4_ 2a...» 131.7,. .. .. ‘ . .. gtrfi Lunar. . d. ‘ .. ......mw.vmmma~ y. . , . . n ,rrbuaflflhfir H. .045» ‘ , . , ‘ . 9.. , f . .. . .. V . . . . .uu... ‘ henna... 14%...“- .. £12): .1! tier-$33.; . . J. x q ., , . . ‘ 01‘=:IQM,‘JV¢ .n. , . V . . , K , ‘ . . It. It}? t :- .4..S!.hn. g . . , ....: ‘nwhhhshgg . . g a I {I .. "z... .§...u.. 3.1!.”— Ilflgs t). {3.5.3. . .5: Nwfinrx: a... . is $8-. :71» {Amati .. . .. .1. . . 3.51.3”. . A . . .. . . . u {3.15. vstg'D-:lo‘ J u»... wgilynil 13.23.55; . , its: a q :1..b.t..233. ti ‘\7‘3%\dlu’"ofi\d¥\‘rv r. . {Fist :1. .. rig-x: .Soafl \ .6 tvtz‘ .15.! I It}... 1’. , t. L... 2.. .. t 5.1.517: L11; 1251!! .9 I\. A hut. Ag‘i..l,l\1!.3 ( t. .Jatoaldnivll ,rv.. {lvsht ti 1%: T.).~I!’AJ 0,. ‘60 u. 5.. .«33m cmwmfiusadhuflfiafl mu: «at $ v‘ . I in no .2623 E.lr..?.s . A Li}, .2 a. , ‘Ju ..r..mw.4t . ‘ . . V Q4hm¥fi . A , . A .‘ :23! :t. LIBRARY Michigan State University C? b3 ‘,\\ C) This is to certify that the thesis entitled ROLE OF JAZ PROTEINS IN THE REGULATION OF JASMONATE SIGNALING IN ARABIDOPSIS presented by Hoo Sun Chung has been accepted towards fulfillment of the requirements for the PhD. degree in Biochemistry and Molecular Biology 44 {A / ‘ m M310} Professor’s Signature — " ‘1" ‘ Aug, 2‘43 7,000! Date MSU is an Affirmative Action/Equal Opportunity Employer u---------:--------------.—--------—--------.----n-o--.---a---o--------------------n--.-—-- PLACE IN RETURN BOX to remove this checkout from your record. . To AVOID FINES retum on or before date due. i MAY BE RECALLED with earlier due date if requested. . DATE DUE DATE DUE DATE DUE 5/08 K:IProi/Aoc&Pres/CIRCIDateDue.indd ROLE OF JAZ PROTEINS IN THE REGULATION OF JASMONAT E SIGNALING IN ARABIDOPSIS By [-100 Sun Chung A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Biochemistry and Molecular Biology 2009 ABSTRACT ROLE OF JAZ PROTEINS IN THE REGULATION OF JASMONATE SIGNALING IN ARABIDOPSIS By Hoo Sun Chung The plant hormone jasmonate (JA) regulates a wide range of growth, developmental, and defense-related processes. Recently, JASMONATE ZIM-domain (JAZ) proteins were identified as negative regulators of transcription factors that control the expression of JA- response genes. Upon perception of bioactive JA derivatives by the F-box protein CORONATINE INSENSITIVE 1 (C011), JAZ proteins are degraded via the SCFCOH/ubiquitin/26S proteasome pathway, thereby relieving the restraint on JA- response genes. A highly conserved “dergon” sequence, referred to as the Jas domain, mediates the JA-dependent interaction of JAZ proteins with C01]. The broad aim of this thesis research is to further understand the molecular mechanism by which J AZ proteins regulate the IA signal transduction pathway. To address this objective, the effect of wounding, insect herbivory, and cycloheximide on the expression level of Arabidopsis JAZ genes was determined. The results showed that most JAZ genes are rapidly induced by tissue damage, and that JAZs are primary response genes whose expression is activated upon turnover of labile repressor proteins. Transgenic plants that overexpress a truncated form of JAZl lacking the Jas domain were compromised in resistance to the generalist herbivore Spodoptera exigua, suggesting a role for JAZ proteins in the regulation of anti-insect defenses. It was also shown that the plant-specific ZIM domain mediates JAZ homo-and heteromeric interactions, and that these interactions are essential for the repressive function of JAZ. Molecular characterization of three JAZlO splice variants showed that alternative splicing events affecting the Jas domain alter the in vivo stability of JAZ proteins by modifying their ability to interact with SCFCO”. Comparative analysis of JAZ genes from diverse plant species identified an evolutionarily conserved intron whose retention during pre-mRNA splicing contributes to functional diversification of J AZ proteins through altered protein stability. ACKNOWLEDGEMENTS There are many people to whom I wish to convey thanks and admiration towards in their myriad of roles in shaping not only this dissertation but my academic career. First I would like to thank Dr. Gregg Howe for providing the opportunity to join his lab, for giving me a chance to work on the exciting project, for his inspiration and guidance, and for his strong support throughout my career at Michigan State University. I sincerely thank my committee members Dr. Christoph Benning, Dr. Sheng Yang He, Dr. Beronda Mongomery-Kaguri, and Dr. Rob Larkin for their support, their criticisms and for the opportunity to discuss issues throughout the course of this dissertation. I also like to thank Dr. Brad Day for his critical comments on this thesis. Thanks to all the present and past lab members who provided a stimulating environment that supported open discussions, troubleshooting, and constant feedback on projects throughout my Ph.D. Many thanks go out to Eliana for being an wonderful company in the Howe lab girls’ room and Leron for all the joyful moments in the lab and for initiating the JAZ study. Also thanks to Abe Koo, Lalita Patel, Yuki Yoshida, Christine Shyu, Marco Herde, Javier Moreno, Jinho Kang who have all helped in one way or another for my Ph.D study Besides the people that have contributed directly to my research, there are many people through the years that have helped me reach this point through their support and encouragement. Thank you to my family, to my mom and dad who constantly supported and loved me and who encouraged me through life to be my best. iv And now I would like to give my special thanks to my fiancee Woo Jun Sul who has been supporting me for seven years including the last five years in Michigan. I could not have accomplished this PhD. study without his continuous encouragement and love. TABLE OF CONTENTS LIST OF TABLES ................................................................................ viii LIST OF FIGURES ................................................................................ ix CHAPTER 1: LITERATURE REVIEW ...................................................... l l. BIOLOGICAL ROLE OF JASMONATE .................................................... 2 1.1. Defense against insect herbivores ...................................................... 2 1.2. Defense against necrotrophic pathogens ............................................... 4 1.3. Plant growth and development ........................................................ 5 2. BIOSYNTHESIS OF JA ....................................................................... 7 2.1. The octadecanoid pathway for JA synthesis ......................................... 7 2.2. JA metabolism ........................................................................... 8 2.3. Coronatine is a structural mimic of JA-Ile ......................................... 10 2.4. Regulation of IA biosynthesis ........................................................ lO 3. REGULATORY COMPONENTS OF THE JA SIGNALING PATHWAY .......... 12 3.1. The F-box protein, COIl ............................................................. 12 3.2. JAZ repressor proteins ................................................................. 13 3.2.1. Discovery of JAZ genes ..................................................... 14 3.2.2. The ZIM domain ........................................................... 15 3.2.3. The Jas domain ............................................................. 16 3.2.4. Functional diversification of JAZ proteins by alternative splicing. 19 3.3. (3R,7S)-JA-Ile is the active stereoisomer of JA-Ile ............................... 21 3.4. MYC2 is a JAZ-interacting transcription factor .................................. 22 CURRENT MODEL OF JA SIGNALING ................................................... 25 AIM OF THIS STUDY .......................................................................... 26 FIGURES AND TABLES .................................................................... 28 REFERENCES ..................................................................................... 36 CHAPTER 2: REGULATION AND FUNCTION OF ARABIDOPSIS JASMONA TE ZIM-DOMAIN GENES IN RESPONSE TO WOUNDING AND HERBIVORY ....................................................................................... 49 ABSTRACT ....................................................................................... 50 INTRODUCTION ................................................................................ 5 1 RESULTS .......................................................................................... 55 DISCUSSION ..................................................................................... 63 MATERIALS AND METHODS ................................................................ 72 FIGURES AND TABLES ..................................................................... 77 CHAPTER 3: A CRITICAL ROLE OF TIF Y MOTIF IN REPRESSION OF JASMONATE SIGNALING BY A STABLE SPLICE VARIANT OF JAZ10 ....... 98 ABSTRACT ....................................................................................... 99 INTRODUCTION ............................................................................... 1 00 vi RESULTS ........................................................................................ 104 DISCUSSION .................................................................................... 1 15 MATERIALS AND METHODS ............................................................. 121 FIGURES AND TABLES ..................................................................... 127 REFERENCE .................................................................................... 148 CHAPTER 4: ROLE OF THE JAS INTRON IN FUNCTIONAL DIVERSIFICATION OF JAZ PROTEINS ................................................. 154 ABSTRACT ..................................................................................... 155 INTRODUCTION .............................................................................. 1 56 RESULTS ........................................................................................ 160 DISCUSSION .................................................................................... 166 MATERIALS AND METHODS ............................................................. 172 FIGURES AND TABLES ...................................................................... 176 REFERENCES .............................................................................. 194 CHAPTER 5: SUMMARY AND FUTURE PERSPECTIVES ......................... 200 vii LIST OF TABLES Table 1.1. The TIFY family in Arabidopsis. .................................................. 35 Supplemental Table 2.1. JAZ genes are co-expressed with JA biosynthetic genes. ....87 Supplemental Table 2. 2. Forward (F) and reverse (R) primers and the corresponding sizes of the RT-PCR products used in this study. ............................................ 89 Table 3.1. Summary of JAZ-JAZ interactions in yeast. ................................... 139 Supplemental Table 3.1. List of constructs and primers used in this study. ............. 144 Table 4.1. Identification of JAZ genes and the conserved Jas intron in various plant species. ............................................................................................ 188 Supplemental Table 4.1. List of JAZ proteins identified in various species. ........... 189 Supplemental Table 4.2 List of primers used in this study. ............................ 193 viii LIST OF FIGURES Figure 1.1. The jasmonate biosynthetic pathway in Arabidopsis. .......................... 28 Figure 1.2. Chemical structure of JA-Ile diastereomers and coronatine. 30 Figure 1.3. The TIFY protein family in Arabidopsis. ....................................... 31 Figure 1.4. Sequence similarity between the Jas and CCT domains. ...................... 33 Figure 1.5. Current model of jasmonate signaling. ........................................... 34 Figure 2.1. Expression of JAZ genes in response to herbivore feeding and mechanical wounding ............................................................................................. 77 Figure 2.2. Effect of the coil-1 mutation on wound-induced expressionofJAZs. ....... 79 Figure 2.3. Rapid induction of JAZ transcripts and accumulation of JAs in response to mechanical wounding. ........................................................................... 80 Figure 2.4. Wound-induced expression of JA-responsive genes in the jarl-l mutant. ..82 Figure 2.5. Effect of cycloheximide treatment on JA-responsive genes. .................. 83 Figure 2.6. coil-1 plants are deficient in wound-induced accumulation of JA. ..........84 Figure 2.7 JAZIA3A plants are compromised in resistance to feeding by S. ex'igua. ...85 Supplemental Figure 2.1. Phylogenetic tree of Arabidopsis JAZ family. ................ 86 Figure 3.1. Homo- and heteromeric interaction of Arabidopsis JAZ proteins. ......... 127 Figure 3.2. Requirement for the TIFY motif in homo- and heteromeric interaction of J AZ proteins. .......................................................................................... 128 Figure 3.3. Identification of three JAZIO variants produced by alternative splicing. ..129 Figure 3.4. Protein-protein interaction characteristics of JAZIO isoforms. ............. 130 Figure 3.5. Subcellular localization of JAZIO splice variants. ............................. 132 Figure 3.6. Overexpression of J AZIO splice variants differentially affects JA responses. ....................................................................................................... 133 Figure 3.7. Stability of JAZIO splice variants in viva. ...................................... 134 Figure 3.8. The TIFY motif is required for repression of JA responses by JAZIO.4. ..135 ix Figure 3.9. The 1107A mutation does not affect the accumulation or localization of JAZIO.4. ........................................................................................... 137 Supplemental Figure 3.1. Immunoblot analysis of wild-type and mutant forms of JAZ3 and JAZIO.1. ..................................................................................... 140 Supplemental Figure 3.2. Yeast two-hybrid analysis of homo- and heteromeric interactions between Arabidopsis ZIM, ZIM-like] (ZMLl), and ZIM-like2 (ZML2) proteins. ........................................................................................... 141 Supplemental Figure 3.3. Detection of transcripts encoding JAZIO.3 and JAZIO.4 using transcript-specific primers. ..................................................................... 142 Supplemental Figure 3.4. Yeast two-hybrid analysis of the effect of 1107A and G11 1A TIFY mutations on the interaction of J AZlO.4 with other members of the JAZ family.143 Figure 4.1. Structural organization of Arabidopsis JAZ genes. ........................... 176 Figure 4.2. Retention of the Jas intron during pre-mRNA splicing is predicted to alter the Jas domain. ....................................................................................... 177 Figure 4.3. PCR-based detection of Jas intron retention in various JAZ transcripts. ..178 Figure 4.4. Differential ligand-dependent interaction of JAZ proteins with COIl. ..... 179 Figure 4.5. Overexpression of JAZIO genomic DNA attenuates JA signaling in Arabidopsis. ...................................................................................... 180 Figure 4.6. Phylogenetic tree of JAZ proteins from higher and lower in land plants. ..181 Supplemental Figure 4.1. Amino acid sequence alignment of Arabidopsis .JAZS and JAZ6. .............................................................................................. 183 Supplemental Figure 4.2. Western blot analysis of JAZ proteins expressed in the yeast strains used for the Y2H experiment shown in Figure 4.4A. ............................... 184 Supplemental Figure 4.3. Amino acid sequence alignment of the ZIM and Jas domains of JAZ proteins from various land plants. ................................................... 185 Supplemental Figure 4.4. Structural organization of selected JAZ genes in subclade A. ....................................................................................................... 187 CHAPTER 1: LITERATURE REVIEW l. BIOLOGICAL ROLE OF JASMONATE Jasmonate (JA) was first identified as a volatile fragrance from the jasmine flower and has been used as a scent in many commercial products. In plants, however, JA and its bioactive derivatives are signaling compounds that regulate diverse aspects of plant growth and development. Mutants that are deficient in JA synthesis and signaling have provided powerful tools to identify the many roles of jasmonate in various plant processes. 1.1. Defense against insect herbivores Among various plant processes regulated by JA, defense responses against insect herbivores are among the most extensively explored area. In contrast to other plant responses involving complex crosstalk between several hormone pathways (Koornneef and Pieterse, 2008), numerous studies by independent research groups suggest that defense responses against insect herbivores is largely dependent on the IA pathway (Howe and lander, 2008). JA—mediated defenses provide effective protection against a wide range of arthropod herbivores, including caterpillars (Lepidoptera), beetles (Coleoptera), thrips (Thysanoptera), leaflioppers (Homoptera), spidermites (Acari), and fungal gnats (Diptera) (Browse and Howe, 2008). I Mutants in Arabidopsis and other plants (e.g., tomato) that are defective in IA perception and synthesis have demonstrated JA’s key role in plant resistance to herbivores. Coronatine insensitive 1 (coil) mutants that are defective in JA perception are impaired in all jasmonate responses and, as a consequence, are highly susceptible to herbivore attack (Stintzi et al., 2001; Mewis et al., 2005; Reymond et al., 2004; Zarate et al., 2007). Similarly, constitutive expression of truncated JASMONATE ZIM-DOMAIN (JAZ) proteins (see below) confers strong JA insensitive phenotypes, including compromised plant resistance against the lepidopteron herbivore Spodoptera. exigua (Chung et al., 2008). The JA synthetic mutant acyl—coA oxidase I/acyl-coA oxidase 5 acxIacx5, which is impaired in a peroxisomal B-oxidation step of IA synthesis, is susceptible to the lepidopteran insect Trichoplusia. ni, suggesting a strict requirement for JA synthesis in defense against herbivores (Schilmiller et al., 2007). An intermediate product in octadecanoid pathway for JA synthesis, 12-oxo-phytodienoic acid (OPDA), was also reported to have a role in anti-insect defense responses (Stintzi et al., 2001). However, current studies provide evidence that the OPDA signaling pathway is distinct from the IA signaling pathway (Taki et al., 2005; Thines et al., 2007; Ribot et al., 2008). Most herbivorous chewing insects activate JA synthesis and subsequent large- scale reprogramming of defense gene expression (De Vos et al., 2005; Devoto et al., 2005; Reymond et al., 2000, 2004). Reymond et al. (2004) reported that transcriptional changes induced in Arabidopsis leaves by the crucifer specialist (Pieris. rapae) and the generalist (Spodoptera. littoralis) are nearly identical, suggesting that different lepidopteran insects activate the IA pathway in a similar way. Among the JA-regulated genes induced by insect attack are those encoding defense proteins that directly affect insect performance (e.g., cysteine protease that disrupts the membrane of insect gut), enzymes involved in the synthesis of glucosinolates that are toxic secondary metabolite to insects, and JA biosynthetic enzymes that positively regulate the JA response (Stenzel et al., 2003; Konono et al., 2004; Mewis et al., 2005, 2006; Zhang et al., 2006; Mohan et al., 2006; Wasternack, 2007). Interestingly, phloem-feeding insects such as the whitefly Bemisia tabaci appear to avoid JA-regulated defenses by minimizing tissue damage (Zarate et al., 2008). 1.2. Defense against necrotrophic pathogens Plant defense against bacterial pathogens involves several hormone signaling pathways, including salicylic acid (SA), abscisic acid (ABA), ethylene (ET), etc (Grant and Jones, 2009). However, plant resistance to necrotrophic pathogens, such as Alternaria brassicicola and Botrytis cinerea, depends on the IA signaling pathway and on the production of camalexin (Glazebrook, 2005). Camalexin is an antimicrobial compound (Le, a phytoalexin) that is synthesized by the cytochrome P450 monoxygenase CYP7 IBIS, PHYTOALEXIN DEFICIENT 3 (PAD3). Susceptibility of pad3 is reduced by exogenous JA, further suggesting that JA signaling is required for plant resistance (Thomma et al., 1998). Several independent lines of evidence have established a critical role for JA in plant defense against nectotrophic pathogens. The observation that mutations in the C011 locus, lead to enhanced susceptibility indicates that perception of a JA signal is required for resistance (Thomma et al., 1998). Furthermore, pathogen-induced expression of BOT RYT IS SUSCEPTIBLE 1 (8081), which encodes an R2R3 Myb transcription factor, is dependent on COIl (Mengiste et al., 2003). The bosl mutation compromises resistance of Arabidopsis to both A. brassicicola and B. cinerea (Mengiste et al., 2003; Veronese et al., 2004). Mutations in JASMONA TE RESISTANT 1 (JARI), which encode the enzyme catalyzing the final step in the production of JA-Ile (conjugation of IA to Isoleucine by adenylation) (see below), also compromises plant resistance to B. cinerea, indicating that a JA signal is necessary for the pathogen resistance response (Ferrari et al., 2003). The JA and ET signaling pathways serve antagonistic roles in plant resistance to B. cinerea. MYC2/JA-INSENSITIVE1 (JINl) is a basic helix-loop-helix (bHLH) transcription factor that differentially regulates two branches of the IA signaling pathway in a COIl-dependent manner (Lorenzo et al., 2004). MYC2 positively regulates the expression of wound-response genes, but negatively regulates genes involved in pathogen responses, including PLANT DEFENSIN 1.2 (PDF1.2) gene and several PA THOGENESIS RELATED (PR) genes (Lorenzo et al., 2004). As a consequence, jin] mutant plants are more resistant to B. cinerea, supporting the idea that MYC2 negatively regulates pathogen responses (Lorenzo et al., 2004). Overexpression of the ETHYLENE RESPONSE FACTOR] (ERFI) transcription factor, which regulates the two branches of the JA pathway in a manner opposite to that of MYC2, enhances resistance to B. cinerea (Berrocal-Lobo et al., 2002; Lorenzo et al., 2003, 2004). Furthermore, disruption of the ET pathway by ethylene insensitive 2 (ein2) results in increased susceptibility to B. cinerea (Ferrari et al., 2003; Thomma et al., 1999a). These results exemplify the complexity by which different hormone signaling pathways are integrated for plant defense responses against pathogen attack. 