. 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 mRN A instability. RNA 2, 308-315. 35 Suzuki, A., Shirata, Y., Ishida, H., Chiba, Y., Onouchi, H., and Naito, S. (2001). The first exon coding region of cystathionine gamma-synthase gene is necessary and sufficient for downregulation of its own mRNA accumulation in transgenic Arabidopsis thaliana. Plant Cell Physiol 42, 1174-1180. Taylor, C. B., and Green, P. J. (1995). Identification and characterization of genes with unstable transcripts (GUTS) in tobacco. Plant Mol Biol 28, 27-38. Theodorakis, N. G., and Cleveland, D. W. (1992). Physical evidence for cotranslational regulation of beta-tubulin mRNA degradation. Mol Cell Biol 12, 791-799. Thomashow, M. F. (1998). Role of cold-responsive genes in plant freezing tolerance. Plant Physiol 118, 1-8. Tourriere, H., Chebli, K., and Tazi, J. (2002). mRNA degradation machines in eukaryotic cells. Biochimie 84, 821-837. Tucker, M., and Parker, R. (2000). Mechanisms and control of mRNA decapping in Saccharomyces cerevisiae. Annu Rev Biochem 69, 571-595. Vakalopoulou, E., Schaack, J ., and Shenk, T. (1991). A 32-kilodalton protein binds to AU-rich domains in the 3' untranslated regions of rapidly degraded mRNAs. Mol Cell Biol 11, 3355-3364. van Hoof, A., and Parker, R. (1999). The exosome: a proteasome for RNA? Cell 99, 347- 350. Vasudevan, S., and Peltz, S. W. (2001). Regulated ARE-mediated mRNA decay in Saccharomyces cerevisiae. Mol Cell 7, 1 191-1200. Vorst, 0., van Dam, F., Weisbeek, P., and Smeekens, S. (1993). Light-regulated expression of the Arabidopsis thaliana ferredoxin A gene involves both transcriptional and post-transcriptional processes. Plant J 3, 793-803. Wang, Y., Liu, C. L., Storey, J. D., Tibshirani, R. J ., Herschlag, D., and Brown, P. O. (2002). Precision and functional specificity in mRNA decay. Proc Natl Acad Sci U S A 99, 5860-5865. Waterhouse, P. M., Graham, M.W., Wang, M.-B. (1998). resistance and gene silencing in plants can be induced by Simultaneous expression of sense and antisense RNA. Proc Natl Acad Sci USA 95, 13959-13964. Weigel, D., Ahn, J. H., Blazquez, M. A., Borevitz, J. O., Christensen, S. K., Fankhauser, C., Ferrandiz, C., Kardailsky, 1., Malancharuvil, E. J ., Neff, M. M., et a1. (2000). Activation tagging in Arabidopsis. Plant Physiol 122, 1003-1013. 36 Weinmann, P., Gossen, M., Hillen, W., Bujard, H., and Gatz, C. (1994). A chimeric transactivator allows tetracycline-responsive gene expression in whole plants. Plant J 5, 559-569. Williamson, J. D., Quatrano, R. S., and Cuming, A. C. (1985). Em polypeptide and its messenger RNA levels are modulated by abscisic acid during embryogenesis in wheat. Eur J Biochem 152, 501-507. Wolfraim, L. A., Langis, R., Tyson, H., and Dhindsa, R. S. (1993). cDNA sequence, expression, and transcript stability of a cold acclimation-specific gene, cas18, of alfalfa (Medicago falcata) cells. Plant Physiol 101, 1275-1282. Yu, S. M., Lee, Y. C., Fang, S. C., Chan, M. T., Hwa, S. F ., and Liu, L. F. (1996). Sugars act as signal molecules and osmotica to regulate the expression of alpha-amylase genes and metabolic activities in germinating cereal grains. Plant Mol Biol 30, 1277-1289. Zamore, P. D., Bartel, D. P., Lehmann, R., and Williamson, J. R. (1999). The PUMILIO- RNA interaction: a single RNA-binding domain monomer recognizes a bipartite target sequence. Biochemistry 38, 596-604. 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 mRN A degradation. Mol Cell Biol 15, 2219-2230. 37 CHAPTER 2 NEW MOLECULAR PHENOTYPES OF THE dst] MUTANT REVEALED BY MICROARRAY ANALYSIS Part of this chapter was published in “Pérez-Amador, M. A., Lidder, P., Johnson, M. A., Landgraf, J., Wisman, E., and Green, P. J. (2001). New molecular phenotypes in the dst mutants of Arabidopsis revealed by DNA microarray analysis. Plant Cell 13, 2703-2717. Experiments done by Pérez-Amador, M.A. or in collaboration with Pe’rez-Amador, M.A. are clearly marked. 38 INTRODUCTION Cells must be able to adjust gene expression patterns quickly in order to respond to intra- and extracellular stimuli; therefore, it is necessary for certain transcripts to reach new steady-state levels rapidly. The steady state levels of eukaryotic mRNAs are determined by both their rate of synthesis and rate of degradation. The control of mRNA stability is a major determinant of steady-state messenger RNA levels in the cell and ofien has a great impact on the level at which a particular gene is expressed. Further, mRNA stability affects the rate at which new steady state RNA levels are achieved after changes in transcription; the more unstable the mRN A, the more rapidly it reaches a new steady state (Abler and Green, 1996). Most recent studies on mRNA stability in eukaryotes have focused on transcripts that are relatively unstable, generally with half-lives of less than 60 minutes (Ross, 1995; Johnson et al., 1998). Unstable transcripts are particularly interesting because they often correspond to genes that must be rapidly or stringently controlled, such as those involved in regulating cell growth and differentiation. Transcripts that fall into this category include phytochrome mRN A (Higgs etal., 1995) and several auxin-induced transcripts (McClure and Guilfoyle, 1989) in plants, mating-type transcripts in yeast (Peltz and Jacobson, 1992) and several proto-oncogene transcripts in mammalian cells (Greenberg and Belasco, 1993). The majority of mRNAs fall in the stable range of mRNA half-lives for a given organism (Taylor and Green, 1995; Ross, 1996). Thus, within the body of unstable mRNAs, specific sequence motifs must be present that can either act constitutively to 39 establish the inherent instability of a particular transcript or modulate the Stability of an mRNA in response to certain physiological, developmental, or environmental cues. The most widely studied instability determinants are the mammalian AU-rich elements (ARES). ARES are found in the 3' untranslated region (UTR) of several of the most unstable mammalian transcripts, such as lymphokine, cytokine and proto-oncogene mRNAs (Chen and Shyu, 1995). Repeats of the pentamer AUUUA are often found in these AU-rich elements and are important for their instability function (Shyu et al., 1991; Vakalopoulou et al., 1991). AUUUA repeats are likely to be of broad significance in higher eukaryotes, since they can target transcripts for rapid decay in plants (Ohme- Takagi et al., 1993), and recent evidence suggests that the ARE-mediated decay pathway is firnctional in yeast as well (Vasudevan and Peltz, 2001). It seems that all functional ARES mediate deadenylation as the first step in mRNA decay, although different classes of ARES exhibit different reaction kinetics (Chen and Shyu, 1995). In plants, besides ARES, an instability determinant called DST (Newman et al., 1993) has been studied in detail. The DST (Qowngtream) element was first identified as a highly conserved sequence in the 3' UTRS of the soybean SA UR genes (McClure et al., 1989). The SA UR (small-auxin-up-RNAS) genes encode unstable transcripts whose half- lives have been estimated to be on the order of 10-50 minutes (McClure and Guilfoyle, 1987; Franco et al., 1990). Although the function of the SAUR proteins is unknown, the temporal and spatial expression of SA UR genes correlates with auxin-induced cell elongation (McClure and Guilfoyle, 1989). The prototype DST, from SA UR15A of soybean, is about 45 base pairs in length and consists of three highly conserved subdomains separated by two variable regions (Newman et al., 1993). Mutagenesis 40 studies have demonstrated that residues within two of the conserved subdomains, the ATAGAT and GTA regions, are necessary for the instability function (Sullivan and Green, 1996). In order to gain insights into the cellular components involved in the DST- mediated mRNA degradation pathway, we isolated Arabidopsis mutants defective in their ability to recognize DST elements. dst] and dst2 were isolated based on their ability to stabilize specific DST-containing transgene mRNAs (Johnson et al., 2000). For this strategy, two transgenes, hygromycin phosphotransferase (HPH) and B-glucuronidase (GUS), both containing a tetramer of the consensus DST element in the 3' UTR, were introduced into Arabidopsis as selectable and screenable markers, respectively. The presence of the DST elements decreased the mRNA levels for both transgenes, HPH-DS T and G US-DST , resulting in decreased resistance to hygromycin and low GUS activity. The dst mutants were isolated as hygromycin resistant plants with increased GUS activity due to stabilization of the corresponding mRNAs. In addition, dst mutants have elevated levels of SA UR-A C] mRNA, the only endogenous DST-containing unstable transcript characterized thus far (Gil and Green, 1996; Johnson et al., 2000). The only reported phenotype of the dst mutants is the elevation of these three mRNAs due to increased mRNA stability. Genetic analysis of two dst mutants isolated via this selection showed that they are incompletely dominant and represent two independent loci. The dst mutants exhibit no obvious morphological or developmental phenotype (Johnson et al., 2000). The main objective of the present work was to search for additional genes that were directly or indirectly regulated by this pathway. 41 Genes differentially regulated in mutant backgrounds have been identified in the past by a variety of techniques, such as differential display and subtractive hybridization (Kehoe et al., 1999). Now, with the application of plant genomics, monitoring gene expression can be achieved on a scale much larger than previously possible. DNA microarray technology (Schena et al., 1995) takes advantage of the vast amount of information generated by the genome sequencing projects, as well as the large number of expressed sequence tag (EST) clones isolated and sequenced from different organisms. In addition to dramatic changes in mRNA levels that can be identified by differential display or other techniques, DNA microarrays allow the detection of more subtle changes in gene expression (Kehoe et al., 1999). By comparing global patterns of gene expression in a mutant compared to wild type, we now have the unique opportunity to identify changes in biochemical processes due to the mutation. This approach may be especially fruitful when studying a mutation that affects the regulation of mRNA abundance and may be of particular interest when the mutation does not generate an obvious morphological phenotype. Recently published studies have demonstrated how microarrays containing a large number of Arabidopsis genes can provide a powerful tool for plant gene discovery, functional analysis, and elucidation of genetic regulatory networks (Kreps et a1, 2002; Mandaokar et al, 2003; Martzivanou and Hampp, 2003; Seki et al., 2001; Schaffer et al., 2001; Schenk et al., 2000). Plant microarray experiments have begun to focus on mutant analyses in addition to gene expression in wild-type plants. A pioneering microarray study carried out by Reymond et al. (2000) highlighted the promise of this approach, in which the coronatine-insensitive coil-I Arabidopsis mutant was analyzed on DNA 42 microarrays, resulting in the classification of a large number of C01] -dependent and C01] -independent wound-inducible genes. Other studies have been carried out with Arabidopsis mutants since then providing further insight into various biological pathways (Goda et a1, 2002; Scheible et al, 2003; Wang et al, 2002). In this report, we describe the use of DNA microarrays containing more than 11,000 Arabidopsis expressed sequence tags (ESTS), representing approximately 7,800 unique genes, to examine gene expression in the DST-mediated mRNA degradation mutant dst] . Our results indicate that DNA microarrays are a powerful tool to identify new molecular markers affected by DST-mediated mRNA decay, some of which are likely direct targets of the pathway. Furthermore, our results suggest new experimental directions to pursue the biological significance of the pathway. RESULTS Generation of a 600-element DNA microarray AS a first step towards analyzing the dst] mutant for additional changes in gene expression, a DNA microarray representing approximately 600 Arabidopsis genes was assembled. The clones included were mainly ESTS from the MSU collection (Newman et al., 1994). Table 2.1 includes a description of these genes, grouped according to function or sequence similarity. Because the dst] mutant affects RNA metabolism, half of the clones included on the microarray were those predicted to be associated with some aspect of RNA metabolism. The criteria for the selection of many of these EST clones was their sequence Similarity to potential RNA metabolism genes previously described in Arabidopsis or other organisms. 43 Table 2.1. Genes included on the 600-element DNA microarray (done by Pérez- Amador, M.A.) Category Number of genes Pathogen-related 70 Lipid metabolism 40 Wounding/jasmonic response 25 MAPK 25 Protein transport vacuole/chloroplast 30 Senescence-associated 10 Auxin-induced 10 Polyamine metabolism 5 RNA metabolism 300 RNA binding proteins 70 Helicases 20 Transcription/translation 20 RNases 15 Splicing factors/snoRNP 15 Nonsense decay 8 Putative DST-containing genes 140 AU-rich genes 10 Other 20 Controls 76 Highly expressed/well studied 17 Ribosomal proteins 15 rRNA 2 Transgenes 5 Human clones 12 Other 25 44 Thus far, the SA UR-A C1 transcript is the only known target for the DST- dependent mRNA degradation pathway in Arabidopsis (Johnson et al., 2000). To identify additional direct targets of this pathway, genes with possible DST elements were included. The strategy to identify genes containing potential DST elements was to combine the results of several pattem searches using different degenerate criteria, since our knowledge of what constitutes a fimctional DST element is still, rather limited (Newman et al., 1993; Sullivan and Green, 1996). Because ARES also function in plants as mRNA instability elements (Ohme-Takagi et al., 1993), genes with AU-rich elements were selected in a similar manner, by searching for variations of the element AUUUA in an AU-rich context in the 3' UTR. As a positive control, the HPH and GUS coding sequences were also included so that the mutant phenotype of dst] could be confirmed. Approximately 300 additional clones were included on the microarray to expand coverage beyond genes associated with RNA metabolism. A complete list of the clones on this array and the raw microarray expression data can be found at http://www.bch.msu.edu/pamgreen/Perez-Amador_etal/600_list.htm. The expression levels for most genes are similar in dstI and the parental plants The dst] mutant has a subtle phenotype; it exhibits a 3- to 4-fold elevation, relative to the parental plants, of HPH-DST and GUS-DST transgene mRNAs, which was the basis for its isolation. Each of these transgenes contains a tetramer of the prototype DST element (Newman et al., 1993; Sullivan and Green, 1996). The dst] mutant also Shows a 3-fold elevation of the unstable SA UR-A C] mRNA, an endogenous Arabidopsis transcript known to contain a DST element. Prior to this work, these were the only known 45 dst] phenotypes; dst] mutant plants appear normal with respect to morphology and development when compared to parental plants (Johnson et al., 2000). To determine whether additional molecular differences existed, poly(A)+ RNA from leaves of 40-day-old dst] and parental plants was extracted and used in a DNA microarray experiment. After hybridization and normalization of data, several genes could be identified with changes in mRNA levels greater than 1.5-fold. As expected, HPH-DST and GUS-DST were among genes with elevated mRNA levels in dst] . SA UR- AC] was also detected on the microarray to be elevated in the mutant. In addition, a DEAD box RNA helicase RH15 (At5g11200) and EST 12509T7 were identified as being elevated at the mRNA levels in the dst] mutant. When the sequence from EST 12509T7 was compared to the TAIR database (http://www.arabidopsis.org/), the translation Showed high similarity (63.7%) to the SAUR-AC1 protein and several other SAUR-like proteins as depicted in Figure 2.1A. AS a result, this gene was named SA UR-like] . Most interestingly, although the 3' UTR of this gene lacks a classical DST element, it contains multiple DST-like subdomains, i.e., two ATAGAT-like and three GTA-like subdomains, as shown in Figure 2.18. RNA gel blot analysis was used to confirm the differential regulation of these genes identified by DNA microarray. In previous experiments, RNA gel blot analysis had proven an effective method to measure subtle differences in mRNA abundance between dst] and parental plants (Johnson et al., 2000). In addition, this approach has other advantages. RNA gel blot analysis has traditionally been relied upon for accurate analysis of gene expression at the level of mRN A abundance. Moreover, once differentially abundant mRNAs are identified using one combination of mutant and wild type, these 46 .boéooamoc .5303— EE EbOmQ Hmm .«o moocoscom .m can .m 2: :09. woman 83 MED .m 05 mo conga—.886 2E. $33838. £39 98 mach“ .3 32865 2a SEEoBE ox:-<.ro can -H 1519 dsfl dst2 u — HPH “IF“ - - - a .- z: B :2 a 3 RH15 .. . “m U U C a 1- c 2 E 2; SAUR-AC1 . . . SA UR-llko1 .1 - H elF4A . . - Figure 2.2. RNA gel blot analysis to confirm DNA microarray data (in collaboration with Perez-Amador, M.A.) Lanes contained 10 pg of total RNA extracted from WT (1519), dst], and dst2 plants. Each blot was hybridized sequentially with 3'zP-labeled eIF 4A, and with A) HPH, B) RH15, and C) SA UR-A C1 and SA UR-likel . 49 carried out using four independently grown sample sets to control for biological variability. Each sample set consisted of the dst] mutant and parental plants. For each set, two microarrays were used, each with reverse labeling, for a total of eight microarray slides (raw data is available on http://genome-www4.stanford.edu/cgi- bin/SMD/cluster/QuerySetup.pl, under the experimenter name Green). The DNA microarray expression data was normalized as indicated in the “Methods” section to calculate the ratios of the fluorescence intensities of the two probes. The number of clones with a ratio greater than 1.5 fold was determined. Although a relatively low cut-off ratio of 1.5 was used, the data were reproducible and were confirmed by RNA gel blot analysis (see below). When the expression data, represented as the median of eight microarray slides, was plotted, 36 clones (31 ESTs and 5 clones for HPH and GUS) with 2 1.5-fold elevated or decreased mRNA levels could be identified (Figure 2.3). Most of these clones showed a difference of 1.5 fold or greater in at least seven of the eight slides analyzed. Each EST was mapped to the Arabidopsis genome by conducting a BLAST search against the Arabidopsis genome sequence database. Multiple ESTS representing the same gene were also identified in this manner and revealed that the 31 ESTS corresponded to 25 genes. The redundancy amongst the ESTS present on the array fiirther allowed us to verify the microarray data, since ESTs for the same genes showed similar changes in expression. In some cases, ESTs corresponding to the same gene showed slightly different expression ratios (e.g., the four ESTS corresponding to RAP2.4 ranged from 2- to 5-fold higher in dst], with a mean of 3.9-fold); this variability could be due to either variable probe or target length, genetic redundancy, or a combination thereof. The 50 dst1 - RAP2.4“. RAP2.4 g 5‘ \' HPH RAP2.4 -§ 4. 18702311 . §§ /"P” \ GUS / . mama-n n: w 37 ° GUS .° ° .4/ Es / o o 0 ° “6:” 2‘ o a: o 3 ° 0 0 £3 1th” . _ _,_. m .' 1. ° '0 .' 90° 0 . .2‘ ° 0\ /° )0 f 111G9T7 COR-like xylwd... COR-like Parental Figure 2.3. Comparison of mRNA levels in the dst] mutant and parental plants using the 11,521 element DNA microarray (in collaboration with Perez-Amador, M.A.). The graph was generated from the data obtained with ScanAnalyze software. Ratio values below 1 were transformed to - l/ratio in order to plot them as fold-difference (y-axis) against arbitrary clone number (x-axis). Five-point scale indicates genes with increased/decreased mRNA levels in the dst] mutant. 51 transgenes HPH and GUS, as well as the DEAD box RNA helicase gene RH15, were again detected with elevated mRNA levels in dst] . Twenty genes, including the transgenes HPH and GUS, showed elevated mRN A levels while seven displayed decreased levels of mRNA in the dst] mutant compared with the parental plants, as listed in Tables 2.2 and 2.3, respectively. After correction for redundancy in the ESTs and excluding the transgenes HPH and GUS, a total of 25 endogenous genes whose expression levels were altered in dst] were identified. The probable biological functions of these genes, listed in Tables 2.2 and 2.3, were based on annotation by the MIPS Arabidopsis thaliana database. RNA gel blot analysis confirms DNA microarray data To test whether the transcript changes identified in the dst] mutant by microarray analysis were reliable, total RNA was obtained from the same plants that were used for each of the four microarray experiments and examined by RNA gel blot analysis. Several genes exhibiting altered mRN A levels in the dst] mutant as determined from the 11,521 microarray analysis were tested. All of the genes exhibited increased or decreased mRN A abundance as expected. Although mRNA changes were relatively small (1.5-3.9 fold), in all cases, differences in gene expression detected by the microarray translated into similar fold differences as determined by RNA gel blot analysis (Figure 2.4). Again, RNA levels in dst2 were also monitored in these experiments. Using this approach, molecular markers specific for each dst mutant were identified. Most of the genes that had increased mRN A abundance in the dst] mutant were unaffected in the dst2 mutant. As demonstrated in Figure 2.5A, RAPZ. 4, RH15, and ESTS 120E6T7, F2H12T7, and 141P19T7 showed mRNA levels similar to the parental plants in the dst2 mutant. 52 62% 82365 05 .20 "EC 326% 222 .«o :88 OH< 222 .20 88.228: mcmza 8:258 Pmm A3 No a 3 2222.23 2.2.22; 2 82 2223 2:22 as22....2 <72 2822 32mm 22.22 a 2.2.2 3885. 