1.3. Plant growth and development In addition to its essential role in plant defense responses, JA is also involved in a wide range of plant growth and developmental processes. Historically, one of the first- described biological effects of 1A was its senescence-promoting properties (Ueda and Kato, 1980). Subsequent studies showed that exogenous JA induces senescence of Arabidopsis leaves in a COIl-dependent manner (He et al., 2002). This study employed gene expression analysis to provide additional evidence for a role of JA in senescencing leaves. The results showed that leaf senescence is accompanied by increased levels of J A and induction of JA-response genes. Among the induced genes were JA-biosynthetic genes and SENESCENCE ASSOCIATED GENES (SAGs), including those involved in photosynthesis and defense responses. JA and wounding up-regulate SAGs in defense responses is and dbwn-regulate house-keeping genes (Wastemack and Hause, 2002; Wastemack, 2007). Root growth inhibition is another well-defined physiological effect of 1A treatment (Dathe et al., 1981). Addition of picomolar amounts of JA or its methyl ester, methyl-1A (MeJA), to the growth medium effectively inhibits root elongation (Staswick et al., 1992). Because of the facile nature of this assay, JA-induced root growth inhibition has been commonly used in many forward genetic screens for mutants that are impaired in IA signaling. Examples of JA-insensitive mutants isolated in this manner include coil (Xie et al., 1998) and myc2/jin1 (Lorenzo et al., 2004). The constitutive expression of VSPI (cev1), constitutive expression of JA-response genes 1 (cexl), constitutive expression of Thi mutants (cet), and bestatin resistant mutants (ber) exhibit constitutive JA signaling and thus have roots that are shorter than wild type plants (Hilpert et al., 2001; Xu et al., 2001; Ellis et al., 2002; Zheng et al., 2006). The effect of JA on growth inhibition is not restricted to roots but rather extends to above-ground tissues as well, indicting that the hormone has a general role in negatively regulating plant growth. An inhibitory effect of MeJA on cell division and specific phases of the cell cycle has been demonstrated (Saniewski, 1988; Swiatek et al., 2002). More recently, Yan et a1. (2007) and Zhang et a1. (2008) reported convincing genetic evidence showing that the IA signaling pathway is involved in wound-induced inhibition of plant growth. Flower development relies on the plant’s ability to synthesize and perceive JA. Several Arabidopsis mutants that are defective in IA biosynthesis (Fig. 1.1), as well as the mutants such as coil that are impaired in JA signaling, are male sterile as a result of defects in anther dehiscence and filament elongation (Feys et al., 1994; McConn and Browse, 1996; Ishiguro et al., 2001; Park et al., 2002; von Malek et al., 2002; Sanders et al., 2000; Stinizi and Browse, 2000). Exogenous JA complements the male sterile phenotype of all JA synthetic mutants, indicating that JA or a derivative of IA directly regulates this developmental program in Arabidopsis (Browse, 2009). Recent studies in maize further indicate that JA is an essential signal in determining male identity (Acosta et al., 2009). It is interesting to note that the C011 homolog in tomato is essential for female fertility, suggesting that the hormone exerts species-specific effects on plant reproductive development (Li et al., 2004). Other developmental processes controlled by JA include trichome patterning (Yoshida et al., 2009) and shade avoidance responses (Moreno et al., 2009). 2. BIOSYNTHESIS OF JA 2.1. The octadecanoid pathway for JA synthesis The octadecanoid pathway for JA biosynthesis is initiated in the chloroplast and completed in the peroxisome (Figure 1.1.). Many enzymes and corresponding genes involved in JA biosynthesis have been identified by forward and reverse genetic screens (Wastemack, 2007). Subcellular localization studies have confirmed the intracellular compartmentation of the various pathway enzymes. The first step of IA biosynthesis is initiated by a chloroplastic lipase that releases linolenic acid (18:3) from membrane lipids (Ishiguro et al., 2001; Hyun et al., 2008). The released linolenic acid is converted by a l3-lipoxygenase to l3-hydroxyperoxylinolenic acid (l3-HPOT), which is then metabolized to lZ-oxo-phytodienoic acid (OPDA) by the sequential action of allene oxide synthase (AOS) and allene oxide cyclase (AOC) (Vick and Zimmerman, 1987; Lee et al., 2008; Ziegler et al., 2000). Although it is unclear how OPDA is transported to the peroxisome, several studies have implicated the ATP-binding cassette transporter PXAl in this transport process (Theodoulou et al., 2005; Footitt et al., 2007). The peroxisomal enzyme OPDA reductase3 (OPR3) reduces OPDA to 3-oxo-2(2’ [Z]-pentenyl)- cyclopentane-l-octanoic acid (OPC-8:0), which is then activated by the carboxyl CoA ligase, OPCLl (Sanders et al., 2000; Schaller et al., 2000; Stintzi and Browse, 2000; Koo et al., 2006). Three cycles of B-oxidation in the peroxisome ultimately complete production of the (3R,7S) stereoisomer of .IA (Fig. 1.1) (Vick and Zimmerman, 1984; Li et al., 2005; Schilmiler et al., 2007). 2.2. JA metabolism Newly synthesized JA can be further metabolized in various ways, including: 1) methylation by JA methyl transferase (JMT) to yield MeJA; 2) decarboxylation to cis- jasmone (Koch et al., 1997); 3) reduction of the keto group in the cyclopentanone ring to yield cucurbic acid (Sembdner and Parthier, 1993); 4); hydroxylation of the C11 or C12 position to produce derivatives that are subsequently modified by glycosylation (Sembdner and Parthier, 1993; Swiatek et al., 2004) or sulfation (Gidda et al., 2003); and 5) JARl-mediated conjugation to amino acids, including the ethylene precursor l- aminocyclopropane-l-carboxylic acid (ACC), to the corresponding JA-amino acid conjugates (Staswick and Tiryaki, 2004). Of particular importance to the mechanism of JA signaling is conjugation of 1A to isoleucine to produce JA-Ile. This reaction is catalyzed by JARl, and likely by other members of the large acyl-activating family of enzymes (Staswick and Tiryaki, 2004). JA-Ile is the most abundant naturally occurring JA-amino acid conjugates in Arabidopsis leaves (Staswick and Tiryaki, 2004; Suza and Staswick, 2008; Koo et al., 2009). Genetic and biochemical analysis of the jar] mutant has shown that JA-lle synthesis is essential for plant defense responses against soil pathogens, lepidopteran insects, and various abiotic stresses as well (Staswick et al., 1998; Rao et al., 2000; Kang et al., 2006). The level of JA-Ile in jar] mutant plants is ~15-fold lower than that in wild type, indicating that JA-Ile synthesis by JARl is essential for activation of JA signaling (Suza and Staswick, 2008). Consistent with this interpretation, JA signaling defects in jar] plants can be complemented with exogenous JA-Ile (Staswick and Tiryaki, 2004). Endogenous levels of J A-Ile rapidly increase in response to mechanical wounding (Chung et al., 2008; Glauser et al., 2008). This response is tightly correlated with transcriptional induction of early JA-response genes (e.g., JAZs). The wound-induced expression pattern of JA-response genes in jar] leaves is very similar to the wild-type response, suggesting that the amount of JA-Ile produced in jar] is sufficient for activation of expression (Chung et al., 2008). This observation, together with the fact that jar] plants are fertile, indicates that JARl is not required for all jasmonate responses. 2.3. Coronatine is a structural mimic of JA-Ile Coronatine is a phytotoxin produced by several plant pathogenic strains of Pseudomonas syringae (Bender et al., 1999). Coronatine and JA induce very similar effects when applied to plant tissues (Feys et al., 1994; Uppalapati et al., 2005). A wealth of genetic, molecular, and physiological data support the hypothesis that coronatine exerts its virulence effects by activating the JA signaling pathway, which subsequently suppresses the SA signaling pathway normally responsible for host immunity to P. syringae (Feys et al., 1994; Zhao et al., 2003; Uppalapati et al., 2005; Nomura et al., 2005; Melotto et al., 2008a). More recent work suggests that coronatine also facilitates bacterial entry into host plants by inducing stomatal opening (Melotto et al., 2008b). Coronatine action, like that of IA, depends on COIl. Indeed, JA-insensitive coil mutants of Arabidopsis were isolated in a screen for plants that are resistant to coronatine (Feys et al., 1994). The structure of coronatine supports the current view that this bacterial toxin is a structural and functional mimic of (3R,7S)-JA-Ile (Figure 1.2; see below). 2.4. Regulation of JA biosynthesis 1A biosynthesis is regulated by a positive feedback mechanism. Most genes encoding JA biosynthetic enzymes, including LOX2, LOX3, AOS, AOCX, OPR3, and OPCLl, are induced via the JA/COIl pathway in response to JA treatment or stress conditions (e.g., wounding) that activate JA synthesis (Reymond et al., 2000; Sasaki et al., 2001; Stenzel IO et al., 2003; Devoto and Turner, 2005; Koo et al., 2006; Chung et al., 2008). Many JA biosynthetic genes are also highly induced in response to treatment with the translation inhibitor cycloheximide (Chung etal., 2008). Thus, JA-induced expression of these genes biosynthetic genes does not require de novo synthesis of transcription factors or other regulatory components, but rather is likely to involve JA-induced removal of a labile repressor. Genetic analysis in Arabidopsis has provided additional evidence for positive feedback regulation of IA biosynthesis. The mutant cevl constitutively overproduces JA and exhibits constitutive expression of A 0C and other JA-response genes (Ellis et al., 2002. The cetl and cet3 mutants also produce elevated levels of OPDA and JA, which is associated with constitutive activation of JA signaling (Hilpert et al., 2001). Conversely, decreased levels in the expression of IA biosynthetic genes in mutants that are deficient in JA synthesis or perception also supports the existence of a positive feedback loop for regulating JA synthesis (Stintzi et al., 2001; Stenzel etal., 2003; Chung et al., 2008). Several lines of evidence indicate that JA biosynthesis is limited by substrate availability rather than by induced expression of biosynthetic genes. For example, constitutive expression of 1A biosynthetic enzymes does not lead elevated production of J A in the absence of wounding or other inductive cues (Laudert et al., 2000; Stenzel et al., 2003). It has also been observed that coil mutants are deficient in wound-induced JA production (Chung et al., 2008; Glauser et al., 2008; Paschold et al., 2008). This observation may be explained by the fact that coil mutant plants contain lower levels of OPDA and other JA biosynthetic precursors (Buseman et al., 2006; Kourtchenko et al., 2007). ll 3. REGULATORY COMPONENTS OF THE JA SIGNALING PATHWAY 3.1. The F-box protein COIl Like other phytohormones, identification of mutants that are defective in IA perception provides essential clues to understand the molecular mechanism of J A signal transduction. The Arabidopsis coil mutant that was isolated in a forward genetic screen to identify mutations that confer resistance to coronatine-inhibited root elongation, are impaired in every aspect of J A responses indicating C011 is an essential component for mediating J A signal (Feys et al., 1994; Browse, 2009). Map-based cloning experiments revealed that C011 encodes an F-box protein, which is a component of the E3-type of ubiquitin ligase known as the SCF (Skp/Cullin/F-box) complex (Xie et al., 1998). F-box proteins impart specificity to the SCF complex by recognizing a specific substrate (typically a regulatory protein), which is then ubiquitinated and subsequently degraded by the 26S proteasome (Moon et al., 2004). Identification of C01] as an F-box protein, together with strong JA- insensitive phenotypes of coil mutants, suggested that protein degradation via ubiquitin/26S proteasome pathway is a critical step in the IA signaling cascade (Xie et al., 1994). Biochemical studies demonstrated the association of COIl with the ASKl/Z, CULl, and RBXl components of the SCFC011 complex (Xu et al., 2002). It was also shown that mutations in the genes encoding these and other components of the SCFCOIl complex impair plant responses to JA (Devoto et al., 2002; Xu et al., 2002; Gray et al., 2003; Lorenzo and Solano, 2005; Ren et al., 2005). The ubiquitin/26S proteasome pathway plays a central role in the mechanism of action of many plant hormones (Santner and Estelle, 2009). In the same year that C011 12 was identified, the F-box protein component of the auxin signaling pathway, TRANSPORT INHIBITOR RESPONSE 1 (TIRI), was cloned and subsequently shown to function as an auxin receptor (Ruegger et al., 1998; Dharmarsiri et al., 2005; Kepinski and Leyser, 2005; Tan etal., 2007). This work led to a model for auxin response in which binding of auxin to TIRl induces recruitment of Aux/1AA repressor proteins to SCFTIRI. Ubiquitin-medaited degradation of the Aux/1AA substrate allows pre-existing ARF transcription factors to promote expression of early auxin response genes (Dharmasiri and Estelle, 2004). Sequence similarity beyween C011 and TIRl led to the hypothesis that JA signaling may involve SCFCOH/26S proteasome-mediated degradation of proteins that repress expression of early JA-response genes. Testing of this hypothesis required knowledge of the SCFCO” substrates, which were not discovered until 2007. . 3.2. JAZ repressor proteins 3.2.1. Discovery of JAZ genes Extensive genetic screening and searches for C011-interacting partners failed to identify targets of SCF C0”. These key components of the J A signaling were finally discovered by transcriptional profiling experiments aimed at identifying JA-response genes in Arabidopsis stamens (Thines et al., 2007). Among 31 genes showing rapid induction afier JA treatment, 7 of them were found to encode ZIM-domain containing proteins with unknown functions. Based on sequence similarity, 5 additional genes encoding ZIM domain-containing proteins were identified in the Arabidopsis genome. These twelve genes were designated as JASMONA TE ZIM-DOMAIN (JAZ) genes (Thines et al., 2007; Chini et al., 2007). As described below and throughout this thesis, the discovery of JAZ l3 proteins has greatly improved our understanding on molecular mechanism of JA signaling. The term ZIM is derived from an Arabidopsis gene (At4g24470) named Zinc- finger protein expressed in Inflorescence Meristem (N ishi et al., 2000). ZIM and ZIM-like (ZML) genes encode putative transcription factors that contain a C-terminal GATA-type zinc-finger domain presumably involved in DNA binding, a CCT (CONSTANS/CO- like/T 0C1) domain implicated in protein-protein interaction (Robson et al., 2001), and a novel ~36-amino-acid sequence motif located near the N-terminus (Shikata et al., 2004). Pfam (http://pfam.sanger.ac.uk/) and InterPro (http://www.ebi.ac.uk/Databases/) databases annotated the latter conserved sequence as a plant-specific domain called ZIM, after the founding member (At4g24470) in which the sequence motif was described. Database searches show that the ZIM domain is present in many plant proteins, including J AZ and PEAPOD (PPD) proteins, which lack a GATA-type zinc—finger or any other recognizable DNA binding motif. Use of the term ZIM to describe both the transcription factor encoded by At4g24470 and the ZIM domain has led to the inclusion of JAZ and PPD proteins in various plant transcription factor databases (Guo et al., 2005; Rushton et al., 2008; Riano-Pachon et al., 2007). Available evidence indicates that JAZs exert their effects on gene expression through physical interaction with DNA binding-type transcription factors in the nucleus (Chini et al., 2007; Thines et al., 2007; Yan et al., 2007; Browse, 2009). These considerations prompted Vanholm et a1. (2007) to rename the family as TIFY, a term that reflects the ZIM domain’s most highly conserved TIFYXG motif (Figure 1.3.B). l4 Sequences encoding TIFY proteins are found in higher and lower (i.e., moss) plants but not in green algae or non-photosynthetic eukaryotes (Vanholme et al., 2007; Katsir et al., 2008a; Chico et al., 2008). In Arabidopsis, TIFY proteins are encoded by 18 genes (Table 1.1 and Figure 1.3.A). Family members can be classified into two major subgroups depending on the presence or absence of the GATA-type zinc finger domain (Vanholme et al., 2007). Based on existing knowledge of member function, phylogeny, and domain architecture, family members may also be distinguished according to whether they contain a CTT domain (3 ZIM/ZMLs) or a divergent CCT domain (see below) previously referred to as Domain 3 (Thines et al., 2007) or the C-terminal (CT) domain (Chini et al., 2007), but now referred to as the Jas domain (Browse, 2008; Yan et al., 2007). Arabidopsis proteins containing the Jas domain or slight variations of this motif include 12 JAZ and 2 PPD proteins (Figure 1.3.A). The protein encoded by At4g32570 is unique in that it contains a ZIM domain but lacks a recognizable CCT or J as motif. 3.2.2. The ZIM domain Conservation of the ZIM domain in JAZ, PPD and ZIM/ZML proteins suggests that this sequence serves an important role in regulating gene expression during plant growth and development. Until recently, however, a molecular function for the ZIM domain had not been described for any member of the family. Yeast two-hybrid and bimolecular fluorescence complementation assays showed that JAZ proteins have the capacity to homodimerize (Table 1.1) (Chung and Howe, 2009; Chini et al., 2009). These studies also revealed a network of JAZ-JAZ heteromeric interactions. JAZ-JAZ partnering may provide combinatorial diversity that modulates the specificity or strength of J A responses. 15 Systematic analysis of truncated JAZ proteins, together with site-directed mutagenesis of the conserved TIFYxG motif, showed that the ZIM domain is necessary and sufficient for JAZ-JAZ interaction. The ability of ZIM and ZIM-like (ZML) proteins to homo- and heterodimerize in yeast indicates that protein-protein interaction involving the ZIM domain extends to other members of the TIFY family (Chung and Howe, 2009). This observation raises the posSibility that JAZ proteins functionally interact with PPD and ZIM/ZML proteins. A physiological role for the ZIM domain in IA signaling was established by experiments showing that the strong lA-insensitive phenotype of JAZIO.4- overexpressing plants is suppressed by point mutations within the TIFY motif that block JAZlO.4 interaction with other JAZs (Chung and Howe, 2009). The dominant negative action of JAZIO.4 therefore depends on ZIM domain-mediated interaction with another TIFY protein, possibly JAZlO.4 itself. The ability of dimerization-defective forms of JAZIO.4 to accumulate in the nucleus argues against the possibility that JAZ-JAZ interaction is required for nuclear import of the repressor. The importance of the ZIM domain for JAZlO.4-mediated repression of IA signaling provides new insight into the mechanism by which J AZ proteins negatively regulate the J A pathway. 3.2.3. The Jas domain A distinguishing feature of JAZ proteins is the 27-amino-acid Jas domain located near the C-terminus (Figure 1.3.C). Interestingly, this sequence is similar to the N-terminal portion of the CCT domain that was first identified in the plant proteins TOCl and CONSTANS (CO) (Strayer et al., 2000; Robson et al., 2001) (Figure 1.4.). Based on this 16 similarity, the Jas domain is described in the Pfam database as a divergent CCT motif called CCT2 (PF09425). CONSTANS and other CCT domain-containing proteins have a well-established role in regulating plant responses to light, temperature, and the circadian clock. Although the C-terminal portion of the CCT domain has been implicated in protein-protein interactions that regulate gene expression, the biochemical function of the N-terminal CCT sequence resembling the Jas domain is not known (Wenkel et al., 2006). It will be interesting to explore the evolutionary and functional relationship between the J as and CCT domains. Several independent lines of evidence indicate that the Jas domain is involved in destabilizing JAZ repressors by a mechanism that directly couples hormone perception to 1 AZ degradation. First, JAZ proteins fused to GUS or GF P/YF P reporters are degraded in JA-treated cells in a manner that depends on C011 and the 26S proteasome (Thines et al., 2007; Chini et al., 2007). Second, truncated JAZ proteins (referred to as JAZAJas) that lack the Jas domain are stable in the presence of IA (Chini et al., 2007; Thines et al., 2007; Shoji et al., 2008; Chung and Howe, 2009). Several studies have shown that overexpression of JAZAJas proteins results in hallmark JA-insensitive phenotypes, including reduced expression of IA response genes and secondary metabolites, resistance to JA-mediated inhibition of root growth, susceptibility to insect feeding, enhanced resistance to coronatine-producing strains of Pseudomonas syringae, and male sterility (Thines et al., 2007; Chini et al., 2007; Melotto et al., 2008; Shoji et al., 2008; Chung and Howe, 2009). These initial observations led to the hypothesis that JAZ proteins may be FCOII substrates for SC and that the Jas domain mediates protein destabilization. 17 Yeast two-hybrid and in vitro protein pull-down experiments have shown that JAZ proteins bind directly to COD in a hormone-dependent manner, and that the Jas domain is necessary and sufficient for this interaction (Thines et al., 2007; Melotto et al., 2008; Katsir et al., 2008b; Chung and Howe, 2009; Chini et al, 2009). Melotto et a1. (2008) identified two adjacent basic amino acids near the N-terminal end of the Jas domain of JAZl and JAZ9 (R205R206 and R223K224, respectively; Figure 1.3.C) that are required for interaction with C01]. Alanine substitution of these residues blocked hormone-induced interaction with C011 and, in the case of JAZI , conferred dominant JA insensitivity in transgenic plants. A JAZIO splice variant (JAZlO.3) harboring a truncated Jas domain ending in the conserved KRKER motif (Figure 1.3.C) was shown to interact weakly with COIl in comparison to the full-length splice variant (JAZ10.1) (Chung and Howe, 2009). Overexpression of JAZlO.3 resulted in partial insensitivity to JA, whereas overexpression of JAZlO.l had no apparent effect on JA responsiveness (Yan et al., 2007; Chung and Howe, 2009). These structure-function studies provide initial insight into regions of the Jas domain that have a role in C011-JAZ interaction and JAZ destabilization. Additional work is needed to define sequence determinants that are necessary and sufficient for degron activity. X-ray crystallography studies, similar to those performed with the auxin receptor TIRl (Tan et al., 2007), are expected to provide important insight into the molecular details of how the Jas domain interfaces with C01] in a hormone-dependent manner. 3.2.4. Functional diversification of JAZ proteins by alternative splicing 18 Alternative splicing is a fundamental process for expanding protein diversity and the functional complexity of eukaryotic organisms. Recent studies indicate that 95% of multiexon transcripts encoded by the human genome undergo alternative splicing (Pan et al., 2008). Although our understanding of alternative splicing in plants is still in its infancy, increasing evidence indicates that this process promotes plant adaptation to stress (Barbazuk et al., 2008; Reddy et al., 2007). The recent identification of functionally distinct splice variants of JAZlO/JASI has broadened our appreciation of the role of post- transcriptional regulation in IA signaling (Yan et al., 2007; Chung and Howe, 2009). Alternative splicing of JAZIO pre-mRNA generates three splice variants that differ in the sequence of the Jas domain. The full-length JAZlO.l protein contains an intact Jas domain similar to that of other JAZ proteins that are destabilized in the presence of JA. JAZ10.1 strongly interacts with C011 in a ligand-dependent manner and is degraded via the 26S proteasome pathway in response to JA treatment (Chung and Howe, 2009). JAZlO.3, which lacks seven amino acids from the C-terminal end of the of Jas motif, interacts weakly with COIl in a ligand-dependent manner and is degraded in planta in response to high concentrations of exogenous JA. Consistent with the intermediate stability of JAZlO.3, overexpression of this splice variant in Arabidopsis confers partial insensitivity to JA (Yan et al., 2007; Chung and Howe, 2009). A third JAZIO splice variant (JAZIO.4) lacks the entire Jas domain. As predicted, this protein does not interact with C011 and is highly resistant to JA-mediated degradation. JAZlO.4-overexpressing lines exhibit strong JA insensitive phenotypes similar to those observed in coil mutants. These findings demonstrate that alternative splicing events affecting the Jas domain expand the functional diversity of JAZ proteins in Arabidopsis. It will be interesting to 19 determine whether alternative splicing of JAZIO pre-mRNA is regulated in a manner that favors the accumulation of a particular transcript in specific cell types, and whether alternative splicing events affecting the Jas domain are widespread in the plant kingdom. A common feature of hormone signaling pathways in all eukaryotes is that prolonged stimulation decreases responsiveness to the signal, a phenomenon called desensitization. JAZlO.3 and JAZIO.4 appear to function in this capacity. According to this hypothesis, synthesis of JA-Ile in response to inductive cues would trigger rapid destruction of unstable JAZs (e.g., JAZIOJ) that repress JA response genes in unstimulated cells. Depletion of these J AZs would lead to transcriptional activation of the JAZ/0 gene, which itself is controlled by the SCFCOH/JAZ pathway (Yan et al., 2007; Chung et al., 2008). Alternative splicing of JAZIO pre-mRNA generates transcripts for de novo synthesis of all three JAZIO splice variants but, owing to differences in protein stability, only the JAZlO.3 and JAZlO.4 proteins are predicted to accumulate and eventually attenuate the signal output. In this manner, JA—stimulated cells would become desensitized to elevated hormone levels through the synthesis of JAZlO.3/JAZIO.4 and perhaps other stable JAZ proteins. This model predicts that the increased sensitivity of jazIO loss-of-function mutants (Yan et al., 2007) results from the absence of JAZlO.3/ 10.4 rather than from reduced production of the more labile JAZlO.1. The ability of JAZlO.3/10.4 to desensitize the signaling pathway may be important for curtailing JA responses that are not beneficial to the plant. For example, JAZIO.4 may attenuate the virulence activity of the P. syringae toxin coronatine. It is also possible that stabilized JAZs play a role in promoting plant growth under environmental conditions in which plants are faced with the “dilemma to grow or defend” (Herms and Mattson, 1992). 20 3.3. (3R,7S)-JA-Ile is the active stereoisomer of JA-Ile Bioactive JAs can be defined as JA derivatives that directly promote the formation of C011-JAZ complexes (Katsir et al., 2008a). Non-bioactive JAs are either precursors or deactivated forms of the bioactive compounds. Protein-protein interaction studies have shown that binding of several JAZ proteins to C01] is stimulated by the JA-amino acid conjugate JA-Ile (Thines et al., 2007; Katsir et al., 2008b; Melotto et al., 2008). Significantly, JA, methyl-IA (MeJA), and the JA precursor OPDA do not promote these interactions. Identification of JA-Ile as a causal signal for COIl-JAZ binding extends the groundbreaking work by Staswick and colleagues showing that IA is activated by JARl- mediated conjugation to Ile (Staswick and Tiryaki, 2004; Staswick, 2008). The combined genetic analysis of jarl mutants (Staswick and Tiryaki, 2004) and use of cell-free and yeast-based assays to study JA-Ile action (Thines et al., 2007) definitively establishes this conjugate as the bioactive form of the hormone. Structurally related JA-amino acid conjugates, including JA-Val and JA—Leu, also promote interaction between tomato C011 and JAZl/3 proteins to varying degrees in vitro (Thines et al., 2007; Katsir eta1., 2008b). These conjugates did not stimulate interaction between Arabidopsis C011 and JAZ3/9 proteins (Forseca et al., 2009), suggesting possible differences in the binding characteristics of CO“ proteins from different plant species. Several studies have provided evidence that non-conjugated JA derivatives elicit distinct responses (Hopke et al., 1994; Blechert et al., 1999; Miersch et al., 1999; Stintzi et al., 2001; Wang et al., 2008; Ribot et al., 2008). However, all C011-JAZ interaction studies reported to date show that these compounds do not promote C011-J AZ interactions. 21 Plants contain two stereoisomers of JA-Ile, referred to as (3R,7S)-JA-Ile and (3R,7R)-JA-lle, that differ with respect to the orientation of the pentenyl side chain (Figure 1.2.) (Creelman and Mullet, 1997; Wastemack, 2007). Initial C011-JAZ interaction studies (Thines et al., 2007; Katsir et al., 2008b; Melotto et al., 2008) used a standard synthetic preparation of JA-Ile that contains predominantly (3R,7S) and minor amounts of the thermodynamically less stable (3R,7R) isomer, in equilibrium (Miersch et al., 1986; Creelman and Mullet, 1997; Wastemack et al., 2007). These studies established the specificity of JA-Ile as the chemical mediator of COIl-JAZ binding but, because of the mixed stereoisomeric composition of the preparation, did not draw conclusions about the relative activity of the (3R,7S) and (3R,7R) isomers. Because (3R,7S)-JA-Ile is stereochemically similar to coronatine (Figure 1.2), this isomer was generally assumed to be the active form of JA-lle (Staswick, 2008). This was recently confirmed in studies performed with purified stereoisomers of (3R,7S)-JA-Ile and (3R,7R)-JA-Ile (Fonseca et al., 2009). In showing that (3R,7R)-JA-Ile is unable to promote interaction of C011 and JAZ proteins from Arabidopsis, Fonseca et al. (2009) proposed that (3R,7S)-JA-Ile may be inactivated by epimerization to (3R,7R)-JA-Ile in vivo. 3.4. MYC2 is a JAZ-interacting transcription factor JA exerts its many effects through large-scale reprogramming of gene expression. DNA microarray studies employing various treatments and mutant backgrounds have identified hundreds of JA-response genes in Arabidopsis and other plants (Schenk et al., 2000; Reymond et al., 2000; Sasaki, 2001; Goossens et al., 2003; Sasaki-sekimoto et al., 2005; 22 Suzuki et al., 2005; Devoto et al., 2005; Uppalapati et al., 2005; Mandaokar et al., 2006; Reymond et al., 2004). Meaningful interpretation of this expression data requires identification of the relevant transcription factors and their biological functions in IA signaling. Forward and reverse genetic screens have identified a number of transcription factors involved in JA-mediated transcriptional changes. These include a bHLH transcription factor MYC2, members of ApetalaZ(AP2)/ERF family including ERFl, ERF2, ERF4, OCTADECANOID RESPONSIVE ARABIDOPSIS AP2/ERF 37 (ORA37), ORA47, ORA59, and several WRKY proteins including WRKY18, WRKY70, (Fujimoto et al., 2000; Lorenzo et al., 2003, 2004; Atallah, 2005; Wastemack, 2007; Chico et al., 2008; Wang et al., 2008). The bHLH protein MYC2 is the only transcription factor known to directly interact with JAZ proteins and to regulate the expression of early JA-response genes. M YC2 mutants were identified in two independent screens for JA-insensitive plants; these mutants were named fin] and jai1 (jasmonate-insensitivel) (Berger et al., 1996; Lorenzo et al., 2004). jinI/jail mutants exhibit reduced sensitivity to JA-mediated root growth inhibition, a typical JA-resistant phenotype. As mentioned above, MYC2 differentially regulates two branches of JA-mediated responses (Lorenzo et al., 2004). Mutant analysis in Arabidopsis has also revealed that MYC2 acts as an integrator of light, ABA, and JA signaling pathways (Abe et al., 1997; Abe et al., 2003; Lorenzo et al., 2004; Yadav et al., 2005; Dombrecht et al., 2007). JAZ proteins do not contain a known DNA-binding domain, but rather are hypothesized to regulate gene expression through interaction with DNA-binding transcription factors. Because MYC2 is a key transcription factor in JA responses, Chini 23 et al. (2007) considered MYC2 to be a potential candidate for direct interaction with JAZ proteins. Indeed, MYC2 physically associates with most JAZ proteins in yeast two- hybrid and pull-down assays (Table 1.1) (Chini et a1, 2007; Chung and Howe, 2009; Chini et al., 2009). Promoter analysis revealed that the MYC2-binding G-box motif, or a variant called the T/G-box, is over-represented in JAZ promoters. Protein-DNA binding studies confirmed that MYC2 binds directly to the JAZ3 promoter region that contains these motifs (Chini et al., 2007). Furthermore, many JAZ genes are repressed in myc211;in_1 mutants and constitutively expressed in MYC2-overexpressing plants. These and other recent results suggest that MYC2-mediated expression of JA-early response genes, including JAZ genes, is controlled by JAZ3 and other JAZs via direct protein interaction (Chini et al., 2007; Chung et al., 2008; Melotto et al., 2008; Chung and Howe, 2009). Chini and colleagues (2007) showed that JAZ3 interacts with the N-terminal region of MYC2. Interestingly, this plant-specific region of MYC2 is conserved in only a subgroup of plant bHLH proteins (Heim et al., 2003). Identification of specific sequence determinants in MYC2 that interact with JAZs could be helpful for identifying other transcription factors that are controlled by JAZ proteins. Indeed, MYC2 cannot be the only transcription factor targeted by JAZ repressors because mch/jinl loss-of-function mutants are not deficient in all JA responses; for example, these mutants are not male sterile (Lorenzo et al., 2004). Although information about sequence determinants within JAZ proteins that interact with MYC2 is beginning to emerge, a single MYC2-interacting sequence common to all JAZs has not been identified. Pull-down and yeast two—hybrid assays 24 showed that the Jas domain is necessary and sufficient for interaction of JAZ3 with MYC2 (Chini et al., 2007; Chini et al., 2009). Mutations in the Jas domain that block hormone-dependent interaction of JAZ9 with COIl did not affect MYC2 binding, suggesting that the C011 and MYC2 interaction surfaces of JAZ9 are not identical (Melotto et al., 2008). Yeast two-hybrid assays showed that the Jas domain of JAZIO is not required for interaction MYC2, indicating that sequences outside the Jas domain are responsible for repression of IA signaling by JAZlO.4 (Chung and Howe, 2009). A greater understanding of how JAZ proteins interact with MYC2 (and other transcription factors) will likely shed light on the molecular mechanism by which JAZs repress the expression of JA-responsive genes. In contrast to the initial model put forth by Chini et al. (2007), recent studies suggest that repression of IA signal output by JAZlO.4 and JAZ3AJas may involve ZIM domain-mediated interaction with another JAZ protein (Chung and Howe, 2009; Chini et al., 2009). The fact that bHLH transcription factors typically function as homo- or heterodimers (Heim et al., 2003) is consistent with the idea that a JAZ dimer is the functional unit for MYC2 repression. JAZ-MYC2 complexes may directly repress MYC2 activity, for example by preventing MYC2 from binding target promoter regions. It is also possible that JAZ-MYC2 complexes recruit transcriptional co-repressors to the pre-initiation complex (Browse, 2009). 4. CURRENT MODEL OF JA SIGNALING Identification of JAZ repressor proteins has provided many new insights into the molecular mechanism of IA signaling. In healthy (e.g., unwounded) plants, low levels of JA-Ile allow JAZ proteins to accumulate and repress expression of JA-response genes 25 (Figure 1.5.). This repressed stage is established by direct interaction of JAZ proteins with transcription factors such as MYC2 that promote the expression of target genes. In response to various biotic or abiotic stresses or developmental cues, accumulation of JA- Ile promotes direct interaction of JAZ with SCFCO”. Formation of the resulting COIl-JA- JAZ complex leads to ubiquitination and degradation of JAZ proteins via the 26S proteasome, which releases MYC2 from repression. Among the early JA-response genes activated by MYC2 are those encoding JAZ proteins and several JA biosynthetic enzymes. Increased expression of the latter may reflect a positive feedback loop to enhance JA biosynthetic capacity. In contrast, induced expression of JAZ proteins appears to provide the cell with a means to rapidly attenuate JA signaling. Such fine tuning of the IA signaling pathway may allow the plant to mount a response that is commensurate with the nature of the threat encountered. 5. AIMS OF THE THESIS RESEARCH The research described in this thesis was initiated shortly after the discovery of JAZ proteins as negative regulator of IA signaling in Arabidopsis (Thines et al., 2007). Although Thines et al (2007) clearly established a role for JAZ proteins in repressing JA signaling, very little was known about the general mechanism of JAZ action. Two initial aims of this thesis research were to determine: 1) how the JAZ gene family in Arabidopsis is regulated by JA and wound stress, and 2) whether J AZ proteins play a role in plant defense responses to insect herbivores. The results of experiments designed to address these questions are described in Chapter 2 (Chung et al., 2008). Another central aim of this thesis was to investigate the biochemical function of the conserved ZIM and 26 Jas domains in JAZ proteins. The results of experiments aimed at addressing this issue are described in Chapter 3 (Chung and Howe, 2009). During the course of these studies, it was discovered that several JAZ genes (e.g., JAZIO) are subject to alternative splicing events that modify the Jas domain. Following up on this observation, a detailed comparative analysis of JAZ genes from diverse plant species was undertaken. This research, which is described in Chapter 4, identified an evolutionarily conserved intron whose retention during pre-mRNA splicing contributes to functional diversification of J AZ proteins through altered protein stability. 27 Figure 1.1. The jasmonate biosynthetic pathway in Arabidopsis. Enzymes of the pathway are boxed. Arabidopsis mutants that are defective in the corresponding enzyme activity are indicated to the right (italicized symbols). fad, fatty acid desaturase; dad], defective in anther dehiscence]; 10x2, lipoxygenaseZ; aos, allene oxide synthase; opr3, opda reductase3; opcll ; opc-830 CoA ligase] ; acx, acyl CoA oxydase. 28 radsradr t... I dad1 a-llnolenlc acld I FS-Lipoxygenase ] 10x2 OOH \ _ 13(S)-hydroperoxyllnolenlc acld COOH I Fellene oxide synthasrfl aos o \ __ 12, 13(8)-epoxyl|nolenlc acld COOI'I I [Allene oxide cyclasej o coon (98,136)-12-oxo-cls-10,15-phytodlenolc acld (OPDA) I [OPDA reductaseJ opr3 o coon 3-oxo-2v(cls-2’~pentenyl)-cyclopentane-1-octanolc acld (CFC-8:0) I (CPO-8:0 CoA ligase I cpcl1 O O 'é-S-COA 3-oxo-2~(cls-2’-pentenyl)-cyclopentane-1-octanoyl CoA (OPC-8:0-CoA) * B-oxidation x 3 acxt acx5 coon (3R,78)-jasmonlc acld (JA) 29 \ \I \\ Chloroplastlc steps 1' l I \ s Peroxlsomal steps I! l (3R,7S)—JA—L-lle (3R,7R)-JA-L-lle Coronatine Figure 1.2. Chemical structure of JA-Ile diastereomers and coronatine. (3R, 7S)-JA-Ile can epimerize to (3R, 7R)-JA-Ile. 30 Figure 1.3. The TIFY protein family in Arabidopsis. (A) The phylogenetic tree includes all known Arabidopsis TIFY proteins, including 14 members of the JAZ subfamily. Full-length amino acid sequences were aligned using clastalW and the tree was constructed by Neighbor Joining (NJ) method. The relative position of the conserved domains in each protein is shown in color. (B-C) Sequence logo (Crooks et al., 2004) of the ZIM (B) and Jas (C) domains found in Arabidopsis. PPD proteins contain a diverged Jas domain that lacks the conserved PY at the C- terrninal end (Fig. 1.4). Sequences used to create the ZIM and Jas domain logos were 36 and 27 amino acids, respectively, in length. Secondary structure (or—helix, red; B-sheet, yellow) in each domain was predicted by Jpred3 (Cole et al., 2008; http://www.compbio.dundee.ac.uk/~www- jpred/) and is depicted below the logo. 31 ZIM Jas CCT GATA ZInc-finger J— JAZ11 cm L- JAZ12 21:21: T ZIM |_‘{: ZML1 :2-2-2-2 ZML2 1:21:31: 'llWl'U AT4G32570 £22.22: co €011 €012 €013 €010 (301.5 €016 COL? €010 €013 €01.10 €01.11 €01.12 €01.13 €01.14 €01.15 €01.16 ".1 m2 m1 m2 m1 m2 me ms ms “5 m7 m9 mm mu £13312 CCT domain 20 30 so 50 WIHm-mt IIII ---1 LS “1’ n In l-I:I_II;H I ‘1 .-I4 II . -‘I I tit 11 It" I-I’I 'I-—- LSP “15:11 [In] ”Hz-t: :1 :IHII . - WEFII'I:I$I.I’11H——— LT :I."?! I t-l I: :I: .l-‘il aI-HII I —-~l Ignml il --l-:t.I.-l I ——— LSP III Afll- t-l KIIHI Iii mum-l — ~l- I Air-drum III rn——- LI :Itll‘tfil ..l I: III-Z le‘l led-CHI In .- It. nil-um ltl rn—-- AS ‘ t- II.“ II t.l I: Itlvil nil III: II". — urtEIzilugxmzl ra——— ADD tilt-t RES: [.1 III I I-ihill’i it HI-iIr'Izmrl" S--— SS!) :I.-n ~14! 1l :II'II-i'l‘ n. tilt-imam ' tS——- in wl-L‘HI-II-II‘II-l iIIi-I-Ir't' *I'IfidI-"P-HIJ-‘I SL-—— first -I I: -I: I:.'l -I:- I". . : -rt ‘1: I.- I:l.l.-l cz—-— .. 151.1 I:-I:!l.-!l ‘Kflll‘f - .- 15.x“ :‘lHKtJH ' an--- ..I—. ".l“.. ,ql. I‘ll. . 1 ‘14! t-gLI.$It-I '1’; I I! t - ‘I-‘Ifi¥'lt 11.1-ilmtlll- SOB—-- 1 I"; II: t 'IP~‘I~ FELIPY : art It Ir I:l.ItI ' —-— l cuff-d H-vrlgI-'HK:I It". - Igfl . IJ-l-Nil—hl-‘l —-- MI’T:I I::K:I-’EH lI<"I' . 1*n. trip-Emu I'D--— mo Irmlll-"l'.'l P2113“ unnullw - T - -‘ -I* AlAS!—--— t8!) - :I: n It't' - l- I:Iz It Ink -I In I S--- CSLLII;:IIIi :KIIA’REI:H‘-I-Yr SSfl—u- mu? :1 I I-' I:ItI:.-I2*l iHHIItTT— I SAKS—~— rSIPIIr A .l l. I.III:=.-i -I< 'lR'l’ SAKS—- ZGI Paar : - ———- 28! I: - ‘-.I !———— BL? Itinxl CL--— 31.? H.713] SIP——r—NPAILSSR-— ALP I‘m: —ssrncuflsuszcxsssr.s—— sto II'A‘J *SSIDCR'I'IJSBCVSCPPAwe VIII! Irl'nm OMYPPKPBIVTGQPL—~— .t- t-n “.1 8811— -LPPKPBIVAPSIKS9— 1| :1 - t‘-.<,l NP! ~I"~.II SW 1'. 1'in I _ : _. I LDLSTOBSSG—IDIISSTSPT—— nut It- I ~HI ”i=1 I::ItI I-LVSTS 15A ————————————————————— WP I I'qul KII‘PM ‘SATrsnAnmm'l‘SPm , DLP i:+- I III. lI;ll n ~LVNKNP -rprsnmrmrows1mrpr-— Jas domain Figure 1.4. Sequence similarity between the Jas and CCT domains. Sequence alignment of the Jas domain of Arabidopsis JAZ and PPD proteins with the CCT domain of Arabidopsis ZIM, ZIM-like (ZML), CONSTANS (CO), and CO-like (COL) proteins using Clustal W. Sequences used for the alignment consisted of 50 amino acids that span the domain. Sequences less than 50 amino acids in length were used for those JAZ proteins that terminate shortly after the J as domain 33 Developmental cues Bioticl Ablotlc stress JAZ Unstable JAZs SCFCO" JA.||° "‘ JAZ ‘ Stable JAZS v @ ° o 0 00° JA7 \. "JAZ“ — , .----> JAZ "a ’JAZI'L‘ l \w/ Early genes Early genes o I " ‘i' x km/ Early genes I . _____ I d' JAZs, _J O $58231 G biosynfijefic genes) Negative regulation Jasmonate responses Figure 1.5. Current model of jasmonate signaling. Low levels of JA in the cell allow JAZ repressors to inhibit the activity of transcription factors (TF, e.g., MYC2) and repress expression of early JA-response genes. Bioactive JAs (e.g., JA-Ile) are synthesized in response to developmental cues and biotic/abiotic stresses. Increased levels of JA-Ile in the cell stimulate degradation of stable JAZ proteins via the ubiquitin/26S proteasome pathway mediated by SCFCO”. Transcription of JAZ genes and alternative splicing of primary JAZ mRNAs results in production of JAZ variants that are stabilized against JA-mediated turnover. Accumulation of stable JAZ proteins in the cell desensitizes JA responses. 34 Table 1.1. The TIFY family in Arabidopsis If TIFY 1»le Homo- Interaction Interaction Protein name 1 AGI number motlf Locallzatlon Dlmerlzatlon 2’3 wlth COI1 wlth MYC2 ZIM TIFY1 AT4624470 TISFRG Nuclear Yes/nd nd Nd ZML1 TIFY2a AT1651600 TLSFQG Nd Yes/nd nd Nd ZML2 TlFY2b AT3621175 TLSFQG Nd Yes/nd nd Nd JAZ1 TlFY10a AT1619180 TIFYAG Nuclear Yes/Y es Yes Yes JAZZ TIFY10b AT1674950 TIFYGG Nd Yes/No nd Yes JAZ3 TlFY6b AT3617860 TIFYAG Nuclear Yes/Yes Yes Yes JAZ4 TIFY6a AT1648500 TIFYAG Nd Yes/Yes nd Yes JA25 TIFY11a AT1617380 TIFFGG Nd Yes/No nd Yes JA26 TlFY11b AT1672450 TIFFGG Nuclear Yes/No nd Yes JAZ7 TlFY5b AT2634600 TIFYNG Nd No/No nd No JAZ8 TIFY5a AT1630135 TIFYNG Nd No/No nd Yes JAZQ TIFY7 AT1670700 TIFYGG Nd No/Y es Yes Yes JAZ10.1 TIFY9 AT5613220.1 TIFYNG Nuclear Yes/No Yes Yes JAZ10.3 AT5613220.3 TIFYNG Nuclear Yes/nd Yes (weak) Yes JAZ10.4 AT5613220.4 TIFYN6 Nuclear Yes/nd No Yes JAZ11 TlFY3a AT3643440 TllF66 Nd No/No nd Yes JAZ12 TlFY3b AT5620900 TIFFGG Nd NolNo nd Yes PPD1 TIFY4a AT4614713 TIFYSG Nd nd/nd nd Nd PPD2 TlFY4b AT4614720 TIFYSG Nd nd/nd nd Nd Unknown TlFY8 AT4632570 TIFYGG Nd nd/nd nd Nd 1 Vanholme et al. (2007) 2/3 Homodimerization capacity as determined by yeast two-hybrid analysis in two different studies (Chung and Howe, 2009/Chini et al., 2009). Differences in the JAZ-JAZ interactions reported in these studies may reflect differences in the stringency of the yeast two-hybrid assays employed. 4 Indicates JA-Ile- or coronatine-dependent JAZ interaction with COIl in yeast two-hybrid and/or in vitro pull-down assay (Thines et al., 2007 ;Melotto et al., 2008; Chung and Howe, 2009; Chini et al., 2009). 5 As determined by yeast two-hybrid and/or in vitro pull-down assay (Chini et al., 2007; Melotto et al., 2008; Chung and Howe 2009; Chini et al, 2009). nd, not determined. 35 REFERENCES Abe, H., Yamaguchi-Shinozaki, K., Urao, T., Iwasaki, T., Hosokawa, D., Shinozaki, K. (1997). Role of arabidopsis MYC and MYB homologs in drought- and abscisic acid-regulated gene expression. Plant Cell 9: 1859-1868. Abe, H., Urao, T., Ito, T., Seki, M., Shinozaki, K., Yamaguchi-Shinozaki, K. (2003). Arabidopsis AtMYCZ (bHLH) and AtMYBZ (MYB) function as transcriptional activators in abscisic acid signaling. Plant Cell 15: 63-78. Acosta, I.F., Laparra, H., Romero, S.P., Schmelz, E., Hamberg, M., Mottinger, J.P., Moreno, M.A., Dellaporta, S.L. (2009). tasselseedl is a lipoxygenase affecting jasmonic acid signaling in sex determination of maize. Science. 323: 262-265. Barbazuk, W.B., Fu, Y., and McGinnis, K.M. (2008). Genome-wide analyses of alternative splicing in plants: Opportunities and challenges. Genome Res. 18: 1381-1392. Bender, C.L., Alarcon-Chaidez, F., Gross, D.C. (1999). Pseudomonas syringae phytotoxins: mode of action, regulation, and biosynthesis by peptide and polyketide synthetases. Microbiol Mol Biol Rev. 63: 266-292. Berrocal-Lobo, M., Molina, A., Solano, R. (2002). Constitutive expression of ETHYLENE-RESPONSE-FACTORI in Arabidopsis confers resistance to several necrotrophic fungi. Plant J. 29: 23-32. Blechert, S., Bockelmann, C., Fusslein, M., Von Schrader, T., Stelmach, B., Niesel, U., Weller, E.W. (1999). Structure-activity analyses reveal the existence of two separate groups of active octadecanoids in elicitation of the tendril-coiling response of Bryom'a dioica J acq. Planta 207: 470—479. Browse, J. (2009). Jasmonate Passes Muster: A Receptor and Targets for the Defense Hormone. Annu. Rev. Plant Biol. 60: 183-205. Browse, J., Howe, G.A. (2008). New weapons and a rapid response against insect attack. Plant Physiol. 146: 832-838. Buseman, C.M., Tamura, P., Sparks, A.A., Baughman, E.J., Maatta, S., Zhao, J., Roth, M.R., Esch, S.W., Shah, J., Williams, T.D. et al. (2006). Wounding stimulates the accumulation of glycerolipids containing oxophytodienoic acid and dinor-oxophytodienoic acid in Arabidopsis leaves. Plant Physiol 142: 28—39 Chico, J.M., Chini, A., Fonseca, S., Solano, R. (2008). JAZ repressors set the rhythm in jasmonate signaling. Curr. Opin. Plant Biol. 11: 486-494. 36 Chini, A., Fonseca, S., Fernandez, G., Adie, B., Chico, J.M., Lorenzo, 0., Garcia- Casado, 6., Lopez-Vidriero, 1., Lozano, F.M., Ponce, M.R., Micol, J.L., and Solano, R. (2007). The JAZ family of repressors is the missing link in jasmonate signalling. Nature 448: 666-671. Chini, A., Fonseca, S., Chico, J., Fernandez-Calvo, P., Solano, R. (2009). The ZIM domain mediates homo- and heteromeric interactions between Arabidopsis JAZ proteins. Plant J. (in press) Chung, H.S., Koo, A.J.K., Gao, X., Jayany, S., Thines, B., Jones, A.D., Howe, G.A. (2008). Regulation and function of Arabidopsis JASMONA TE ZIM-domain genes in response to wounding and herbivory. Plant Physiol. 146: 952-964. Chung, H.S., Howe, G.A. (2009). A critical role for the TIFY motif in repression of jasmonate signaling by a stabilized splice variant of the JASMONATE ZIM- domain protein JAZIO in Arabidopsis. Plant Cell 21: 131-145. Creelman, R.A., Mullet, J.E. (1997). Oligosaccharins, brassinolides, and jasmonates: nontraditional regulators of plant growth, development, and gene expression. Plant Cell 9: 1211-1223. Dathe, W., R6nsch, H., Preiss, A., Schade, W., Sembdner, G., Schreiber, K. (1981). Endogenous plant hormones of the broad bean, Vicia faba L. ( - )-Jasmonic acid, a plant growth inhibitor in pericarp. Planta 155: 530-535. Dharmasiri, N., Estelle, M. (2004). Auxin signaling and regulated protein degradation. Trends Plant Sci. 9: 302-308. Dharmasiri, N., Dharmasiri, S., Estelle, M. (2005). The F-box protein TIRl is an auxin receptor. Nature 435: 441-445. De Vos, M., Van Oosten, V.R., Van Poecke, R.M., Van Pelt, J.A., Pozo, M.J., Mueller, M.J., Buchala, A.J., Métraux, J.P., Van Loon, L.C., Dicke, M., Pieterse, CM. (2005). Signal signature and transcriptome changes of Arabidopsis during pathogen and insect attack. Mol Plant Microbe Interact. 