2.52% 2 8223223, 2505225 2.2.22 a 2.2 38$ £9.23 2 82$»? $2.222 a 322.25 2.22 a 22 22825 E2232 2 3 2522ng 22305225 no a 22.2 2.2222me EEENNN 2 22322225 no 2” 22.2 2 £2 52288 2 82.2232 2233225 28 2" a2 $5.32 E38 2 822%? 2230522222 no 2” m2 222.3% r2882 2 3 82322322 cease: 22 a 2.~ 228222 SSS 2 2283223. 822228282 8228.23 328 «2 a 3 222223: £2.38 2 2833 8 2222222 28222 2822 «.22 a 2: -- -- N .So 22.22 a 3 8222522 522292 2 232.2522 a22222222252225 no 2. 2.~ R2? 3928 2 88mm 22 2282 82222228522 22232.22 S2 2" 3 38$ E2922: 2 8833, 2255225 :2 « 2..~ 223% Pa: 222 2 22822223222 2222222220 055222222 22522.2 Omaha—Cofl m2 2" 22 23:. $22 2 2 2238322 82295: 228222.22”2 9 222228222 f 2. 22.2“ 223222 £82222 2 2282223 0822.22.22 2282220 2.2 a an -- -- N -- 5: 22228.8 23 22.2 2. 2% $222.82 $282 2 3 822w 22 252222222 822282.28 2 22222.22% 222.2232 ESE 8mm 2 >2 EASE 32222 $258 2.2 a 3 283 2.82% 2. 222222wa 22 2.22228 23 2.352 Om a owfio>< 2228222 9 mEE 280 Hmm 2.22.: 835-35 .222; 822282.228 :22 352222 28:83 .m> 2% E 232 (2728 25238222 523 850 .N.~ ozah 53 .ocow c8865 05 mo :0 326% 05 no cocoa OH< 05 mo anbmacmzm 8:258 Hmm A3 22 a 22.2 SSS/2 52 2320 2 3 o2m€wa< 2565225 2222 #8225: $232 PEEQN 28222 22 a 22.2 2823.2 Emom m 88%? .32 2222222 ~28 2 222222225 22.22 a m2 2 23% E2222: 2 82 2M2 22 222825.22 3.822222 :5 2 232 £82 0 no « o2 $235 E2222 N Smovwmé 88282? 23 2” S SE: E2028 2 223832 2223222 2228229535 9 22222225 22 2" 3. o 2 £32 52222: 2 252va 2222222 Basemgéaooasm 22.2 a 2 882; E32 22 2 SBNwNE 2223222 2222322 222552 282389222 9 DmE 8.820 .8 am 2“ owns; 2228222 B we: 2200 Hmm 2.222.: 28.25.22-322 22223 8228322228 :22 Bum—q ~85qu .m> 2% E 232 SAME wommeoou at? 8:00 .m.~ «Ema. 54 .2220 :08 2o.“ m_o>o_ SAME 28 35.20.293-28 05 $5852 £28-» 05 223 @832 8:20 Hmm 32.29» 2: 88:8 €8.22 2:. duh—22a 83 Em <72 can max—28 3.2.3838 wfima 853cc mos—9 568298 28m 05 mataqEOo BE Efiwemi in 9.53,.— mmcoa hwm LlJvOI-OZ LU IVVQ LLO ld I-lJ L16 IvNOLIv -Ll.|-393 SFHH LLBLd ”Iv LLZI-HZd “.930er LLSIGS L 1.1.9 K)” L HdH i'ZdVH m s - “.69le H m m 5 523.32 I 3:33.: U L. 55 aauapunqe VNHW HPH and GUS mRNA levels were elevated in both mutants. The EST for one gene, 168C16T7, which shows high similarity to a cadmium-induced AP2 protein, had mRNA levels that were elevated in both mutants, although the elevation was higher in dst] than in dst2 (approximately 3.5-fold versus 2-fold compared to the parental type). EST 13913T7 which encodes an unknown protein may have marginally reduced mRNA levels in the dst2 mutant. Of particular interest was the analysis of the seven genes with lower mRNA abundance in dst] compared to parental plants. Most of these genes were either elevated (ESTS lllG9T7, E6E1T7, 170N19T7, and H5A10T7) or unaffected (ESTS 171P10T7 and 20101T7) in the dst2 mutant, as is evident from the histograms shown in Figure 2.58. The only exception was EST G4A11T7; this EST codes for an unknown protein, and the mRN A levels for this gene were also diminished in dst2. SA UR-A C1 and SA UR-likel were also analyzed by RNA gel blot analysis and were elevated in both the dst] and dst2 mutants (as shown in Figure 2.2), although they were beneath our detection limit on the majority of the slides used for the microarray experiments. These signals were below detection because, in most experiments, the signals were too close to background to calculate a valid ratio. However, in the experiments in which we could detect a valid ratio, elevation in dst] was observed as expected (data not shown). Identification of primary targets of the dst] mutation The 3' UTRS of all the genes identified as being differentially regulated in dst] were analyzed for the presence of either a classical DST element or DST-like 56 50 3 3.0 :2.“ 120E617 mRNA level 1510 dc" am 1510 dc” den 4.0 3.0 2.0 1 .0 0.0 1510 d!" (182 F2H12T7 mRNA level RAP2.4 mRNA level > '2‘ 0 E 2 l: HPH mRNA level 168C1617 mRNA level 141P1917 mRNA level _ 3.0. 3 E 2.5‘ 5 < 20 z a... g 5 1.0 p D 0.5 g (D o.ol '“ ' :2 .o 1519 dat1 dst2 1619 do" d w 1519 do" dst2 2.5 E 2.5 E 3.0 2 .2 l 2.0 2.0 l < 2.5: ‘ g z 2.ol .5 z 1.5 a: l E E 1.5. 1 0 t 1.0 p 1 0 a: . 0.5 E 0.5 E 0.5 0-0 a 0.0 2 0.0. 2 1519 den den 1519 d." dst2 v- 1519 d5" dat2 ‘d-ld” COON‘GOO O 1 . . 0. > ' 0.4 - 0.2 > 0.0 M 1519 do" den 7 1519 as" dst2 1.2 1.2 1.0 1.0 l 04. 0.0 . 0-0 0.6 . 0.4 0.4 , 2010117 mRNA level H5A10T7 mRNA level 1116917 mRNA level m G4A11T7 mRNA level 171P1017 mRNA level 1.2 . . 1519 68” £182 1519 dsfi dst2 Figure 2.5. Histograms of the mutant (dst] or ds12)/parental ratios of mRNA expression for the indicated clones. A) Increased levels of mRNA in dst] B) Decreased levels of mRNA in dst] 57 subdomains. Seven genes were identified that contain possible DST-like sequences in their 3' UTRS (listed in Table 2.4). Therefore, these genes could be hypothesized to be primary targets of the DST-mediated decay pathway which is deficient in the mutant. To determine whether the decay of some of the hypothesized primary targets required a functional DST] pathway, mRN A turnover rates were measured in the dst] mutant and 1519 parental lines. Northern blot analysis of cordycepin time courses indicated that RAP2.4 mRNA was more stable in dst] (Figure 2.6A and B) while Ccr-like mRNA decayed faster in the mutant as compared to 1519 parental Arabidopsis plants (Figure 2.6C and D ). SEN] expression is not high enough to measure mRNA decay accurately in mature leaves, the material used for the microarrays and the decay curves in Figures 2.6A and C. However, half-life measurements of SEN] mRNA in seedlings (Figure 2.6E and F) showed that it too was less stable in dst] than in the parental 1519. This data indicates that RAPZ. 4, Ccr-like and SEN] mRNAs are indeed bona fide targets of the DST- mediated mRNA decay pathway in Arabidopsis. Multiple replicates remove non-reproducible changes An important aspect of our analysis of the dst] mutant was the ability to detect subtle changes in gene expression. As shown in Figure 2.4, using eight microarray slides, all of the changes that met our criteria were highly reproducible in independent RNA gel blots. However, it was of interest to examine the degree of reproducibility using a single pair of technical replicate slides and whether fewer than eight slides are sufficient because the number of slides to use for microarray experiments is not yet routine. This could be particularly significant for future experiments, since the 1.5-fold cut-off used 58 EmEovpsw oVEédb imp—£53 omeH I Reproducible (96 clones In final set) 100 - E] Non-reproducible 3 1: O 3 '6 o u 8 E o O. 0 "314.1131? 11334.11337 11340.11341 11329.11321 Slide pairs 100 o I Best pair to worst pair g 80 ‘ Cl Worst pair to best pair h - o o as so- e s '0 § 0 40 - o a “ $ c 20 r O r: 0 2 4 8 8 Number of slides Figure 2.7. Assessment of the reproducibility of the microarray data using fewer replicates. A. Histogram plot comparing the percentage of reproducible and non-reproducible clones. The percentage of clones (y-axis) is plotted against slide pairs (x-axis) used for the microarray experiments. Each slide pair denotes microarray slides that were technical replicas, i.e., reciprocal labeling was performed with each pair of targets. The slide numbers denote the ExptID in the Stanford Microarray Database. B. Graph showing the percentage of non-reproducible clones (y-axis) plotted against the number of slides used for the microarray analysis (x-axis). 62 Stanford Microarray Database (SMD) (http://genome-www.stanford.edu/Microarray) at the time of the analysis. A portion of the resulting clusters is shown in Figure 2.8. As expected, when EST redundancy was not removed, ESTS representing the same gene clustered next to or in the immediate vicinity of each other, thus validating the observed pattern of gene expression (data not shown). Although the cluster analysis was carried out using the limited number of genes that were identified in this study, certain common expression patterns could be found. The most prominent cluster is comprised of three genes that are regulated in a diurnal fashion and share several other expression characteristics. This cluster is characterized by genes whose transcripts are decreased in abundance in the dst] mutant and include genes encoding a protein similar to a stem-specific protein, senescence-associated protein, and xylosidase. These genes show tissue-preferential expression, with a high level of expression in leaves and a low level in flowers and roots, and are elevated by light. Two of the three genes in this cluster, senescence-associated protein and xylosidase, have unstable transcripts (see discussion), contain possible DST-like sequences in their 3' UTRS, and are also elevated in the dst2 mutant. Apart from the aforementioned cluster, a second cluster of coordinately expressed genes was identified. This cluster includes genes for chalcone synthase and a protein similar to chalcone flavonone isomerase, both of which show increased mRNA levels in the dst] mutant. These genes are regulated in a diurnal fashion as well but are decreased by light and have higher levels of expression in flowers compared to leaves or roots. In addition, these mRNAs are elevated in a mutant defective for protein import into chloroplasts. 63 :28 E @8885 mm 8:23.536 m3“ :2. 0mg: 5%. E 20>": a= 2.8.29 8 ._£_E_w At 3 520.5 oona «as: M...“ 5505...: + Esocxca E .532. is e83... 5352.8 2 .225.» + 92.2 F .88. 3328034100 T. 5505...: 3 2:22.83 s «.252 . :26:ch T. 5505:: 3 8.2. 52:36.8 + 34 352256 + 5.65:: 3 firm 88:2. <2: .23 n 2026: .6535 :2 26.29: 25.2.3220 muoo. E 26.29... uommofiotuoosufi mu. No other common characterizing gene expression patterns could be seen. However, eight genes (senescence-associated protein, putative patatin protein, protein similar to CCR, putative membrane channel protein, protein similar to APR2, calmodulin-like protein, an unknown protein, and chalcone synthase) are regulated in a circadian manner, which points to a possible connection between the DST-mediated decay pathway and circadian rhythms. Four of these eight genes also contain DST-like subdomains in their 3' UTRs and could be direct targets of the dst] mutation. DISCUSSION Microarray technology has great potential for characterization of mutants, especially those with no visible aberrant phenotype. Here we report on the dst] mutant, which fits into this category. To evaluate molecular phenotypes of this mutant, DNA microarray technology was used, which allowed the identification of primary as well as secondary effects, and also revealed clues towards the identification of possible relationships among genes. Transcripts altered in dst] have predominantly increased levels, although some decrease in abundance The dst] mutant is known to elevate mRNA levels of HPH-DST and GUS-DST transgenes by 3- to 4-fold (Johnson et al., 2000). The DNA microarray experiments carried out in this report were sensitive enough to detect reproducibly the elevations of these transcripts in multiple replicates of the experiment. Moreover, additional changes of similar magnitudes were detected for transcripts corresponding to 25 genes from among 65 those represented by the 11,521 clone array. Such a limited number of changes is not surprising, since the mutant does not exhibit any morphological abnormalities. It has been suggested that dst] and dst2 correspond to weak alleles or affect genes that are part of gene families and therefore have only partial defects in mRNA stability (Johnson et al., 2000). The results of our microarray experiments are certainly consistent with this idea, since we detected only moderate changes in mRNA levels for a few genes in dst] . Increased transcript levels for some genes were expected, since dst] is known to increase levels of DST-containing transgene mRNAs and the endogenous DST-containing SA UR- ACI , presumably due to a defect in a component of the cellular machinery involved in the recognition and degradation of DST-containing transcripts. Our analysis showed that expression of several genes decreases in dst] . Decreases in gene expression could be explained by the same defect if the DST recognition component acts as both a repressor and activator in a context-dependent manner similar to some transcriptional regulators. For example, Drosophila Drapl and dCtBP are bifunctional transcription factors with distinct activation and repression functions (Willy et al., 2000; Phippen et al., 2000). Further support for this hypothesis comes from studies conducted on AUF 1. It has been demonstrated that AUF 1 acts as an RNA-destabilizing protein in the ARE-mediated mRNA decay pathway (Loflin et al., 1999; Buzby et al., 1999), but it has also been implicated as being part of the complex that mediates the stabilization of a-globin mRNA (Kiledjian et al., 1997). Similarly, the DST recognition component could be part of different multiprotein complexes that carry out separate functions. 66 The genes identified in this study encode for proteins involved in a variety of cellular processes. The DEAD box RNA helicase, RH15, is a part of a large gene family in Arabidopsis. RNA helicases are primarily RNA unwinding enzymes and have been implicated in a variety of molecular processes, including mRNA splicing, ribosome assembly, and translation initiation (Aubourg et al., 1999). Of the 18 endogenous genes that are elevated in the dst] mutant, two encode for proteins that have an amino acid motif known as the AP2 domain (see Table 2.2). The AP2 domain is essential for APETALAZ functions and contains an 18-amino acid core region that is predicted to form an amphipathic a-helix (Okamuro et al., 1997). The four ESTS that were most highly elevated in dst] correspond to the gene RAP2.4, which belongs to the RAP2 (related to AP2) gene family. The AP2 polypeptide is distinct from known fungal and animal regulatory proteins, and it has been proposed that the RAP2 proteins may function as plant sequence-specific DNA binding proteins. Another interesting EST, 222C9T7, that was elevated in dst] corresponds to a protein that is similar to the CCR4 associated factor 1 protein (CAFl). CAFl is a transcription-associated protein and is a member of the RNase D family of 3' to 5' exonucleases (Moser et al., 1997). Recently, it was shown that Caflp is a critical component of the major cytoplasmic deadenylase in yeast (Tucker et al., 2001), suggesting a potential link between mRNA deadenylation and the DST- mediated mRNA degradation pathway. Most transcripts that are affected in the dst] mutant relative to the parental type are expected to fall into one of two categories. One category consists of mRNAs that are the direct targets of the DST-mediated decay pathway, the primary transcripts, while the second category consists of transcripts, referred to as secondary transcripts, that are 67 elevated or diminished as a result of secondary effect of changes in the primary transcripts. A number of genes showing altered levels of gene expression in dst] may be due to secondary effects, since these genes do not appear to contain a consensus DST element or DST-like subdomains in their 3' UTR. Secondary effects would be expected if altered levels of DST-containing mRNAs influence the abundance of other RNAS. This situation is not unique to dst] ; secondary effects would be expected in all microarray experiments comparing mutants with the wild type. However, this study allowed us to predict and subsequently confirm some of the primary and secondary effects of the dst] mutation based on the presence or absence of DST-like sequences, respectively. Although the precise sequences required for a functional DST element have not been fully elucidated, our data provide us with the advantage of identifying at least some of the most likely primary effects. For example, none of the 140 possible DST-containing sequences found by motif searching tools on the 600 element array, were affected in dst] . This suggests that the DST element is more complex in Arabidopsis than in soybean and that it is difficult to identify this element by simple sequence search tools. It appears that multiple subdomains of the DST element are sufficient for its function as an instability determinant, since 7 of the 25 genes identified contain DST-like subdomains. Earlier mutagenesis studies of individual subdomains are also consistent with this hypothesis (Feldbrugge et al., 2002). Further experiments are required to define the DST element, which would ultimately help refine our understanding of the requirements for DST recognition. 68 Molecular markers to expedite characterization of and differentiation between dst] and dst2 The most immediate utility of these results is the identification of new molecular markers for dst] that could be used to enhance mapping and prompt additional biological experiments. For example, the analysis of dst] showed that RAP2.4 mRNA levels are more highly elevated than HPH or GUS mRNA levels. Thus RAP2.4 could be a more useful marker for the detection of homozygous dst] mutants in future experiments. Prior to this study, it was difficult to distinguish between dst] and dst2, because the known phenotypes of the two mutants were identical, and F2 progeny of a cross between the mutants do not show additive increases in the abundance of DST-containing mRNAs (Johnson et al., 2000). By examining mRNAs in dst2 that show altered levels in dst], differences between the two mutants were identified. dst] and dst2 represent mutations in independent genes (Johnson et al., 2000). The results from these experiments indicate that some targets of the decay pathway mediated by the two genes could be different. Further, there could be some degree of overlap leading to the coordinate regulation of common targets, as exemplified by SA UR-A CI and the transgenes. The cataloging of molecular markers specific for each mutant is also significant because specific markers should allow us to identify the double mutant without relying on map positions and subsequent crosses. This emphasizes another potential use of microarrays in identifying genes that can be used to examine and compare related mutants without necessitating individual microarray experiments for each. For example, a new dst mutant was recently obtained in our laboratory by activation tagging. The molecular markers identified in this study will facilitate 69 characterization of this mutant as well as subsequent mutants that affect this sequence- specific mRNA degradation pathway. Circadian association of the dst] mutation Utilizing the technique of cluster analysis, parallel expression profiling over many experiments was conducted. The clusters generated revealed co-expression characteristics of the genes affected in the dst] mutant that would not have been evident otherwise. Eight genes, which correspond to 32% of the 25 genes described in this analysis, were regulated in a circadian manner, while only 2% of 7,800 genes were found to cycle with a circadian rhythm in a study conducted by Schaffer et al. (2001). In another study by Harmer et al. (2000), 6% of 8000 genes showed circadian changes in mRNA levels. Based on comparison with their results, 4 of 25 (16%) genes in our experiments exhibited a circadian pattern of expression. These data point to a higher representation of circadian- regulated genes than would be expected by chance. Despite differences in results and experimental conditions between the two studies (e.g. only about 1500 genes were in common), four genes with altered expression in dst] were found to be circadian regulated in both studies adding more credence to our data. The significance of several genes in the cluster being regulated in a circadian manner is not yet clear, but it suggests an association between the dst] mutation and circadian rhythms. Further, Harmer et al. (2000) showed that SA UR-A CI, which is a target of the DST-mediated mRNA degradation pathway, is also circadian regulated. It is thought that a number of circadian- regulated transcripts are unstable and are expressed during a narrow window of time. It is possible that in the dst] mutant, the level of a potential regulatory factor or signal molecule that plays some part in the circadian clock function is altered, which leads to a 70 cascade effect and changes the expression of several circadian-regulated genes. If this is true, the circadian effects uncovered here would provide the first insight into the biological or physiological significance of the DST-mediated mRNA decay pathway. Elucidating the exact nature of this association and fiirther testing of this hypothesis will require the cloning of DST 1 . CONCLUSIONS AND FUTURE PROSPECTS In addition to reproducing the known elevation of two transgene mRNAs in dst] , 25 additional transcripts were found to increase or decrease in abundance relative to the parental plant. Many of the corresponding genes were subsequently examined by RNA gel blot analysis to confirm the dst] microarray results and to evaluate their expression in the dst2 mutant. In this way, several interesting and useful differences were uncovered between the two dst mutants. These new molecular markers should enhance subsequent analysis of the dst mutants, provide insight into their biological significance, and help identify other targets of the DST-mediated mRN A decay pathway. The 25 genes identified in this study are most likely an underrepresentation of the molecular phenotypes of dst] due to the stringency of our parameters to avoid false positives. For example, ratios derived by the microarray results from weakly expressed genes, such as SA UR-A CI and SA UR-likel , which have low levels of expression in the absence of auxin (McClure and Guilfoyle, 1987), were significantly different from those determined by RNA gel blot analysis. The channel intensity values for these transcripts are close to the background fluorescence intensity levels and therefore are more susceptible to variation. Also, there may be additional targets of the DST-mediated decay 71 pathway not identified because the present generation of slides do not contain all of the genes of Arabidopsis. In the future, comprehensive microarrays with improved sensitivity should result in the discovery of a greater number of molecular phenotypes for dst] . Beyond extending our knowledge of the DST-mediated decay pathway in plants, the current study provides additional general insights. Our study shows that subtle changes in gene expression can be measured reliably using multiple microarrays, which should enhance the global investigation of gene expression patterns under several conditions. This analysis demonstrates that new molecular phenotypes for mutants without a visible phenotype can be identified using DNA microarray technology. Also with full genome microarrays, it should be possible to catalog all the molecular phenotypes for any mutant. Further, new hypotheses and associations, such as the potential link between the dst] mutation and circadian rhythms, can be developed by employing the publicly available databases. Future mutant analyses via DNA microarray analysis should have even greater utility, particularly for the analysis of mutants identified by reverse genetic approaches (e. g., T-DNA insertions) that have no apparent mutant phenotype. MATERIALS AND METHODS Plant Material All Arabidopsis thaliana plants described in this report are from the accession Columbia, grown in growth chambers under 16 hr light and 60% relative humidity at 20°C. Tissue from the parental line (p1519-31) and dst] and dst2 homozygous mutants (Johnson et al., 2000) were harvested from 35- to 40-day-old plants. The dst] and dst2 72 lines used were from the second backcross to the parental line. All tissue was harvested from plants grown in parallel under the same conditions in different growth chambers. Generation of a 600-element DNA microarray EST clones were selected based on sequence similarity by BLAST analysis (Altschul et al., 1997) to known genes involved in RNA metabolism in Arabidopsis, in other plants, or in other systems, such as bacteria, yeast, and mammals. Selected Arabidopsis EST clones were obtained from the PRL2 EST collection (Newman et al., 1994). These ESTS were cloned in lambda Zip-Lox (pZLl clones, GibcoBRL, Rockville, MD) or pBluescript SK' vector (pBSK clones, Stratagene, La J olla, CA). To amplify the ESTS by PCR, we designed universal primers corresponding to the 3' and 5' ends flanking the cloning site of each vector backbone. pZLl clones were amplified using the 5' primer 5' CGACTCACTATAGGGAAAGCTGG 3', and 3' primer 5' ATTGAATTTAGGTGACACTATAGAAGAGC 3'. pBSK clones were amplified using the 5' primer 5' CGACTCACTATAGGGCGAATTGG 3' and 3' primer 5' GGAAACAGCTATGACCATGATTACG 3'. cDNA clones isolated in our laboratory were cloned in pBluescript (primers as above) or pGEM-T (Promega, Madison, WI) Vectors. To amplify clones in pGEM-T, we designed the 5' primer 5' CGACTCACTATAGGGCGAATTGG 3' and 3' primer 5' A'I‘TTAGGTGACACTATAGAATACTCAAGC 3'. Plasmid DNA from the MSU collection was diluted to a final concentration of 1-3 mg. 11L" in TE (10 mM TrisHCl pH 8.0, 1 mM EDTA). PCR reactions, in a final volume of 1 00 pL, contained 40 pmol of the corresponding primers, 0.2 mM of each dNTP, 2-5 “8 of plasmid DNA and 2 units of Taq DNA polymerase in l x reaction buffer (IOmM 73 Tris-HCl pH 8.0, 50 mM KCl, 2mM MgClz). After PCR, DNA was precipitated in ethanol and resuspended in 20 uL of 3 x SSC (1 x SSC = 0.15 M NaCl, 0.015 M NaCitrate, pH 7.0). To check the quality and quantity of the amplified DNA, 1 uL of the final resuspension was analyzed by electrophoresis in 1.0% agarose gels. PCR products with low concentration (<200 ng-uL'l) or showing multiple bands were discarded and replaced with a new PCR amplification product derived from a different EST clone corresponding to the same gene. If no alternative EST was available, a new PCR amplification was performed using a gene-specific primer designed for the 5' end of the EST clone along with the corresponding 3' vector primer. In this way, we ensured that only high quality PCR products were included on the 600-element DNA microarray. To confirm the identity of the amplified DNAs, we sequenced 12 randomly selected clones. In each case, the sequence obtained matched the EST sequence deposited in the database. For clones obtained from genomic DNA, DNA was extracted from total above ground tissue of mature Arabidopsis plants using the method of Dellaporta et al. (1983), and 100 ng was amplified as described for EST clones. In all cases, a single PCR product was obtained. Low abundance PCR products were re-amplified under the same Conditions using 1 uL of the first PCR reaction as template. DNA from the PCR reactions was transferred to master DNA plates and stored at 4°C until printing. The final concentration of DNA for printing was estimated to be 200- 400 ng-uL". DNAs were arranged as four subgrids of 12 x 13 and printed twice on poly- L”lysine-coated slides. Printing, handling, and use of the 600-element DNA microarray Was as for the 11,521 MSU DNA microarray as indicated below. 74 11,521 AFGC DNA Microarray Microarrays were generated at the Arabidopsis Functional Genomics Consortium (AF GC) Microarray facility at Michigan State University. A total of 11,521 ESTS were spotted on super-aldehyde glass slides (Telechem International, Inc.; Sunnyvale, CA). Slides were washed and blocked according to the Telechem protocol. Half-life measurements, total RNA Extraction, Poly(A)+ RNA Purification, and RNA Blot Hybridization Half-lives were determined as described by Seeley et al. (1992) with the following modifications. Two-week old Arabidopsis seedlings or rosette leaves from Arabidopsis plants were transferred to a flask with incubation buffer. Afier a 30 min incubation, 3'-deoxyadenosine (cordycepin) was added to a final concentration of 0.6 mM (time 0). Tissue samples were harvested at regular intervals thereafter and frozen in liquid nitrogen. Total RNA from leaf samples was extracted as previously described (Newman et al., 1993). Poly(A)+ RNA was purified from 200 to 400 ug of total RNA using the Oligotex mRNA kit (Qiagen, Valencia, CA). RNA ( 10 ug of total RNA or 2 pg of poly(A)+ RNA) was analyzed by electrophoresis on 2% formaldehyde/1.2% agarose gels and blotted onto nylon membrane (Nytran Plus, Schleicher & Schuell, Keene, NH). DNA probes were labeled with [a-32P]dCTP by the random primer method (Feinberg and Vogelstein, 1983) and purified from unincorporated nucleotides using probe purification columns (NucTrap, Stratagene, La Jolla, CA). The RNA blots were hybridized as described in Taylor and Green (1991) using the indicated 32P-labeled probes. For a loading control, RNA blots were hybridized with a 32P-labeled cDNA probe for the Arabidopsis translation initiation factor eIF 4A (Taylor et al., 1993). Blots were stripped 75 between hybridizations in 0.1% SDS at 90 to 95°C for 1 hour. Quantification of hybridization signals was achieved using a phosphorimager (Molecular Dynamics, Sunnyvale, CA). Labeling of Poly(A)+ RNA For first strand cDNA synthesis, 1 pg of poly(A)+ RNA and 1 pg oligo dT (GibcoBRL) in a total volume of 25 pl of DEPC-treated water was denatured at 70°C for 10 min and cooled down on ice. On ice, 8 pL of 5 x RT buffer, 4 pL of 0.1 M DTT, 2 pL of dNTPs (10 mM each), and 1 pL of Superscript II (GibcoBRL) were added, and the mixture was incubated at 42°C for 1 hr. To remove RNA, 0.5 pL of RNase H (4 units-pL’ 1) was added, and the reaction was incubated for 30 min at 37°C. Single strand cDNA was purified in a Microcon YM-30 column (Millipore, Bedford, MA), concentrating to a volume of 10 pL TE (10 mM Tris-HCl, 1 mM EDTA, pH 8.0). Single-stranded cDNA was divided into two aliquots for Cy3- and Cy5-dCTP labeling, allowing for two hybridizations for each sample. The second-strand synthesis reaction contained 5 pL of single-stranded cDNA, 22 pL water, 4 pL Klenow buffer (500 mM Tris-HCl, pH 8.0, 500 mM NaCl, 100 mM MgClz), and 2 pL random hexamers (3 pg-pL'l, GibcoBRL). The mixture was denatured at 95 °C for 3 min and annealed at room temperature for 5 minutes. Next, 4 pL of dNTP mix (250 pM dATP, dGTP, and dTTP, and 90 pM dCTP), 1 pL of dCTP-Cy3 or dCTP-CyS (Amersham, Arlington Heights, IL), and 2 pL Klenow (5 units-pL") (GibcoBRL) were added, and the reaction was incubated for 2 hr at 37°C. All incubations were carried out in thin-wall 0.5 mL tubes in a RoboCycler 40 Temperature Cycler (Stratagene). Labeled dsDNA was purified using a QIAquick PCR purification kit (Qiagen) and eluted in 50 pL 2 mM Tris-HCl, pH 8.0. The probe was 76 dried down to 10 pL. To test the quality and quantity of the product, 1 pL was separated on a 1% agarose gel using a miniprotean gel electrophoresis system (Bio-Rad Laboratories, Hercules, CA). The portion of the gel containing the DNA sample was placed on a glass microscope slide, dried on a heat block at 70°C, and scanned using a ScanArray 3000 or 5000 (GSI Lumonics). The rest of the probe was used to prepare the hybridization mixture. DNA Microarray Hybridization and Analysis For a single DNA microarray, 30 pL of hybridization solution was prepared by mixing 5.2 pL of 20 x SSC, 4.5 pL 2% SDS, 2.4 pL of tRNA (20 pg-pL'l), 9 pL Cy3- labeled probe, and 9 pL CyS-labeled probe. The mixture was denatured at 100°C for 1 min and spun for 1 min to recover any condensate. The mixture was then hybridized to the array under a glass coverslip (24 mm x 40 mm, Corning) that had been washed in 95% ethanol, then 0.2% SDS, and rinsed in distilled water. The slide was then placed in a microarray hybridization chamber (Arraylt Hybridization Cassette, TeleChem International, Inc.; Sunnyvale, CA) with 200 pL of 3 x SSC to ensure high humidity conditions. Hybridization was carried out in a water bath at 65°C for 12 to 20 hours. After hybridization, the microarray was washed for 5 min in 1 x SSC/0.2% SDS, 5 min in 0.1 x SSC/0.2% SDS, and 30 sec in 0.1 x SSC without SDS, and finally dried by centrifugation at 600 rpm for 5 min. The slide was scanned once in a ScanArray 3000 or 5000 (G81 Lumonics) for both channels 1 and 2 (corresponding to Cy3- and Cy5-labeled probes, respectively) at 10 pm resolution. The image files obtained were analyzed using ScanAnalyze sofiware (v. 77 2.32, M. Eisen, Stanford University, http://genome- ww4.stanford.edu/MicroArray/SMD/restech.html). To ensure that only spots of high quality were used in the analysis, quality control measurements produced by the ScanAlyze software were employed. For example, the GTB2 value represents the fraction of pixels within each spot that are more than 1.5 x the background measurement. Spots with GTB2 values lower than 0.50 for either channel were removed and not considered for further analysis. Data from each channel was transformed to the natural logarithm, and a Z-score was calculated to normalize the channel values in order to account for variation in RNA labeling. For the Z-score calculation [Z = (x—pYo, where x = channel value, p = mean of channel data, and o = standard deviation of channel data], the trimmed mean and standard deviation using the middle 93% of the channel values were used in order to not bias the calculation due to extremely high or low values. The new data set has, by definition, a normal distribution with zero mean and unit variance. Values were retransformed from the natural logarithm by raising to the power e, and the channel ratio was calculated. From each of the four replica samples used with the 11,521 MSU DNA microarray, we generated two slides, with direct and reverse labeling. Ratios from reverse labeling were reversed in order to compare with the ratios from direct labeling. Therefore, for all slides, ratios above 1 and below 1 indicated elevated and decreased mRN A levels in dst] vs. parental plants, respectively. An average, standard deviation, and median of these eight ratios were determined and used as final ratio of mRNA levels. 78 Hierarchical clustering was performed using Cluster and Treeview software (Eisen et al., 1998; available at http://genome- www4.stanford.edu/MicroArray/SMD/restech.html). 79 REFERENCES Abler, M. L., and Green, P. J. (1996). Control of mRNA stability in higher plants. Plant Mol Biol 32, 63-78. Altschul, S. F ., Madden, T. L., Schaffer, A. A., Zhang, J ., Zhang, 2., Miller, W., and Lipman, D. J. (1997). Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res 25, 3389-3402. Aubourg, S., Kreis, M., and Lechamy, A. (1999). The DEAD box RNA helicase family in Arabidopsis thaliana. Nucleic Acids Res 27, 628-636. Buzby, J. S., Brewer, G., and Nugent, D. J. (1999). Developmental regulation of RNA transcript destabilization by A + U-rich elements is AUF l-dependent. J Biol Chem 274, 33973-33978. Chen, C. Y., and Shyu, A. B. (1995). AU-rich elements: characterization and importance in mRNA degradation. Trends Biochem Sci 20, 465-470. Dellaporta, S. L., Wood, J. and Hicks, J .B. (1983). A plant DNA minipreparation: version II. Plant Mol Biol Rep 1, 19-21. DeRisi, J. L., Iyer, V. R., and Brown, P. O. (1997). Exploring the metabolic and genetic control of gene expression on a genomic scale. Science 278, 680-686. Eisen, M. B., Spellman, P. T., Brown, P. 0., and Botstein, D. (1998). Cluster analysis and display of genome-wide expression patterns. Proc Natl Acad Sci U S A 95, 14863-14868. Feinberg, A. P., and Vogelstein, B. (1983). A technique for radiolabeling DNA restriction endonuclease fragments to high specific activity. Anal Biochem 132, 6-13. Feldbrugge, M., Arizti, P., Sullivan, M. L., Zamore, P. D., Belasco, J. G., and Green, P. J. (2002). Comparative analysis of the plant mRNA-destabilizing element, DST, in mammalian and tobacco cells. Plant Mol Biol 49, 215-223. Franco, A. R., Gee, M. A., and Guilfoyle, T. J. (1990). Induction and superinduction of auxin-responsive mRN As with auxin and protein synthesis inhibitors. J Biol Chem 265, 15845-15849. Gil, P., and Green, P. J. (1996). Multiple regions of the Arabidopsis SAUR-AC1 gene control transcript abundance: the 3' untranslated region functions as an mRN A instability determinant. Embo J 15, 1678-1686. Goda, H., Shimada, Y., Asami, T., Fujioka, S., and Yoshida, S. (2002). Microarray analysis of brassinosteroid-regulated genes in Arabidopsis. Plant Physiol 130, 1319-1334. 80 Greenberg, M. E. and Belasco, J. G. (1993). Control of the decay of labile protooncogene and cytokine mRNAs. Control of Messenger RNA Stability (San Diego: Academic Press), 199-218. Harmer, S. L., Hogenesch, J. B., Straume, M., Chang, H. S., Han, B., Zhu, T., Wang, X., Kreps, J. A., and Kay, S. A. (2000). Orchestrated transcription of key pathways in Arabidopsis by the circadian clock. Science 290, 2110-2113. Higgs, D. C., Barnes, L. J ., and Colbert, J. T. (1995). Abundance and half-life of the distinct oat phytochrome A3 and A4 mRNAs. Plant MolBiol 29, 367-377. Johnson, M. A., Baker, E.J., Colbert, J .T. and Green, P]. (1998). Determinants of mRNA stability in plants. A look beyond transcription: Mechanisms determining mRNA stability and translation in plants (Rockville: American Society of Plant Physiologists Press), 40- 53. Johnson, M. A., Perez-Amador, M. A., Lidder, P., and Green, P. J. (2000). Mutants of Arabidopsis defective in a sequence-specific mRNA degradation pathway. Proc Natl Acad Sci U S A 97, 13991-13996. Kehoe, D. M., Villand, P., and Somerville, S. (1999). DNA microarrays for studies of higher plants and other photosynthetic organisms. Trends Plant Sci 4, 38-41. Kiledjian, M., DeMaria, C. T., Brewer, G., and Novick, K. (1997). Identification of AUFl (heterogeneous nuclear ribonucleoprotein D) as a component of the alpha-globin mRNA stability complex. Mol Cell Biol 17, 4870-4876. Kreps, J. A., Wu, Y., Chang, H. S., Zhu, T., Wang, X., and Harper, J. F. (2002). Transcriptome changes for Arabidopsis in response to salt, osmotic, and cold stress. Plant Physiol 130, 2129-2141. Loflin, P., Chen, C. Y., and Shyu, A. B. (1999). Unraveling a cytoplasmic role for hnRN P D in the in vivo mRNA destabilization directed by the AU-rich element. Genes Dev 13, 1884-1897. Mandaokar, A., Kumar, V. D., Amway, M., and Browse, J. (2003). Microarray and differential display identify genes involved in jasmonate-dependent anther development. Plant Mol Biol 52, 775-786. Martzivanou, M., and Hampp, R. (2003). Hyper-gravity effects on the Arabidopsis transcriptome. Physiol Plant 1 18, 221-231. McClure, B. A. and Guilfoyle, T. (1987). Characterization of a class of small auxin- inducible soybean polyadenlyated RNAS. Plant MolBiol 9, 611-623. 81 McClure, B. A., and Guilfoyle, T. (1989). Rapid redistribution of auxin-regulated RNAS during gravitropism. Science 243, 91-93. McClure, B. A., Hagen, G., Brown, C. S., Gee, M. A., and Guilfoyle, T. J. (1989). Transcription, organization, and sequence of an auxin-regulated gene cluster in soybean. Plant Cell 1, 229-239. 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, 5110-5118. Newman, T. C., Ohme-Takagi, M., Taylor, C. B., and Green, P. J. (1993). DST sequences, highly conserved among plant SAUR genes, target reporter transcripts for rapid decay in tobacco. Plant Cell 5, 701-714. Newman, T., de Bruijn, F. J ., Green, P., Keegstra, K., Kende, H., McIntosh, L., Ohlrogge, J., Raikhel, N., Retzel, E., Thomashow, M., and Somerville, S., (1994). Genes galore: a summary of methods for accessing results from large-scale partial sequencing of anonymous Arabidopsis cDNA clones. Plant Physiol 106, 1241-1255. Ohme-Takagi, M., Taylor, C. B., Newman, T. C., and Green, P. J. (1993). The effect of sequences with high AU content on mRN A stability in tobacco. Proc Natl Acad Sci U S A 90,11811-11815. Okamuro, J. K., Caster, B., Villarroel, R., Van Montagu, M., and Jofiiku, K. D. (1997). The AP2 domain of APETALAZ defines a large new family of DNA binding proteins in Arabidopsis. Proc Natl Acad Sci U S A 94, 7076-7081. Peltz, S. W., and Jacobson, A. (1992). mRNA stability: in trans-it. Curr Opin Cell Biol 4, 979-983. Phippen, T. M., Sweigart, A. L., Moniwa, M., Krumm, A., Davie, J. R., and Parkhurst, S. M. (2000). Drosophila C-terminal binding protein functions as a context-dependent transcriptional co-factor and interferes with both mad and groucho transcriptional repression. J Biol Chem 275, 37628-37637. Reymond, P., Weber, H., Damond, M., and Farmer, E. E. (2000). Differential gene expression in response to mechanical wounding and insect feeding in Arabidopsis. Plant Cell 12, 707-720. 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. 82 Schaffer, R., Landgraf, J ., Accerbi, M., Simon, V., Larson, M., and Wisman, E. (2001). Microarray analysis of diurnal and circadian-regulated genes in Arabidopsis. Plant Cell 13, 113-123. Scheible, W. R., Fry, 3., Kochevenko, A., Schindelasch, D., Zimmerli, L., Somerville, S., Loria, R., and Somerville, C. R. (2003). An Arabidopsis mutant resistant to thaxtomin A, a cellulose synthesis inhibitor from Streptomyces species. Plant Cell 15, 1781-1794. Schena, M., Shalon, D., Davis, R. W., and Brown, P. O. (1995). Quantitative monitoring of gene expression patterns with a complementary DNA microarray. Science 270, 467- 470. Schenk, P. M., Kazan, K., Wilson, I., Anderson, J. P., Richmond, T., Somerville, S. C., and Manners, J. M. (2000). Coordinated plant defense responses in Arabidopsis revealed by microarray analysis. Proc Natl Acad Sci U S A 97, 11655-11660. Seeley, K. A., Byrne, D. H., and Colbert, J. T. (1992). Red Light-Independent Instability of Oat Phytochrome mRNA in Vivo. Plant Cell 4, 29-38. Seki, M., Narusaka, M., Abe, H., Kasuga, M., Yamaguchi-Shinozaki, K., Caminci, P., Hayashizaki, Y., and Shinozaki, K. (2001). Monitoring the expression pattern of 1300 Arabidopsis genes under drought and cold stresses by using a fiill-length cDNA microarray. Plant Cell 13, 61-72. 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 mRN A decay. Genes Dev 5, 221-231. 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. Taylor, C. B., and Green, P. J. (1991). Genes with homology to fungal and S-gene RNases are expressed in Arabidopsis thaliana. Plant Physiol 96, 980-984. 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. Taylor, C. B., and Green, P. J. (1995). Identification and characterization of genes with unstable transcripts (GUTS) in tobacco. Plant Mol Biol 28, 27-38. Tucker, M., Valencia-Sanchez, M. A., Staples, R. R., Chen, J ., Denis, C. L., and Parker, R. (2001). The transcription factor associated Ccr4 and Cafl proteins are components of the major cytoplasmic mRNA deadenylase in Saccharomyces cerevisiae. Cell 104, 377- 386. 83 Vakalopoulou, E., Schaack, J ., and Shenk, T. (1991). A 32-kilodalton protein binds to AU-rich domains in the 3' untranslated regions of rapidly degraded mRNAs. Mol Cell Biol 11, 3355-3364. Vasudevan, S., and Peltz, S. W. (2001). Regulated ARE-mediated mRNA decay in Saccharomyces cerevisiae. Mol Cell 7, 1191-1200. Wang, H., Ma, L., Habashi, J., Li, J ., Zhao, H., and Deng, X. W. (2002). Analysis of far- red light-regulated genome expression profiles of phytochrome A pathway mutants in Arabidopsis. Plant J 32, 723-733. Wildsmith, S. E., and Elcock, F. J. (2001 ). Microarrays under the microscope. Mol Pathol 54, 8-16. Willy, P. J ., Kobayashi, R., and Kadonaga, J. T. (2000). A basal transcription factor that activates or represses transcription. Science 290, 982-985. Yue, H., Eastman, P. S., Wang, B. B., Minor, J ., Doctolero, M. H., Nuttall, R. L., Stack, R., Becker, J. W., Montgomery, J. R., Vainer, M., and Johnston, R. (2001). An evaluation of the performance of cDNA microarrays for detecting changes in global mRNA expression. Nucleic Acids Res 29, E41-41. 