18: 923-937. Devoto, A., Nieto-Rostro, M., Xie, D., Ellis, C., Harmston, R., Patrick, E., Davis, J., Sherratt, L., Coleman, M., Turner, J. G. (2002). C011 links jasmonate signalling and fertility to the SCF ubiquitin-ligase complex in Arabidopsis. Plant J. 32: 457-466. Devoto, A., Turner, J.G. (2005). Jasmonate-regulated Arabidopsis stress signalling network. Physiol Plant 123: 161—172. 37 Devoto, A., Ellis, C., Magusin, A., Chang, H.S., Chilcott, C., Zhu, T.,and Turner, J.G. (2005). Expression profiling reveals COIl to be a keyregulator of genes involved in wound- and methyl jasmonate-inducedsecondary metabolism, defence, and hormone interactions. Plant Mol. Biol. 58: 497—513. Dombrecht, B., Xue, G.P., Sprague, S.J., Kirkegaard, J.A., Ross, J .J., Reid, J.B., Fitt, G.P., Sewelam, N., Schenk, P.M., Manners, J.M. et al. (2007). MYC2 differentially modulates diverse jasmonate-dependent functions in Arabidopsis. Plant Cell 19: 2225—2245. Ellis, C., Karafyllidis., I., Turner, J.G. (2002). Constitutive activation of jasmonate signaling in an Arabidopsis mutant correlates with enhanced resistance to Erysiphe cichoracearum, Pseudomonas syringae, and Myzus persicae. Mol Plant Microbe Interact. 15: 1025-1030. Ferrari, S., Plotnikova, J.M., De Lorenzo, G., Ausubel, EM. (2003). Arabidopsis local resistance to Botrytis cinerea involves salicylic acid and camalexin and requires EDS4 and PADZ, but not SID2, EDS5 or PAD4. Plant J. 35: 193—205. Feys, B., Benedetti, C.E., Penfold, C.N., Turner, J.G. (1994). Arabidopsis mutants selected for resistance to the phytotoxin coronatine are male sterile, insensitive to methyl jasmonate, and resistant to a bacterial pathogen. Plant Cell 6: 751—759. Fonseca, S., Chini, A., Hamberg, M., Adie, B., Porzel, A., Kramell, R., Miersch, 0., Wastemack, C., Solano R. (2009). (+)-7-iso-Jasmonoy1-L-isoleucine is the endogenous bioactive jasmonate. Nat Chem Biol. 5: 344-350 . Footitt, S., Dietrich, D., Fait, A., Fernie, A.R., Holdsworth, M.J., Baker, A., Theodoulou, EL. (2007). The COMATOSE ATP-binding cassette transporter is required for full fertility in Arabidopsis. Plant Physiol. 144:1467-1480. Fujimoto, S.Y., Ohta, M., Usui, A., Shinshi, H., Ohme—Takagi, M. (2000). Arabidopsis ethylene-responsive element binding factors act as transcriptional activators or repressors of GCC box-mediated gene expression. Plant Cell. 12: 393-404. Gidda, S.K., Miersch, 0., Levitin, A., Schmidt, J., Wastemack, C., Varin, L., (2003). Biochemical and molecular characterization of a hydroxyjasmonate sulfotransferase from Arabidopsis thaliana. J Biol Chem. 278: 17895-17900. Glazebrook, J. (2005). Contrasting mechanisms of defense against biotrophic and necrotrophic pathogens. Annu. Rev. Phytopathol. 43: 205—227. 38 Glauser, G., Grata, E., Dubugnon, L., Rudaz, 8., Farmer, E.E., Wolfender, J.L. (2008). Spatial and temporal dynamics of jasmonate synthesis and accumulation in Arabidopsis in response to wounding. J Biol Chem. 283: 16400-16407. Goossens, A., Hakkinen, S.T., Laakso, 1., Seppanen-Laakso, T., Biondi, S., De Sutter, V., Lammertyn, F., Nuutila, A.M., Soderlund, H., Zabeau, M., Inze, D., and Oksman-Caldentey, K.M. (2003). A functional genomics approach toward the understanding of secondary metabolism in plant cells. Proc. Natl. Acad. Sci. USA 100: 8595—8600. Grant, M.R., Jones, J.D. (2009). Hormone (dis)harmony moulds plant health and disease. Science. 324: 750-752. Gray, W. M., deI Pozo, J. C., Walker, L., Hobbie, L., Risseeuw, E., Banks, T., Crosby, W. L., Yang, M., Ma, H. and Estelle, M. (1999). Identification of an SCF ubiquitin-ligase complex required for auxin response in Arabidopsis thaliana. Genes Dev. 13: 1678-1691. Guo, A., He, K., Liu, D., Bai, S., Gu, X., Wei, L., Luo, J. (2005). DATF: a database of Arabidopsis transcription factors. Bioinformatics 21: 2568-2569. He, Y., Fukushige, H., Hildebrand, D.F., Gan, S. (2002). Evidence supporting a role of jasmonic acid in Arabidopsis leaf senescence. Plant Physiol. 128: 876-884. Heim, M. A., Jakoby, M., Werber, M., Martin, C., Weisshaar, B. and Bailey, P. C. (2003). The basic helix-loop-helix transcription factor family in plants: a genome- wide study of protein structure and functional diversity. Mol. Biol. Evol. 20: 735- 747. Howe, G.A., Jander, G. (2008). Plant immunity to insect herbivores. Annu Rev Plant Biol. 59: 41-66. Hilpert, B., Bohlmann, H., op den Camp, R.O., Przybyla, D., Miersch, 0., Buchala, A., Apel, K. (2001). Isolation and characterization of signal transduction mutants of Arabidopsis thaliana that constitutively activate the octadecanoid pathway and form necrotic microlesions. Plant J. 26: 435-446. Hopke, J., Donath, J., Blechert, S., Boland, W. (1994). Herbivore-induced volatiles: the emission of acyclic homoterpenes from leaves of Phaseolus lunatus and Zea mays can betriggered by a B-glucosidase and jasmonic acid. FEBS Lett. 352: I46—150. Hyun, Y., Choi, S., Hwang, H.J., Yu, J., Nam, S.J., Ko, J., Park, J.Y., Seo, Y.S., Kim, E.Y., Ryu, S.B., Kim, W.T., Lee, Y.H., Kang, H., Lee, I. (2008). Cooperation 39 and functional diversification of two closely related galactolipase genes for jasmonate biosynthesis. Dev Cell. 14: 183-192. Kepinski, S. and Leyser, O. (2005). The Arabidopsis F-box protein TIRl is an auxin receptor. Nature 435, 446-451. Koch, T., Bandemer, K., Boland, W., (1997). Biosynthesis of cis-jasmone: a pathway for the inactivation and the disposal of the plant stress hormone jasmonic acid to the gas phase? Helvetica Chimica Acta 80: 838—850. Ishiguro, S., Kawai-Oda, A., Ueda, J., Nishida, 1., Okada, K. (2001). The DEFECTIVE IN ANTHER DEHISCIENCE gene encodes a novel phospholipase A1 catalyzing the initial step of jasmonic acid biosynthesis, which synchronizes pollen maturation, anther dehiscence, and flower opening in Arabidopsis. Plant Cell. 13: 2191-2209. Kang JH, Wang L, Girl A, Baldwin IT. (2006) Silencing threonine deaminase and JAR4 in Nicotiana attenuata impairs jasmonic acid-isoleucinemediated defenses against Manduca sexta. Plant Cell 18: 3303—3320. Katsir, L., Chung, H.S., Koo, A.J.K., Howe, G.A. (2008a.) Jasmonate signaling: a conserved mechanism of hormone sensing. Curr. Opin. Plant Biol. 11: 428-435. Katsir L., Schilmiller A.L., Staswick P.E., He S.Y. Howe GA. (2008b). C011 is a critical component of a receptor for jasmonate and the bacterial virulence factor coronatine. Proc. Natl. Acad. Sci. USA 105: 7100-7105. Kepinski, S. and Leyser, O. (2005). The Arabidopsis F-box protein TIRI is an auxin receptor. Nature 435: 446-451. Koo, A.J.K., Chung, H.S, Kobayashi, Y., Howe, G.A. (2006). Identification of a peroxisomal acyl-activating enzyme involved in the biosynthesis of jasmonic acid in Arabidopsis. J Biol Chem 281: 33511—33520. Koo, A.J., Gao, X., Daniel Jones, A., Howe, G.A., (2009). A rapid wound signal activates the systemic synthesis of bioactive jasmonates in Arabidopsis. Plant J. (In press) Koornneef, A., Pieterse, CM. (2008). Cross talk in defense signaling. Plant Physiol. 146: 839-844. Kourtchenko, 0., Andersson, M.X., Hamberg, M., Brunnstrom, A., Gobel, C., McPhaiI, K.L., Gerwick, W.B., Feussner, 1., Ellerstrom, M. (2007). Oxophytodienoic acid containing galactolipids in Arabidopsis: jasmonate signaling dependence. Plant Physiol 145: 1658—1669.. 40 Laudert, D., Schaller, F., Weiler, E.W. (2000). Transgenic Nicotiana tabacum and Arabidopsis thaliana plants overexpressing allene oxide synthase. Planta. 211: 163-165. Lee, D.S., Nioche, P., Hamberg, M., Raman, C.S. (2008). Structural insights into the evolutionary paths of oxylipin biosynthetic enzymes. Nature. 455: 363-368. Li, L., Zhao, Y., McCaig, B.C., Wingerd, B.A., Wang, J., Whalon, M.E., Pichersky, E., Howe, G.A. (2004). The tomato homolog of CORONATINE-INSENSITIVEI is required for the maternal control of seed maturation, jasmonate-signaled defense responses, and glandular trichome development. Plant Cell 16: 126-143. Li, C., Schilmiller, A.L., Liu, C., Lee, G.I., Jayanty, S., Sageman, C., Vrebalov, J., Giovannoni, J.J., Yagi, K., Kobayashi, Y., Howe, G.A. (2005). Role of beta- oxidation in jasmonate biosynthesis and systemic wound signaling in tomato. Plant Cell. 17: 971-986. Lorenzo, 0., Piqueras, R., Sénchez-Serrano, J.J., Solano, R. (2003). ETHYLENE RESPONSE FACTOR] integrates signals from ethylene and jasmonate pathways in plant defense. Plant Cell. 15: 165-178. Lorenzo, O., Chico, J. M., Sanchez-Serrano, J. J. and Solano, R. (2004). JASMONATE-INSENSITIVEI encodes a MYC transcription factor essential to discriminate between different jasmonate-regulated defense responses in Arabidopsis. Plant Cell 16: 1938-1950. Lorenzo, O. and Solano, R. (2005). Molecular players regulating the jasmonate signalling network. Curr. Opin. Plant Biol. 8: 532-540. Mandaokar, A., Thines, B., Shin, B., Lange, B. M., Choi, 6., K00, Y. J., You, Y. J., Choi, Y. D. and Browse, J. (2006). Transcriptional regulators of stamen development in Arabidopsis identified by transcriptional profiling. Plant J. 46: 984-1008. McConn, M., Browse, J. (1996). The critical requirement for linolenic acid is pollen development, not photosynthesis, in an Arabidopsis Mutant. Plant Cell. 8: 403- 416. Melotto, M., Mecey, C., Niu, Y., Chung, H.S., Katsir, L., Yao, J., Zeng, W., Thines, B., Staswick, P.E., Browse, J., Howe, G.A., He, S.Y. (2008a). A critical role of two positively charged amino acids in the Jas motif of Arabidopsis JAZ proteins in mediating coronatine- and jasmonoyl isoleucine-dependent interactions with the C011 F -box protein. Plant J. 55: 979-988. 41 Melotto, M., Underwood, W., He, S.Y. (2008b). Role of stomata in plant innate immunity and foliar bacterial diseases. Annu. Rev. Phytopathol. 46: 101-122. Mengiste T, Chen X, Salmeron J, Dietrich R. (2003). T heBOTRYTIS SUSCEPTIBLEI gene encodes an R2R3MYB transcription factor protein that is required for biotic and abiotic stress responses in Arabidopsis. Plant Cell 15: 2551—2265. Miersch, O., Kramell, R., Parthier, B., Wastemack, C. (1999). Structure-activity relations of substituted, deleted or stereospecifically altered jasmonic acid in gene expression of barley leaves. Phytochemistry 50: 353—361. Mewis, 1., Appel, H.M., Hom, A., Raina, R., Schultz, J.C. (2005). Major signaling pathways modulate Arabidopsis glucosinolate accumulation and response to both phloem-feeding and chewing insects. Plant Physiol. 138: 1149—1162. Moon, J., Parry, G. and Estelle, M. (2004). The ubiquitin-proteasome pathway and plant development. Plant Cell 16: 3181-3195. Moreno, J.E., Tao, Y., Chory, J., Ballaré, CL. (2009). Ecological modulation of plant defense via phytochrome control of jasmonate sensitivity. Proc Natl Acad Sci U S A. 106: 4935-4940. Nishii, A., Takemura, M., Fujita, H., Shikata, M., Yokota, A., Kohchi, T. (2000). Characterization of a novel gene encoding a putative single zinc-finger protein, ZIM, expressed during the reproductive phase in Arabidopsis thaliana. Biosci. Biotechnol. Biochem. 64: 1402-1409. Nomura, K., Melotto, M., He, S.Y., (2005). Suppression of host defense in compatible plant-Pseudomonas syringae interactions. Curr. Opin. Plant Biol. 8: 361-368. Park, J.B., Halitschke, R., Kim, H.B., Baldwin, I.T., Feldmann, K.A., Feyereisen, R. (2002). A knock-out mutation in allene oxide synthase results in male sterility and defective wound signal transduction in Arabidopsis due to a block in jasmonic acid biosynthesis. Plant J. 31: 1-12. Paschold, A., Bonaventure, G., Kant, M.R., Baldwin, LT. (2008). Jasmonate perception regulates jasmonate biosynthesis and JA-Ile metabolism: the case of C01] in Nicotiana attenuata. Plant Cell Physiol. 49: 1165-1175. Rao, M.V., Lee, H., Creelman, R.A., Mullet, J.E., Davis, KR. (2000). Jasmonic acid signaling modulates ozone-induced hypersensitive cell death. Plant Cell 12: 1633—1646. 42 Reddy, A.S.N. (2007). Alternative splicing of pre-messenger RNAs in plants in the genomic era. Annu. Rev. Plant Biol. 58: 267-294. Ren, C., Pan, J., Peng, W., Genschik, P., Hobbie, L., Hellmann, H., Estelle, M., Gao, B., Peng, J., Sun, C. and Xie, D. (2005). Point mutations in Arabidopsis Cullinl reveal its essential role in jasmonate response. Plant J. 42: 514-524. Reymond, P., Weber, H., Damond, M., Farmer, E.E. (2000). Differential gene expression in response to mechanical wounding and insect feeding in Arabidopsis. Plant Cell 12: 707-720. Reymond, P., Bodenhausen, N., Van Poecke, R.M., Krishnamurthy, V., Dicke, M., Farmer, E.E. (2004). A conserved transcript pattern in response to a specialist and a generalist herbivore. Plant Cell 16: 3132—31347. Riafio-Pachon, D.M., Ruzicic, S., Dreyer, 1., Mueller-Roeber, B. (2007). PlnTFDB: an integrative plant transcription factor database. BMC Bioinformatics 8: 42. Ribot, C., Zimmerli, C., Farmer, E.E., Reymond, P., Poirier, Y. (2008). Induction of the Arabidopsis PHOl;H10 gene by 12-oxo-phytodienoic acid but not jasmonic acid via a CORONATINE INSENSITIVEl-dependent pathway. Plant Physiol. 147: 696-706. Robson, F., Costa, M.M., Hepworth, S.R., Vizir, 1., Pifieiro, M., Reeves, P.H., Putterill, J., Coupland, G. (2001). Functional importance of conserved domains in the flowering-time gene CONSTANS demonstrated by analysis of mutant alleles and transgenic plants. Plant J. 28: 619-631. Ruegger, M., Dewey, E., Gray, W.M., Hobbie, L., Turner, J., Estelle, M. (1998). The TIRI protein of Arabidopsis functions in auxin response and is related to human SKPZ and yeast grrlp. Genes Dev. 12: 198-207. Rushton, P.J., Bokowiec, M.T., Laudeman, T.W., Brannock JF, Chen X, Timko MP. (2008). TOBFAC: the database of tobacco transcription factors. BMC Bioinformatics 9: 53. Sanders, P.M., Lee, P.Y., Biesgen, C., Boone, J.D., Beals, T.P., Weiler, E.W., Goldberg, RB. (2000). The arabidopsis DELAYED DEHISCENCEI gene encodes an enzyme in the jasmonic acid synthesis pathway. Plant Cell. 12: 1041- 1061. Saniewski, M. (1988). The effect of methyl jasmonate and benzyladenine of epiphyllous bud formation in Bryophyllum diagremontianum. Bull. Acad. Pol. Sci. Biol. 36: 39—44. 43 Santner, A., Estelle, M. (2009). Recent advances and emerging trends in plant hormone signalling. Nature. 459: 1071-1078. Sasaki, Y., Asamizu, E., Shibata, D., Nakamura, Y., Kaneko, T., Awai, K., Amagai, M., Kuwata, C., Tsugane, T., Masuda, T. et al. (2001). Monitoring of methyl jasmonate-responsive genes in Arabidopsis by cDNA macroarray: selfactivation of jasmonic acid biosynthesis and crosstalk with other phytohorrnone signaling pathways. DNA Res 8: 153—161. Sasaki-Sekimoto, Y., et a1. (2005). Coordinated activation of metabolic pathways for antioxidants and defence compounds by jasmonates and their roles in stress tolerance in Arabidopsis. Plant J. 44: 653-668. Schaller, F., Biesgen, C., Miissig, C., Altmann, T., Weiler, E.W. (2000). 12- Oxophytodienoate reductase 3 (OPR3) is the isoenzyme involved in jasmonate biosynthesis. Planta. 210: 979-984. 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. USA 97: 11655—11660. Schilmiller, A.L., Koo, A.J., Howe, G.A. (2007). Functional diversification of acyl- coenzyme A oxidases in jasmonic acid biosynthesis and action. Plant Physiol. 143: 812-824. Shoji, T., Ogawa, T., Hashimoto, T. (2008). Jasmonate-induced nicotine formation in tobacco is mediated by tobacco C011 and JAZ genes. Plant Cell. Physiol. 49: 1003-1012. Staswick, P.E., Su, W., Howell, S.H. (1992). Methyl jasmonate inhibition of root growth and induction of a leaf protein are decreased in an Arabidopsis thaliana mutant. Proc Natl Acad Sci U S A. 89: 6837-6840. Staswick, P.E., Yuen, G.Y., Lehman, CC. (1998). Jasmonate signaling mutants of Arabidopsis are susceptible to the soil fungus Pythium irregulare. Plant J 15: 747—754. Stintzi, A., Weber, H., Reymond, P., Browse, J., Farmer, E.E. (2001). Plant defense in the absence of jasmonic acid: the role of cyclopentenones. Proc. Natl. Acad. Sci. USA 98: 12837—12842. Stintzi, A., Browse, J. (2000). The Arabidopsis male-sterile mutant, opr3, lacks the 12- oxophytodienoic acid reductase required for jasmonate synthesis. Proc Natl Acad Sci USA. 97: 10625-10630. 44 Staswick, P.E., Tiryaki, I. (2004). The oxylipin signal jasmonic acid is activated by an enzyme that conjugates it to isoleucine in Arabidopsis. Plant Cell 16: 2117—2127. Strayer, C., Oyama, T., Schultz, T.F., Raman, R., Somers, D.E., Mas, P., Panda, S., Kreps, J.A., Kay, S.A. (2000). Cloning of the Arabidopsis clock gene TOCI, an autoregulatory response regulator homolog. Science 289: 768-771. Sembdner, G., Parthier, B. (1993). The biochemistry and the physiological molecular actions of Jasmonates. Annual Review of Plant Physiol. and Molecular Biol. 44: 569-589. Stenzel, 1., Hause, B., Miersch, 0., Kurz, T., Maucher, H., Weichert, H., Ziegler, J., Feussner, 1., Wastemack, C. (2003). Jasmonate biosynthesis and the allene oxide cyclase family of Arabidopsis thaliana. Plant Mol Biol. 51: 895-911. Suza, W.P., Staswick, P.E. (2008). The role of JARI in jasmonoyl-L-isoleucine production in Arabidopsis wound response. Planta 227: 1221-1232. Suzuki, H., Reddy, M.S., Naoumkina, M., Aziz, N., May, G.D.,Huhman, D.V., Sumner, L.W., Blount, J.W., Mendes, P., and Dixon,R.A. (2005). Methyl jasmonate and yeast elicitor induce differentialtranscriptional and metabolic rc- programming in cell suspensioncultures of the model legume Medicago truncatula. Planta 220: 696—707. Swiatek, A., Lenjou, M., Van Bockstaele, D., Inzé, D., Van Onckelen, H. (2002). Differential effect of jasmonic acid and abscisic acid on cell cycle progression in tobacco BY-2 cells. Plant Physiol. 128: 201-211. Taki, N., Sasaki-Sekimoto, Y., Obayashi, T., Kikuta, A., Kobayashi, K., Ainai, T., Yagi, K., Sakurai, N., Suzuki, H., Masuda, T., Takamiya, K., Shibata, D., Kobayashi, Y., Ohta, H. (2005). 12-oxo-phytodienoic acid triggers expression of a distinct set of genes and plays a role in wound-induced gene expression in Arabidopsis. Plant Physiol. 139: 1268-1283. Tan, X., Calderon-Villalobos, L.I., Sharon, M., Zheng, C., Robinson, C.V., Estelle, M., Zheng, N. (2007). Mechanism of auxin perception by the TIRl ubiquitin ligase. Nature 446: 640-645. Theodoulou, F.L., Job, K., Slocombe, S.P., Footitt, S., Holdsworth, M., Baker, A., Larson, T.R., Graham, LA. (2005). Jasmonic acid levels are reduced in COMATOSE ATP-binding cassette transporter mutants. Implications for transport of jasmonate precursors into peroxisomes. Plant Physiol. 137: 83 5-840. 45 Thomma, B.P., Eggermont, K., Tierens, K.F., Broekaert, W.F. (1999). Requirement of functional ethylene-insensitive 2 gene for efficient resistance of Arabidopsis to infection by Botrytis cinerea. Plant Physiol. 121: 1093—1102. Thomma, B.P., Nelissen, 1., Eggermont, K., Broekaert, W.F. (1999). Deficiency in phytoalexin production causes enhanced susceptibility of Arabidopsis thaliana to the fungus Alternaria brassicicola. Plant J. 19: 163—171. Thomma, B.P.H.J., Eggermont, K., Penninckx, I.A.M.A., Mauch-Mani, B., Vogelsang, R. et al. (1998). Separate jasmonate-dependent and salicylate- dependent defense-response pathways in Arabidopsis are essential for resistance to distinct microbial pathogens. Proc. Natl. Acad. Sci. USA 95: 15107—15111. Thines, B., Katsir, L., Melotto, M., Niu, Y., Mandaokar, A., Liu, C., Nomura, K., He, S.Y., Howe, G.A., and Browse, J. (2007). JAZ repressor proteins are targets of the SCFCO” complex during jasmonate signalling. Nature 448: 661-665. Ueda, J., Kato, J. (1980). Isolation and Identification of a Senescence-promoting Substance from Wormwood (Artemisia absinthium L.). Plant Physiol. 66: 246- 249. Uppalapati, S.R., Ayoubi, P., Weng, H., Palmer, D.A., Mitchell, R.E., Jones, W., and Bender, CL. (2005). The phytotoxin coronatine and methyl jasmonate impact multiple phytohormone pathways in tomato. Plant J. 42: 201—217. Vanholme, B., Grunewald, W., Bateman, Kohchi, T., Gheysen, G. (2007) The tify family previously known as ZIM. Trends Plant Sci. 12: 239-244. van Wees, S.C., Chang, H.S., Zhu, T., Glazebrook, J. (2003). Characterization of the early response of Arabidopsis to Alternaria brassicicola infection using expression profiling. Plant Physiol. 132: 606—617. Veronese, P., Chen, X., Bluhm, B., Salmeron, J., Dietrich, R., Mengiste, T. (2004). The BOS loci of Arabidopsis are required for resistance to Botrytis cinerea infection. Plant J. 40: 558—574. Vick, B.A., Zimmerman, D.C. (1984). Biosynthesis of Jasmonic Acid by Several Plant Species. Plant Physiol. 75: 458-461. von Malek, B., van der, Graaff E., Schneitz, K., Keller, B. (2002). The Arabidopsis male-sterile mutant dde2-2 is defective in the ALLENE OXIDE SYNTHASE gene encoding one of the key enzymes of the jasmonic acid biosynthesis pathway. Planta. 216: 187-192. 46 Wastemack, C., Hause, B. (2002). Jasmonates and octadecanoids: signals in plant stress responses and development. Prog Nucleic Acid Res Mol Biol. 72: 165—221. Wastemack, C. (2007). Jasmonates: an update on biosynthesis, signal transduction and action in plant stress response, growth and development. Ann Bot (Lond) 100: 681-697. Wang, L., Allmann, S., Wu, J., Baldwin, LT. (2008). Comparisons of LOX3- and JAR4/6-silenced plants reveal that JA and JA-AA conjugates play different roles in herbivore resistance of Nicotiana attenuata. Plant Physiol. 146: 904-915. Wenkel, S., Turck, F., Singer, K., Gissot, L., Le Gourrierec, J., Samach, A., Coupland, G. (2006). CONSTANS and the CCAAT box binding complex share a functionally important domain and interact to regulate flowering of Arabidopsis. Plant Cell 18: 2971-2984. Xie, D. X., Feys, B. F., James, S., Nieto-Rostro, M. and Turner, J. G. (1998). C011: an Arabidopsis gene required for jasmonate-regulated defense and fertility. Science 280: 1091-1094. Xu, L., Liu, F., Wang, Z., Peng, W., Huang, R., Huang, D., Xie, D. (2001). An Arabidopsis mutant cexl exhibits constant accumulation of jasmonate-regulated AtVSP, Thi2.l and PDFI.2. FEBS Lett. 494: 161-164. Xu, L., Liu, F., Lechner, E., Genschik, P., Crosby, W. L., Ma, H., Peng, W., Huang, D. and Xie, D. (2002). The SCF(COII) ubiquitin-ligase complexes are required for jasmonate response in Arabidopsis. Plant Cell 14: 1919-1935. Yadav, V., Mallappa, C., Gangappa, S. N., Bhatia, S. and Chattopadhyay, S. (2005). A basic helix-loop-helix transcription factor in Arabidopsis, MYC2, acts as a repressor of blue light-mediated photomorphogenic growth. Plant Cell 17: 1953- 1966. Yan, Y., Stolz, S., Chetelat, A., Reymond, P., Pagni, M., Dubugnon, L., and Farmer, E.E. (2007). A downstream mediator in the growth repression limb of the jasmonate pathway. Plant Cell 19: 2470-2483. Yoshida, Y., Sano, R., Wada, T., Takabayashi, J., Okada, K. (2009). Jasmonic acid control of GLABRA3 links inducible defense and trichome patterning in Arabidopsis. Development. 136: 1039-1048. Zarate, S.l., Kempema, L.A., Walling, LL. (2007). Silverleaf whitefly induces salicylic acid defenses and suppresses effectual jasmonic acid defenses. Plant Physiol. 143: 866-875. 47 Zhao, Y., Thilmony, R., Bender, C.L., Schaller, A., He, S.Y., Howe, G.A. (2003). Virulence systems of Pseudomonas syringae pv. tomato promote bacterial speck disease in tomato by targeting the jasmonate signaling pathway. Plant J. 36: 485- 499. Zheng, W., Zhai, 0., Sun, J., Li, C.B., Zhang, L., Li, H., Zhang, X., Li, S., Xu, Y., Jiang, H., Wu, X., Li, C. (2006). Bestatin, an inhibitor of aminopeptidases, provides a chemical genetics approach to dissect jasmonate signaling in Arabidopsis. Plant Physiol. 141: 1400-1413. Zhou, N., Tootle, T.L., Glazebrook, J. (1999). Arabidopsis PAD3, a gene required for camalexin biosynthesis, encodes a putative cytochrome P450 monooxygenase. Plant Cell 11: 2419—2428. Ziegler, J., Stenzel, 1., Hause, B., Maucher, H., Hamberg, M., Grimm, R., Ganal, M., Wasternack, C. (2000). Molecular cloning of allene oxide cyclase. The enzyme establishing the stereochemistry of octadecanoids and jasmonates. J Biol Chem. 275: 19132-19138. 48 CHAPTER 2: REGULATION AND FUNCTION OF ARABIDOPSIS JASMONA TE ZIM-DOMAIN GENES IN RESPONSE TO WOUNDING AND HERBIVORY The work presented in this chapter has been published: Hoo Sun Chung, Abraham J.K. Koo, Xiaoli Gao, Sastry Jayanty, Bryan Thines, A. Daniel Jones, and Gregg A. Howe (2008) Plant Physiology 146, 952-964 Author’s contributions: Abraham J .K K00 and Xiaoli Gao measured J A-Ile levels in wounded Arabidopsis leaves and helped with manuscript review. Sastry Jayanty measured wound-induced accumulation of JA in coil and wild type plants. Hoo Sun Chung conducted the remainder of the research and wrote the manuscript. Bryan Thines was involved in initiation and development of the project. A. Daniel Jones supervised the quantification of JA-Ile. Gregg A. Howe supervised the entire project and was involved in all aspects of the manuscript writing. 49 ABSTRACT Jasmonic acid (JA) and its amino acid conjugate, jasmonoyl-isoleucine (JA-Ile), play important roles in regulating plant defense responses to insect herbivores. Recent studies indicate that JA-Ile promotes the degradation of JAsmonate Zim-domain (JAZ) transcriptional repressors through the activity of the E3 ubiquitin-ligase SCFCO”. Here, we investigated the expression pattern and function of JAZ genes during the interaction of Arabidopsis with the generalist herbivore Spodoptera exigua. Most members of the JAZ gene family were highly expressed in response to S. exigua feeding and mechanical wounding. JAZ transcript levels increased within 5 min of mechanical tissue damage, coincident with a large (~25-fold) rise in JA and JA-Ile levels. However, wound-induced expression of JAZ and other C011-dependent genes was not significantly impaired in the jarI-I mutant that is deficient in the conversion of JA to JA-Ile. Experiments performed with the protein synthesis inhibitor cycloheximide provided evidence that JAZs, M YC2, and genes encoding several JA biosynthetic enzymes are primary response genes whose expression is de-repressed upon C011-dependent turnover of a labile repressor protein(s). We also show that overexpression of a modified form of JAZl (JAZIA3A) that is stable in the presence of jasmonate compromises host resistance to feeding by S. exigua larvae. These findings establish a role for JAZ proteins in the regulation of plant anti-insect defense, and suggest that signals other than (or in addition to) JA-Ile activate COIl- dependent wound responses in Arabidopsis. Our results also indicate that the timing of jasmonate-induced transcription in response to wounding is more rapid than previously realized. 50 INTRODUCTION Jasmonic acid (JA) and its bioactive derivatives, collectively known as jasmonates, control many aspects of plant protection against biotic and abiotic stress. Jasmonates play a central role in regulating immune responses to arthropod herbivores and necrotrophic pathogens, as well as stress responses to UV light and ozone (Devoto and Turner, 2005; Glazebrook, 2005; Schilmiller and Howe, 2005; Gfeller et al., 2006; Wastemack et al., 2006; Balbi and Devoto, 2007; Wastemack, 2007; Howe and Jander, 2008). Jasmonates also exert control over various developmental processes, including pollen maturation, anther dehiscence, embryo maturation, and trichome development (Li et al., 2004; Browse, 2005; Schaller et al., 2005). In general, jasmonates appear to promote defense and reproduction while inhibiting growth-related processes such as photosynthesis and cell division (Devoto and Turner, 2005; Giri et al., 2006; Yan et al., 2007). These contrasting activities of the hormone imply a broader role for the jasmonates in regulating shifts between growth- and defense-oriented metabolism, thereby optimizing plant fitness in rapidly changing environments. CORONATINE-INSENSITIVE 1 (C011) is a Ieucine-rich repeat (LRR)/F-box protein that determines the substrate specificity of the SCF-type E3 ubiquitin ligase SCFCO" (Xie et al., 1998; Xu et al., 2002). The importance of C01] in jasmonate signaling is exemplified by the fact that null mutations at this locus abolish jasmonate responses in diverse plant species (Feys et al., 1994; Li et al., 2004). JAsmonate ZIM- domain (JAZ) proteins are targeted by SCFCO“ for degradation during jasmonate signaling (Chini et al., 2007; Thines et al., 2007). JAZ proteins belong to the larger family of tify proteins that share a conserved TIFYxG sequence within the ZIM motif 51 (Vanholme et al., 2007). A second defining feature of JAZs is the highly conserved Jas motif, which has a SLXZFXZKRX2RX5PY consensus sequence near the C-terminus (Chini et al., 2007; Thines et al., 2007; Yan et al., 2007). Recent studies indicate that JAZ proteins act as repressors of jasmonate—responsive genes. For example, JAZ proteins are degraded in a C011- and 26S proteasome-dependent manner in response to JA treatment. Also, dominant mutations affecting the conserved Jas motif stabilize JAZ proteins against degradation and, as a consequence, reduce the plant’s responsiveness to jasmonate (Chini et al., 2007; Thines et al., 2007; Yan et al., 2007). Current models indicate that, in the presence of low jasmonate levels, JAZ proteins repress the expression of jasmonate- responsive genes by interacting directly with the basic helix-loop-helix (bHLH) transcription factor MYC2 (also known as JINI), which is a positive regulator of jasmonate responses (Lorenzo et al., 2004; Chini et al., 2007). Increased jasmonate levels promote binding of JAZs to C011 and subsequent degradation of JAZ repressors via the ubiquitin/26S proteasome pathway, resulting in de-repression of early response genes. The JAZ-mediated transition between repressed and de-repressed states of gene expression is likely subject to several additional layers of regulation. It is well established, for example, that the expression of JA biosynthetic genes in Arabidopsis and other plants increases in response to jasmonate treatment and wounding (Ryan, 2000; Sasaki et al., 2001; Reymond et al., 2000; Stenzel et al., 2003; Delker et al., 2006; Farmer, 2007; Wastemack, 2007). This observation implies the existence of a positive feedback loop that reinforces or amplifies the plant’s capacity to synthesize jasmonate in response to continuous tissue damage, such as that associated with biotic stress. JAZ genes are also upregulated in response to jasmonate treatment. Because at least some JAZ proteins act 52 as negative regulators, it was suggested that jasmonate-induced JAZ expression constitutes a negative feedback loop in which newly synthesized JAZ repressors dampen the response by inhibiting the activity of transcription factors such as MYC2 (Thines et al., 2007; Chini et al., 2007). This idea is analogous to the explanation for why auxin rapidly induces the expression of Aux/1AA genes, which encode negative regulators of the auxin signaling pathway (Abel et al., 1995; Abel, 2007). Indeed, the emerging picture of jasmonate action is remarkably similar to that of the auxin signaling pathway in which auxin promotes the degradation of the Aux/1AA transcriptional repressors by the E3 ubiquitin-ligase SCFTIRI (Dharmasiri et al., 2005; Kepinski and Leyser, 2005; Tan et al., 2007) Many plant antiherbivore defense responses are activated upon wound-induced accumulation of JA. The initial steps in the octadecanoid pathway for JA synthesis occur in the chloroplast, whereas the latter half of the pathway operates in the peroxisome (Schilmiller and Howe, 2005; Schaller et al., 2005; Wastemack et al., 2006; Wastemack, 2007). Analysis of mutants impaired in peroxisomal B-oxidation enzymes has shown that JA production is strictly required for defense against herbivorous caterpillars and thrips (Schilmiller et al., 2007; Li et al., 2005). It is now clear that metabolism of JA plays a critical role in regulating jasmonate-based defense repsonses. In particular, synthesis of the jasmonoyl-isoleucine (JA-Ile) conjugate by JASMONATE RESISTANT 1 (JARI) (Staswick and Tiryaki, 2004) and related JA-conjugating enzymes is required for plant protection against necrotrophic soil pathogens (Staswick et al., 1998), lepidopteran insects (Kang et al., 2006), and various abiotic stresses as well (Rao et al., 2000). Recent work by Thines et al. (Thines et al., 2007) showed that JA-Ile, but not JA, MeJA, or their 53 chloroplastic C18 precursor 12-oxo—phytodienoic acid (OPDA), stimulates COII binding to JAZ proteins. Collectively, these results support the hypothesis that JA-Ile is a primary jasmonate signal, which we define here as a compound that evokes a physiological response upon binding to a jasmonate receptor. In addition to regulation by exogenous jasmonate (Chini et al., 2007; Thines et al., 2007), JAZ expression is also induced by high salinity and other environmental stress conditions (Jiang and Deyholos, 2006; Vanholme et al., 2007). Transcript profiling experiments in Arabidopsis (Yan et al., 2007) and hybrid peplar (Major and Constabel, 2006) showed that JAZ genes are upregulated in response to wounding and simulated herbivory. Yan et al. (2007) also demonstrated that JASMONA T E—ASSOCIA T EDI (JASI), which is identical to JAZIO, is induced by mechanical wounding in a C011-dependent manner. Moreover, a splice variant of JASI/JAZIO that encodes a C-terminally-truncated protein (JASl.3) acts as a repressor of JA-mediated growth inhibition (Yan et al., 2007). This finding provides new mechanistic insight into jasmonate’s dual role in promoting defense and inhibiting growth. A role for JAZ proteins in promoting plant defense against herbivory, however, remains to be established. Here, we show that mechanical wounding and herbivory increase the expression of 11 of the 12 JAZ genes in Arabidopsis. We employed the protein synthesis inhibitor cycloheximide, jasmonate measurements, and two well-defined jasmonate mutants (coiI- 1 and jar] -1) to study the mechanism by which tissue damage activates the expression of JAZ and other early response genes. Our results support a model in which wound-induced ll FCO jasmonate synthesis triggers SC -mediated degradation of JAZ repressors and subsequent expression of early genes that further regulate the response both positively 54 and negatively. These regulatory circuits have the potential to orchestrate host defenses that are commensurate with the intensity and duration of herbivore attack. We also provide evidence that J AZ proteins play a role in plant defense against insect herbivores. RESULTS Feeding by Lepidopteran Herbivores Induces JAZ Expression The central role of j asmonate signaling in plant resistance to lepidopteran insects led us to investigate whether members of the Arabidopsis JAZ family are differentially regulated in response to feeding by the generalist Spodoptera exigua. S. exigua larvae were allowed to feed on rosette leaves for either 2 or 24 h. Damaged (local) and undamaged (systemic) leaf tissue was harvested for RNA extraction and gel blot analysis with gene-specific probes for each of the 12 members (JAZ1 -— JAZ12) in the Arabidopsis JAZ family (Chini et al., 2007; Thines et al., 2007; Vanholme et al., 2007). Insect feeding resulted in increased expression of all JAZ3, except JAZ11, in damaged leaves (Fig. 2.1A). Various members of the JAZ family were expressed at different levels in herbivore-challenged plants. For example, JAZ1, JAZZ, JAZS, JAZ6, JAZ9, JAZIO and JAZIZ transcripts accumulated to relatively high levels in damaged leaves, whereas JAZ3, JAZ4, JAZ 7, and JAZ8 showed weaker expression. Herbivore-induced expression of JAZ4 was very weak, and detection of these transcripts required prolonged exposure of autoradiographic films. In damaged leaves, transcript levels of most of the inducible JAZs at the 24 h time point were similar to or slightly greater than those at the 2 h time point. Several JAZs (e.g., JAZ1) were also systemically expressed within 2 h of the onset of insect feeding, indicating that both the local response is relatively rapid (i.e., < 2 h). These results 55 demonstrate that feeding by a lepidopteran insect results in major re-programming of JAZ expression, and that different JAZ genes exhibit distinct patterns of herbivore-induced expression. The Jasmonate Pathway Mediates Rapid Induction of JAZ Genes in Response to Mechanical Wounding We next performed RNA blot analyses to determine the JAZ expression pattern in rosette leaves subject to mechanical wounding with a hemostat. Similar to the results obtained with insect feeding, all JAZ mRNAs except JAZ] I accumulated in mechanically damaged leaves (Fig. 2.18). Expression of JAZ], JAZZ, JAZS, JAZ6, JAZ7, JAZ8, and JAZ9 was strongly induced within 30 min of wounding, with mRNA levels declining at later time points. In contrast to these genes, wound-induced accumulation of JAZ3, JAZ4, JAZIO, and JAZ/2 mRNAs was delayed and weaker. Although the overall JAZ expression patterns elicited by mechanical wounding and herbivory by S. exigua were qualitatively similar, some reproducible quantitative differences were also apparent. For example, we reproducibly observed that JAZ 7 and JAZ8 mRNAs accumulated to lower levels (relative to other JAZ transcripts) in insect-damaged leaves compared to mechanically damaged leaves. Because the specific activity of radiolabeled JAZ probes and autoradiographic film exposure times were similar for each JAZ analyzed, this observation suggests that JAZ7 and JAZ8 expression is either enhanced by mechanical damage or suppressed by insect feeding. Plants harboring null mutations in C01] provide a useful tool to determine the contribution of the jasmonate pathway to the expression of wound-responsive genes. 56 Previous studies with the jasmonate-insensitive Arabidopsis coil mutant (Feys et al., 1994) have defined general classes of genes whose wound-induced expression is either strictly dependent on C011, partially dependent on C011, or completely independent of C011 (Titarenko et al., 1997; Reymond et al., 2000; Cruz Castillo et al., 2004; Devoto et al., 2005; Koo et al., 2006). To determine the extent to which COI] regulates the wound- induced expression of JAZ genes, we assessed the expression pattern of selected JAZs in wild-type (WT) and coil plants (Fig. 2.2). MYC2 expression was also analyzed in these experiments because this gene is known to be induced by wounding in a C011-dependent manner (Lorenzo et al., 2004). The results showed that accumulation of all wound- inducible JAZ mRNAs and MYC2 was largely dependent on COIl (Fig. 2.2). Prolonged exposure times of autoradiographic film (data not shown), however, indicated that all JAZs were expressed at low levels in the coil mutant. This experiment also showed that wound-induced accumulation of M YCZ and several JAZ transcripts occurred within 15 min of leaf damage, which prompted us to further investigate the timing of the response. Rapid Activation of JAZ Genes is Correlated with JA and JA-Ile Accumulation To define more precisely the timing of the wound response, we assessed the expression level of various genes at very early time points after wounding. The steady-state level of JAZ], JAZ5, JAZ 7, and MYC2 transcripts increased within 5 min of wounding (Fig. 2.3A), as did the expression of JAZZ, JAZ6, and JAZ9 (data not shown). Quantification of 32P- Iabeled probe intensities on RNA blots showed that the level of JAZ 7 mRNA increased ~ l3-fold during the first 5 min afier wounding. The strong dependence of wound-induced JAZ expression on COIl (Fig. 2.2) indicated that increased expression of these genes is 57 likely triggered by elevated jaSmonate levels. We used liquid chromatography-mass spectrometry (LC-MS) to measure JA and JA-Ile levels at early time points after mechanical damage (Fig. 2.38 and C). The levels of JA and JA-Ile in undamaged leaves were 29.5 :I: 11.2 and 4.5 d: 1.3 pmol/g fresh weight (FW) tissue, respectively. These levels increased by ~25-fold (to 784 :t 99 and 111 d: 4, respectively) within the first 5 min after wounding. At the 30-min time point, JA and JA-Ile levels increased to 4402 d: 499 and 972 i 132 pmol/g FW, respectively. The steady increase in JA and JA-Ile levels during the first 30 min after wounding was tightly correlated with changes in gene expression. In the latter half of the time course, decreased JAZ 7 and MYC2 mRNA levels were correlated with waning J A-Ile accumulation. Wound-induced JAZ Expression Does Not Require JARl To test further the hypothesis that wound-induced, C011-dependent expression of JAZ genes is mediated by JA-Ile, we analyzed the pattern of wound-induced gene expression in the jar] -1 mutant that is impaired in the conversion of JA to JA-Ile (Staswick et al., 2002; Staswick and Tiryaki, 2004). As shown in Fig. 2.4, the level of JAZ5, JAZ7, and MYC2 transcripts in wounded jarI-l plants was comparable to that observed in WT plants. Parallel analysis of gene expression in the coil -1 mutant confirmed that the induced expression of these genes is dependent on an intact jasmonate signaling pathway. Similar results were obtained for two JA biosynthesis genes, ALLENE OXIDE SYNTHASE (AOS) and 12-0PDA REDUCTASE3 (OPR3), whose wound-induced expression is also C011-dependent (Reymond et al., 2004; Devoto etal., 2005; Koo et al., 58 2006). These findings indicate that JARl activity is not strictly required for wound- induced expression of these jasmonate-responsive genes. JAZ, MYC2, and JA Biosynthetic Genes are Primary Response Genes in the J asmonate Signaling Pathway The current model of jasmonate signaling indicates that JAZ genes are transcribed by MYC2 following jasmonate-induced degradation of one or more J AZ repressors (Chini et al., 2007; Thines et al., 2007). This model implies that JAZs are primary response genes in the jasmonate signaling pathway, which is consistent with their rapid induction following mechanical wounding (Figs. 2.2 and 2.3A). To test directly whether JAZs are primary response genes, we used the protein synthesis inhibitor cycloheximide (CHX) to determine whether jasmonate-induced expression of JAZs and MYC2 requires de novo protein synthesis. Treatment of liquid-grown seedlings with MeJA induced the expression of JAZs and MYC2, as expected (Fig. 2.5A). CHX treatment resulted in the accumulation of MYC2, JAZ], JAZIO and all other JAZ transcripts except JAZII (Fig. 2.5A and data not shown). Induction of JAZs and M YC2 by MeJA was not inhibited by CHX. Rather, seedlings treated with both MeJA and CHX accumulated higher levels of these mRNAs than seedlings treated with either compound alone (Fig. 2.5A). These results support the notion that JAZs and MYC2 are primary response genes. VSPI and LOXZ were used as markers for secondary response genes. In agreement with previous reports (Rojo et al., 1998; Jensen et al., 2002), we found that MeJA-induced expression of VSPI and LOXZ was blocked by CHX. The conclusion that JAZ/MYC2 and VSPI/LOXZ are primary and secondary response genes, respectively, is supported by 59 differences in their temporal expression patterns: JAZ and M YC2 transcript levels peaked early (e.g., 0.5 h) after hormone treatment, whereas VSPI and LOXZ expression was delayed and more gradual. We used the Expression Angler data-mining tool (Toufighi et al., 2005) to identify genes that are co-regulated with JAZs. Among the genes that were consistently identified as being co-expressed with JAZs and MYC2 in both hormone and pathogen data sets were several JA biosynthetic genes, including LIPOXYGENASE3 (LOX3), LOX4, ALLENE OXIDE SYNTHASE (AOS), ALLENE OXIDE CYCLASE3 (AOC3), 12- OPDA REDUCTASE3 (OPR3), and OPC-8:0 CoA LIGASE (OPCLI) (Table SI). LOXZ was not identified in this list of co-regulated genes. We therefore hypothesized that, like JAZ and MYC2, co-regulated JA biosynthesis genes are primary response genes in the jasmonate signaling pathway. To test this idea, we compared the effects of MeJA and CHX treatments on the expression of IA biosynthesis genes to those of JAZ and MYC2. As shown in Fig. 2.5A, the MeJA- and CHX-induced expression patterns of LOX3, LOX4, AOS, AOC3, OPR3, and OPCLI were very similar to those of JAZ], JAZIO, and MYC2. Specifically, the MeJA-induced expression of these JA biosynthesis genes was not inhibited by CHX, and the effects of MeJA and CHX were additive. Moreover, the timing of MeJA-induced expression of these JA biosynthesis genes, iwith the exception of AOC3, was similar to that of M YC2 and JAZI/JAZIO. The JAZ repressor model predicts that CHX-induced expression of primary response genes results from cellular depletion of one or more JAZ repressors. Because CHX blocks de novo synthesis of JAZ proteins, the ability of CHX alone to activate primary response genes (Fig. 2.5A) suggests that JAZ repressors are highly unstable in 60 WT seedlings, even in the absence of exogenous jasmonate. To test the hypothesis that SCFCO” contributes to JAZ turnover in the absence of exogenous MeJA, we determined the expression pattern of JAZ, MYC2, and JA biosynthesis genes in WT and coil seedlings treated with either CHX or a mock control (Fig. 2.58). CHX-induced accumulation of primary gene transcripts was severely attenuated in coil compared to WT seedlings. Interestingly, the accumulated level of JAZ] and M YCZ mRNAs in CHX- treated coil plants was much greater than that of other genes tested (JAZ 7, JAZIO, AOS, and OPR3). CHX-induced expression of JAZZ, JAZ5, and JAZ9 was also strongly suppressed in coil plants (data not shown). These results are consistent with the idea that COII promotes the turnover of JAZ repressors even in the absence of exogenous j asmonate. Wound-Induced JA Accumulation is Dependent on C01] The finding that CHX-induced expression of A OS and OPR3 is dependent on COII (Fig. 2.5B) is consistent with other studies showing that wound- and jasmonate-induced expression of these genes requires COII (Reymond et al., 2000; Cruz Castillo et al., 2004; Devoto et al., 2005; Koo et al., 2006). To directly test the idea that CO“ activity promotes JA production in response to wounding, we used gas chromatography-mass spectrometry (GC-MS) to measure JA levels in unwounded (control) and mechanically damaged leaves of WT and coil plants (Fig. 2.6). The basal level of JA in unwounded WT and coil plants was not significantly different (0.20 i 0.07 and 0.19 d: 0.09 nmol/g F W tissue, respectively). The JA content in WT plants increased rapidly after wounding, with peak levels (6.94 3: 0.42 nmol JA/g FW) attained 1 h after treatment. In comparison 61 to this robust response, wounded coil leaves were severely deficient in JA accumulation. Mechanical wounding increased the IA content in WT and coil leaves by approximately 35-fold and 4-fold, respectively, at the I h time point. The amount of JA in coil leaves at all time points after wounding ranged between 9 and 14% of WT levels. These results demonstrate that COII activity plays an important role in promoting the accumulation of JA in wounded Arabidopsis leaves. Disruption of Jasmonate Signaling by a Truncated Form of JAZI Compromises Resistance to S. exigua Feeding JAZ proteins that lack the C-terminal J as motif reduce the plant’s sensitivity to jasmonate and, as a consequence, cause several jasmonate-related phenotypes (Chini et al., 2007, Thines et al., 2007, Yan et al., 2007). To test whether such truncated JAZ derivatives alter host resistance to herbivory, we compared the defense response of S. exigua- challenged WT plants to that of a previously characterized transgenic line (Thines et al., 2007) expressing a Jas-motif-deleted form (JAZIA3A) of JAZl. As shown in Fig. 2.7A and B, larvae reared on JAZIA3A plants gained significantly more weight than larvae grown on WT plants (Student’s t test, P < 0.0001). Thus, perturbation of jasmonate signaling by overexpression of JAZIA3A decreases host resistance to S. exigua feeding. RNA blot analysis was used to determine the effect of JAZIA3A on the expression of various wound-response genes. In WT plants subjected to insect