84 CHAPTER 3 CIRCADIAN CONTROL OF mRNA STABILITY: IMPACT OF THE dst MUTANTS 85 INTRODUCTION Circadian clocks control rhythmic biological processes in many organisms and persist in the absence of environmental cues (Allada et al., 2001; Harmer et al., 2001; McClung, 2001). The Drosophila clock has been very well characterized and has been a model system to investigate circadian mechanisms. The transcription factors Clock (CLK) and Cycle (CYC) activate the transcription of the period (per) and timeless (tim) genes; PER and TIM proteins in turn negatively regulate the transcription of per and tim thus inhibiting the activity of CLK and CYC (Allada et al., 2001). Genomic approaches to study circadian gene expression on a global basis in Arabidopsis, Drosophila and mammalian systems have revealed that 2-10% of mRNAs show circadian oscillations (Harmer et al., 2000; Claridge-Chang et al., 2001; McDonald and Rosbash, 2001; Miyazaki et al., 2003; Schaffer et al., 2001; Ueda et al., 2002). A recent study by Michael and McClung (2003) proposes that this could be an underestimate and using enhancer trapping, the authors showed that 36% of the Arabidopsis genome is under transcriptional control by the circadian clock. For all the circadian clocks studied thus far, the core circadian oscillator is comprised of transcriptional feedback loops. However numerous lines of evidence in various systems suggest that posttranscriptional regulatory mechanisms are also required for clock function (Edery, 1999). The oscillation in mRNA levels of all clock controlled genes cannot be accounted for by transcription alone. The best studied example to date corresponds to the per gene, one of the components of the central oscillator in Drosophila where comparison of per transcription rates and mRNA levels implicated a temporal regulation of mRNA half-life 86 (So and Rosbash, 1997). Further support for posttranscriptional regulation of per comes from transgenic flies that demonstrate daily cycling of per mRNA expressed from a constitutive promoter (F risch et al., 1994). It has also been shown that sequences within per’s 5’ UTR are required for regulating temporal RNA expression (Stanewsky et al., 2002) and this event depends on functional PER and TIM proteins (Suri et al., 1999). In addition, posttranscriptional regulation is thought to be involved in the adaptation of Drosophila to cold. An alternatively spliced form of the per transcript is generated at lower temperature causing an advance in the phases of both the mRNA and protein cycles (Majercak et al., 1999). Temperature appears to affect the translational control of a central clock component of Neurospora as well. The levels of F RQ (frequency) increase at higher temperatures although very little changes are seen for frq mRNA levels (Liu et aL,l998) Several posttranscriptional mechanisms are responsible for the lag between the peak levels of mRNA and that of the resulting protein. In Drosophila, a kinase called DBT (doubletime) is believed to be responsible for phosphorylation of PER thus targeting PER for rapid degradation (Price et al., 1998). dbt mutants that have reduced kinase activity show defective PER degradation (Suri and Rosbash, 2000). Another kinase SGG (shaggy) is thought to be important for the phosphorylation of TIM (Martinek et al., 2001). Recent experiments indicate that SLMB (slimb), a component of the ubiquitin proteasome pathway, participates in controlling the levels of PER and TIM (Grima et al., 2002). Interestingly, TIM is photosensitive and is degraded by the proteasome in the presence of light (Naidoo et al., 1999). 87 The Drosophila CLK protein is also posttranscriptionally controlled, possibly due to changes in stability induced by phosphorylation at specific times during the day (Kim et al., 2002). Similar daily oscillations in the phosphorylated state of the Neurospora F RQ protein have been observed (Garceau et al., 1997), and frq mutants in which the phosphorylation site has been abolished have reduced rates of FRQ degradation (Liu et al., 2000). The Neurospora homolog of Drosophila Slimb, FWDl, regulates the degradation of FRQ through the ubiquitin-proteasome pathway (He et al., 2003). A recent report indicates that antisensefrq transcripts are partly responsible for the entrainment of the Neurospora clock (Kramer et al., 2003). Antisensefrq RNA is induced by light and in the dark it cycles in antiphase to sense frq RNA. The existence of antisensefrq RNA could suggest the involvement of RN Ai in the control of circadian gene expression. Even though higher plants do not contain orthologs of clock proteins from other systems, oscillatory feedback loops seem to be conserved. The Arabidopsis circadian clock is suggested to be comprised of the MYB transcription factors LH Y (Late Elongated Hypocotyl) and CCA] (Circadian Clock Associated 1) that are negative elements of a transcriptional feedback loop (Schaffer et al., 1998; Wang and Tobin, 1998) and have partially redundant functions (Mizoguchi et al., 2002). A pseudoresponse regulator T 0C1 (Timing Of CAB expression 1) might function as a positive element activating the transcription of CCA I/LH Y (Alabadi et al., 2001). Posttranscriptional modifications are also an integral part of the plant circadian clock. In Arabidopsis, a casein kinase II (CK2) phosphorylates CCA] and influences CCAl binding to DNA (Sugano et a1; 1998). Moreover CK2 phosphorylates LHY in vitro and its overexpression leads to the shortened period of various circadian rhythms 88 (Sugano et al., 1999). The tej mutant results in a light-independent period lengthening of some clock genes and is caused by a mutation in a poly (ADP-Rib) glycohydrolase (Panda et al., 2002). Two Arabidopsis clock associated proteins, ZTL (Zeitlupe) and FKFl (Flavin-binding, Kelch repeat, F box), contain an F-box motif that promotes ubiquitination of substrate proteins (Somers et al., 2000; Nelson et a1; 2000). Kim et al. (2003) used Arabidopsis cell suspension cultures to demonstrate that the rhythmic changes in the levels of ZTL are caused by circadian phase-specific differences in protein degradation by the proteasome. Transcript stability has been hypothesized to be responsible for the oscillations of the CAB] mRNA in Arabidopsis (Millar and Kay, 1991). Increased transcript stability has also been implicated in the accumulation of high steady state levels of CA T3 mRNA in continuous dark (Zhong et al., 1997). Furthermore, based on nuclear run-on assays, the cycling of NIA2 (Nitrate Reductase 2) is thought to occur through posttranscriptional regulation (Pilgrim et al., 1993). In rice, the circadian regulation of CatA (CatalaseA) expression has been postulated to be at the level of pre-mRNA stability (Iwamoto et al., 2000) Microarray analysis has shown that a subset of unstable transcripts in Arabidopsis are controlled by the circadian clock (Gutierrez et al., 2002). Additional microarray studies led to the finding that an unexpectedly high percentage of transcripts that were changed in abundance in a mutant deficient in DST-mediated decay, dst], were circadian regulated, indicating that the biological significance of the DST-mediated mRNA decay pathway may be associated with the circadian clock (Pérez-Amador et al., 2001). In wild type plants, for two transcripts SEN] and Ccr-like, mRNA decay is regulated by the 89 circadian clock (Gutierrez, 2003) and these transcripts are direct targets of the dst] mutant (see Chapter 2). These observations prompted us to test whether the dst] mutation affects the control of Cer-like and SEN] mRNA stability differently at different times of the day. The results presented in this chapter demonstrate that the circadian clock regulated stability of specific plant mRNAs, Ccr-like and SEN], is altered in the dst] mutant and the DST 1 locus is associated with circadian control. Furthermore our data indicates that control of mRNA stability adds another layer of regulation which impacts at the whole plant level for certain circadian processes and has a prominent role in clock controlled gene expression in plants. RESULTS Diurnal control of Ccr-like and SEN] mRNA stability is affected in the dst] mutant It was previously shown that DSTl function is important for the normal diurnal oscillatory expression of Ccr-like and SEN] transcripts (Gutiérrez, 2003). To determine if this alteration in diurnal oscillation was caused by defective mRN A degradation in the dst] mutant, mRN A decay rates were measured at two times during the day. Cordycepin time courses were carried out one hour after dawn, ZTl (zeitgeber 1) and 8 hours after dawn (ZT8) in two week old 1519 and dst] plants grown in 16 hours light/8 hours dark conditions. In the morning (ZTl), Ccr-like mRNA was more unstable in dst] , relative to the parental (p-value = 0.0003), as expected from the microarray studies (Figure 3.1A and B). However this effect was reversed in the aftemoon (ZT8) such that the transcript was now more stabilized in the mutant (Figure 3.1C and D; p-value < 0.0001). A similar effect was seen on SEN] mRNA decay kinetics with the transcript being rapidly degraded 9O Morning (ZTl) 0 15 30 60 90 120 0 15 30 60 90 120 “1) OF? I . . = Ccr-like ital v M N "r .u. H I; 5 eIF4A-.p--4-—ouw- uuuuuuu I- < .1519 (106:1: 15 min) p1519 .1er 5 E 0 dst] (62 :1: 7 min) 0'1 I I I I I 15 45 75 105 135 time (min) C. D. Afternoon (ZT8) 1. an 0 15 30 60 90 120 0 15 30 60 90 120 .2 '3 Ccr-like..“"h .."?“ E h eIF4Auflm--M'- < 0 E 0 1519(48102 min) ”1519 ‘1‘” 0E1 u dst] (86:1: 12 min) l I I I l 15 45 75 105 135 time (min) Figure 3.1. Regulation of Ccr-like mRNA stability is altered in the dst] mutant during the day. Representative northern blot analysis of cordycepin time courses performed in 1519 and dst] plants (A) 1 hour after dawn (zeitgeberl or ZTl), and (C) 8 hours after dawn (ZT8) for Ccr-like and ez'f4A mRNAs. Samples consisted of 10 pg of total RNA isolated from the indicated time points. Quantitation of the decrease in mRNA abundance and half-life estimation for Ccr-like (B) in the morning and (D) in the afternoon. The stable eIF 4A transcript was used as a reference for equal loading. Half-life values are representative of 2 independent cordycepin time courses. 91 in the afiemoon and stabilized in the morning in parental plants (Figure 3.28 and D). The opposite trend was seen for the dst] mutant with SEN] mRNA being less stable in the am. (p-value < 0.0001) and more stable in the p.m. (Figure 3.2B and D; p-value = 0.0061). Statistical analysis, as indicated above, showed that the differences in half-lives seen for Ccr-like and SEN] mRNAs in the mutant and parental plants are significant. These results suggest that normal DST 1 function is required for the proper timing of degradation of Ccr-like and SEN] transcripts under diurnal conditions. DST] function is required for normal circadian expression of SEN] and Ccr-like mRNAs To evaluate the impact of the dst] mutation on the circadian oscillation of Ccr- like and SEN], mRNA levels were examined under free running conditions. Arabidopsis seedlings were grown for 12 days in 16/8 LD cycles and on the morning of the 12th day transferred to continuous light. Seedling tissue for mRN A isolation was harvested every 3 hours starting on the morning of the 12th day (circadian time 0/CTO) up to the 14th day (CT45). As shown in Figure 3.3A and B, Ccr-like mRNA peaked approximately 3 hours later in the dst] mutant than in the parental plants. This effect of the mRNA peak lagging in the dst] was even more pronounced on the second day in constant light. A reduction in amplitude for Cor-like mRNA was also seen in addition to the delay in phase. dst] influence on the circadian oscillation of SEN] was slightly more complicated since SEN] is known to be induced during the dark (Oh et al., 1996) possibly at the level of transcription (Chung et al., 1997). As is evident in Figure 3.4A, the dark induction of SEN] was not seen during the second day in continuous light. However, the SEN] 92 Morning (ZTl) 1 015306090120 015306090120 a” .. '3 mm a----- ~---- g 2 ,1 < . 31p,“ 5 0 1519 (97 :1: 16 mm) {.3 p1519 dst] E dst] (41 :1: 7 min) 0.0] I I I r l 15 45 75 105 135 time (min) C. D. 1 l Afternoon (ZT8) \ DD 0 . \ .E - 015306090120 015306090120 é o E 0 SEN] OU"'"' """""‘ E E o 1519 (53 :1 3 min) 811’“ E U dst] (75 1: 5 min) p1519 dst] 0'1 I I I j I 15 45 75 105 135 time (min) Figure 3.2. Regulation of SEN] mRNA stability is altered in the dst] mutant during the day. Representative northern blot analysis of cordycepin time courses performed in 1519 and dst] plants (A) 1 hour after dawn (zeitgeberl or ZTl), and (C) 8 hours afier dawn (ZT8) for SEN] and eif4A mRNAs. Samples consisted of 10 pg of total RNA isolated from the indicated time points. Quantitation of the decrease in mRN A abundance and half-life estimation for SEN] (B) in the morning and (D) in the afternoon. The stable eIF 4A transcript was used as a reference for equal loading. Half-life values are representative of 2 independent cordycepin time courses. 93 A. 1519 0 3 6 9 12 15 18 2124 27 30 33 36 39 42 45 (CThrs.) Car-like .-... ...... ..,.. a muoOO¢ppudubobuuuo dst] 0 3 6 9 12 15 18 2124 27 30 3336 39 42 45 (CThrs.) Ccr-like «I Q... Q d an O... Q o .m4-¢00¢0.09000999Q B. +1519 - ' dst] 1.2 Relative mRNA levels 0 F 1 T T 1 I I f f 1 l T 1 l 1 0 3 6 9121518212427303336394245 CT (hrs) Figure 3.3. Circadian oscillation of Ccr-like mRNA is altered in the dst] mutant. A) Representative northern blot analysis of time courses performed throughout two entire days in continuous light for Ccr-like mRN A in dst] mutant and parental 1519 plants. Samples consisted of 10 pg of total RNA isolated from the indicated times of the day after subjective dawn (CT=0). B) Quantitation of mRN A levels for Ccr-like. All values are the average of two independent experiments and are made relative to the highest mRNA accumulation in either of the two genetic backgrounds. The signal for eIF 4A was used as a reference for equal loading. 94 A. 1519 0 3 6 9 12 15 18 2124 27 30 3336 39 42 45 (CThrs.) SEN] .O“""' ...."“. eIF4A uuu~~~~“"“““““* dst] 3 6 9 12 15 18 2124 27 30 33 36 39 42 45 (CThrs.) s15~1....... QIOCIO. eIF4A -uuuouuuuuu—udoéd B. +1519 ' dst] 1.2 5 > 2 L 7 7 _ 2 E 0 L C: 7‘ 3 . ‘ 7. " , ." ‘ > 0 .. .. 0 3 6 9 121518212427303336394245 CT (hrs) Figure 3.4. Circadian oscillation of SEN] mRNA is altered in the dst] mutant. A) Representative northern blot analysis of time courses performed throughout two entire days in continuous light for SEN] mRNA in dst] mutant and parental 1519 plants. Samples consisted of 10 pg of total RNA isolated from the indicated times of the day after subjective dawn (CT=0). B) Quantitation of mRNA levels for SEN] . All values are the average of two independent experiments and are made relative to the highest mRN A accumulation in either of the two genetic backgrounds. The signal for eIF 4A was used as a reference for equal loading. 95 transcript had reduced accumulation in the afternoon in the dst] mutant and this difference in p.m accumulation between the dst] and parental plants was more apparent on the second day (Figure 3.4A and B). Also the circadian expression of Cor-like and SEN] transcripts under free running conditions was comparable to that seen previously under diurnal conditions (Gutierrez, 2003). In order to study the effect of dst] on general clock controlled gene expression, circadian oscillation of AtGRP7/CCR2, which functions downstream of the master clock (Staiger, 2001), was tested. Oscillation of C CR2 mRN A was unchanged in the dst] mutant compared to the parental (Figure 3.5A and B). Taken together, this indicates that the circadian oscillations of a subset of clock controlled genes, Ccr-like and SEN, are dependent on DST 1 function. dst] affects circadian control of mRNA stability We hypothesized that the stabilization of Ccr—like and SEN] mRNAs caused by dst] in the afiemoon under diurnal conditions would also occur in the subjective aftemoon under free running conditions. To test this hypothesis, mRNA half-lives were measured in Arabidopsis seedlings that were transferred to continuous light for 2 days. Transcription was inhibited 1 hour (CTl) and 8 hours (CT8) after the subjective morning to determine mRNA decay rates. The impact of dst] was recapitulated for Ccr-like in continuous light with the transcript being more unstable in dst] in the subjective morning (Figure 3.68; p-value < 0.0001) and more stable in dst] in the subjective afternoon (Figure 3.6D; p-value = 0.0171) relative to the parental. Even though some dampening in mRNA half-lives was seen under circadian conditions relative to the diurnal conditions, the differences in mRNA decay rates between the mutant and parental were statistically 96 A. 1519 3 6 9 15 18 2124 27 30 33 36 39 42 45 (CThrs.) ‘99.""'.9"' elm .O...gouuuou..00 AtGRP7 d5” 3 6 9 12 15 18 2124 27 30 33 36 39 42 45 (CThrs.) AtGRP7 0‘..‘...‘....6. eIF4A ........QUCCQQQQ 3' + 1519 dst] 1.2 Relative mRNA levels 0 3 6 9 121518212427303336394245 CT (hrs) Figure 3.5. Circadian oscillation of AtGRP7 mRNA is unaltered in the dst] mutant. A) Representative northern blot analysis of time courses performed throughout two entire days in continuous light for AtGRP7 mRN A in dst] mutant and parental 1519 plants. Samples consisted of 10 pg of total RNA isolated from the indicated times of the day after subjective dawn (CT=0). B) Quantitation of mRN A levels for AtGRP7. All values are representative of at least two independent experiments and are made relative to the highest mRNA accumulation in either of the two genetic backgrounds. The signal for eIF 4A was used as a reference for equal loading. 97 Subjective Morning (CTI) ‘ ’1 “— 0 15 30 60 90 120 Ccr-like .‘llfi" .UIIH elF4A “.F". 1...-.- p1519 Subjective Afternoon (ZT8) 015 3060 90120 Ccr-like . . . - _ .. BIFM Conn.- p1519 an 0 15 30 60 90 120 .5 .3 (x E 2 g 9 1519 (129 1 21 min) dst] E 'J dst] (70 :1: 10 min) 0'1 I I T r I 15 45 75 105 135 time (min) 1). 1 l. g0 \‘n‘,\ 0 15 30 60 90 120 IE - < .....- g °1519(50:l:3min) o E a dst] (64 1 2 min) dst] 0 l T I I I l 15 45 75 105 135 time (min) Figure 3.6. Circadian regulation of Ccr-like mRNA stability is altered in the dst] mutant. Representative northern blot analysis of cordycepin time courses performed in 1519 and dst] plants (A) 1 hour after subjective dawn (circadian time 1 or CTl), and (C) 8 hours afier subjective dawn (CT8) for Ccr-like and eif4A mRNAs. Samples consisted of 10 pg of total RNA isolated from the indicated time points. Quantitation of the decrease in mRNA abundance and half-life estimation for Cor-like (B) in the subjective morning and (D) in the subjective afternoon. The stable eIF 4A transcript was used as a reference for equal loading. Half-life values are representative of 2 independent cordycepin time courses. 98 significant. In the subjective morning, SEN] mRNA degrades faster in dst] than in parental (Figure 3.7A and B), whereas in the afternoon no significant differences in mRNA stability could be seen in dst] compared to the parental (Figure 3.7C and D). This could be attributed in part to the dampening observed under circadian conditions since the effect of dst] on SEN] mRNA stability under diurnal conditions was less dramatic than for Ccr-like (Figure 3.1 and 3.2). Opposite effect of dst2 on SEN] mRNA stability For Ccr-like and SEN] , the dst2 mutation is known to have an effect opposite to that of dst] in RNA gel blots of samples harvested in the am. (see Chapter 1). To confirm this at the level of mRNA stability, mRNA half-lives were measured at ZTl and ZT8 in dst2 and parental plants. As shown in Figure 3.8B, SEN] transcript was more stabilized in the dst2 mutant in the morning. In the aftemoon, SEN] mRNA decayed at a more or less similar rate in dst2 and parental plants (Figure 3.8D), but on comparing the am. and p.m. half-lives for the transcript in dstZ, it was observed that the transcript decayed at a much faster rate than in the parental (Figure 3.8). Similar experiments with Ccr-like are in progress. Impact at the whole plant level: Classical circadian phenotypes are altered in the dst mutants To address the question if the link between the DST-mediated decay pathway and the circadian clock impacted at the whole plant level, leaf movement, a classical circadian phenotype, was monitored in the dst mutants. The oscillation in the position of leaves can be monitored by video imaging and this technique has been used to study 99 Subjective Morning (CTl) 015306090120 015306090120 14,. :0 SEN] -----.. III-I... .5 .a .. eIF4A ..---- ...UU- 5 p1519 dst] g 0 1519 (159 1 33 min) E C dst] (85 1 8 min) 0.1 y I I I I 15 45 75 105 135 time (min) C D. Subjective Afternoon (ZT8) l 015306090120 015306090120 ”1X fl SEN] .u.--- III..-” 5 ' ‘-\: "" V . . p g \ eIF4A hobo-69“ .16...- 2 '\ < 0 1519 (54 1 3 min) ' 9'5” ‘1‘” E a dst] (54 1: 03 min) 0.1 I I I I I 15 45 75 105 135 time (min) Figure 3.7. Circadian regulation of SEN] mRNA stability is altered in the dst] mutant. Representative northern blot analysis of cordycepin time courses performed in 1519 and dst] plants (A) 1 hour after subjective dawn (circadian time 1 or CTl), and (C) 8 hours after subjective dawn (CT8) for SEN] and eif4A mRNAs. Samples consisted of 10 pg of total RNA isolated from the indicated time points. Quantitation of the decrease in mRNA abundance and half-life estimation for SEN] (B) in the subjective morning and (D) in the subjective afternoon. The stable eIF 4A transcript was used as a reference for equal loading. Half-life values are representative of 2 independent cordycepin time courses. 100 Morning (211) a: F- :1 IE 015306090120 015306090120 2 SEN] ....._ . ---....- 3 < 6 E o 1519 (51 min) 0 eIF4A ~ --_..__ ""“'"'- E Udst2(10lmin) p1519 dstZ 0.1 . . . 1 r 15 45 75 105 135 time (min) C. D Afternoon (ZT8) LLK g “L. .. 0 15 30 60 90 120 0 15 30 60 90 120 § \ ‘ ~~ E ..._ SENI C.‘”" VIC-'Hr 2 \‘ . < E o 1519 (35 min) eIF4A ufi'“-- ------ E ads!2(53min) p1519 dst2 0.01 U U , , , 15 45 75 105 135 time (min) Figure 3.8. Regulation of SEN] mRNA stability is altered in the dst2 mutant during the day. Representative northern blot analysis of cordycepin time courses performed in 1519 and dst2 plants (A) 1 hour after dawn (zeitgeberl or ZTl), and (C) 8 hours after dawn (ZT8) for SEN] and eif4A mRNAs. Samples consisted of 10 pg of total RNA isolated from the indicated time points. Quantitation of the decrease in mRN A abundance and half-life estimation for SEN] (B) in the morning and (D) in the afternoon. The stable eIF 4A transcript was used as a reference for equal loading. 