feeding for 13 d, M YC2, JAZ5, OPR3, and VSPl transcripts were highly elevated in comparison to untreated control plants (Fig. 2.7C). Herbivore-induced levels of MYC2, JAZ5, and OPR3 mRNAs in JAZIA3A plants were significantly less than those in the WT. The expression level of 62 VSPI in insect-damaged JAZIA3A plants, however, was similar to that in WT plants (Fig. 2.7C). These findings indicate that decreased resistance of JAZIA3A plants to S. exigua feeding is correlated with reduced expression of some, but not all, jasmonate responsive genes. DISCUSSION Wound-Induced Expression of JAZ Genes in Arabidopsis The recent discovery of J AZ proteins as negative regulators of jasmonate signaling marks an important advance in our mechanistic understanding of how plants respond to biotic stress through changes in growth- and defense-related processes (Chini et al., 2007; Thines et al., 2007; Yan et al., 2007). Given the central role of jasmonates in the control of induced resistance to insect attack, we initiated the present study with the goal of characterizing the expression pattern of the Arabidopsis JAZ family in response to mechanical wounding and herbivory. With the exception of JAZII, all JAZs exhibited increased expression in response to both mechanical wounding and feeding by S. exigua larvae. The various wound-responsive JAZs showed differences in the timing and amplitude of expression. Most JAZs (e.g., JAZI) were expressed strongly and rapidly (i.e., <0.5 h) in response to mechanical wounding. Induced expression of other JAZs, including JAZ3, JAZ4, JAZIO, and JAZIZ, was temporally delayed and relatively weak by comparison. Our results are in good agreement with previous studies showing that most JAZ genes are rapidly induced by jasmonate treatment (Chini et al., 2007; Thines et al., 2007), and that some Arabidopsis JAZs are wound responsive (Yan et al., 2007). Wound- induced expression of JAZ genes in poplar (Major and Constabel, 2006) and tomato (L. 63 Katsir and G.A. Howe, unpublished results) has also been observed, indicating that this phenomenon is conserved in the plant kingdom. All JAZ genes induced by mechanical wounding were also induced by S. exigua feeding. This finding is consistent with studies showing that mechanical tissue damage and herbivory (or simulated herbivory) elicit similar, although not identical, changes in gene expression (Reymond et al., 2000; Major and Constabel, 2006; Mithofer et al., 2005; Ralph et al., 2006). We cannot exclude the possibility that mechanical wounding and herbivory elicit quantitative differences in JAZ expression. It is interesting to note, for example, that JAZ 7 and JAZ8 mRNAs accumulated to lower levels in insect-damaged leaves compared to mechanically damaged leaves, which suggests that JAZ7 and JA28 expression may be suppressed by insect feeding. Previous studies have provided evidence for compounds in insect oral secretions that suppress the expression of host plant defenses (Schittko et al., 2001; Musser et al., 2005). The physiological significance of wound-induced JAZ expression remains to be determined. Based on the function of JAZ proteins as repressors of jasmonate-responsive genes, however, it was suggested that rapid synthesis of new JAZ proteins serves to attenuate the transcriptional response soon after it is initiated (Thines et al., 2007). In the context of plant defense responses to herbivory, wound-induced production of JAZ proteins may provide a mechanism to restrain the expression of energetically demanding defensive processes. Such restraint may be particularly important when jasmonate levels decline, for example upon cessation of insect feeding. This putative mechanism of negative feedback control suggests that jasmonate-based defenses operate more as a dynamic continuum than as discrete induced and uninduced states. 64 Rapid Wound-Induced Expression of JAZs is Mediated by the J asmonate Pathway We found that wound-induced expression of JAZs and MYC2 occurs within the first 5 min after tissue damage. Moreover, the accumulation of these transcripts was tightly correlated with large (~25-fold) increases in the level of JA and JA-Ile. Based on the magnitude (i.e., >10-fold induction) of the increase in both hormone and transcript (e.g., JAZ7) levels at this early time point, the actual time between wound trauma and transcriptional activation is likely much shorter than 5 min. The timing of this wound response is thus comparable to other rapidly induced hormonal effects, including auxin- induced expression of AUX/1AA genes (Abel et al., 1995). An important physiological function of both JAZ and Aux/1AA repressor proteins is their ability to promote rapid changes in gene expression by inhibiting the activity of transcription factors that are poised to act immediately upon hormone-induced removal of the repressor. Degradation of JAZ repressors following perception of a wound-induced jasmonate signal would enable the host plant to respond rapidly to insect attack. The dependence of wound-induced JAZ expression on COIl (Fig. 2.2; Yan et al., 2007) indicates that a primary jasmonate signal(s) produced in wounded tissue triggers SCFCOH/26S proteasome-mediated destruction of JAZ repressors and subsequent transcription of early response genes. The correlation between gene expression and accumulation of JA and JA-Ile in damaged leaves (Fig. 2.3) suggests that JA and/or JA- Ile could function as a primary wound signal. The ability of JA-Ile, but not JA/MeJA, to promote COIl interaction with JAZI argues in favor of JA-Ile as the active signal, as does the established role of this conjugate in plant responses to biotic stress (Staswick et al., 1998; Kang et al., 2006). Surprisingly, however, wound-induced expression of C011- 65 dependent genes in the JA-Ile-deficient jarl -1 mutant was not significantly impaired (Fig. 2.4). One interpretation of this result is that JA is indeed inactive and that the jarl-I mutant produces a sufficient amount of JA-Ile to promote the expression of wound responsive genes. It was recently shown that jarI-l reduces JA-Ile accumulation in wounded leaves to approximately 10% of WT levels and that the pool of JA-Ile in jar] -1 plants results from the activity of another JA-conjugating enzyme (P. Staswick, personal communication). Thus, JARI is not strictly required for wound-induced expression of C011-dependent genes. An alternative explanation for our results is that the primary jasmonate signal is J A or a J A derivative whose synthesis does not depend on JARl. This hypothesis is supported by recent work indicating that JA or one of its precursors or metabolites complements the function of JA-Ile in promoting defense responses in N. attenuata (Wang et al., 2008). The hypothesis that JA is active per se as a signal predicts the existence of one or more JAZ proteins whose interaction with C011 is promoted by JA. It will thus be important to determine the jasmonate specificity of the complete repertoire of JAZ proteins in plants such as Arabidopsis that have a well-defined JAZ family. Positive Feedback Regulation of JA Biosynthesis is a Primary Response of Jasmonate Signaling Hormone-induced changes in physiology typically involve the expression of primary response genes that, in turn, control secondary transcriptional responses. The protein synthesis inhibitor CHX provides a useful reagent to identify primary and secondary response genes in the jasmonate signaling pathway (van der Pits and Memelink, 2001; 66 Pauw and Memelink, 2005). The ability of CHX to block MeJA-induced expression of LOX2 and VSPI indicates that these genes are secondary response genes, in agreement with previous studies (Rojo et al., 1998; Jensen et al., 2002). In contrast to LOX2 and VSPI, the insensitivity of MeJA-induced MYC2 and JAZ expression to CHX indicates that these genes can be classified as primary response genes. This interpretation is consistent with the ability of MYC2 to recognize the G-box motif found in the promoter of JAZ genes, and the proposed direct inhibitory action of JAZ3 on MYC2 (Chini et al., 2007). There is also evidence to indicate that MYC2 binds to a G-box motif in the MYC2 promoter, thereby regulating its own transcription (Dombrecht et al., 2007). We thus suggest a scenario in which CHX-induced turnover of JAZ repressors releases JAZ- mediated inhibition on MYC2, which would then transcribe JAZ, MYC2 and other target genes. Our results differ from those of Dombrecht et al. (2007), who reported that MYC2 is a secondary response gene. These workers also reported that the expression of VSPl, although a secondary response gene, is induced by CHX, whereas our results and those of Rojo et al. (1998) indicate that VSPI is not induced by CHX. These discrepancies may reflect differences in the methodology of the plant treatments or transcript quantification. Several studies have shown that Arabidopsis genes encoding JA biosynthetic enzymes are upregulated via the jasmonate/C011 pathway in response to wounding and jasmonate treatment (Sasaki et al., 2001; Reymond et al., 2000; Stenzel et al., 2003; Devoto and Turner, 2005; Koo et al., 2006). The generally accepted view of this regulatory phenomenon is that it provides a positive feedback mechanism to reinforce or amplify the plant’s capacity to synthesize J A in response to long-term environmental (e.g., herbivory) or developmental cues (Stenzel et al., 2003; Farmer, 2007; Wastemack, 2007). 67 Although the sensitivity of MeJA-induced LOX2 expression to CHX suggests that this feedback mechanism is a secondary response, our results indicate that many other known or putative JA biosynthetic genes are primary targets of jasmonate signaling. First, we observed that AOS, AOC3, OPR3, OPCLI, LOX3, and LOX4 (but not LOX2) are tightly co-regulated with MYC2 and JAZs (Table S1). Second, these biosynthetic genes were induced by CHX treatment, and superinduced in response to treatment with both MeJA and CHX. Finally, CHX-induced expression of AOS and OPR3 was largely dependent on COIl. We thus conclude that JA biosynthetic genes, like JAZ genes, are negatively regulated by one or more labile proteins whose turnover is dependent on COIl activity. JAZ proteins are obvious candidates for such repressors. Among the five LOXs in Arabidopsis, LOX2 is the only isoforrn known to be involved in IA biosynthesis (Bell et al., 1995). The sequences of LOX3 and LOX4 predict that they are 13-LOXs that, like LOX2, catalyze formation of JA precursors in the plastid (Feussner and Wastemack, 2002; Liavonchanka and Feussner, 2006). The co- expression of LOX3 and LOX4 with other JA biosynthesis genes (Fig. 2.4 and Table S2.l) leads us to speculate that these LOXs may also serve a role in JA synthesis. Rigorous testing of this idea will require analysis of 10x3 and 10x4 mutants. Our observation that coil plants are severely deficient in wound-induced JA accumulation supports the idea that JA biosynthesis is regulated by a positive feedback loop (Farmer, 2007; Wastemack, 2007). Given the jasmonate/COIl-dependent expression of JA biosynthesis genes, one interpretation of this finding is that coil leaves contain limited amounts of one or more JA biosynthetic enzymes. Support for this idea comes from the observation that unwounded leaves of the opr3 mutant contain 68 significantly reduced levels of AOC protein (Stenzel et al., 2003). This scenario is clearly different from WT in plants in which wound-induced JA biosynthesis is limited by substrate availability rather than by the level of octadecanoid pathway enzymes (Stenzel et al., 2003; Wastemack, 2007). It is also possible that the JA deficiency in wounded coil leaves reflects reduced amounts of the initial substrate for JA synthesis, or the increased activity in the mutant of an enzyme that metabolizes JA. The former hypothesis is supported by recent work showing that coil plants are deficient in the accumulation of OPDA- and dinor-OPDA-containing galactolipids that may function as precursors for J A synthesis (Buseman et al., 2006; Kourtchenko et al., 2007). Regulation of Early Response Genes by JAZ Repressors The identification of JAZ proteins as negative regulators that link the action of SCFCOIl to transcription factors such as MYC2 has led to a relatively simple model of jasmonate signaling (Thines et al., 2007; Chini et al., 2007). One prediction of this model is that the jasmonate-insensitive phenotype of coil plants results from the accumulation of JAZ repressors. Our results provide indirect support of this idea. First, wound-responsive JAZ genes exhibit low basal expression in the coil mutant, indicating that JAZ proteins are likely synthesized in the coil mutant. Similar results were obtained for JAZ genes in the C011-deficient jail -1 mutant of tomato (L. Katsir and G.A. Howe, unpublished results). Second, our data showing that CHX-induced accumulation of JAZ transcripts is attenuated in coil seedlings is consistent with the idea that JAZ proteins are destabilized by SCFCOH-mediated ubiquitination (Chini et al., 2007; Thines et al., 2007). Taken together, these findings imply that JAZ proteins are more stable in the absence of 69 SCFCO" ligase activity and, as a consequence, accumulate in coil plants to levels that effectively repress gene expression. This model predicts that JAZ repressors also accumulate in mutants that are deficient in JA synthesis. Measurement of JAZ protein levels in wild-type, coil, and jasmonate synthesis mutants will provide an important test of this hypothesis. It is interesting to note that the coil mutation had a differential effect on CHX- induced expression of various primary response genes. For example, coil nearly abolished CHX-induced accumulation of JAZ7 mRNA, whereas JAZI and MYC2 transcripts persisted to higher levels in CHX-treated coil seedlings. One interpretation of this finding is that different JAZ genes are repressed by different JAZ proteins. For example, rapid accumulation of JAZ 7 transcripts in CHX-treated WT but not coil seedlings suggests that the JAZ repressor of JAZ 7 is relatively stable in the absence of €011. Likewise, the putative JAZ repressor of JAZI and MYC2 would appear to be less stable in the absence of €011. Chini et al. (2007) demonstrated that MYC2 interacts directly with JAZB. Because MYC2 is implicated in the transcriptional regulation of most JAZ genes (Chini et al., 2007), we speculate that JAZ proteins other than JAZ3 also inhibit MYC2 function. Other interpretations, including differential distribution of positively acting transcription factors at various JAZ promoters, or differences in the stability of JAZ mRNAs, may also explain why coil differentially affects CHX-induced expression of different primary response genes. 70 A Role for JAZ Proteins in Defense Against Insect Herbivores We found that S. exigua larvae reared on JAZIA3A plants gained significantly more weight than larvae reared on WT plants (Fig. 2.7). This result provides evidence that JAZ proteins play an important role in regulating the expression of plant processes that confer resistance to insect herbivores. The increased susceptibility of JAZIA3A plants to S. exigua can most likely be attributed to the fact that this mutant exhibits decreased responsiveness to JA and several other coil -Iike phenotypes, including male sterility (Thines et al., 2007). The reduced accumulation of various wound/jasmonate-responsive transcripts in herbivore-challenged JAZIA3A plants is consistent with this interpretation. Moreover, recent studies have shown that S. exigua larvae perform better on coil than WT plants (Mewis et al., 2005; 2006). JAZIA3A mutants are presumably deficient in defensive compounds that normally act to deter S. exigua feeding on WT plants. Some Arabidopsis VSPs are expressed in a C011-dependent manner and are known to function as anti-insect proteins (Benedetti et al., 1995; Liu et al., 2005). However, because herbivore-treated JAZIA3A plants were not significantly affected in VSPl expression, we do not favor this hypothesis. Mewis and co- workers demonstrated that increased performance of S. exigua on the coil mutant correlates with reduced production of glucosinolates, which have a well-established role in defense against generalist herbivores such as S. exigua (Mewis et al., 2005; 2006). This observation raises the possibility that JAZIA3A plants are defective in glucosinolate- based defenses. Transgenic expression of JAZIA3A or other C-terminally-truncated JAZs may provide a useful approach to further understand the many C011-dependent 71 processes that confer plant protection to insect herbivores and other forms of environmental stress. MATERIALS AND METHODS Plant Material and Growth Conditions ArabidOpsis thaliana ecotype Columbia-0 (Col-0) was used as the wild type (WT) for all experiments. Soil-grown plants were maintained in a growth chamber at 21°C under 16 h light (100 uE m'2 s") and 8 h dark. For growth of seedlings in liquid media, seeds were surface-sterilized with 30% (v/v) commercial bleach for 15 min and washed ten times with sterile water. Approximately 100 seeds were placed in 50 mL Murashige and Skoog (MS) medium in a 125-mL Erlenmeyer flask. The flasks were placed at 4°C for four d in darkness, and then incubated under normal growth conditions (described above) for 12 d prior to treatment. Flasks were rotated on an orbital shaker (150 rpm) for the duration of the experiment. Seeds collected from heterozygous coil -1 plants (Feys et al., 1994) were germinated on MS medium containing 50 11M MeJA to select for jasmonate-insensitive coil-l homozygous plants, which were then transferred either to soil or MS liquid medium for further experiments. Seed for the jar] -1 mutant (Staswick et al., 2002) was obtained from the Arabidopsis Biological Resource Center. The jasmonate-insensitive root growth phenotype of jarI-l plants was verified by germinating seeds on MeJA- containing MS medium (Staswick et al., 2002). A male sterile line of Arabidopsis expressing the 35S-JAZIA3A-GUS transgene (Thines et al., 2007) was propagated by outcrossing to wild-type pollen. F1 progeny containing the transgene were selected on MS medium containing kanamycin (50 ug/mL). 72 Plant Treatments Spodoptera exigua eggs were obtained from Benzon Research (Carlisle, PA) and hatched at 27°C. For the insect feeding experiment shown in Fig. 2.1A, newly hatched larvae were transferred to a petri dish and reared on Arabidopsis leaves for 3 to 4 d. Prior to the feeding experiment, second instar larvae were transferred into a new petri dish and starved for 14 h. Approximately 10 larvae were transferred to fully expanded rosette leaves (2 to 3 larvae per leaf) on 5-week-old plants. Insect-challenged and control unchallenged plants were maintained under continuous light at 26°C. Two h after transfer of larvae to the plants, insect-damaged leaf tissue was harvested for RNA extraction. Approximately 5% of the leaf area (local response) was removed by feeding at this time point. A second set of plants was used to collect tissue for the 24 h time point, at which time 20 to 60% of the leaf area was damaged by herbivory. Undamaged leaves from challenged plants were harvested at both the 2 and 24 h time points to determine the effect of insect feeding on systemic expression of JAZ genes. For the herbivore performance assay shown in Fig. 2.7, newly hatched Spodoptera exigua larvae were transferred to five-week-old wild-type and JAZ 1A3A -GUS plants. Eight larvae were reared on each of 48 wild-type and 48 transgenic plants. Plants were maintained under standard grth conditions (see above). The weight of individual larvae was determined 9 d after the start of the feeding trial. Larvae were returned to the same set of plants and, after 4 additional days of feeding, were weighed again. For mechanical wound treatments, fully expanded rosette leaves on 5-week-old plants were wounded three times by crushing the leaf across the mid-rib with a hemostat. This wounding protocol, which resulted in damage to ~40% of the leaf area, was administered 73 to ~six rosette leaves per plant. At various times after wounding, damaged leaves were harvested, immediately frozen in liquid nitrogen, and stored at -80°C until use for RNA extraction. Stock solutions (100 mM) of MeJA and CHX (Sigma) in dimethyl sulfoxide (DMSO) were added to liquid cultures of Arabidopsis seedlings (see above) to a final concentration of 50 M. To determine the effect of CHX treatment on MeJA—induced gene expression, liquid-grown seedlings were pretreated with 50 uM CHX for 1.5 h prior to the addition of MeJA. Seedlings were treated with 0.2% (v/v) DMSO as a mock control. At various times after treatment, seedlings were harvested, frozen in liquid nitrogen, and stored at -80°C until needed for RNA extraction. Quantification of Jasmonate Levels Leaf extracts were prepared essentially as described by Wang et al. (2007), with minor modifications. Briefly, 400 to 500 mg of leaf tissue was frozen in liquid N2 and ground to a fine powder with a mortar and pestle. Dihydro-JA and l3C-JA-Ile were added as internal standards for quantification of JA and JA-Ile, respectively. Following addition of 2.5 mL ethyl acetate, homogenates were mixed and centrifuged at 12,000g for 10 min at 4°C. The supernatant was transferred to a new glass tube and the pellet was re-extracted with 1 mL ethyl acetate. The combined extracts were evaporated at 55°C under a stream of N2 gas. The remaining residue was dissolved in 0.3 mL 70% methanol/water (v/v) and filtered through a 0.2-uM PTFE membrane (Millipore, Bedford, MA). Compounds in the resulting extract (5 uL sample per injection) were separated on a UPLC BEH C18 column (1.7 uM, 2.1 X 50 mm) attached to an Acquity Ultra Performance Liquid 74 Chromatography (UPLC) system (Waters Corporation, Milford, MA). A gradient of 0.15% aqueous formic acid (solvent A) and methanol (solvent B) was applied in a 3-min program with a mobile phase flow rate of 0.4 mL/min. The column, which was maintained at 50°C, was interfaced to a Quattro Premier XE tandem quadrupole mass spectrometer (Waters Corporation, Milford, MA) equipped with electrospray ionization (negative mode). Transitions from deprotonated molecules to characteristic product ions were monitored for JA (m/z 209 > 59), dihydroJA (m/z 211>59), JA-Ile (m/z 322>l30), and l3C6-JA-Ile (m/z 328>136) using a 20 V collision cell potential for each ion. Peak areas were integrated, and the analytes were quantified based on standard curves generated by comparing analyte responses to the corresponding internal standard. Details regarding the performance of this method will be described elsewhere. Because this method does not distinguish J A-Ile from JA-Leu, values reported for J A-Ile represent the sum of JA-Ile plus JA-Leu (Wang et al., 2007). The level of JA-Leu in Arabidopsis seedlings is reported to be <25% of JA-Ile levels (Staswick et al., 2004). l3C-JA-Ile was synthesized by conjugation of cis-(i)-JA (Sigma, St Louis, MO, USA) to ['3c,]-L- isoleucine (Cambridge Isotope Laboratories, Andover, MA, USA) as previously described (Kramell et al., 1988; Staswick et al., 2004). For the experiment shown in Fig. 2.5, total JA was extracted from 200 to 300 mg of leaf tissue using a vapor phase extraction method (Schmelz et al., 2004) and quantified by GC-MS as previously described (Li et al., 2005). 75 RNA Gel-Blot Analysis Primers used to amplify cDNA probes are described in Table S2. cDNAs were obtained by reverse transcription (RT)-PCR of RNA isolated from wounded Arabidopsis (Col-O) leaves. Amplified cDNA fragments were cloned into vector pGEM-T Easy (Promega) and verified by DNA sequencing. These clones were used as templates for PCR reactions with gene-specific primers (Table S2) to generate cDNA fragments that were used as probes in RNA blot hybridization experiments. The nucleotide identity between all pair- wise combinations of the 12 JAZ cDNAs ranged between 11 and 66%. The percent nucleotide identity between the most closely related pairs of JAZ genes is: JAZI and JAZZ, 66%; JAZ5 and JAZ6, 62%; and JAZ7 and JAZ8, 60%. Thus, under the high stringency conditions used for hybridization experiments, full-length cDNA probes were assumed to be gene specific. RNA extraction and gel-blot analyses were performed as described previously (Li et al., 2002). Probed RNA blots were visualized with a phosphorimager and the signal intensities quantified with the Quantity One-4.2.2 program (Bio Rad). Values for each time point were normalized to the AC T8 loading control. 76 Figure 2.1. Expression of JAZ genes in response to herbivore feeding and mechanical wounding. (A) Five-week-oldwild-type plants were challenged with S. exigua larvae. At the indicated times (h) after feeding, damaged local (L) leaves and undamaged systemic (S) leaves were harvested for RNA extraction. A separate set of unchallenged plants was used as a control (C). Five micrograms of total RNAwas loaded in each lane and blots were hybridized with the indicated cDNA probes. A CTIN8 (ACTS) was used as a loading control. JAZ4- and JAZII-probed blots were exposed to autoradiographic film for 16 h, whereas all other blots were exposed for 6 h. The contrast of JAZ4-probed blots was adjusted to facilitate visualization of the JAZ4 signal. (B) Five-week-old wild-type plants were wounded three times across the midrib with a hemostat and damaged leaves were collected for RNA extraction at the indicated times (h) after wounding. Ten micrograms of total RNA was loaded in each lane and blots were hybridized to gene-specific probes for each of the 12 JAZ genes, as well as AC T 8 as a loading control. JAZ4-, JAle-, and AC T 8-probed blots were exposed to autoradiographic film for 16 h, whereas all other blots were exposed for 5 h. 77 JAZ1 JAZ2 JAZ3 JAZ4 JAZS JAZG JAZ7 JA28 JAZQ JAZ10 JAZ11 JAZ12 ACTB Time after feeding 211 2411 C L S C L S rag-am... a ”uflfi'wfiw 78 h: JAZ1 JAZZ JAZ3 JAZ4 JAZ5 JAZG JAZ7 JAZ8 JAZ9 JAZ10 JAZ11 JAZ12 ACTB Time after wounding 00.513 612 prim fig“ W we WT COI1-1 min: 0 15 3O 60 0 15 30 60 Q'W JAZ1 JAZZ JAZ3 JAZS JAZ6 JAZ7 JAZ8 3'4. u» *3 M JAZQ JAZ10 M YCZ ACT8 ..... "r" .. w" 1...... “- ---' w Figure 2.2. Effect of the coil-I mutation on wound-induced expressionof JAZs. Mechanical wound treatments and RNA gel-blot analysis were performed as described in the legend to Figure 18. Damaged leaves were collected for RNA extraction at the indicated times (min) after wounding. Blots were hybridized to gene-specific probes for each of the indicated JAZ genes, M YCZ, and AC T8 as a loading control. All blots were exposed to autoradiographic film for 8 h. 79 Figure 2.3. Rapid induction of JAZ transcripts and accumulation of JAs in response to mechanical wounding. (A) RNA gel-blot analysis of JAZ expression in wounded leaves. Wound treatments and northem- blot analyses were perforrned as described in the Figure 2.13 legend. Damaged leaves were collected for RNA extraction at the indicated times (min) after wounding. M YCZ- and JAZ-probed blots were exposed to autoradiographic film for 4 and 14 h, respectively. (B) and (C) Time course of JA (B) and JA-Ile (C) accumulation in response to mechanical wounding. Leaf tissue from the same set of plants used for RNA blot analysis (A) was harvested at the indicated time points after wounding for extraction of JAs, as described in “Materials and Methods”. JA and JA-Ile (measured as the total of JA-Ile plus JA-Leu) levels were determined by liquid chromatography-mass spectrometry according to the procedure described in “Materials and Methods.” Each data point represents the mean :tSD of four biological replicates. 80 0 Time after wounding (min) JAZ1 l.___ _____ -_ ~.-.-...-..———. --.- .— . 5 10 15 20 30 60 JA 25 JAZ7 M YC2 ”qua-- JA (pmollg FW) l l l 1 l I 10 20 30 40 50 60 70 Time after wounding (min) 1400 O 1 200 1 000 800 600 400 JA-lle (pmollg FW) 200 l l l 1 l L J L 10 20 30 40 50 60 70 Time after wounding (min) 81 WT coi1-1 jar1-1 MYCZ JAZ5 JAZ7 AOS OPR3 ACT8 Figure 2.4. Wound-induced expression of JA-responslve genes in the jarl-I mutant. Five-week-old wild-type, coil-I, and jarI-I plants were mechanically wounded as described in the legend to Figure 2.1B. Damaged leaves were collected for RNA extraction at the indicated times (h) after mechanical wounding. Blots were hybridized to probes for MYC2, JAZ5, JAZ7, two JA biosynthesis genes (AOS and OPR3), as well as ACT8 as a loading control. All blots except AC T8 were exposed to autoradiographic film for 6 h. The AC T8 blot was exposed for 16 h. 82 a v no.“ Baaaomuflofia 203 $03 =< 96% contempt ma vofiotoa 203 Ema—mam 83-6w <79— 93 EoEuwob XIU .mocow o>_m=o%8-<_. mo 2:88on counci¢nzo no 7:8 mo Locum Amy .: m no.“ Bide—36833 82> $65 55o =a macho—LB E 3 L8 EE 9 8898 203 803 concenékbq USN -mkbq ._obuoo wEUmS x mm vow: ma? EUV .93: 83 mecca c0822: 2: 8 33253 225 $03 EB 28— :08 E 893— mg, _m=on_mo..-<_. :o “5:53.: oEEu—oaeomo no Baum .WN v.5»...— wkU< AWE. . Wm) rill L. I . ,Em> ”I .. ,rmmls Hun . In..." moo cl l:i! m «we a rial.“ a... o w w no a: E x192. .20 2.. 095 < 83 8000 + WT —0— coi1-1 3 1.1. g: T: 40004 E .9: 5., 2000 . 492 fie 0 . . . r . 0 1 2 3 4 Time after wounding (h) Figure 2.6. coil-I plants are deficient in wound-induced accumulation of JA. Rosette leaves on 5-week-old wild-type (black circles) and coil-1 mutant (white circles) plants were mechanically wounded at the distal end with a hemostat. Wounded leaves were harvested for JA extraction at the indicated times (h) after wounding. Unwounded leaf tissue was used as a control for the 0-h time point. Each data point represents the mean iSD of three samples from independent sets of plants. 84 5 +wr 5. -o—JAz1A3A Larval weight (mg) § o . Days of feeding B C WT JAZ1A3A C W C W MYC2 O '- JA25 '- AOS '. ' OPR3 ! :- vsp1 ! ' wr JAZ1A3A Ac" Figure 2.7. JAZIA3A plants are compromised in resistance to feeding by S. exigua. (A) Newly hatched S. exigua larvae were reared on wild-type (WT in the image) and JAZIA3A transgenic plants. Larval weights were measured 9 and 13 d after the start of the feeding trial. Values indicate the mean :l:SE. The number of wild-type-reared larvae at the 9- and l3-d time points was 79 and 73, respectively, whereas the number of JAZIA3A-reared larvae was 87 and 111, respectively. (B) Representative S. exigua larvae recovered wild-type and JAZIA3A plants at the 13-d time point. (C) Expression of various wound-responsive genes in undamaged control (C) and S. exigua- damaged (W) wild-type and JAZIA3A plants. The arrow in C denotes a higher-Mr JAZI transcript that presumably is derived from the JAZIA3A -GUS transgene. RNA was extracted from S. exigua-damaged leaves collected at the 13-d time point, or from a set of undamaged plants grown in parallel. Northem-blot analyses were performed as described in the Figure 2.1 legend. 85 100 ._ JAZ1 64 ~-————— JAZZ ~—————JA25 100 JAZG JAZ1 1 99 JAZ12 JAZ10 JAZQ '“7 JAZ4 98 JA23 JAZ7 100 JAZ8 T Supplemental Figure 2.1. Phylogenetic tree of Arabidopsis JAZ family. The phylogenetic tree of twelve JAZ proteins was constructed with the Neighbor Joining (NJ) method using MEGA 4 software (Molecular Evolutionary Genetics Analysis) available at http://www.megasoftware.net/index.htrnl. Values on each node are percentage of bootstrap values (only values greater than 50 are shown). 86 5 £8 E :3 a: 83 as a; a: 89 33 § 3 «.93 c2 2: 2: 2e 2: 2: 2: «5:. ea 88 a: .33 2: at 5:. 2: E: a: 83 SE. 2: 8: ES 2: as 39° 2.. 2: as no; a?» as 83 5 83 s: E «8... 2: 2: ca 9.8 mg 66 80.0 a: SS 2: ea 2: e2 2 c 83 Rs. :9 and 5 82. 89 3:. as «Ed 2: E: as $8 «CNS. 2: E *8... as :2. at A49 88 a: 8a... 1.2 £5. 39 83 § mm; 2: 89 9o... 2: E: a: :2. Rs. 63 86... a: 83 2: 89 Rod 2: 2: c c «8.: «5% E 23 E «:3 89 ES 89 Ed 3. :9 58 S 88... .NS. 288:< 8889‘ 8595. 885m? 86%an 816.? Summer: a :80 Eco 89. mo... .23 .083 we»: a «om Sea ESE—o: amoegxmaooz 05 5 85» 3655on <~ use NE. 50389 nose—oboe 5.5855 mo baa—ham .< 43qu 9305535 «A. .53 38055.3 one «one» NE. .—.N 03:. 32.059395 87 .mocow 8mmo5xoéo 808 2: 06 waoEa 8828 8: ”E: c .85 boss 05 88:2: homo—o 82: 8:5 868538 08:3 beam a 82865 :7. 50 L85 :8 < .3055 9: so 85> 5.858 a mafia 88w 83655-8 50 a: 2: E 08w 2: so L85 x8 2: 3 858:8 885855 5 £882 .5085 86855.8 18:82 8 88:8 o._ so 83> L =< 8:5 93 8238 85855-8 50 858 05 $5865 AEBoEooo 8:20on :89: 83>; an 9 858:8 033. 05 E 89:52 o .85 E86 N3. 05 5:5 55855-8 muse? ionow 085.585 3. can NOE 9.632: .momow o>fi8=omoe8m a .fiom 88 ...85 E 5085 so: 203 ~ NNVN. Ea “4N3. .wwoesxmcogxx a flow 88 Am: 8.555 5 3s 2555: 05 556 5 88—5860 05 85 85% @382 8 beau a as 8a: 083 £3. 8% a as was as Ea s: as 88 s: s: as 3 8S 2: s: s: s: a: s: E a; «5% s: as 8:. s: s: s: s: 2.. EN: as 88 as was as and as as; s: as as; :s mad 8.: A8 5 83 as 58 as 88 s: as and as 38 s: 8% as as; as 8a... as «.85 s: as Ed as :8 2.. 5% as :8 As was as 8a... 2.. a: 88 as as; s: 8% s: a: 28 as 88 s: a. sad as as; 2.. a8 2: cs 8:. a: as 8:. s: as 8:. as 8:. as“ as :8 a. :8 as 8:. as 3:. s: as 3:. as and 8% as 28 as 88 as 83 s: s: as 88 s: as. 2885 888m? 8882 88.82 8892 8.292 8805. / :80 88 8... «.3 sea axes 83 a Low San— :omofiam mmo5xm80~< 05 E 88w 085585 S. was N3. :838 coca—0:8 558536 90 SE55 .m 855:8 —.N flash. _3=ofio_55=w 88 .n8((P0 JAZlO.3 > JAZlO.1, and that JAZlO.4 is highly resistant to JA-induced degradation. Pretreatment of seedlings with MG132, a specific inhibitor of the 26S proteasome, impaired JA-induced turnover of JAZlO.1-YFP (Fig. 3.78). This finding indicates that JA-induced degradation of this Jas domain-containing isoform is mediated by 26S-proteasome pathway. The Dominant Negative Effect of JAZ10.4 Requires a Functional TIFY Motif We next investigated whether the strong repression of JA responses by JAZIO.4 is dependent on a functional TIFY motif. To address this question, we generated transgenic lines that overproduce JAZlO.4I'>A and JAZIO.4G'>A proteins harboring 1107A or G11 1A point mutations, respectively, in the TIFY motif of JAZIO.4. Although these mutations abrogate JAZlO.1 interaction with other JAZs in the Y2H system (Fig. 3.2F), it was necessary to determine the effect of these mutations on JAZlO.4-JAZ interactions. As shown in Figure 3.8A, JAZIO.4">A lost the ability to interact with all four JAZ partners 112 tested, including JAZ6, JAZlO.1, JAZlO.3, and JAZIO.4. JAZIO.4G’>A also failed to bind JAZIO.4 but did retain ability to interact with JAZ6, JAZlO.1, and JAZlO.3. Western blot analysis of protein extracts from yeast strains co-transformed with JAZlO.1 and mutant forms of JAZIO.4 showed that the JAZ fusion proteins were expressed (Fig. 3.8B). These findings, together with results of assays to test JAZIO.4G'>A and JAZIO.4">A interaction with other members of the JAZ family (Fig. 53.4), indicate that 1107A is generally more effective than G11 1A at disrupting JAZIO.4 heterodimerization with other JAZs. Having established the effect of 1107A and G1 1 1A on JAZlO.4-JAZ interactions in yeast, we assessed JA-mediated fertility phenotypes in transgenic plants that overexpress JAZIO.4I'>A and JAZlO.4G'>A. Of 150 independent 35S-JAZIO.4">A plants (Tt generation) examined, none showed defects in filament elongation, anther dehiscence, silique development, or viable seed production (Fig. 3.8C and D), indicating that I107A completely suppresses JAZlO.4-mediated male sterility. In a population of 40 independent Tt 35S-JAZ10.4G'>A plants, 12 individuals exhibited reduced fertility; these plants produced only a few mature siliques during the latter stage of reproductive development (Fig. 3.8C). Some flowers produced by this group of semi-fertile 35S- JAZ10.4G'>A plants were indistinguishable from wild-type flowers (data not shown), whereas other 35S-JAZ10. 40'” flowers had anther filaments that failed to fully elongate (Fig. 3.8E and F). We also studied the effect of TIFY point mutations (1107A and GI 1 1A) on the JA-resistant root growth phenotype that results from JAZIO.4 overexpression (Fig. 3.6E and F). In agreement with the ability of 1107A to strongly suppress JAZlO.4-medaited sterility, 35S-JAZ10.4">A roots were as sensitive to JA-induced growth arrest as wild-type 113 roots (Fig. 3.8G and H). 35S-JAZ10.4G'>A roots, by contrast, were less sensitive to JA than wild-type roots, but significantly more sensitive (P < 0.0001; Student’s t-test) than 35S-JAZ1 0.4 roots. These results show that the ability of JAZIO.4 to repress JA signaling in various tissues is strongly suppressed by 1107A and moderately suppressed by G11 1A. l->.4 To test the possibility that normal JA responsiveness of 35S-JAZIO.4 plants results from low expression of the JAZIO.4I'>A mutant protein in planta, we compared the level of B-glucuronidase activity in stable transgenic lines that constitutively express either JAZlO.4-GUS or JAZlO.4I'>A-GUS fusion proteins. Seedlings expressing the 35S- JAZ10.4">’4-GUS transgene exhibited a pattern of strong blue staining (in roots, cotyledons, and leaves) that was very similar to that of 35S-JAZIO.4-GUS seedlings (Fig. 3.9A). 3SS-JAZIO.4-GUS plants were male sterile and highly insensitive to JA-induced root growth arrest (Fig. 3.9B and C; Fig. 3.6H), indicating that JAZlO.4-GUS, like JAZIO.4, strongly represses JA signal output. Plants overexpressing JAZIO.4">"-GUS, however, were fully fertile and exhibited normal sensitivity to exogenous JA (Fig. 3.98 and C). We also examined the subcellular localization of JAZlO.4-YFP and JAZlO.4"M3 YFP fusion proteins to determine whether the 1107A mutation affects subcellular protein distribution. Following transient Agrobacterium-mediated expression of these proteins in tobacco leaf epidermal cells, laser seaming confocal microscopy revealed that both JAZlO.4-YFP and JAZIO.4">A-YFP are targeted to the nucleus (Fig. 3.9D). These results show that the 1107A mutation does not affect the stability or subcellular localization of JAZIO.4. 114 DISCUSSION Functional Diversification of JAZ Proteins by Alternative Splicing Alternative splicing plays an important role in increasing protein diversity and, ultimately, biological complexity. Our characterization of functionally distinct JAZIO splice variants adds to a growing body of literature indicating that alternative splicing of pre-mRNAs provides a general mechanism to alter the proteome in ways that optimize plant adaptation to stress (Reddy, 2007; Barbazuk et al., 2008). In addition to the JAZlO.1 and JAZlO.3 isoforrns described by Yan et al. (2007), we identified a third JAZIO splice variant (JAZlO.4) that lacks the entire Jas domain. Several observations indicate that differential association of JAZIO variants with COIl affects the stability and action of each isoform in different ways. We showed that JAZlO.] interacts with C01] in a ligand- dependent manner, whereas JAZIO.4 fails to bind COIl. JAZlO.3, which contains a partially truncated Jas domain, interacted weakly with C01]. Analysis of JAZIO-YFP reporter lines provided direct evidence that JAZlO.] is degraded via the 26S proteasome pathway in response to JA treatment, whereas JAZ10.4 is highly resistant to hormone- induced destruction (Fig. 3.7). Consistent with C011-interaction studies performed in yeast, high concentrations of JA were required to induce turnover of JAZlO.3-YFP in vivo. The differential physical interaction of the three splice variants with C01] can explain the JA-related phenotypes associated with ectopic expression of each isoform. Specifically, overexpression of JAZlO.3 and JAZIO.4 conferred weak and strong JA- insensitive phenotypes, respectively, whereas 35S-JAZ10.1 plants that produce the full- length protein did not exhibit altered sensitivity to the hormone. We conclude that functional differences between JAZIO splice variants can be attributed in large in part to 115 differences in the extent to which these proteins interact with COIl in JA-stimulated cells, which is a direct consequence of splicing-induced sequence changes in the Jas domain. Based on the strong JA-insensitive phenotype of JAZlO.4-overexpressing plants and the inability of JAZIO.4 to interact with COIl, we suggest that the function of this protein in wild-type plants is to attenuate signal output in the presence of bioactive JAs. Previously characterized JAZ proteins appear to repress JA responses in the absence of JA, and are degraded by regulated proteolysis in the presence of the hormone (Thines et al., 2007; Chini et al., 2007). In contrast, JAZIO.4 appears to be a potent negative regulator of signal output in hormone-stimulated cells. This view is consistent with the observation that JAZIO RNAi-silenced plants are hypersensitive to JA (Yan et al., 2007), whereas null mutations in several other JAZ genes do not cause such a phenotype (Thines et al., 2007; Chini et al., 2007). Our results indicate that the increased sensitivity of JAZIO silenced lines to JA likely reflects a deficiency in JAZIO variants (e.g., JAZlO.4) that are stabilized against JA-induced turnover. The ability of JAZIO.4 to restrain signal output may be particularly important for reigning in IA responses that are energetically demanding and potentially damaging to the cell. It is also possible that COIl- noninteracting JAZs attenuate JA-mediated growth inhibition under environmental conditions where rapid growth is required to compete for limited resources (Izaguirre et al., 2006; Yan et al., 2007). A more precise understanding of the normal function of JAZlO splice variants will require additional information about the downstream targets and tissue/cell-type-specific expression of each isoform. Our results are consistent with a scenario in which negative feedback control of signal output by JAZIO.4 and JAZlO.3 is initiated upon transcriptional activation of 116 JAZIO in response to wounding or other inductive cues that trigger JA-Ile synthesis (Yan et al., 2007; Chung et al., 2008). Although it is possible that wounding regulates alternative splicing of JAZ10 pre-mRNA in a manner that favors the accumulation of a specific transcript (Bove et al., 2008), our data and those of Yan et al. (2007) show that all three transcripts accumulate in wounded leaves. Our results also indicate that protein stability is likely to be a major factor in determining the relative abundance of each JAZIO isoform in cells that express JAZ10. JAZ isoforms that fail to interact (JAZIO.4) or weakly interact (JAZlO.3) with COIl are predicted to accumulate to higher levels in JA-stimulated cells than JAZ proteins (e.g., JA210.1) that interact strongly with C01]. Differential accumulation of JAZIO-YFP reporter proteins in JA-treated cells (Fig. 3.7) provides direct support for this idea. The ZIM Domain Mediates Protein-Protein Interaction The ZIM domain is the defining feature of the tify protein family that contains J AZ, PPD, and ZIM/ZML proteins (Vanholme et al., 2007). Notably, however, a biochemical function for this conserved sequence motif has not previously been described. We present several lines of evidence to indicate that the ZIM domain mediates homo- and heteromeric protein-protein interactions among JAZ proteins. We have identified nine Arabidopsis JAZs (including the three JAZIO splice variants) that self-associate in yeast. BiFC experiments showed that JAZ3 homomeric complexes accumulate in the plant nucleus. Our results also revealed an extensive network of heteromeric JAZ interactions. In a Y2H experiment to test all 66 possible heterodimeric combinations between 12 full- length Arabidopsis JAZs (Table 3.1), we identified 38 heteromeric interactions involving 117 most but not all JAZs. Given that protein dimerization is a common regulatory mechanism in signal transduction (Klemm et al., 1998), it is conceivable that combinatorial interactions among various JAZs plays a role in generating diverse JA signal outputs. Additional work is needed to determine the physiological relevance of these interactions, and to assess whether differences in the interaction potential of various JAZs have functional significance. Site-directed mutagenesis studies showed that the TIFY motif is a key determinant of JAZ homo- and heteromerization. Because the ZIM domain contains several highly conserved residues in addition to the TIFY motif (Fig. 3.2D), it is likely that other regions of the domain are also important for protein-protein interaction. Y2H assays with JAZ3 deletion constructs and JAZIO.4 showed that JAZ-JAZ binding does not require the N-terminal region or the C-terminal Jas domain. Although this finding supports the notion that the ZIM domain is the primary determinant for intermolecular contact between JAZ proteins, it remains to be determined whether the ZIM “domain” is a self-stabilizing structure that is sufficient for protein-protein interaction. The ability of ZIM and ZML proteins to homo- and heterodimize in yeast (Fig 83.2) shows that protein- protein interaction is not specific for JAZ proteins, but rather extends to other members of the tify family. This finding raises the possibility that JAZs functionally interact with PPD and ZIM/ZML proteins. Finally, we note that our results extend the functional analogy between JAZ proteins and Aux/1AA proteins that are substrates for the SCFTIR' E3 ligase component of the auxin response pathway. JAZ proteins appear to share several key features with Aux/1AA repressor proteins, including hormone-induced rapid degradation by the 118 SCF/ubiquitin/26S proteasome pathway, interaction with transcription factors in the nucleus, and rapid induction of their corresponding genes in response to increased hormone levels (Thines et al., 2007; Chini et al., 2007; Chico et al., 2008; Katsir et al., 2008b; Chung et al., 2008). A role for the ZIM domain in JAZ-JAZ interactions is reminiscent of the ability of Aux/1AA protein domains III and IV to mediate homo- and heterodimer formation among Aux/1AA proteins (Kim et al., 1997). Domains III and IV are conserved in ARF transcription factors that control expression of auxin response genes, which enables Aux/IAAs to bind to and repress ARF activity (Kim et al., 1997; Ulmasov et al., 1999). The ZIM domain does not appear to be conserved in MYC2, and available experimental evidence indicates that this domain is not required for JAZ-MYC2 partnering (Chini et al., 2007). Nevertheless, the possibility that the ZIM domain mediates JAZ interaction with proteins outside the tify family cannot be excluded. A Role for the ZIM Domain in Regulating JAZ Function We took advantage of the strong JA-insensitive phenotype of 35S-JAZ1 0.4 plants to study the functional significance of the ZIM domain in JA signaling. The main conclusion of these experiments is that mutations (e.g., 1107A) in the TIFY motif that disrupt JAZlO.4 interaction with other JAZs also abrogate the dominant-negative effects of JAZIO.4 on JA signal output. Importantly, functional analysis of JAZIO.4 TIFY mutants (JAZIO.4">A and JAZlO.4G'>A) showed that the capacity of JAZlO.4 to interact with other JAZs in yeast is tightly correlated with the strength of the JAZlO.4-mediated dominant-negative effect. Because mutation of the TIFY motif did not appear to affect the accumulation or nuclear localization of JAZIO.4, the most straightforward interpretation of these results is 119 that the repressive activity of JAZIO.4 depends on the ability of its ZIM domain to interact with another protein. Our Y2H analyses support the hypothesis that JAZIO.4 interacts in planta either with itself or with another tify protein to repress JA signaling. Interestingly, Ouellet and colleagues (2001) showed that a mutation (in domain III) that impairs the ability of an Aux/1AA protein to homo- and heterodimerize suppresses auxin- related phenotypes conferred by an intragenic mutation that stabilizes the Aux/1AA against auxin-induced degradation. Thus, the repressive activity of stabilized JAZ and Aux/1AA variants depends on their ability to interact with other family members or with downstream transcription factors. Our results raise the possibility that a network of JAZ-JAZ complexes differentially interact with transcription factors to control the specificity and strength of JA signal output. Among the important factors that would determine the composition and stiochiometry of JAZ-JAZ dimers (or