10] mutations in the circadian clock (Millar et el., 1995). Seedlings were grown for 5 days in 12 hr light / 12 hr dark, transferred to 24-well plates and released into continuous light. Cotyledon movement was then recorded for 7 days The potential significance of our mRNA steady state and half-life results was emphasized by the finding that dst] was phase lagging (Figure 3.9A) and dstZ was phase leading (Figure 3.9B) in these experiments (in collaboration with Salome, P. and McClung, R.). DISCUSSION In this study, we have shown that the DST-mediated decay pathway is required for the normal oscillation of selected clock controlled genes in Arabidopsis and is critical for proper circadian regulation of the stability of these DST-containing transcripts. The stability of the SEN] and Ccr-like mRNAs was altered in the mutant under diurnal conditions and the effect in the morning and afternoon was different (Figure 3.1 and 3.2). From these findings it can be concluded that dst] is required for the proper diurnal regulation of the stability of these DST-containing transcripts. A cycling mRNA with a longer half-life will show a phase delay and reduced amplitude when compared with an mRNA with a shorter half-life, assuming that the transcription rates are similar (Wuarin et al., 1992). Under circadian conditions, the Ccr-like transcript has an increased half-life in dst] compared to the parental (Figure 3.6) in the subjective aftemoon and this posttranscriptional effect probably regulates the phase delay seen at the level of mRN A oscillation and leaf movement (Figure 3.3 and Figure 3.9A). The dampening in the mRNA decay rates seen under circadian conditions could be due to the fact that the clock is not reset in constant light leading to desynchronization such that the individual seedlings drift out of phase from each other. 102 +1519 . dstl 29; j " . ’iii" 1 23 1 l l j 4 l a * 1 + z: 27 ‘ ' ' l r - l J '8 261 ' i ' ' CL 5 “ ' j 25' ' ' ‘ WA ’EL 24 ‘ é . . [HHVJ o ' ‘ ‘ - i ; £5 23 . l ' l I j 5' 22 : l , F l i l 21 I ( I I L I 20 5.1.1.1 1 id 0 12 24 36 48 60 72 84 96 108120 Hours in continuous light Relative pixel position 4- _. , . __ , __ 0 12 24 ‘3.“45 '60 72— 84 96 108 120 Hours in continuous light Figure 3.9. Circadian rhythm of leaf movement is altered in the dst mutants. The graphs represent Arabidopsis leaf movement data taken by video cameras over a period of five days. Each peak represents the "up" position of the cotyledon as the underside of the petiole grows more than the upper side. Each trough is the "down" position as the upper side of the petiole grows more than the underside. The traces for A) dst] and parental are an average of 6 seedlings while the traces for B) dst2 and parental are average of 12 seedlings. 103 An opposite effect was seen with dst2 being phase leading in the leaf movement studies (Figure 3.98). This observation is in agreement with the half-life changes seen in dst2 and the parental plants. Even though the SEN] transcript is more stabilized in the mutant compared to the parental in the morning, the mRN A decays at a much faster rate in dst2 (Figures 3.8, compare am. and p.m.). This results in a shift in phase with the peak being reached earlier in dst2. It is evident that cis-acting elements as well as trans-acting factors are involved in the posttranscriptional regulation of clock controlled mRN As. The 3’ UTR of Drosophila per gene is AU-rich and it is known that AU-rich elements (ARES) are found in the 3'UTRs of several of the most unstable mammalian transcripts (Chen and Shyu, 1995). Both Ccr-like and SEN] contain DST-like elements in their 3 ’UTR and are primary targets of the dst] mutation. In addition, they belong to a unique subset of transcripts that are decreased in abundance in dst] but are increased in abundance in dst2 (Pérez-Amador et al., 2001). The exact nature of the DST 1 mutation is not known as yet but it is possible that it corresponds to a sequence-specific RNA-binding protein. Clock controlled RNA- binding proteins have been identified in the algae Gonyaulax polyedra (Morse et al., 1989) and Chlamydomonas reinhardtii (Mittag et al., 1994) that bind to the 3’ UTRS of several mRNAs that contain a UG-repeat region (Waltenberger et al., 2001). The circadian clock also controls the binding activity of these proteins which are hypothesized to function as translational suppressors (Mittag, 2003). Another example is the putative RNA-binding LARK protein of Drosophila that oscillates in abundance and regulates adult eclosion (McNeil et al., 1998). 104 Precedence for clock-controlled RNA-binding proteins in Arabidopsis has also been documented. AtGRP7 mRNA as well as the protein undergoes circadian oscillations with slight delay of the protein peak relative to the RNA peak (Heintzen et al., 1997). Overexpression studies have shown that the transcript and protein are linked in a negative autoregulatory circuit (Staiger, 2001). This negative autoregulation was shown to be mediated through the binding of the protein to its own pre-mRNA resulting in the formation of an alternatively spliced transcript with a premature stop codon which is rapidly degraded (Staiger et al., 2003). The role of mRNA stability in circadian control of gene expression has been further highlighted by the recent characterization of a Xenopus deadenylase, noctumin that is expressed in a rhythmic manner (Baggs and Green, 2003). A nocturnin homolog is also present in Arabidopsis (Dupressoir et al., 2001). Since noctumin is expressed in a circadian fashion, it is possible that this enzyme is responsible for the deadenylation of clock controlled mRNAs. A relevant example is the oscillation of vasopressin transcript levels in mammals where two species of mRNAs with differences in their poly(A) tail length are present at different times of the day (Robinson et al., 1988). dst] plants do not exhibit a severe effect on the clock at the molecular level since other clock controlled genes such as AtGRP7 show normal oscillation patterns in dst] (Figure 3.5). Our data predicts that DSTl probably functions downstream of the master clock and affects a subset of clock controlled genes at the level of mRNA stability. This hypothesis is further supported by the fact that DST 1 maps to a region on chromosome 1 (see Chapter 4) that does not seem to contain any genes known to be involved in clock function. Cloning of the DST I gene should help us elucidate the precise relationship 105 between the DST-mediated decay machinery and the circadian clock. The link between sequence-specific decay and the circadian clock uncovered in our experiments could be extended to other systems as a powerful tool towards unraveling circadian clock mechanisms at the posttranscriptional level. MATERIALS AND METHODS Arabidopsis strains and growth conditions All Arabidopsis thaliana plants described are from the accession Columbia. dst] mutant (from the second backcross to the parental line) and 1519 parental seeds were plated on agar plates containing 1x Murashige and Skoog salts, 1x Gamborg's vitamins, 1% sucrose and 50 ug/ml kanamycin for two weeks in an incubator set at 16 hours light (125 uE/mz) /8 hours dark and 21°C. For the circadian (free running) experiments, Arabidopsis seedlings were grown for 12 days in 16 hours light/8 hours dark and on the morning of the 12th day were transferred to continuous light for 48 hours. Lighting was provided by fluorescent light bulbs. Half-life measurements, RNA preparation and analysis Half-lives were determined as described by Seeley et al. (1992) with the following modifications. Two-week old Arabidopsis seedlings were transferred to a flask with incubation buffer. After a 30 min incubation, 3'-deoxyadenosine (cordycepin) was added to a final concentration of 0.6 mM (time 0). Tissue samples were harvested at regular intervals thereafter and frozen in liquid nitrogen. Total RNA was isolated using standard techniques. RNA was analyzed by electrophoresis on 2% formaldehyde/ 1.2% 106 agarose gels and blotted onto nylon membrane (Nytran Plus, Schleicher & Schuell, Keene, NH). DNA probes were labeled with [a-32P]dCTP by the random primer method (Feinberg and Vogelstein, 1983) and purified from unincorporated nucleotides using probe purification columns (NucTrap, Stratagene, La Jolla, CA). The RNA blots were hybridized as described in Taylor and Green (1991) using the indicated 32P-labeled probes. For a loading control, RNA blots were hybridized with a 32P-labeled cDNA probe for the Arabidopsis translation initiation factor eIF 4A (Taylor et al., 1993). SEN] is a single gene in the genome of Arabidopsis thaliana (Oh et al., 1996). Ccr-like protein has no similarity to other protein sequences in the Arabidopsis genome as determined by BLASTCLUST (ftp://Ptp.ncbi.nlm.nih.gov/blast/documents/README.bcl). In addition, BLASTN analysis (nucleotide vs. nucleotide comparison) using SEN] and Cor-like transcribed sequences as query revealed no significant similarity to other transcribed sequences in the Arabidopsis genome. Hence, northern blot probes used in this study were made by the polymerase chain reaction using EST clones as template, 111D3T7 for SEN] and 245H16T7 for Ccr-like, and SP6 and T7 vector primers. Statistical comparison of the half-lives measured at the different times of the day was performed using a repeated measure model by Xue Lan (Statistics Department, Michigan State University). Leaf movement assay Assessment of rhythmicity in leaf movement was carried out as described (Millar et al., 1995). For light entrainment, seedlings were grown under white light for 5 days in a 12 hr light / 12 hr dark photoperiod. For temperature entrainment, seedlings were grown 107 under white light for 7 days in a 12 hr 22°C / 12 hr 12°C temperature regime. On the 5th or 7th day, seedlings were transferred to 24-well cloning plates and the plates were released into continuous white light and constant temperature of 22°C. Leaf movement was recorded every 20 min over 7 days by Panasonic CCTV cameras, model WV-BP120 (Matsushita Communications Industrial, Laguna, Philippines). Post-run analysis was performed using the Kujata software program [Millar et al., 1995), and traces were analyzed by FFT-NLLS [Plautz et al., 1997). 108 REFERENCES Alabadi, D., Oyama, T., Yanovsky, M. J., Harmon, F. G., Mas, P., and Kay, S. A. (2001). Reciprocal regulation between TOC 1 and LHY/CCA] within the Arabidopsis circadian clock. Science 293, 880-883. Allada, R., Emery, P., Takahashi, J. S., and Rosbash, M. (2001). Stopping time: the genetics of fly and mouse circadian clocks. Annu Rev Neurosci 24, 1091-1119. Baggs, J. E., and Green, C. B. (2003). Noctumin, a Deadenylase in Xenopus laevis Retina. A Mechanism for Posttranscriptional Control of Circadian-Related mRNA. Curr Biol 13, 189-198. Chen, C. Y., and Shyu, A. B. (1995). AU-rich elements: characterization and importance in mRNA degradation. Trends Biochem Sci 20, 465-470. Chung, B. C., Lee, S. Y., Oh, S. A., Rhew, T. H., Nam, H. G., and Lee, C. H. (1997). The promoter activity of SENl, a senescence-associated gene of Arabidopsis, is repressed by sugars. J Plant Physiol 151, 339—345. Claridge-Chang, A., Wijnen, H., Naef, F ., Boothroyd, C., Rajewsky, N., and Young, M. W. (2001). Circadian regulation of gene expression systems in the Drosophila head. Neuron 32, 657-671. Dupressoir, A., More], A. P., Barbot, W., Loireau, M. P., Corbo, L., and Heidmann, T. (2001). Identification of four families of yCCR4- and M g2+-dependent endonuclease- related proteins in higher eukaryotes, and characterization of orthologs of yCCR4 with a conserved leucine-rich repeat essential for hCAF 1/hPOP2 binding. BMC Genomics 2, 9. Edery, I. (1999). Role of posttranscriptional regulation in circadian clocks: lessons from Drosophila. Chronobiol Int 16, 377-414. Feinberg, A. P., and Vogelstein, B. (1983). A technique for radiolabeling DNA restriction endonuclease fragments to high specific activity. Anal Biochem 132, 6-13. Frisch, B., Hardin, P. E., Hamblen-Coyle, M. J ., Rosbash, M., and Hall, J. C. (1994). A promoterless period gene mediates behavioral rhythmicity and cyclical per expression in a restricted subset of the Drosophila nervous system. Neuron 12, 555-570. Garceau, N. Y., Liu, Y., Loros, J. J ., and Dunlap, J. C. (1997). Alternative initiation of translation and time-specific phosphorylation yield multiple forms of the essential clock protein FREQUENCY. Cell 89, 469-476. Grima, B., Lamouroux, A., Chelot, E., Papin, C., Limbourg-Bouchon, B., and Rouyer, F. 109 (2002). The F-box protein slimb controls the levels of clock proteins period and timeless. Nature 420, 178-182. Gutierrez, R. A. (2003). Inherent and regulated mRNA stability in A. thaliana. PhD Dissertation, Michigan State University, 67-98. Gutierrez, R. A., Ewing, R. M., Cherry, J. M., and Green, P. J. (2002). Identification of unstable transcripts in Arabidopsis by cDNA microarray analysis: rapid decay is associated with a group of touch- and specific clock-controlled genes. Proc Natl Acad Sci U S A 99,11513-11518. Harmer, S. L., Hogenesch, J. B., Straume, M., Chang, H. S., Han, B., Zhu, T., Wang, X., Kreps, J. A., and Kay, S. A. (2000). Orchestrated transcription of key pathways in Arabidopsis by the circadian clock. Science 290, 2110-2113. Harmer, S. L., Panda, 8., and Kay, S. A. (2001). Molecular bases of circadian rhythms. Annu Rev Cell Dev Biol 17, 215-253. He, Q., Cheng, P., Yang, Y., Yu, H., and Liu, Y. (2003). FWDl-mediated degradation of FREQUENCY in Neurospora establishes a conserved mechanism for circadian clock regulation. Embo J 22, 4421-4430. Heintzen, C., Nater, M., Apel, K., and Staiger, D. (1997). AtGRP7, a nuclear RNA- binding protein as a component of a circadian-regulated negative feedback loop in Arabidopsis thaliana. Proc Natl Acad Sci U S A 94, 8515-8520. Iwamoto, M., H.Higo, and K.Higo. (2000). Differential diurnal expression of rice catalase genes: the 5'- flanking region of CatA is not sufficient for circadian control. Plant Science 151, 39—46. Kim, B. Y., Bae, K., Ng, F. S., Glossop, N. R., Hardin, P. E., and Edery, I. (2002). Drosophila CLOCK protein is under posttranscriptional control and influences light- induced activity. Neuron 34, 69-81. Kim, W. Y., Geng, R., and Somers, D. E. (2003). Circadian phase-specific degradation of the F-box protein ZTL is mediated by the proteasome. Proc Natl Acad Sci U S A 100, 4933-4938. Kramer, C., Loros, J. J ., Dunlap, J. C., and Crosthwaite, S. K. (2003). Role for antisense RNA in regulating circadian clock function in Neurospora crassa. Nature 421, 948-952. Liu, Y., Loros, J ., and Dunlap, J. C. (2000). Phosphorylation of the Neurospora clock protein FREQUENCY determines its degradation rate and strongly influences the period length of the circadian clock. Proc Natl Acad Sci U S A 97, 234-239. 110 Liu, Y., Merrow, M., Loros, J. J., and Dunlap, J. C. (1998). How temperature changes reset a circadian oscillator. Science 281, 825-829. Majercak, J ., Sidote, D., Hardin, P. E., and Edery, I. (1999). How a circadian clock adapts to seasonal decreases in temperature and day length. Neuron 24, 219-230. Martinek, S., Inonog, S., Manoukian, A. S., and Young, M. W. (2001). A role for the segment polarity gene shaggy/GSK-3 in the Drosophila circadian clock. Cell 105, 769- 779. McClung, C. R. (2001). Circadian Rhythms in Plants. Annu Rev Plant Physiol Plant Mol Biol 52, 139-162. McDonald, M. J ., and Rosbash, M. (2001). Microarray analysis and organization of circadian gene expression in Drosophila. Cell 107, 567-578. McNeil, G. P., Zhang, X., Genova, G., and Jackson, F. R. (1998). A molecular rhythm mediating circadian clock output in Drosophila. Neuron 20, 297-303. Michael, T. P., and McClung, C. R. (2003). Enhancer trapping reveals widespread circadian clock transcriptional control in Arabidopsis. Plant Physiol 132, 629-639. Millar, A. J ., and Kay, S. A. (1991). Circadian Control of cab Gene Transcription and mRNA Accumulation in Arabidopsis. Plant Cell 3, 541-550. Millar, A. J ., Straume, M., Chory, J., Chua, NH, and Kay, S. A. (1995). The regulation of circadian period by phototransduction pathways in Arabidopsis. Science 267, 1163- 1166. Mittag, M. (2003). The function of circadian RNA-binding proteins and their cis-acting elements in microalgae. Chronobiol Int 20, 529-541. Mittag, M., Lee, D. H., and Hastings, J. W. (1994). Circadian expression of the luciferin- binding protein correlates with the binding of a protein to the 3' untranslated region of its mRNA. Proc Natl Acad Sci U S A 91, 5257-5261. Miyazaki, T., Kuwano, H., Kato, H., Ando, H., Kimura, H., Inose, T., Ohno, T., Suzuki, M., Nakajima, M., Manda, R., et al. (2003). Correlation between serum melatonin circadian rhythm and intensive care unit psychosis afier thoracic esophagectomy. Surgery 133, 662-668. Mizoguchi, T., Wheatley, K., Hanzawa, Y., Wright, L., Mizoguchi, M., Song, H. R., Carre, I. A., and Coupland, G. (2002). LHY and CCAl are partially redundant genes required to maintain circadian rhythms in Arabidopsis. Dev Cell 2, 629-641. 111 Morse, D., Pappenheimer, A. M., Jr., and Hastings, J. W. (1989). Role of a luciferin- binding protein in the circadian bioluminescent reaction of Gonyaulax polyedra. J Biol Chem 264, 11822-11826. Naidoo, N., Song, W., Hunter-Ensor, M., and Sehgal, A. (1999). A role for the proteasome in the light response of the timeless clock protein. Science 285, 1737-1741. Nelson, D. C., Lasswell, J ., Rogg, L. E., Cohen, M. A., and Bartel, B. (2000). FKF l, a clock-controlled gene that regulates the transition to flowering in Arabidopsis. Cell 101, 331-340. Oh, S. A., Lee, S. Y., Chung, I. K., Lee, C. H., and Nam, H. G. (1996). A senescence- associated gene of Arabidopsis thaliana is distinctively regulated during natural and artificially induced leaf senescence. Plant Mol Biol 30, 739-754. Panda, S., Poirier, G. G., and Kay, S. A. (2002). tej defines a role for poly(ADP- ribosyl)ation in establishing period length of the arabidopsis circadian oscillator. Dev Cell 3, 51-61. Perez-Amador, M. A., Lidder, P., Johnson, M. A., Landgraf, J ., Wisman, E., and Green, P. J. (2001). New molecular phenotypes in the dst mutants of Arabidopsis revealed by DNA microarray analysis. Plant Cell 13, 2703-2717. Pilgrim, M. L., Caspar, T., Quail, P. H., and McClung, C. R. (1993). Circadian and light- regulated expression of nitrate reductase in Arabidopsis. Plant Mol Biol 23, 349-364. Plautz, J .D., Straume, M., Stanewsky, R., Jamison, C.F., Brandes, C., Dowse, H.B., Hall, J .C., and Kay, SA. (1997). Quantitative analysis of Drosophila period gene transcription in living animals. J Biol Rhythms 12, 204-217. Price, J. L., Blau, J., Rothenfluh, A., Abodeely, M., Kloss, B., and Young, M. W. (1998). double-time is a novel Drosophila clock gene that regulates PERIOD protein accumulation. Cell 94, 83-95. Robinson, B.G., Frim, D., M., Schwartz, W., J. and Majzoub, J .A. (1988). Vasopressin mRNA in the suprachiasmatic nuclei: daily regulation of polyadenylate tail length. Science 241, 342—344. Schaffer, R., Landgraf, J ., Accerbi, M., Simon, V., Larson, M., and Wisman, E. (2001). Microarray analysis of diurnal and circadian-regulated genes in Arabidopsis. Plant Cell 13, 113-123. Schaffer, R., Ramsay, N., Samach, A., Corden, S., Putterill, J ., Carre, I. A., and Coupland, G. (1998). The late elongated hypocotyl mutation of Arabidopsis disrupts circadian rhythms and the photoperiodic control of flowering. Cell 93, 1219-1229. 112 Seeley, K. A., Byme, D. H., and Colbert, J. T. (1992). Red Light-Independent Instability of Cat Phytochrome mRNA in Vivo. Plant Cell 4, 29-38. So, W., V., and Rosbash, M. (1997). Post-transcriptional regulation contributes to Drosophila clock gene mRNA cycling. Embo J 16, 7145-7155. Somers, D. E., Schultz, T. F., Milnamow, M., and Kay, S. A. (2000). ZEITLUPE encodes a novel clock-associated PAS protein from Arabidopsis. Cell 101, 319-329. Staiger, D. (2001). RNA-binding proteins and circadian rhythms in Arabidopsis thaliana. Philos Trans R Soc Lond B Biol Sci 356, 1755-1759. Staiger, D., Zecca, L., Kirk, D. A., Apel, K., and Eckstein, L. (2003). The circadian clock regulated RNA-binding protein AtGRP7 autoregulates its expression by influencing alternative splicing of its own pre-mRNA. Plant J 33, 361-371. Stanewsky, R., Lynch, K. S., Brandes, C., and Hall, J. C. (2002). Mapping of elements involved in regulating normal temporal period and timeless RNA expression patterns in Drosophila melanogaster. J Biol Rhythms 17, 293-306. Sugano, S., Andronis, C., Green, R. M., Wang, Z. Y., and Tobin, E. M. (1998). Protein kinase CK2 interacts with and phosphorylates the Arabidopsis circadian clock-associated 1 protein. Proc Natl Acad Sci U S A 95, 11020-11025. Sugano, S., Andronis, C., Ong, M. S., Green, R. M., and Tobin, E. M. (1999). The protein kinase CK2 is involved in regulation of circadian rhythms in Arabidopsis. Proc Natl Acad Sci U S A 96, 12362-12366. Suri, V., Hall, J. C., and Rosbash, M. (2000). Two novel doubletime mutants alter circadian properties and eliminate the delay between RNA and protein in Drosophila. J Neurosci 20, 7547-7555. Suri, V., Lanjuin, A., and Rosbash, M. (1999). TIMELESS-dependent positive and negative autoregulation in the Drosophila circadian clock. Embo J 18, 675-686. 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. Taylor, C. B., and Green, P. J. (1991). Genes with homology to fungal and S-gene RNases are expressed in Arabidopsis thaliana. Plant Physiol 96, 980-984. Ueda, H. R., Matsumoto, A., Kawamura, M., lino, M., Tanimura, T., and Hashimoto, S. (2002). Genome-wide transcriptional orchestration of circadian rhythms in Drosophila. J Biol Chem 277, 14048-14052. 113 Waltenberger, H., Schneid, C., Grosch, J. 0., Bareiss, A., and Mittag, M. (2001). Identification of target mRNAs for the clock-controlled RNA-binding protein Chlamy 1 from Chlamydomonas reinhardtii. Mol Genet Genomics 265, 180-188. Wang, Z. Y., and Tobin, E. M. (1998). Constitutive expression of the CIRCADIAN CLOCK ASSOCIATED l (CCAl) gene disrupts circadian rhythms and suppresses its own expression. Cell 93, 1207-1217. Wuarin, J ., Falvey, E., Lavery, D., Talbot, D., Schmidt, E., Ossipow, V., Fonjallaz, P., and Schibler, U. (1992). The role of the transcriptional activator protein DBP in circadian liver gene expression. J Cell Sci Suppl 16, 123-127. Zhong, H. H., Resnick, A. S., Straume, M., and Robertson McClung, C. (1997). Effects of synergistic signaling by phytochrome A and cryptochromel on circadian clock- regulated catalase expression. Plant Cell 9, 947-955. 114 CHAPTER 4 GENETIC MAPPING OF dst] 115 INTRODUCTION Insertional mutagenesis techniques have been widely used to identify the function of unknown genes using T-DNAs (Alonso et al., 2003; Krysan et al., 1999) and transposable elements (Aarts et al., 1995; Weigel et al., 2000). However these methods cannot create partial loss-of-function or gain-of-function alleles. In contrast, chemical mutagenesis causes point mutations generating different types of alleles; this could be an advantage, especially if the genes in the regulatory pathway are essential. Furthermore, because EMS mutagenesis can result in a higher mutational frequency than T-DNA mutagenesis, fewer plants need to be screened in order to find a mutation in any given gene. The mutated genes then have to be isolated based on their map position. Map-based cloning is an indirect approach; mapping narrows down the genetic interval that contains a mutation by sequentially excluding all other parts of the genome. Mapping with a high resolution requires a high density of genetic markers around the region of interest. This has been greatly facilitated by the sequencing of Columbia and Landsberg erecta (Ler) ArabidOpsis ecotypes which are sufficiently divergent to support the design of molecular markers at this high density. Also, the sequence of the Arabidopsis genome has been completed which should expedite pinpointing the genes after fine mapping. Some molecular markers commonly used in mapping experiments include simple sequence length polymorphisms (SSLPs), cleaved amplified polymorphic sequences (CAPS) and derived CAPS (dCAPS) (Lukowitz et al., 2000). These markers are co- dominant and are PCR-based. SSLP (Bell and Ecker, 1994) markers are based on the variability of short repetitive sequences. These markers are amplified using PCR and 116 exhibit ecotype-specific polymorphisms on agarose/polyarylamide gels. CAPS markers detect polymorphisms that occur in restriction sites (Konieczny and Ausubel, 1993) while dCAPS are capable of exploiting all single nucleotide changes whereby mismatches in PCR primers are used to create restriction sites (Neff et al., 1998). It is also possible to generate additional markers in the region of interest by utilizing single nucleotide polymorphism (SNP) changes as well as insertion-deletion (InDel) differences. On comparing Arabidopsis Columbia and Ler genomic sequences, there is one SNP every 3.3 kb and one InDel every 6.6 kb (Jander et al., 2002). A total of 56,670 polymorphisms have been deposited on The Arabidopsis Information Resource web site (http://www.arabidopsis.org/Cereon) including 37,344 SNPs, 18,579 InDels, and 747 Large Indels. To better understand the molecular mechanisms involved in sequence-specific mRNA decay, specially designed transgenic plants were used to select for mutants that are defective in DST-mediated mRNA degradation using an ethyl methane-sulfonate- mutagenized population (Johnson et al., 2000). The mutants that were isolated were extremely rare (3/~800,000), were all incompletely dominant and lacked any visible aberrant phenotype. This could be because the DST genes are essential and the dst mutants are weak alleles, or it is possible that the DST genes are part of gene families that have redundant members. Cloning of the DST genes should help in the elucidation of the components of the DST-recognition and decay machinery. dst] was chosen for cloning since it is the best characterized but these experiments should be applicable to the other dst mutants as well. The results described in this chapter demonstrate that RAP2.4, first identified in the microarray experiments (see Chapter 2), is as an excellent endogenous 117 molecular marker for scoring dst] . Using RAP2.4 expression levels, homozygous dst] mutants were followed independent of the transgene and the mutation has been mapped to a 107 kb interval on the bottom of chromosome 1. RESULTS dst] mutation does not appear to be linked to the transgene The dst] phenotype is a 3-4 fold increase in HPH mRN A levels and heterozygous plants exhibit half this increase (Johnson et al., 2000). The requirement of the transgene to score the dst] mutants had several drawbacks since both the dst] mutation and the transgene were segregating in the mapping population. Consequently, in the F2, HPH- DST mRN A abundance reflected the dosage of the transgene and the dst] mutation. As a result larger number of plants had to be screened to identify homozygous dst] mutants. To overcome this dependence on the transgene, inheritance of RAP2.4 mRNA abundance was examined for the mutant. Progeny of the second backcross to WT (1519-31) (F 1) and the progeny of self-fertilizations of these plants (F2) were used in these studies. The F1 plants showed intermediate levels of RAP2.4 mRN A abundance as expected (data not shown). Segregation of increased RAP2.4 mRNA abundance in the F2 populations was also consistent with semi-dominance. Of 240 F2 plants segregating dst], 57 were observed with high RAP2.4 mRNA abundance (9.3 i 3.3 fold higher than 1519-31), 111 had intermediate RAP2.4 mRNA abundance (2.07 i 0.8) and 72 showed RAP2.4 mRNA abundance that was similar to parental (0.92 i 0.2) (Figure 4.1). Segregation in the dst] F2 populations is most easily explained by a ratio of 1:2:1 as would be expected for the 118 # 329 # 563 # 431 # 363 1519 1519 # 489 # 411 # 421 # 424 RAP2. 4 eIF4A “n. -_. ~~ - .. no. 5.... out. u...- Figure 4.1. mRNA abundance of RAP2.4 in F2 plants from the second back-cross of dst] relative to WT (1519-31). 8 representative plants are shown. Total RNA was prepared from rosette leaves and analyzed by Northern blot hybridization. Radiolabeled probes, prepared against the indicated transcripts, were used in sequential hybridizations of the same blot. The numbers indicate individual plants. #411 is an example of an F2 plant with high RAP2.4 mRNA levels, #431 with intermediate and #329 with RAP2.4 mRNA levels similar to WT (1519-31). 119 inheritance of a semi-dominant single gene, indicating that the mutation is not linked to the transgene (Table 4.1). Use of RAP2.4 as an endogenous marker to score homozygous mutants It is known that the dst] mutation is associated with circadian rhythms (Pérez- Amador et al., 2001). In order to determine if RAP2.4 could be used as a phenotypic marker for scoring the dst] mutation in the mapping population, it was critical to establish that RAP2.4 was not circadian regulated. Towards this end, mRNA half-life experiments were carried out in dst] and 1519 plants in the morning and afiemoon. A comparison of Figures 4.2B and 4.2D shows that RAP2.4 mRNA is more stable in dst] than in the parental 1519 line when mRNA decay is monitored in the morning (zeitgeber time 1/ZT1) and in the afternoon (ZT 8) and all of the decay curves are nearly identical under the two conditions. These mRN A stability measurements indicate that the effect of dst] on RAP2.4 is insensitive to the time of day. It is also of interest to note that RAP2.4 was not regulated by circadian rhythms in the dst] microarray experiments. Fine mapping of the dst] locus The first step in a mapping experiment is to generate a mapping population. Since the mutation is in a Columbia (glI) background, dst] was crossed to Landsberg erecta (gl] -1 ). F2 plants from this cross were scored by Northern blot analysis for S-fold or higher RAP2.4 mRNA abundance compared to parental 1519 plants. 78 dst]/dst] F2 plants were identified and genomic DNA was extracted from them for PCR. Mapping of the DST] gene to chromosome 1 was accomplished by linkage analysis to SSLP markers 120 Table 4.1. Segregation of increased RAP2.4 mRNA abundance in the progeny of crosses between dst1 and 1519-31 (DST1). Cross Class RAP2.4' n 12, P’ DST1/dst1 X DSTT/dst 1 F2 High 9.3 i 3.3 57 Int 2.07 i 0.8 111 1.9 P>0.3 WT 0.92 i 0.2 72 Int, intermediate. WT, wild-type. * RAP2.4 mRNA abundance: (RAP2.4/eif4A) segregating class/ (RAP2.4/eif4A) WT i standard error. f 12 calculated for 1:2:1 segregation of WT/lnt/High RAP2.4 mRNA abundance DST/DST, WT (1519-31). 121 Morning (ZTl) 1 15 30 60 90 120 15 30 60 90 120 on E . II as: 1‘ d .E RAP24 ' Q . ' . g h .2 eIF4A fi_,._.__ .................. 2 <1 p1519 dst1 E ° 1519(53 min) a a dst1 (76 min) 0.1 U I Y I I 15 45 75 105 135 time (min) C- Afternoon (ZT8) D- ] Rh. 15 30 60 90 120 15 30 60 90 120 5D RAP2.4 $4.! .. is: u an a. .5 '3 eIF4A . . 1.. ... .. 5 i- 4: p1519 (15!! E E °1519(51 min) ‘3 dst] (70 min) 0.1 . 15 45 75 105 135 time (min) Figure 4.2. RAP2.4 mRNA is stabilized in the dst1 mutant and the decay kinetics are similar in the am and p.m. Representative northern blot analysis of cordycepin time courses performed in 1519 and dst] plants (A) 1 hour afier dawn (zeitgeberl or ZTl), and (C) 8 hours after dawn (ZT8) for RAP2.4 and eif4A mRNAs. Samples consisted of 10 pg of total RNA isolated from the indicated time points. Quantitation of the decrease in mRN A abundance and half-life estimation for RAP2.4 (B) in the morning and (D) in the aftemoon. The stable eIF 4A transcript was used as a reference for equal loading. 122 (Figure 4.3). Recombination events between molecular markers and the dst1 locus in the homozygous mutants were analyzed to confirm that dst1 was located between markers f16p17 and f1n192. Additional polymorphic SSLP and CAPS markers were identified in this interval and two markers (f22c121 and flnl9l) were found that were 5.8% recombination apart (Figure 4.4). A larger F2 population was then generated for fine-resolution mapping. The closely linked SSLP markers (107 kb), f22c121 and f1n191, were used to screen a mapping population of 970 F2 plants. 24 recombinants within the interval defined by the flanking molecular markers were identified. RNA gel blot analysis was used to determine whether the recombinants were homozygous mutant, homozygous wild type, or heterozygous at the DST 1 locus. This information together with additional markers is now being used to further narrow down the interval. DISCUSSION The strategy for scoring dst] was tailored in order to avoid dependence on the transgene and thereby expedite cloning. The reliance of the dst] phenotype on the transgene is an immense drawback for the generation of mapping populations due to the small magnitude of change in HPH-DST mRNA levels. This scenario is further complicated by the semi-dominance of dst1. A locus like dst] presents unique challenges as it does not exhibit any visible phenotype and therefore various approaches were devised to overcome this difficulty. Early on in the mapping process, scoring only the very highest HPH-DS T mRN A levels as dst] homozygotes was tried but this proved to be very tedious because a number of homzygotes were lost due to low transgene dosage. 123 I II III IV V lw 5 nga 63 (54%) — l— ciw z c —. — nga 162 —— CTR] ciw12_— ciw3(56%) "'— ciw ll -— ciw6 __ ciw8 —— Cb ciw7 —— d) nga1126 (52%) “'- ciw4 J— ciw 1 (12%) q— a 280 (9°/) nga 6 - — nga o _— PHYC —— nga I68 —— “331107 -— ciw 9"— nga 111 (14%) —— ciw 10 -- Figure 4.3. DST] maps to chromosome I. Bulked segregant analysis was used to assign a rough map position to DST 1 on chromosome 1. SSLP marker positions and the recombination frequencies in the mapping experiment are indicated. 124 nga 63 ‘- ciw 12 '1'- ciw 1 (12%) nga 128 (10%) t20n2 (9.2%) nga 280 (9%) fl4j16 (9.4%) __ fl4j9 (8.2%) -_ f25p12 (9%) 5.2 Mb -_ t2k10(7%) -_ tl3d8 (5.1%) .. f8a5 (6.2%) f11p17 (6.3%) nfllp17 (6.3%) ' f‘ 6P” (49%) _ tl3mll (7.1%) .. t8k4 (7%) l- f2kll (3.9%) _ nf19k23 (8.3%) __ 124d7 (3.9%) - fl6p17 (4.9%) ngalll'l I “P 1220122 (3.9%) - 12ch21 (2.9%) :] -” flnl91 (2.9%) 107 kb 720 kb :- r1n192 (3.9%) - flnl92 (3.9%) Figure 4.4. Map-based cloning of DST 1 . SSLP and CAPS marker positions with the respective recombination frequencies are indicated. 125 Microarray technology was used as a novel tool towards the cloning of dst] since it identified a robust endogenous molecular marker. The experiments described in this chapter suggested that the RAP2.4 transcript has several features which make it an ideal phenotypic marker for dst]. The 3’ UTR of RAP2.4 contains DST-like sequences and subsequent half-life analysis indicated that it is a direct target of DST-mediated mRN A decay. Also, RAP2.4 mRNA levels exhibited greater increases in abundance than HPH- DST mRNA levels in dst] plants (5-fold compared to 3-4 fold). An added advantage was that RAP2.4 could be used to screen mapping populations and follow the dst] phenotype even in plants lacking the transgene. A larger number of homozygous mutants could thus be scored in a more accurate manner because the presence of the transgene was not necessary. Finally, an appealing feature of RAP2.4 was that even though it is a primary target of the DST-mediated decay pathway, it is not regulated in a circadian fashion. Harvesting hundreds of samples for RNA preparation from a mapping population is a lengthy and time-consuming process and using RAP2.4 as a marker alleviated concerns about circadian variations occurring during that interval. The 107 kb interval containing DST] spans 27 genes, including one t-RNA gene. Three genes in this region encode transcription factors, one of which belongs to the RAP2 (related to Apetala 2) family and RAP2.4 mRN A levels are known to be elevated in dst1. Three genes encode proteins which could be functioning in protein degradation. There are several evolutionary links that seem to exist between RNA metabolism, protein degradation and ubiquitin signaling pathways suggesting that these processes interact with one another (Anantharaman et al., 2002). Numerous ubiquitin, ubiquitin E3 ligases and other proteins involved with the proteasome show fusions with various domains of 126 proteins functioning in RNA metabolism (Aravind and Ponting, 1998; Fang et al., 2000) implicating a role for RNA binding proteins in bringing the proteolytic machinery to the RNA-bound complexes. Once a small interval containing DST] is defined, candidate genes in that region and/or the entire interval will be sequenced to identify the mutation. Construction of a cosmid library is in progress to aid in sequencing. Standard complementation tests to confirm the identity of DST 1 might not be straightforward due to the semi-dominant nature of the mutation. As such, the mutant gene might have to be introduced into wild type plants to phenocopy the mutation, which would also be greatly facilitated by the cosmid library being generated. MATERIALS AND METHODS Plant Material dst] mutant plants described in this report are from the accession Columbia (gll) derived from EMS mutagenized Arabidopsis populations (Johnson et al., 2000). 1519 and dst1 plants, from the second backcross to the parental line, were grown in grth chambers under 16 hr light and 60% relative humidity at 20°C. Rosette leaves were harvested from 35- to 40-day-old plants. All tissue was harvested from plants grown in parallel under the same conditions in different growth chambers. Half-life measurements, total RNA Extraction and RNA Blot Hybridization Half-lives were determined as described by Seeley et al. (1992). Rosette leaves from Arabidopsis plants were transferred to a flask with incubation buffer. After a 30 min 127 incubation, 3'-deoxyadenosine (cordycepin) was added to a final concentration of 0.6 mM (time 0). Tissue samples were harvested at regular intervals thereafter and frozen in liquid nitrogen. Total RNA from leaf samples was extracted as previously described (Newman et al., 1993). 10 ug of total RNA or was analyzed by electrophoresis on 2% formaldehyde/1.2% agarose gels and blotted onto nylon membrane (Nytran Plus, Schleicher & Schuell, Keene, NH). DNA probes were labeled with [or-32P]dCTP by the random primer method (F einberg and Vogelstein, 1983) and purified from unincorporated nucleotides using probe purification columns (N ucTrap, Stratagene, La Jolla, CA). The RNA blots were hybridized as described in Taylor and Green (1991) using the indicated 32P-labeled probes. For a loading control, RNA blots were hybridized with a 32P-labeled cDNA probe for the Arabidopsis translation initiation factor eIF 4A (Taylor et al., 1993). Genomic DNA extraction and markers for mapping Genomic DNA was prepared from 1519 and dst1 plants as previously described (Saghai-Maroofet al., 1984). Initial mapping of the dst] locus was carried out using SSLP markers (Bell and Ecker, 1994; http://genome.bio.upenn.edu/SSLP_info/SSLP.html). Primers were designed from available sequences and used to amplify genomic DNA by PCR. The primers used for genotyping the various markers are listed in Table 4.2. The CAPS PCR products were digested with the appropriate restriction enzymes according to the manufacturer’s instructions. All PCR products were analyzed on high resolution agarose gels (4%) or 15% polyacrylamide gels. 128 Table 4.2. Oligonucleotides and restriction enzymes used to detect various SSLP and CAPS markers Marker Left oligonucleotide Right oligonucleotide Enzyme 464030 ggcccattcacaacagagat ctcgcgaaagatcgagattc Alul 464305 tcaggttccttctccgtcat ccaagaacaaatggctgtcc Tan 464424 atcgacggaaaacaaaaagc caaagcgcgtgaataacaac Tsp509I 464476 tgggttatgtcgcgagagtt gtcaattccttggctgcatt AluI 464477 gcatccgttgaggattttgt ccatgaacattgtgtatgtgagaa Alul 464507 gcctaaacacagtgagggaga tgcaaatgcaaacaacaaaga Sau3AI 479633 actttaaagggcccaaatgc cagctgtgttggtcatggag HindIII 479657 tgccactattgactaggttttctg cagcttctgcaaaacgaaga Tsp509I 479671 acaacaggagcaggaaccac gccctgatgcttacgacaat Hian 479760 ccactctcaacaattcccaag acagatcgtgcttgatgtcc Sau3AI 479762 atatccgggacatcaagcac tggatccatctgtcttttaacg Tsp509I 479788 gccaatgtctagccaccaaa tctaccaccgtttagccactg BalI 479790 ggctaatcagccacaaattca gtcttcgaatcgggttgaaa Tsp509I 4798 14 ttgccagtttggagatgaga caacaaaaatcacataacgatttca SpeI 4798 15 cctgcatgggaagaaaaaga ggtcggtttgaccttcttca HindIII 479833 tctgcttcggtttcgttctt gctgttgaatcagagcacca Hinfl 47987 1 gctaagcacgcataggtgaa ttcacgaggaaaattcaaacg pr l 881 23 853 l 99 tggtccaaggatacaaatcaca aacgacattgtttcacctgct Bstz 1 71 23950913 agtctgcgacgagtgaaggt tcttccacctcttcacactca pr1881 2395278 1 ac gcccatctaattcccatt ttttggcctcaacaaggttc Tsp509I 23953329 gcttgaggaatccaaacaatg ggttcgtcccagtcagagtc BerI 23954354 ggtaaaaagaaaatgccattcg cctcacggttatggatctgg NheI 23954862 ccgtgagggatatagccaaa tccaaggccattaagaccac NspI 23955033 ggtagcgacagcgactgaat gccattcagccgtaaactgt NdeI SNP17905 tttatttcggcccaagtctg gacctcgcaacaattggact Sau3AI SNP17907 tgaagtggtgacggtaaaatga catggaatcattttgtttagttcg SphI CER464688 atgaactgtggtacgcgaat aaaatttctttcctttccgtttt - CER464689 cgccgttttcgttgataaat ccgctctcctccattgatag - CER464692 tggttatggtttgacattattgc attggcccatttgaagagtg - CER464693 gatggtcggctctcactctc ttgagttggacggtggagat — F 1301 l_1 1m aaaagaaaaggcttgggattg tgggacacagaacttgttgc - F 16G l 6_1 Om aaaatc gacacatcactaagtc g tgatttcgcaaaaacgaatg - F16M 1 9_1 6b cggtagatgatttccgatgtt atgcgtttttcgtgaattgg - F 16M 1 9_8m cgctttttcacgaatttaaacc aactcgccattgacacaaca - F 16P1 7_8b acacgagagagcaatgatcc ttcgcactgcaaattgctta - F1N19__10e aacgaaacaggggactgaga ttcttggctttcctcttcca - F 1N1 9_9b cgtctagtttcgcggtgttt gcgtcatcaccatcatcatc - F22C12_11e tgaaaacatgcaaagggaaa ttcatattttcaatctcttgacttttt - F 22C 1 2_l 2e ttgtatcttgtgtcaccgtcaa tgatttgagtttaaaccatgttc g - F22C12_l 2ve tcctcttcgtcttcttcgtca aaagaccggccaaagaaaat - F22C12_14b gatttcgccggtgatgttac gtcggcccaattgattttt - 129 F 22C 12_1 5b F 22C 12_l 8b F23N19_6m F23N19_7e F24D7_17m F24D7_18m F24D7_9e F 240 1_l 7m F2Kl l_l 1m FSIl4_1 1b T12P18_6b T12P18_8m T13D8_15m T13D8_9b T1F9__4e T3P l 8_6b ccagtccacaaggagtcca cctgctcaaatgcgttacaa agatggctgtgccttctagg tgaagatctgaaatgaccctaaaa ttcagcatgcttattttattagcc aaccagttttcaaatcaactgaag tccctctcctcctttctcatt tgtattgtcacaaaaatgcaaca cggttgtgactcccctaaaa tgcaacgaccaagaaattga tgcgcagttatctccttttt agataaatgcatcaacaaattgac gaaaacggaatctgcaaaaca agaaaaggcccatggaatct gtaggaggagccatggattg ggtggaaaaggtggtggtaa tgcccacacaaacaatttca tccacgagtcagaaggatga cagcattccccaactctttc tttctcaagcttttgaatgttttct cgagcgccttactctgtgat tggtggtgttgaggtaccaa gcaggcgatcagacattttt ctcggacccgttacaagaaa ttggctaagcatcttttccatt gaagaagtgcattgcttgaca aaagagtataatgtcatgactcacgta acccacctcacactctctcc aattggcttttaaaacaatgtca agcagaaacagttggaaacga aacaaaagcaaatctatagttgttcac accactacgggaggaggtct 130 REFERENCES Aarts, M. G., Corzaan, P., Stiekema, W. J ., and Pereira, A. (1995). A two-element Enhancer-Inhibitor transposon system in Arabidopsis thaliana. Mol Gen Genet 247, 555- 564. Alonso, J. M., Stepanova, A. N., Leisse, T. J., Kim, C. J ., Chen, H., Shinn, P., Stevenson, D. K., Zimmerman, J ., Barajas, P., Cheuk, R., et al. (2003). Genome-wide insertional mutagenesis of Arabidopsis thaliana. Science 301, 653-657. Anantharaman, V., Koonin, E.V., and Aravind, L. (2002). Comparitive genomics and evolution of proteins involved in RNA metabolism. Nucl Acids Res 30, 1427-1464. Aravind, L., and Ponting, GP. (1998). Homologues of 26S proteasome subunits are regulators of transcription and translation. Protein Sci 7, 1250-1254. Bell, C. J ., and Ecker, J. R. (1994). Assignment of 30 microsatellite loci to the linkage map of Arabidopsis. Genomics 19, 137-144. Fang, S., Jensen, J .P., Ludwig, R.L., Vousedn, K.H., and Weismann, A.M. (2000). Mdm2 is a RING finger-dependent ubiquitin protein ligase for itself and p53. J Biol Chem 275, 8945-8951. Feinberg, A. P., and Vogelstein, B. (1983). A technique for radiolabeling DNA restriction endonuclease fragments to high specific activity. Anal Biochem 132, 6-13. Jander, G., Norris, S. R., Rounsley, S. D., Bush, D. F ., Levin, I. M., and Last, R. L. (2002). Arabidopsis map-based cloning in the post-genome era. Plant Physiol 129, 440- 450. Johnson, M. A., Perez-Amador, M. A., Lidder, P., and Green, P. J. (2000). Mutants of Arabidopsis defective in a sequence-specific mRN A degradation pathway. Proc Natl Acad Sci U S A 97, 13991-13996. Konieczny, A., and Ausubel, F. M. (1993). A procedure for mapping Arabidopsis mutations using co-dominant ecotype-specific PCR-based markers. Plant J 4, 403-410. Krysan, P. J ., Young, J. C., and Sussman, M. R. (1999). T-DNA as an insertional mutagen in Arabidopsis. Plant Cell 11, 2283-2290. Lukowitz, W., Gillmor, C. S., and Scheible, W. R. (2000). Positional cloning in Arabidopsis. Why it feels good to have a genome initiative working for you. Plant Physiol 123, 795-805. Neff, M. M., Neff, J. D., Chory, J ., and Pepper, A. E. (1998). dCAPS, a simple technique 131 for the genetic analysis of single nucleotide polymorphisms: experimental applications in Arabidopsis thaliana genetics. Plant J 14, 387-392. Newman, T. C., Ohme-Takagi, M., Taylor, C. B., and Green, P. J. (1993). DST sequences, highly conserved among plant SAUR genes, target reporter transcripts for rapid decay in tobacco. Plant Cell 5, 701-714. Perez-Amador, M. A., Lidder, P., Johnson, M. A., Landgraf, J ., Wisman, E., and Green, P. J. (2001). New molecular phenotypes in the dst mutants of Arabidopsis revealed by DNA microarray analysis. Plant Cell 13, 2703-2717. Saghai-Maroof, M. A., Soliman, K. M., Jorgensen, R. A., and Allard, R. W. (1984). Ribosomal DNA spacer-length polymorphisms in barley: mendelian inheritance, chromosomal location, and population dynamics. Proc Natl Acad Sci U S A 81, 8014— 8018. Seeley, K. A., Byme, D. H., and Colbert, J. T. (1992). Red Light-Independent Instability of Cat Phytochrome mRNA in Vivo. Plant Cell 4, 29-38. Taylor, C. B., and Green, P. J. (1991). Genes with homology to fungal and S-gene RNases are expressed in Arabidopsis thaliana. Plant Physiol 96, 980-984. Taylor, C. B., Bariola, P. A., DelCardayré, S. B., Raines, R. T., and Green, P. J. (1993). RN S2: A senescence-associated RNase of Arabidopsis that diverged from the S-RNases before speciation. Proc Natl Acad Sci USA 90, 5118-5122. Weigel, D., Ahn, J. H., Blazquez, M. A., Borevitz, J. 0., Christensen, S. K., Fankhauser, C., Ferrandiz, C., Kardailsky, I., Malancharuvil, E. J ., Neff, M. M., et a1. (2000). Activation tagging in Arabidopsis. Plant Physiol 122, 1003-1013. 132 CHAPTER 5 CHARACTERIZATION OF dst3, A NEW GENE IN THE DST- MEDIATED mRNA DEGRADATION PATHWAY 133 INTRODUCTION Rates of mRNA degradation are highly variable in eukaryotic cells and these differences allow for precise control of gene expression (Gutierrez et al., 1999). Cis- regulatory elements have been identified that act as either instability or stability determinants in mammalian systems (Chen and Shyu, 1995; Holcik and Liebhaber, 1997). Several sequences that target transcripts for rapid turnover in plants have also been identified. These include the DST element (Newman et al., 1993), the AUUUA repeat (Ohme-Takagi et al., 1993), and premature stop-codons (van Hoof and Green, 1996). It is critical to understand the mechanisms by which the cell recognizes these sequence elements and targets the transcripts that contain them for rapid mRNA degradation. Most studies aimed at understanding the cellular factors in mRNA decay pathways have involved characterization of proteins that bind the instability sequences in vitro. Isolation of sequence-specific RNA-binding proteins from plant cells has not been as successful probably due to the difficulty associated with preparing cytoplasmic protein extracts that are free of non-specific ribonucleases. In order to gain insights into the molecular basis of sequence-specific recognition and degradation of unstable mRNAs, a genetic approach was devised to isolate Arabidopsis mutants defective in DST-mediated mRNA degradation (Johnson et al., 2000). Such a strategy has several distinct advantages. First, a gene that is identified by a mutant phenotype is, by definition, affecting the process in vivo. Second, basic information about the mechanisms of mRN A degradation may be obtained by studying mutants. Third, genetic analysis is not limited by pre-conceived mechanistic ideas. For 134 example, the cellular factor that recognizes an RNA degradation sequence may itself be an RNA molecule, not a protein. Finally, experimental complications that have limited the success of biochemical approaches, such as unstable or low abundance proteins or the presence of non-specific RNA degrading activities in protein extracts, are eliminated. The mutagenesis strategy involved the generation of transgenic plants expressing HPH (hygromycin phosphotransferase) and GUS (B-glucuronidase) reporter genes (line 1519-31). The transcripts from both genes were destabilized by insertion of a tetramer of the DST instability determinant into their 3’ UTRS (Johnson et al., 2000). Mutants in the mRNA decay pathway mediated by the DST element are expected to have increased HPH and GUS mRNA abundance and therefore, it should be possible to isolate them on the basis of these increased expression levels. dst1 and dst2 were isolated based on their ability to stabilize specific DST-containing transgene mRNAs, indicating that they harbor mutations that diminish the function of the DST-mediated decay pathway. Genetic analysis has demonstrated that dst] and dst2 are semi-dominant mutations in independent single genes (Johnson et al., 2000). In this chapter, the isolation and characterization of a third gene, dst3, in the DST-mediated mRN A decay pathway is described. RESULTS dst3 exhibits increased HPH-DST and GUS-DST mRNA levels A third putative mutant was isolated during the selection of 794,000 mutagenized M2 seeds (Johnson et al., 2000). The abundance of the HPH-DST mRNA for dst3 is shown in Figure 5.1. The level of this transcript is approximately 3.5 fold higher relative to p1519-31 and this increase in mRNA abundance is similar to that observed in dst] and 135 l 5 1 9 dst3 l 493 dst] dst2 HPH-DST ' - Q ~ ~ eIF4A inn-w»... .w . E“ E” P UINUuw'J-Au: All... HPH/eIF4A/av par _ l 0.5“ l l 151 9-3 1 dst3 dst] dst2 Plant Line Figure 5.1 HPH-DST mRNA levels are elevated in dst3. A) HPH-DST mRNA abundance was measured in the rosette leaves of dst3, p1519-31, the non-destabilized control 1493 and in previously isolated mutants dst1 and dst2. Each lane contained 10 pg of total RNA. The abundance of HPH-DS T mRNA was normalized to that of the loading control, eIF 4A. B) Quantitation of the increase in HPH-DS T mRNA levels in the dst mutants. 136 dst2 (Figure 5.1B; Johnson et al., 2001). Increased HPH-DST mRNA abundance has been observed consistently in the progeny of two backcrosses of dst3 to 1519-31. An analogous increase in GUS-DST mRNA levels was also observed, although the elevation of this transcript is higher in dst3 than in dst] and dst2 (Figure 5.2). Two transcripts have always been detected for GUS in the parental line and the dst mutants. The higher molecular weight transcript is probably due to the recognition of a cryptic downstream polyadenylation signal. Both transcripts are regulated in a similar manner in the dst mutants, although for quantitation purposes, the abundance of the lower molecular weight transcript (the expected size) was measured. Since both the HPH-DST and GUS-DST transcripts were elevated in the mutant, 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 reporter transcripts. To examine this possibility, genomic DNA was isolated from the mutant plants and the DST tetramer was amplified and sequenced. The DST tetramer was found to be unaltered for both the genes. Genetic analysis of dst3 For the genetic analysis of dst3, the inheritance of HPH-DST mRNA abundance was examined. Progeny of the first backcross to WT (1519-31) (F1) and the progeny of self-fertilizations of these plants (F2) were used in these studies. Ten Fl plants were examined. HPH-DST mRNA abundance was on average 1.9 i 0.53 fold higher in dst3 F1 compared to WT (1519-31). These levels of HPH-DST mRNA abundance are intermediate between WT (1519-31) and mutant levels as would be expected if heterozygotes showed a semi-dominant phenotype. Segregation of increased HPH-DST 137 3 a a '12 fl *8 48 -§ 1.33: GUS-DST ti r-u I e! F 4A Figure 5.2 GUS-DST mRNA levels are elevated in dst3. GUS-DS T mRN A abundance was measured in the rosette leaves of dst3, p1519-31, and in previously isolated mutants dst] and dst2. Each lane contained 10 pg of total RNA. The abundance of GUS-DST mRNA was normalized to that of the loading control, eIF 4A. 138 mRNA abundance in the F2 populations was also consistent with semi-dominance. Of 80 F2 plants segregating dst3, 20 were observed with high HPH-DST mRNA abundance (3.2 i 0.59 fold higher than WT[1519-31]), 38 had intermediate HPH-DS T mRNA abundance (1.7 :t 0.22) and 22 showed HPH-DST mRNA abundance that was similar to WT (1519- 31) (0.97 i 0.24) (Figure 5.3). This pattern of segregation in the dst3 F2 populations is most easily explained by a ratio of 1:2:1 as would be expected for the inheritance of a semi-dominant single gene (Table 5.1). dst3 is not allelic to dst1 or dst2 To determine whether dst3 was allelic with dst1 or dst2, a complementation test was performed. The data from these experiments are summarized in Table 5.2. Like dst] and dst2, dst3 is partially dominant, a characteristic of these mutations that complicates standard complementation tests. The F 1 progeny of a cross between dst3 and dst1 or dst2 showed intermediate HPH-DST mRNA abundance levels implying the lack of an additive interaction (Table 5.2). The results of DST3-/- crosses to DST 1-/- or DST2-/- were similar to those of the cross to p1519-31 (DST +/+), indicating that dst], d512, and dst3 are likely to affect three distinct genes. If dst] or dst2 had been allelic with dst3, these crosses would have resulted in F1 plants with high levels of HPH—DST mRNA abundance. In addition some F2 plants with HPH-DST mRNA levels similar to 1519-31 were recovered (Table 5.2) which would not be possible if the dst mutations were in the same gene or tightly linked. Increased stability of HPH-DST mRNA in dst3 compared with 1519-31 139 1519-31 1519-31 no N m N at so m c r co r~ — 3 an m In at at: at 44: :4: =3: =34: HPH-0S T ” ” «a - eIF 4A Figure 5.3. mRNA abundance of HPH-DST in F2 plants from the first back-cross of dst3 relative to WT (1519-31). 8 representative plants are shown. Total RNA was prepared from rosette leaves and 10pg was analyzed by Northern blot hybridization. Radiolabeled probes, prepared against the indicated transcripts, were used in sequential hybridizations of the same blot. The numbers indicate individual plants. #56 is an example of an F2 plant with high HPH-DST mRNA levels, #82 with intermediate and #12 with HPH-DST mRNA levels similar to WT (1519-31). 140 Table 5.1. Segregation of increased HPH mRNA abundance in the progeny of crosses between dst3 and 1519-31 (DST3). Cross Class HPH-DST‘ n )8, P’ DST3/DST3 X dSl3/dst3 F1 Int 1.9 i 0.53 10 DST3/dst3 X DST3/dst3 F2 WT 0.97 i 0.24 22 Int 1.70 :t 0.22 38 0. 3 P>0.8 High 3.20 i 0.59 20 Int. intermediate. WT, wild-type. * HPH-DST mRNA abundance: (HPH-DST/eif4A) segregating class/ (HPH-DST/eif4A) WT : standard error. f 12 calculated for 12221 segregation of WT/lnt/High HPH-DST mRNA abundance osr/osr, wr (1519-31). 141 Table 5.2. Segregation of increased HPH mRNA abundance in the progeny of crosses between dst3 and dst1 and dst2. Cross Class HPH-DST' n dst3/dst3 x dst1/dst1 F1 lnt 1.6 _+_ 0.1 15 F2 WT 0.9 i 0.2 16 Int 1.6 :1: 0.3 30 High 3.7 i 0.9 21 dst3/dst3 x dstZ/dst2 F1 Int 2.5 i: 0.2 16 F2 WT 0.9 i 0.2 9 Int 1.8 i 0.4 31 High 3.1 i 0.6 21 lnt, intermediate. WT, wild type. * HPH-DST mRNA abundance: (HPH-DST/eif4A) segregating class/ (HPH- DST/eif4A) WT 3: standard error. 142 The elevation in HPH-DST transcript levels in dst3 is probably due to the DST- mediated decay pathway being deficient in the mutant. To test this hypothesis, mRN A decay rates for HPH-DS T mRNA were measured in dst3 and WT plants. A change in mRNA stability would indicate that the mRNA is a primary target of the decay pathway. Half-life analysis was carried out in rosette leaves using cordycepin as the transcriptional inhibitor. The HPH mRNA abundance was standardized relative to the abundance of eif4A transcript. As is evident in Figure 5.4A and B, the decay of HPH-DST mRNA was slower in dst3 compared to the parental and the differences in steady state mRNA levels became even more prominent at later time points. The half-life of HPH mRN A was calculated to be 1.5-fold greater in dst3 than in 1519-31 suggesting that the dst3 mutation results in an increase in mRNA stability that is responsible for the increased abundance of the HPH-DST transcript. It has been noted before that differences in half-life measurements between the parental and the dst plants are reflective of about 17% of the differences based on steady state mRNA levels (Johnson et al., 2000), and a coordinate dampening in the mRNA half-life value for dst3 relative to the parental line was observed. DISCUSSION dst3 was identified using a targeted genetics strategy and exhibited elevated levels of DST-containing mRN As. mRN A decay kinetics showed that the increased HPH-DS T mRNA abundance is caused by a corresponding increase in message stability. The three mutants isolated by this strategy were extremely rare suggesting that the genes involved in the DST-mediated decay pathway are essential. Further, all three dst mutants are weak 143 0 15 30 60 90 120 0 15 30 60 90120 HPH-DST l! o - t! ,, g :2 . eIF4A “" "' “' w 1519 dst3 B. 1 4.x co = 2?. a E 2 E o E o 1519 (23 min) ° 9 dst3 (35 min) 0.01 l t l l o 30 60 90 120 time (min) Figure 5.4 Analysis of HPH-DST mRNA stability in leaves from 1519 and dst3 plants. A) Representative northern blot analysis of cordycepin time courses over a period of 120 minutes. Samples consisted of 10 pg of total RNA isolated from the indicated time points. B) Quantitation of the decrease in mRNA abundance and half-life estimation for HPH- DST. The stable eIF 4A transcript was used as a reference for equal loading. 144 alleles allowing partial functioning of the DST-mediated decay pathway since they did not restore HPH-DST mRNA levels to those found in plants with non-destabilized mRNAs (Figure 5.1). These findings may indicate that stronger alleles were lethal or that there is redundancy in the DST pathway. If so, the genes may be members of the same gene family. dst] and dst2 are partially dominant mutations (Johnson et al., 2000), and interestingly, so is dst3, the new mutation described in this study. A semi-dominant mutant phenotype could result if the defective DST—recognition factors no longer ! recognized the DST sequence, but still bound the interacting proteins required for mRNA ‘ decay. On the other hand, a semi-dominant phenotype could arise if a defective DST- binding protein retained the ability to bind a DST subdomain but had reduced ability to interact with or recruit proteins required for mRNA degradation. Another alternative would be that a dst gene product encodes a DST-specific ribonuclease that binds the DST sequence but has decreased RNase activity, thereby providing partial protection to mRNAs bound by a mutant version. Studies conducted with the previously identified dst mutants, dst1 and dst2, have indicated that both genes have distinct and overlapping roles in the DST-mediated mRNA decay pathway (see Chapter 2). It will be of interest to examine how dst3 cross talks with dst] and dst2. In addition, the continuing analysis of all possible mutant combinations should conclusively define the roles of each gene as they work alone or in combination. Prior to this work, only two mutants involved in DST-mediated mRNA degradation were known. Considering the rarity of the mutants isolated, the identification of an additional mutant, dst3, is important. In theory it should be possible to generate 145 more mutants in this pathway since the screen was not saturating, in spite of being very tedious. Finally, the cloning of dst3 should lead to a better understanding of the DST- mediated decay pathway in plants. MATERIALS AND METHODS Plant Material dst3 mutant plants described in this report are from the accession Columbia (glI) derived from EMS mutagenized Arabidopsis populations (Johnson et al., 2000). 1519, dst], dst2 and dst3 plants were grown in growth chambers under 16 hr light and 60% relative humidity at 20°C. Rosette leaves were harvested from 35- to 40-day-old plants. All tissue was harvested from plants grown in parallel under the same conditions in different growth chambers. Half-life measurements, total RNA Extraction and RNA Blot Hybridization Half-lives were determined as described by Seeley et al. (1992). Rosette leaves from Arabidopsis plants were transferred to a flask with incubation buffer. After a 30 min incubation, 3'-deoxyadenosine (cordycepin) was added to a final concentration of 0.6 mM (time 0). Tissue samples were harvested at regular intervals thereafier and frozen in liquid nitrogen. Total RNA from leaf samples was extracted as previously described (Newman et al., 1993). 10 pg of total RNA or was analyzed by electrophoresis on 2% formaldehyde/1.2% agarose gels and blotted onto nylon membrane (Nytran Plus, Schleicher & Schuell, Keene, NH). DNA probes were labeled with [a-32P]dCTP by the random primer method (F einberg and Vogelstein, 1983) and purified from 146 unincorporated nucleotides using probe purification columns (NucTrap, Stratagene, La Jolla, CA). The RNA blots were hybridized as described in Taylor and Green (1991) using the indicated 32P-labeled probes. For a loading control, RNA blots were hybridized with a 32P-labeled cDNA probe for the Arabidopsis translation initiation factor eIF 4A (Taylor et al., 1993). Analysis of the HPH-DSTx4 sequence element in dst3 Genomic DNA was prepared from dst3 F2 plants from the first backcross to p1519-31 as described in Saghai-Maroof et al.(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 proof-reading polymerase (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. 147 REFERENCES Chen, C. Y., and Shyu, A. B. (1995). AU-rich elements: characterization and importance in mRNA degradation. Trends Biochem Sci 20, 465-470. Feinberg, A. P., and Vogelstein, B. (1983). A technique for radiolabeling DNA restriction endonuclease fragments to high specific activity. Anal Biochem 132, 6-13. Gutierrez, 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. 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 U S A 94, 2410-2414. Johnson, M. A., Perez-Amador, M. A., Lidder, P., and Green, P. J. (2000). Mutants of Arabidopsis defective in a sequence-specific mRNA degradation pathway. Proc Natl Acad Sci U S A 97, 13991-13996. Newman, T. C., Ohme-Takagi, M., Taylor, C. B., and Green, P. J. (1993). DST sequences, highly conserved among plant SAUR 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 U S A 90,11811-11815. Saghai-Maroof, M. A., Soliman, K. M., Jorgensen, R. A., and Allard, R. W. (1984). Ribosomal DNA spacer-length polymorphisms in barley: mendelian inheritance, chromosomal location, and population dynamics. Proc Natl Acad Sci U S A 81, 8014- 8018. Seeley, K. A., Byrne, D. H., and Colbert, J. T. (1992). Red Light-Independent Instability of Oat Phytochrome mRNA in Vivo. Plant Cell 4, 29-38. Taylor, C. B., and Green, P. J. (1991). Genes with homology to fungal and S-gene RNases are expressed in Arabidopsis thaliana. Plant Physiol 96, 980-984. 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. van Hoof, A. and Green, P.J. (1996). Premature nonsense codons decrease the stability of phtohemeagglutinin mRNA in a position-dependent manner. Plant J 10, 415-424. 148 CHAPTER 6 MICROARRAY ANALYSIS OF dst2 149 INTRODUCTION The expression level of thousands of genes can be monitored simultaneously using DNA microarrays. Therefore, studies using DNA microarray technology can be used to address mechanistic questions on a global basis and identify targets of various metabolic pathways. DNA microarray technology has been exploited to characterize the molecular phenotypes of mutants (Pérez-Amador et al., 2001; Goda et al, 2002; Scheible et al, 2003) and to investigate the intra- and interspecies variations in genome expression patterns (Enard et al., 2002). Genomic approaches have also been particularly useful to discriminate the functions of individual members of a gene family which may have partial redundant functions and would be difficult to dissect using classical genetic techniques (Wang et al., 2002). As pointed out in Chapter 2, microarray analysis of the dst1 mutant was critical in providing the first clue about the physiological significance of the DST-mediated decay pathway in Arabidopsis. In addition, some of the genes identified as being differentially regulated in dst] were subsequently examined by RNA gel blot analysis and found to be either coordinately or oppositely regulated in dst2. In this chapter, preliminary microarray studies were carried out with dst2 to expand our understanding of this mutant and discover new molecular markers specific for dst2. RESULTS 15k element microarray reveals genes with altered gene expression in dst2 15 k slides, prepared by the AF GC microarray facility at MSU, were utilized to examine global gene expression changes in the dst2 mutant. Poly(A)+-selected RNA from 150 the rosette leaves of five-week old dst2 and parental plants was labeled with the fluorescent dyes and used for hybridization. Two biological replicate experiments were performed, each with a reverse labeling technical replicate. After normalization of the data, 33 ESTs and multiple clones for HPH and GUS showed a difference of 1.5-fold or greater in all four slides. Of the 33 ESTS, 29 clones corresponded to 22 annotated Arabidopsis genes while 4 ESTS did not correspond to any annotated open reading frames. Multiple ESTs corresponding to the same gene that behaved in a similar fashion further corroborated the microarray data. Twenty four genes, including the transgenes HPH and GUS, showed elevated mRN A levels while four displayed decreased levels of mRNA in the dst2 mutant compared with the parental plants, as listed in Tables 6.1 and 6.2, respectively. After correction for redundancy in the ESTS and excluding the transgenes HPH and GUS, a total of 26 endogenous genes whose expression levels were altered in dst2 were identified. The probable biological functions of these genes, listed in Tables 6.1 and 6.2, were based on the latest annotation by the TIGR Arabidopsis thaliana database. Identification of the putative primary targets of the dst2 mutation As a first step towards determining the primary targets of dst2, the 3’UTRs of the genes that were differentially regulated were analyzed for the presence of DST-like subdomains. 11 genes were identified that contained possible DST-like sequences in their 3’ UTR (listed in Table 6.3). Direct targets of the decay pathway mediated by DST 2 are expected to be stabilized in the dst2 mutant. It will therefore be of interest to measure mRN A decay rates for these transcripts and determine if the changes in expression levels 151 2.5.30 S a f ENEE N 2232 3:53 £583 muses .28.. uses Be 222 as a S quo _ ONNRw _2 Essa 3336 no a M: £38 _ 2832 £93 saea 8265 82:68 Nd a 3 £28: _ 853.2 N base nannafieaeneo 3 a M: 2.2sz 2 238232 a 30% 2 dose Sentences 92-323an N5 4 M: £2 0N2 _ omega _2 Angela 2&3 dose 8:285 No a 3 SEE _ Rama _2 6.522% .e decades“ e2 32: .552 3 a 3 832% _ 8222 been 522; awe: can 25-28 no a 3 ENE N _ 83.32 35% m 55.8 messes :55 No a 3 2528 _ 2 _ <2 .20 Bees; 2 S a 3 2583.2 _ 22 b5 Bees.“ oz £59. 3 a 3 £2 2N2 N 80322 seen ennneaxm 3 « oN _ saddened? 22o a Beam 2 a 3“ <2 e <2 5: N._ a 2. 22 m 22 So am 9 e232 9 2m: Ema 88: .e 22.5.2 .352 9:: 25c mam—m Egan 3%? N33 E 20>“: <72:— 39885 NE? 3:00 —.e 035. 152 S a f Erma—NM: _ 22 "20 3585 oz 3 a 3 252022 2 <2 "20 eases; ez Na 4 3 NSNoeNN _ ©3822 oar 2983 _ éee 8%. 866828 ENNONS No a 3 N: NmoNN N oNNmmm 2 N_ eased 05083235: e .225 Gm u” owuho>< A: Dm—Z r—imnm 8:30 ac LuDEBZ hon—552 mam—E 0:00 mace—m :35me mam—9» NBS E m_o>o_ SAME 3828c 5:5 3:60 «6 «Sun. 153 Table 6.3. Genes with possible DST-like sequences in their 3’ UTRs Gene 3’ UTR NRAMP metal ion attgaagaattagaacatttcaggtagaagaaatagtgtttacttattgtgatctattgcatagaataa transporter 6, aaatgaagttgtttacttaagtgtaatctattgcatagatcaaacatgaagacttgactacaaaatcc (N RAMP6) atcaaacggtttggccaaagaactgagtgtacagattcataaccagacgagacatattgaaaca gaggtttcacataaccagagacaacaaacaaaagcaacttcaatggttaaggatacttaagggg gatacttgagggaattaataagcttttcaagaccc Expressed protein gaaaaacaaaaggtggtgcaagtgcaaacgaaggtgaaacaattctgaagatgccttttttgtg gaagtctcatcttcgtcacatctaacaaagaagttgtagatttatgggatgatgcttactctgctgg atttaaacacaggtaacaacagatgatgctactaactagttttcctctacctttacacaaactttattg agaagttagatttg atctcgtaatgacagtttcttacacaaaaataaaataaaatacaaatttgtcat gactaatatcttatgaaacatcttctcatatattgtgtttgttct Zn finger (C3HC4- type RING finger) protein family tagttgacttttcttttctcatctttttcatttgtttttgtttgggattgttgataaatacgcaaatacattttc —> aattttttaattgaattatccttgttttgtttaaaattctctgtaacgtacatatgaagtcagatcgaaaa cggattcacttttgaaaactaatg ataaaaaagcttataacaaaacaacggacaatgatgtgag gtgcgctgtggtgcaggtgggtttctcggtttttcaatattaaaaaagtcgtattcgtggacaaaga atatagctatccgatcccctgataaaatatatgctaatagaatata Similar to DNA topoisomerase IV subunit A tcacttaattgaactaatgagattcttgattgtggttaaagcacatgttcaattagttgtggttcttgtg ttttattttcttgtgttgttttgtttgagactttgtttgttgcttatatcaggggagagaggttgaaaaac ctttgagagafitagmgagagatttgtgaaacacaaataatgttttagtcgtcttataataaa acctttttttttcttatatatgctcattgataatactgcaaaaaggtaaaatcatgtttttttttttttaatat gaatcatgatattcagg Zn finger (C3HC4- type RING finger) protein family tgacttgtcacgtgttggtgtctgattggtttaatgttaaccgggagtaaaaaaaggaattactaca agtcaacaggcttttgtctaggtgttgatttcggcgcccaaggacacgtggcgaaactgagctt ccaggaatcaatattcaccgtctattatgattagataggttagatagafifgfgfiacgatgtacaaa gtcatctacaatattgaatctatttccatttattttaccatattctttttttttttataattttcgaagttctac aaactcttttatgtaaaacacaatccaatggtcataattgtgataaagactttgcataatt SKPl interacting partner 5 (SKIPS) tgatttttgctggttttatttatttttcttcttctgtctgt'aacattagfaatctgtgtgatgatgggaaatg atccatcttcatttgtactttgtgtttccccttaagaccagacaggtcctttgttgtatgcttattatgttc agaaataatgtcctttttaggacgaattgccgaattcatttcaattggatttttataatttctgcttagtg tatcttttgtgataaatggtttctttcttcactgttatcaaaatatt Scarecrow-like transcription factor 13 (SCL13) tgatgatggctgggttcacgggttggccggtcagcacatctgcagcgtttgcagcgagtgagat gctgaaagcttatgacaaaaactacaaactgggaggccatgaaggagcgctctacctgttctgg aagagacgacccatggctacatgttccgtgtggaagccaaacccaaafifiéttgggtaagtta tagtgatgatggttacttgagtggataaaFaagagcacaacaaaaacacatctgtcgctgtaaat tttttaggatgtgcaatgatgttttaagtt aacacaacctaagttaWtacaaaccaaacc tggtggttgtttttctcttg aaattgtcatgtggttgtgggtgggttgctagtaatgaaatataacca aaacattgattaggtca Cadmium induced protein (A88) taatgctacccaccacgaggatcttaagcatctgagacgcatttcgagatagttactttgatggaa tttgtgaatgtgactgaaatattcagcacaaatcagaacatgg1tc_ctgtcttgtttggagaagcac tatgaaccaaacctgaagcccttgtgatttagagttacagagatacattaaaagggttcgggattt cacagtagcattcatagacagttcaggttagttaaacgtgcttcacaaaaagaaaaaaaacaga agtgtagttgaggtcacaccacaggtttatgctttgctcttaatcaaatatttatctttctcttttttgtta tttgtcctttttaattccatttgctttattccttacttgtggggattaaatgtttgcagagaaagataattc gactgaataaatacacttttacctgt 154 Identical to Cytochrome P450 tgatgctmtcattaggacgtttctgctgggufifatggcgtgaccaatggttatttttcattg caatatccctttttgttttaatgagtactatgttctcattttaacgaataaaaatg atcagtgctcttgtt tttggactagaaaagaaagtagtccgatgtttaatattcgggtccctttaatattccctctggtttaca atattttaagctatcttagtaaagatctat O-methyl- transferase family 2 taagaggaggacatgaaaga'tatatctttcctttgaggaaaactcaataaatta tgattgttgattt ggtgtttttatacagaacgacgtgt gtgaataattgtgttgtgtgataataag aaggcttgtttc agtatcagcagagttgttttttatttg atttttcacatttttttatcattaatttccttgtgagtggcacct ctaatttacttttttaattgttataggctatataagatagataaccttatacaatgtcaaatcttacaatt acattatgaacatgaa RNA and export factor binding protein, putative tagaaa gggagattaaacctattgcatgthccttcctttagtgccgatatgcttttgt attagtatgatctttagattgaatgtcaatgggtctgacattttcgagctttagcttttgttttttcttctct tggtgagcttcttagctttttgcagttgagttcagagaaacaagcattgtgtaaccgtggaactca aaactcttctaaaatatcaattcaaaacacaattctatctactctctta ATAGAT-like subdomain —’ GTA-like subdomain 155 detected on the microarray slides are indeed due to a corresponding change in mRNA stability. RNA gel blot analysis of previously identified transcripts Prior microarray studies with dst1 followed by Northern blot confirmation demonstrated that some transcripts were affected in both the dst1 and dst2 mutants (see Chapter 2). Intriguingly, only one of these genes showed a greater than 1.5-fold difference on all four slides (chalcone synthase) although some of them could be detected on two out of the four slides. Also some of the ESTS for the relevant genes on the 15k arrays were different from the previously used 11k arrays (see Discussion). In order to validate the quality of the RNA used for labeling, a few of the formerly characterized transcripts were checked by RNA gel blot analysis. As illustrated In Figure 6.1, HPH mRNA levels were elevated in both dst1 and dst2, RAP2.4 mRNA levels were increased in dst1 but unchanged in dst2 and mRNA levels for a putative patatin gene were diminished in dst1 but elevated in dst2. All the transcripts tested increased or decreased in abundance as expected (see Chapter 2). Experiments to confirm additional transcripts are in progress. DISCUSSION Microarray analysis identified 26 genes as being differentially regulated in the dst2 mutant. The genes identified code for proteins involved in variety of metabolic processes. Four ESTS corresponding to a protein similar to DNA topoisomerase IV subunit A showed the highest mRNA elevation in dst2. Topoisomerase IV is a type II 156 a :z a :2 a s HPH-DST 4 no I! - eIF4A B. 5’3 2 9:. :2 s e RAP2. 4 - eIF4A C' 3 a a 2 s s 111G9T7 - eIF4A . ..... ..... Figure 6.1. RNA gel blot analysis of previously identified transcripts. Lanes contained 10 ug of total RNA extracted from WT (1519), dst], and dst2 plants. Each blot was hybridized sequentially with 32P-labeled eIF 4A, and with A) HPH, B) RAP2.4, and C) 111G9T7, EST corresponding to putative patatin protein. 157 isomerase involved in the topological changes of DNA during replication (Zechiedrich and Cozzarelli, 1995). Four transcription factors were identified, three of which seem to contain DST-like sequences in their 3 ’UTR. If these are the primary targets of DST 2, it could be hypothesized that changes in the mRNA levels for these transcripts would lead to downstream secondary effects. Interestingly, one of the transcription factors was a scarecrow-like protein (SCL13) and it was recently shown that a certain species of micro- RNAs (miR171) is responsible for the cleavage of several members of scarecrow-like mRN As in Arabidopsis (Llave et al., 2002). Micro-RNAS interact with perfect complementarity guiding cleavage of the target mRNA (Hammond et al., 2001) and SCL13 seems to be a direct target of dst2, suggesting the possibility of miRNAs being involved in DST-mediated decay. The small number of genes reported in this chapter is most likely a conservative estimate of the actual targets of dst2. Although the high stringency used in the data analysis possibly reduced the number of false positives, some bonafide targets of the dst2 mutant could have been missed. This is especially relevant because two different printings of the 15k microarray slides were used for this study. Even though most of the cDNAs present on the two separately printed slides corresponded to the same genes, different ESTs were used. ESTS corresponding to the same gene can exhibit variable expression ratios, depending on the probe or target length, genetic redundancy, or a combination thereof. It is perhaps pertinent to note that on comparing the technical replicates for each set of printings, many more differentially regulated mRNAs could be detected, some with much greater fold changes and multiple ESTS showing similar changes in expression 158 (data not shown). RNA gel blot analysis to confirm some of these targets, in addition to the ones shown in Tables 6.1 and 6.2 should be useful in ascertaining molecular markers specific for dst2. Once a robust marker is found for dst2, that is unaffected in dst1 and dst3, it could be used to facilitate the mapping of dst2. dstZ- specific markers should also aid in testing double mutants as well as provide unique clues about the possible genetic interactions (epistasis relationships) among the multiple loci involved in the DST- mediated degradation pathway. Finally, additional microarray experiments should be informative about the impact of dst2 at the whole plant level relative to the other dst mutants. MATERIALS AND METHODS Plant Material All Arabidopsis thaliana plants described in this report are from the accession Columbia, grown in growth chambers under 16 hr light and 60% relative humidity at 20°C. Tissue from the parental line (p1519-31) and dst1 and dst2 homozygous mutants (Johnson et al., 2000) were harvested from 35- to 40-day-old plants. The dst1 and dst2 lines used were from the second backcross to the parental line. All tissue was harvested from plants grown in parallel under the same conditions in different growth chambers. 15k AFGC DNA Microarray 15,532 and 15,488 element microarrays were generated at the Arabidopsis Functional Genomics Consortium (AF GC) Microarray facility at Michigan State University. The ESTS were spotted on super-amine glass slides (Telechem International, 159 Inc.; Sunnyvale, CA). Slides were washed and blocked according to the Telechem protocol. Total RNA Extraction, Poly(A)+ RNA Purification, and RNA Blot Hybridization Total RNA from leaf samples was extracted as previously described (Newman et al., 1993). Poly(A)+ RNA was purified from 100 ug of total RNA using the Oligotex mRNA kit (Qiagen, Valencia, CA). RNA (10 pg of total RNA or 2 pg of poly(A)+ RNA) was analyzed by electrophoresis on 2% formaldehyde/1.2% agarose gels and blotted onto nylon membrane (Nytran Plus, Schleicher & Schuell, Keene, NH). DNA probes were labeled with [or-32P]dCTP by the random primer method (Feinberg and Vogelstein, 1983) and purified from unincorporated nucleotides using probe purification columns (NucTrap, Stratagene, La Jolla, CA). The RNA blots were hybridized as described in Taylor and Green (1991) using the indicated 32P-labeled probes. For a loading control, RNA blots were hybridized with a 32P-labeled cDNA probe for the Arabidopsis translation initiation factor eIF 4A (Taylor et al., 1993). Blots were stripped between hybridizations in O. 1% SDS at 90 to 95°C for 1 hour. Quantification of hybridization signals was achieved using a phosphorimager (Molecular Dynamics, Sunnyvale, CA). Labeling of Poly(A)+ RNA Poly(A)+ RNA was labeled using the aminoallyl labeling procedure. 0.5 ug of poly(A)+ RNA and 6 ug random hexamers in a total volume of 18.5 ul of DEPC-treated water was denatured at 70°C for 10 min, set at room temperature for 3 minutes and cooled down on ice. On ice, 6 pl. of5 x RT buffer, 3 uL of 0.1 M DTT, 0.6 uL 50X aminoallyl-dNTP mix, and 2 uL of Superscript II (ZOOU/uL) were added, and the mixture 160 was incubated at 42°C for 3 hrs. The RNA was hydrolyzed by adding 10 uL 1N NaOH and 10 uL 0.5M EDTA and incubating at 65°C for 15 minutes. 10 uL 1N HCL was then added to neutralize the pH. Unincorporated aa-dUTP and free amines were removed with the Qiagen PCR purification kit using phosphate wash and elution buffers. Aminoallyl- labeled cDNA was resuspended in 9 uL 0.1M NazCO3, mixed with the appropriate NH S- ester Cy dye and incubated in the dark for 1 hr. The labeled cDNA was purified using the Qiagen PCR purification kit. To test the quality and quantity of the product, 2 uL of the labeling reaction with 2 uL glycerol was separated on a 1% agarose gel using a miniprotean gel electrophoresis system (Bio-Rad Laboratories, Hercules, CA). The portion of the gel containing the DNA sample was placed on a glass microscope slide, dried on a heat block at 70°C, and scanned using Affymetrix 428TM scanner. The rest of the probe was used to prepare the hybridization mixture. DNA Microarray Hybridization and Analysis For a single DNA microarray, 50 uL of hybridization solution was prepared by mixing 50 uL slide hybridization buffer (Ambion) and 4 uL Cy5- and Cy3-labeled probe. The mixture was denatured at 95°C for 10 minutes. The mixture was then hybridized to the array under a glass coverslip that had been washed in 95% ethanol, then 0.2% SDS, and rinsed in distilled water. The slide was then placed in a microarray hybridization chamber (Arraylt Hybridization Cassette, TeleChem International, Inc.; Sunnyvale, CA) with 200 uL of 3 x SSC to ensure high humidity conditions. Hybridization was carried out in a water bath at 65°C for 12 to 20 hours. After hybridization, the microarray was washed for 5 min in 1 x SSC/0.2% SDS, 5 min in 0.1 x SSC, and 15 sec in 0.05 x SSC without SDS, and finally dried by centrifiJgation at 600 rpm for 5 min. The slide was 161 scanned once in an Affymetrix 428TM scanner for both channels 1 and 2 (corresponding to Cy3- and CyS-labeled probes, respectively). The image files obtained were analyzed using GenePix Pro 3.0 software. Data from each channel was transformed to the natural logarithm, and a Z-score was calculated to normalize the channel values in order to account for variation in RNA labeling. Values were retransformed from the natural logarithm by raising to the power e, and the channel ratio was calculated. 162 REFERENCES Enard, W., Khaitovich, P., Klose, J ., Zollner, S., Heissig, F., Giavalisco, P., Nieselt- Struwe, K., Muchmore, E., Varki, A., Ravid, R., et a1. (2002). Intra- and interspecific variation in primate gene expression patterns. Science 296, 340-343. 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. Goda, H., Shimada, Y., Asami, T., Fujioka, S., and Yoshida, S. (2002). Microarray analysis of brassinosteroid-regulated genes in Arabidopsis. Plant Physiol 130, 1319-1334. Hammond, S. M., Boettcher, S., Caudy, A. A., Kobayashi, R., and Hannon, G. J. (2001). Argonaute2, a link between genetic and biochemical analyses of RNAi. Science 293, 1146-1150. Johnson, M. A., Perez-Amador, M. A., Lidder, P., and Green, P. J. (2000). Mutants of Arabidopsis defective in a sequence-specific mRN A degradation pathway. Proc Natl Acad Sci U S A 97, 13991-13996. Llave, C., Xie, Z., Kasschau, K. D., and Carrington, J. C. (2002). Cleavage of Scarecrow- like mRN A targets directed by a class of Arabidopsis miRNA. Science 297, 2053-2056. Newman, T. C., Ohme-Takagi, M., Taylor, C. B., and Green, P. J. (1993). DST sequences, highly conserved among plant SAUR genes, target reporter transcripts for rapid decay in tobacco. Plant Cell 5, 701-714. Perez-Amador, M. A., Lidder, P., Johnson, M. A., Landgraf, J ., Wisman, E., and Green, P. J. (2001). New molecular phenotypes in the dst mutants of ArabidOpsis revealed by DNA microarray analysis. Plant Cell 13, 2703-2717. Scheible, W. R., Fry, B., Kochevenko, A., Schindelasch, D., Zimmerli, L., Somerville, S., Loria, R., and Somerville, C. R. (2003). An Arabidopsis mutant resistant to thaxtomin A, a cellulose synthesis inhibitor from Streptomyces species. Plant Cell 15, 1781-1794. Taylor, C. B., and Green, P. J. (1991). Genes with homology to fungal and S-gene RNases are expressed in Arabidopsis thaliana. Plant Physiol 96, 980-984. Taylor, C. B., Bariola, P. A., DelCardayré, S. B., Raines, R. T., and Green, P. J. (1993). RNSZ: A senescence-associated RNase of Arabidopsis that diverged from the S-RNases before speciation. Proc Natl Acad Sci USA 90, 5118-5122. Wang, H., Ma, L., Habashi, J ., Li, J., Zhao, H., and Deng, X. W. (2002). Analysis of far- red light-regulated genome expression profiles of phytochrome A pathway mutants in Arabidopsis. Plant J 32, 723-733. 163 Zechiedrich, E. L., and Cozzarelli, N. R. (1995). Roles of topoisomerase IV and DNA gyrase in DNA unlinking during replication in Escherichia coli. Genes Dev 9, 2859-2869. 164 CHAPTER 7 FINAL REMARKS AND FUTURE PROSPECTS 165 An important challenge for plant biologists is to understand all of the mechanisms used by plant cells to regulate gene expression levels. Transcriptional regulation has justifiably enjoyed a great deal of experimental attention over the years. However, there are several important modes of regulation that occur once messenger RNA (mRNA) molecules are generated. The mechanisms responsible for the post-transcriptional regulation of gene expression are only beginning to be understood in plants. Several messenger RNA sequences that control abundance, localization, and translation initiation have been identified, yet the factors that recognize these sequences are largely unknown. The genetic studies coupled with the functional genomic approaches described in this thesis should lead to an enhanced understanding of sequence-specific decay in Arabidopsis and other higher plants. Cloning of the DST 1 gene should be accomplished in the near future and the characterization of its product should provide insight into the decay machinery that recognizes and degrades DST-containing transcripts. The simplest hypothesis is that DST 1 encodes an RNA-binding protein or a ribonuclease. It is also likely that DSTl is a regulatory factor that controls one of these components or has some unanticipated function. An intriguing possibility is that the DST element is recognized by specific complementary noncoding RNAs. If DSTl is a sequence-specific RNA-binding protein, it could be hypothesized to influence deadenylation by modulating the rate at which a single ribonuclease functions or by recruiting kinetically distinct ribonucleases. In addition, if DSTl is indeed a DST- specific RNA binding protein, it would be of interest to examine its interaction with specific subdomains of the DST element. Once the identity of DST 1 is known, knock-out 166 and overexpression lines can be generated to further elucidate the function of the DSTl protein. Microarray studies and subsequent half-life analyses have indicated that DST 1 is most likely involved in the degradation of mRN As in discrete pathways. Identification of the protein complexes that DSTl interacts with will be the next step in determining the biological role of the DST-mediated degradation pathway in Arabidopsis. Protein-protein interactions of DSTl could be examined in vitro with other Arabidopsis proteins predicted to be involved in mRNA decay, such as AtXRN4 (5’-3 exoribonuclease), AtPARN (poly(A) ribonuclease) and components of the exosome. Alternatively, the yeast two-hybrid system could be used to detect novel proteins that associate with DST1. Recent experiments have suggested that DST-mediated mRNA decay might be developmentally regulated. Early in the analysis of dst1 and dst2, it was noted that the mRNA turnover defect of these mutants, at least for HPH-DST mRNA, was more pronounced in older plants compared with seedlings (Johnson and Green, unpublished data). This result indicates that it may be important to assay the impact of mutations that affect posttranscriptional regulatory pathways throughout plant development even when it has been assumed that the element is not developmentally regulated. Future studies will be required to determine whether components of the DST-mediated mRNA decay pathway are under temporal regulation or are modulated in response to other stimuli such as circadian rhythms. Distinct DST-containing mRNAs could also be differentially regulated within the same cell and this regulation might be due to the different mRN A decay pathways mediated by the various DST genes. Likewise, it could be postulated that some of these 167 DST-containing mRNAs require a sequence in addition to the DST element and proper cooperation between multiple RNA elements is necessary for mRNA stabilization or destabilization. Furthermore, specific types or numbers of subdomains present in certain DST—containing mRNAs might be contributing to the modulation of DST-mediated decay. The progress outlined in this thesis opens up fresh avenues for evaluating the association of DST-mediated decay with the circadian clock. The circadian experiments carried out with the dst1 mutant should be applicable to dst2 and dst3 as well. Analysis of the effect of the different dst mutants on circadian gene expression should further aid in unraveling the mechanisms that underlie posttranscriptional control of CCG expression at the level of mRNA stability. The dst mutants have and will continue to serve as excellent tools to address how specific mRNAs are targeted for increased turnover. Although the functioning of the machinery is only partially known, it is clear that the DST-mediated decay pathway can regulate particular mRNAs in a rapid, coordinated fashion. Apart from basic research, detailed knowledge of this pathway should have a potential impact on applied research as well. The inhibition of rapid mRNA turnover through interaction with the DST complex might be a means of increasing the amount of a specific protein in plant cells, a highly desirable trait for commercial applications. 168 ‘ urtmiiiiiitfligti1311311