higher order oligomers) in any given cell type are (l) the rate of JAZ synthesis, (2) the rate of JAZ degradation, and (3) the interaction capacity of each JAZ expressed in that cell type. Inductive cues that increase the synthesis of bioactive JAs are predicted to remodel J AZ-J AZ partnering by increasing the rate of de novo JAZ synthesis and the selective degradation of unstable JAZs. We suggest that complexes containing stabilized JAZs (e.g., JAZIO.4) will accumulate in JA-stimulated cells and, eventually, attenuate expression of JA response genes. Ectopic expression of JAZlO.4 likely perturbs the normal balance of JAZ proteins by creating a situation in which JAZIO.4 predominates over more labile JAZ isoforms. A simple model to explain the requirement for the TIFY motif in JAZlO.4-mediated repression of signal output is that JAZIO.4 homo- or heteromeric complexes are the 120 functional unit for direct repression of MYC2. The ability of JAZIO.4 to interact with MYC2 (Figure 3.4D) in the Y2H system is consistent with this model, as is a site-directed mutagenesis study indicating that the C011 and MYC2 interaction surfaces of JAZ9 are likely not identical (Melotto et al., 2008). This model may also explain the dominant- negative action of JAZIAJas (Thines et al., 2007) and other JAZAJas proteins that have the capacity to form homo- and heterodimers (Table 3.1). This scenario, however, would not appear to explain the repressive action of JAZ3AJas, which was shown to interact with C01] but not with MYC2 (Chini et al., 2007). These workers proposed a model in which binding of JAZ3AJas to SCFCO“ impairs the turnover of endogenous JAZs that would accumulate and repress MYC2 activity. The necessary biochemical and genetic tools are now available to rigorously test these models of JAZAJas action. Further study of naturally occurring JAZ isoforms that evade degradation by the JA/COIl pathway promises to provide new inSight into the molecular mechanisms underlying the control of J A responses. MATERIALS AND METHODS Plant Material and Growth Conditions Arabidopsis thaliana ecotype Columbia (Col-0) was used as the wild type for all experiments. Soil-grown plants were maintained in a growth chamber at 21°C under 16 h light (100 uE m'2 s") and 8 h dark. Arabidopsis was transformed with Agrobacterium tumefaciens (strain C58C1) using the floral dip method (Clough and Bent, 1998). A list of transgenic plants used in this study is provided in Table S1. Transformed lines (T1 generation) were selected on MS plates containing kanamycin (50 ug/mL) and 121 vancomycin (100 ug/mL). At least 40 independent Tl plants per genotype were transferred to soil for subsequent phenotypic analysis. We identified homozygous lines by testing T3 progeny for resistance to kanamycin. We propagated male sterile 355- JAZ10.4 lines by outcrossing to wild-type pollen. F 1 progeny containing the transgene were selected on Murashige and Skoog (MS) medium plates containing kanamycin (50 ug/mL). The 35S—JAZI-GUS transgenic line was described previously (Thines et al., 2007). Nicotiana tabacum (cv Petit Havana) was grown in a growth chamber maintained under 17 h of light (200 uE/m'z/sec) at 28°C and 7 h of dark at 18°C. Yeast Two-hybrid Assays Yeast two-hybrid (Y2H) assays were performed with the Matchmaker LexA system (Clontech). Full-length JAZ cDNAs were isolated as previously described (Chung et al., 2008). cDNAs encoding JAZIO splice variants were generated from RNA prepared from rosette leaves collected 1.5 h after mechanical wounding. RT-PCR reactions were performed with Pfu Turbo DNA polymerase (Stratagene) and the primer sets listed in Table S3.l. JAZ cDNAs were subcloned into the Matchmaker pGILDA vector to generate fusions of the “bait” protein with the LexA DNA-binding domain (DB). These cDNAs were also subcloned into the pB42AD vector to generate fusions of the “prey” protein with the B42 activation domain (AD). Yeast transformation and selection of transformants were conducted as described in Melotto et al. (2008). JAZ-JAZ interaction assays were performed as follows. A l-mL culture of yeast transformants in SD—glucose medium (BD Biosciences) supplemented with —Ura/-His/-Trp dropout solution was grown to an OD600 of 1.0 - 1.2. Cells were recovered by centrifugation and resuspended 122 in 0.4 mL distilled water. Four uL of the resulting cell suspension was placed on SD- galactose/raffinose (-Ura/-His/-Trp)-inducing media (in 96-well plates) containing 80 ug/mL X-gal. To test the interaction of JAZIO variants with COIl, coronatine (Sigma) was added to the medium before the yeast cells were plated. All photographic images of Y2H plates were taken afier 48 h of incubation of the plate at 30°C. Expression of the AD- and BD-fusion proteins was detected by western blot analysis using anti- hemagglutinin (HA; Covance) and anti-LexA (Upstate) antibodies, respectively. B- galactosidase activity in yeast cells was measured with liquid cultures as described by the manufacturer (Clontech), using ortho-nitrophenyl-B-d-galactopyranoside as a substrate. Site-Directed Mutagenesis A PCR-based protocol (Stratagene) was used to perform Ala-scanning mutagenesis of the TIFY motif in JAZ3, JAZlO.1, and JAZIO.4. PCR reactions were performed with JAZ3, JAZlO.1, and JAZ10.4 cDNAs in pB42AD or pGEM-T Easy vectors, and the mutagenic primers listed in Table $3.1. The presence of the desired mutation was confirmed by DNA sequencing. Subcellular Protein Localization The pBI-eYFP binary vector was prepared by subcloning the eYFP coding sequence into the Kpnl and Sacl restriction sites of the pBI-TS binary vector (Koo et al., 2006; Schilmiller et al., 2007), which is a modified version of pB112 l. The PCR-amplified open reading frames (without a stop codon) of JAZlO.1, JAZ1 0.3, JAZ10.4, and JAZIO.4,“ were subcloned into pGEM-T Easy (Promega), followed by re-cloning into the BamHI 123 site of pBI-eYFP. The resulting constructs encode fusion proteins in which the JAZ C- terminus is fused in-frame to the N-terminus of eYFP. Oligonucleotide primers used to generate these constructs are listed in Table 83.1. Agrobacterium tumefaciens (strain C58Cl) harboring binary vector constructs were grown overnight at 28°C in YEP medium containing 50 ug/mL kanamycin and 50 ug/mL rifampicin. Bacterial cells were resuspended in induction medium (10 mM MES, pH 5.6, 10 mM MgC12, and 150 uM acetosyringone) and incubated for 1.5 h at room temperature prior to infiltration. Leaves of five-week-old N. tabacum plants were syringe-infiltrated with the bacterial suspension (ODboo = 0.2), and plants were maintained in a growth chamber for 2 d. Confocal microscopy was performed with a Zeiss LSM 510 META microscope (Carl Zeiss, Jena, Germany) and imaging software provided by the manufacturer. To visualize nuclei, tobacco leaves that were previously infected with Agrobacterium were syringe-infiltrated 2 h prior to imaging with a solution containing 10 ug/mL DAPI (Sigma). YFP and DAPI fluorescence were monitored simultaneously by excitation at 514 nm (argon laser) and 405 nm (diode laser), respectively. YFP and DAPI fluorescence was detected after passage through band pass 530-600 nm and 474-525 nm emission filters, respectively. Blmolecular Fluorescence Complementation Assays The full-length JAZ3 coding sequence was cloned into the pSY728, pSY738, pSY735, and pSY736 vectors (Bracha-Drori et al., 2004) to generate JAZ3-nYFP, JAZ3-cYFP, nYFP-JAZ3, and cYFP-JAZ3, respectively. See Table 83.1 for primer information. These constructs were subcloned into the XhoI and Sail restriction sites of pBl-TS and transformed into Agrobacterium tumefaciens strain C58C1. Mixtures of two 124 Agrobacterium strains were co-infiltrated into tobacco leaves, such that each of the four possible combinations of nYFP and cYFP JAZ3-fusion constructs was co-expressed. Imaging of protein-protein interaction was performed by confocal laser microscopy as described above. In vivo Protein Degradation Assay Seedlings of transgenic plants expressing JAZlO.1-YFP, JAZlO.3-YFP, and JAZIO.4- YFP fusion proteins grown for 7 d on MS plates were transferred to 48-well microtiter plate for JA treatment. Seedlings were incubated with JA (0, l, and 100 uM MeJA) for 2 h at room temperature on an orbital shaker with low speed (70 rpm). To test the effect of proteasome-specific inhibitor on protein stability, seedlings expressing JAZlO.1-YFP were pretreated with water or 50 uM MG132 (Sigma) for 2 h prior to 10 uM MeJA treatment. Fluorescence of JAZlO.1-YFP fusion proteins was analyzed with Zeiss Axio Scope fluorescence microscope (Carl Zeiss, Oberkochen, Germany). Images were taken using the software AxioVision 4.7 provided by the manufacturer. Root Growth Inhibition Assay Seeds were surface-sterilized with 30% (v/v) commercial bleach for 15 min and washed ten times with sterile distilled water. Seeds of each genotype were placed on square Petri plates (Fisher) containing MS medium that was supplemented with 50 uM MeJA. Plates were incubated at 4°C for 4 d in darkness, and then incubated under normal growth conditions for the remainder of the experiment. Plates were oriented in a vertical position 6 to 7 (1 prior to measurement of primary root length. Wild—type (Col-0) and coil -1 seeds 125 (or other appropriate lines) were included as controls for JA-sensitive and JA-resistant phenotypes, respectively. Histochemical Staining Seedlings grown on MS medium (either containing or not containing 50 uM MeJA), as described in the legends to Figs. 3.6 and 7, were transferred to a 48-well microtiter plate for GUS staining. Seedlings were vacuum-infiltrated with GUS staining buffer (50 mM KPO4 buffer, pH 7.2, 2 mM potassium ferricyanide, 2 mM potassium ferrocyanide, 0.1 % (v/v) Triton X-100, and 2 mM X-Gluc) for 15 min and then incubated for 12 h at 37°C. A graded series of ethanol washes (up to 70% ethanol) was used to remove chlorophyll, as described previously (Schilmiller et al., 2007). Photographic images were taken with a Leica M216 dissecting microscope. Strong GUS staining was also observed in several (at least five per construct) independent 35S-JAZ10.4-GUS and 35S-JAZIO.4" )A-GUSlines within 2 to 3 h of incubation with the X-Gluc substrate. 126 B AD-JAZ proteins BD-JAZ proteins 1 2 3 4 5 6 7 8 9 10.11112 1 2 3 4 5 6 7 8 9 10.11112 _.- — - ..u .. ’ ..I-c- ‘= a: - A a-l-IA a-LIXA DAPI YFP Merged JAZ3-nYFP + cYFPslAZ3 Figure 3.1. Homo— and heteromeric interaction of Arabidopsis JAZ proteins. (A) Yeast two-hybrid (Y2H) assay for homomeric JAZ interactions. Yeast strains co-expressing activation domain (AD) and DNA-binding domain (BD) fusions to each of the 12 full-length JAZ proteins were plated on media containing X-gal. Based on the intensity of LacZ-mediated blue- color formation, the strength of each interaction was rated as strong (e.g., JAZI), medium (e.g., JAZS), weak (e.g., JAZ6), or undetectable (e.g., JAZ8). See Table 3.1 for ratings on all interactions. (B) Immunoblot analysis of JAZ proteins in yeast strains used for Y2H assays shown in panel A. Each lane contains total protein extracted from cells expressing AD and BD fusions of the indicated JAZ protein (e.g., “1” corresponds to JAZI). AD—JAZ and BD-JAZ fusion proteins were detected with anti-HA (left panel) and anti-LexA (right panel) antibodies, respectively. (C) Bimolecular fluorescence complementation assay of JAZ3-JAZ3 homomeric interaction in planta. YFP fluorescence was detected in N. tabacum leaves co-infiltrated with Agrobacterium strains expressing JAZ3-nYFP and cYFP-JAZ3. DAPI staining shows the location of nuclei. The merged image shows colocalization of DAPI and YFP fluorescence. 127 A Buzz «210.1 C AD-JAZ3 1 N ZIMITIFY Jas 352 ‘6 ,9 x‘ 9’ .. b 3 9 WT HT m I . . . . 1‘1 352 a. [2111:1121 a. 101 352 Jas 1 as A"ml H [J 1 a-HA D ZIMdomaln E w 11" [WT “30A 1131‘ F1m Y1m 6185A m1 ELTIELLC_.‘JIVE:.DF.ELEKF-KB‘JIIA Jazz ELTIFE": -R‘J’M‘JE’DDZ~’...l-EKI-KE‘JZDLL. JAZ: ,L'TZF'fl-i1"}C‘J'1'DDIF'FEFE-Kéllfllé 4m .L::5‘;L-3:’JL'\J'; DIL: EKL,L.Ii4; 1L. w ,L’TIFEIFEK‘JLV': .EE: JDKL KEI NEW. «3 ,;:::~’E=;:K'J)4'JK.E‘.: :EDK; KEIMEJ‘L F mm M IL'TITY.étiMCV‘iDLTHLEr’ LI:::L. m H mm H Y1 11“ m RI’TIFE 'f-Kr4cr :DiJTh;,L EIIP'IL “A m m° Mn .LT:E"‘-‘"- '-J:~‘ :..:;DKL,'L.:M;CL 4.11.3147: av: JF— v- R KL E. MK-JL. N11,Li"-_‘S‘J:D_I_rEYJE M211 .LTZFKL‘CJ W D LESEKV EILR: L Figure 3.2. Requirement for the TIFY motif in homo— and heteromeric interaction of JAZ proteins. (A) Schematic diagram of JAZ3 deletion constructs analyzed in (B) and (C). The diagram shows the highly conserved J as (blue) and ZIM (pink, with TIFY motif) domains, as well as the weakly conserved sequence (orange, labeled “N”) at the N-terminus (Thines etal., 2007). (B) Y2H assay to assess the interaction of each JAZ3 deletion construct (AD fiJsion) shown in A with full-length JAZ3 and JAZIO.1 (BD fusions). (C) Immunoblot analysis of J AZ3 deletion proteins in yeast strains tested in (B). AD-JAZ3 fusion proteins were detected with an anti-HA antibody. (D) Amino acid sequence alignment of the ZIM domain in 12 Arabidopsis JAZs. The highly conserved TIFY motif (TIFF/YXG) is boxed. (E) Ala-scanning mutagenesis of the JAZ3 TIFY motif. Wild-type (WT) and mutant forms (e.g., T180A) of JAZ3 (AD fusions) were tested in the Y2H system for interaction with WT JAZ3 and JAZIO.1 (BD fusions). (F) Ala-scanning mutagenesis of the JAZIO.1 TIFY motif. Wild-type (WT) and mutant forms (e.g., T106A) of JAZIO.1 (AD fusions) were tested in the Y2H system for interaction with WT JAZ10.1 and JAZ6 (BD fusions). 128 A T GA A a /\ c /\ o A “AI 5 AT5613220.1 _) e in ..as... Arc TM AT561 3220.3 Wmmfiman- '> (- AT TAG G AT5613220.4 _) <- B JA210.1 147 8 VI L PTTL R P K L F G Ci N L E GDLPIARRK§LQRFL§KR5£RLV.31ii:i_:y_‘t P TS A“ 197 JA210.3 147 SVI L PTTLR P K L F6 0 N L E GDLPIARRK§LQBFL§5RK§E 185 JAZ1 0.4 147 SIFPSQGESHCNVFSRSARRD“ 167 c 650 bp-> . - < JAZ10.3 — 4 JAZ10.1 500 bp-> ‘ JAZ10.4 Figure 3.3. Identification of three JAZl0 variants produced by alternative splicing. (A) Schematic diagram of alternatively spliced transcripts At5g13220.l, At5g13220.3, and At5g13220.4 encoding JAZlO.1, JAZlO.3, and JAZIO.4, respectively. Thick lines (black, translated; grey, untranslated) represent exons (labeled A to E). Modified exons in At5g13220.3 and At5g13220.4 result from use of alternative splice sites marked D’ and C’, respectively. (B) Alternative splicing differentially affects the sequence of the C-terrninal Jas domain in the three JAZIO splice variants. Amino acid sequences N-terrninal to Ser147 are identical in the three isoforms and are not shown. Amino acids comprising the Jas domain are underlined, whereas amino acids encoded by exon D are in bold. A frame shift in exon D of At5g13220.4 (resulting from alternative splicing of exon C) removes the Jas domain and adds 20 unrelated amino acid residues C-terrninal to Ser147. Asterisks indicate the stop codon. (C) RT-PCR analysis of JAZ10 transcripts in RNA isolated from wounded leaves. The location of primers used for RT-PCR is indicated by arrows in (A). PCR products were separated by gel electrophoresis and the resulting gel was stained with ethidium bromide. Arrows denote transcripts encoding the three JAZI 0 splice variants. 129 Figure 3.4. Protein-protein interaction characteristics of JAZ10 isoforms. (A) Coronatine-dependent interaction of three JAZIO splice variants with C01] in the Y2H system. Yeast strains co—transformed with AD-JAZIO isoforms (JAZlO.1, JAZlO.3, or JAZIO.4) and BD-COIl were plated on media containing the indicated concentration of coronatine (COR). (B) Quantification of B—galactosidase activity in yeast strains shown in (A). Each strain was grown in the absence of coronatine (“0”), or in the presence of 30 pM or 200 uM coronatine (COR). Data show the mean i SD of triplicate technical replicates. (C) Immunoblot analysis of JAZIO and C011 proteins in yeast cells tested in (A). JAZ10 splice variants and C011 were detected with anti-HA and anti-LexA antibodies, respectively. (D) Interaction of three JAZIO splice variants with MYC2 in the Y2H system. Yeast strains co- transfonned with AD-MYC2 and BD-JAZlO variants (JAZlO.1, JAZlO.3, and JAZIO.4) were plated on the media containing X-gal. A yeast strain co-transformed with AD-MYC2 and an empty BD vector (pGILDA) is shown as a control (empty + MYC2). (E) Homo- and heteromeric interaction of JAZIO splice variants. Yeast strains co-transformed with one of the three JAZIO isoforms (as an AD fusion) and other members (as 3 BD fusion) of the JAZ family were plated on media containing X-gal. (F) Western blot analysis of JAZIO isoforms and JAZ3 in yeast strains used for the Y2H experiment shown in (D). JAZIO splice variants and JAZ3 were detected with anti-HA and anti- LexA antibodies, respectively. 130 COR (psi) 0 1 5 10 30 50 100 200 JAZ10.1 *COH JAZ10.3 +0011 JAZ10.4 +COI1 B 450 N '5 tr _ «9‘ \9' \9‘ g 400 “$210.1 , 5,1. 3,1- 5,1- 3 350 l 1 l 300 "' zoo HA (1- 15° con . 100 z w ’ — - o . 0 30 200 COR (11M) (it-LexA E JAZ proteins 7 8 9 10.1 10.3 10.4 11 12 a N to & 0| 0 JAZ10.1 JAZ10.3 JAZ10.4 - 7 F .9" o9 .9" .4 .w 131 JAZ1 0.1 -YFP JAZ1 0.3-YFP JAZ10.4-YFP YFP DAPI - YFP - Merged . Figure 3.5. Subcellular localization of JAZ10 splice variants. JAZIO-YFP fusion proteins were transiently expressed by Agrobacterium-mediated transformation of tobacco epidermal cells. YFP fluorescence was detected with a confocal laser scanning microscope. DAPI fluorescence was used as a marker for the nucleus. YFP alone (not fiised to JAZ10) is located in both the nucleus and the cytosol. 132 WT JAZ10.4 JAZ10.1 JAZ104 JAZ10.3 COI1—1 4 Root length (mm) WT JAZ10.1 JAZ10.4 JAZ10.3 coi1-1 Figure 3.6. Overexpression of JAZl0 splice variants differentially affects JA responses. (A-D) JAZ10.4-overexpressing plants are male sterile. (A) Wild-type (left) and 35S-JAZ10.4 (right) inflorescence. Close-up view of a wild-type flower (B) and JAZIO.4 flowers (C and D) that have short filaments and non-dehisced anthers. (E) Differential effect of JAZIO isoforms on JA-mediated inhibition of root growth. Germinated seedlings of the indicated genotype were grown for 7 d on MS medium containing 50 pM MeJA. Bar indicates 5 mm. (F) Quantification of JA-induced root grth inhibition of seedlings shown in (B). Data show the mean i SD (n = 15 seedlings per genotype). 133 A JAZ10.1-YFP 10011M JA JAZ10.3-YFP Mock 1 11M JA 100 11M JA JAZ10.4-YFP 1 11M JA 10011M JA B Mock MG132 + JA Figure 3.7. Stability of JAZ10 splice variants in vivo. (A) Differential stability of three JAZIO splice variants in response to JA. Transgenic seedlings expressing JAZlO.1-YFP, JAZlO.3-YFP, or JAZIO.4-YFP fusion proteins were treated either with water (Mock) or with the indicated concentration of MeJA (“IA”). Two h after treatment, YF P signal in root tissue was visualized by fluorescence microscopy. The exposure times for each image in A and B were identical. (B) Proteasome-dependent degradation of JAZlO.1-YFP. Seedlings expressing the 35S-JAZ10.1- YFP transgene were pretreated with water or the 26S proteasome-specific inhibitor MG132 (50 uM) for 2 h, at which time seedlings were treated with either MeJA (“JA”, 10 uM) or water (Mock) for 2 h. YFP signal in root tissue was visualized by fluorescence microscopy. v: ’V‘ A c w, y. B N N QNQ 1} s 3 s * °' ,. . ...L- .. ,, s1- 1 re '8' '9 JA26 i ._. j 3 5. 5 L ~;. __Mh._.._3-.a__j t i ’ 3 Way: JA210.1 ., , 7, 3 \ ‘xgr” I L 1‘3” :fif‘; fifth—TA: “'HA 1 ‘ l , .. JAZ10.1 JAZ10.3‘ " ' . , ‘ .. t... .1 f , Eff-‘1; 15:”; :fi‘: it'- ~ JAZ10.4 :f aw) i 3 'VJ . .xU-’ e --'.~ .2; a'LOXA Figure 3.8. The TIFY motif is required for repression of JA responses by JAZ10.4. (A) Y2H analysis of the effect of 1107A and GlllA TIFY mutations on the interaction of JAZIO.4 with JAZ6, JAZlO.1, JAZlO.3, and JAZIO.4. Yeast strains co-transformed with AD fusions of JAZlO.4 (or JAZ10.4”A and 115.210.4‘3‘)A mutants) and BD-JAZ fusions (JAZ6, JAZlO.1, JAZlO.3, and JAZlO.4) were plated on Xgal-containing medium. (B) Immunoblot analysis of JAZ proteins in yeast cells co-transformed with AD fusions of JAZIO.4 (or JAZlO.4l'>A and JAZIO.4G'>A mutants) and BD-JAZIO.1, which were tested in (A). AD- and BD-J AZ fusions were detected with anti-HA and anti-LexA antibodies, respectively. (C—F) Mutations (1107A and GlllA) in the TIFY motif differentially lack the male-sterile phenotype conferred by overexpression of wild-type JAZI 0.4. Complete lack of sterility is shown by the inflorescence (C, left) and floral (D) phenotype of 35S- JAZ10.4" ” plants. Partial lack of JAZlO.4-mediated sterility is indicated by development of a few mature silques on the inflorescence of 35S- JAZ10. 40' ”4 plants (CD, right). 355- JAZ10.4G"’A plants produced a mixture of wild-type-like flowers (E) and flowers with shortened anther filaments (F). (G) Differential effect of 1107A and GlllA mutations on JAZlO.4-mediated changes in root grth in the presence of JA. Seedlings of the indicated genotype were grown for 6 d on MS medium containing 50 uM MeJA. Bar indicates 5 mm. (H) Measurement of primary root length of seedlings shown in (G). Data are the mean :1: SD (n = 19 seedlings per genotype). 135 Figure 3.8 continued C JAZ10.4'->A JAZlO.4G->A ' . . ‘ ' I ' [1" 1“ JAZIO.4”A JAz1ti.4€*‘->A I N o .3 0'1 Root length (mm) 0' 8 WT JAZiOJ JAZ10.1">‘ JAZ10.1°'>‘ 136 Figure 3.9. The 1107A mutation does not affect the accumulation or localization of JAZ10.4. (A) Histochemical GUS staining of 35S-JAZ10.4-GUS and 35S-JAZ10.4 1' i"”-GUS transgenic seedlings. Seedlings were grown on MS medium containing kanamycin (50 pg/mL) for 14 d prior to transferring to a 48-well microtiter plate for GUS staining. (B) The male sterile phenotype conferred by overexpression of JAZ10.4-GUS (left) is not shown by plants overexpressing the 1107A mutant form of JAZlO.4-GUS (right). (C) The JA-insensitive root growth phenotype conferred by overexpression of JAZlO.4-GUS is not shown by plants overexpressing the 1107A mutant form of JAZlO.4-GUS. (D) Subcellular localization of JAZlO.4-YF P and JAZ10.4 I'>A-YF P fusion proteins that were transiently expressed by Agrobacterium-mediated transformation of tobacco epidermal cells. eYFP and DAPI fluorescence was detected by confocal laser scanning microscopy. 137 ”at! Vd—NS. 80.122 mbwra. . YENS. m30-v.eN<fi .. o 8922 r u U I o— m w an; r or W m I ON IN I ON . . r . \\ :3 - ...f .2 I 8 . v/N an?! «.252. 3.3.25; 0 0 30.... «.852. 804.22.. < 138 28w 5 523:3: 2a 28:85.5 Om v5 Q< £3 E 52030 25:85:: 058055: .3 038555: 5 A+v x53 A++v 825:. A+++v wcobm mm 5:2 53 55825 55 go Swear; 2: .5558 8392.5 5358-53 he .3652: 05 5 58m .55.? fl :OSSEQEGQEO: NS. .5» 3:58 .«o .9253 < .AmcosaEnEoo 058055; NMS Eons—0E 35$ 2 wig—:25. 05 Hmfimwa 22.8510 am 98 Q< Eon E 53.8 53 £5. m_mmou58< £w5_-=_¢ S 05 no 5am . - - . . - +++ - . - - - «5:. t t + + + .. t t t + +++ ++ + FPN<5 + .. +++ + t .. ++ t +++ +++ +++ +++ v.0 FN<—.. t t + .. t t t t . +++ + t ON<__. +++ . +++ + r .. +++ +++ +++ +++ +++ +++ sz. .. - .. r .. r r r - r .. - hN<_.. ++ - ++ - r .. + + - . +++ .. oN<-. .1. t r t t t + ++ t 1 ++ t mN<fi t t t .. t t .. t + +++ .. t VN<—. t t t t t t t t + +++ t .. MN<fi + t t .. t t + + t +++ +++ +++ NN<3 +++ t + + .. t ++ + +++ +++ +++ +++ FN 500 bp—> Supplemental Figure 3.3. Detection of transcripts encoding JAZlO.3 and JAZ10.4 using transcript-specific primers. (A) Schematic diagram of transcripts encoding JAZlO.3 and JAZIO.4. Left-pointing arrows denote the location of Oligonucleotide primers that are specific for each transcript. Primer sequences are listed in Supplemental Table 1. The gene-specific primer for JAZ10.4 hybridizes to a sequence created from splicing of the altematedonor site in exon 3 to the acceptor site in exon 4. (B) Result of an RT-PCR experiment using total RNA isolated from wounded (6 hr post wounding) Arabidopsis leaves and the transcript-specific primers indicated in (A). PCR products were visualized by agarose gel electrophoresis and ethidium bromide staining. 142 JAZproteins 1 2 3 4 5 6 7 8 9 10.1 11 12 JAZ10.4 JAZ10.4 "" JAZ10.4 6" Supplemental Figure 3.4. Yeast two-hybrid analysis of the effect of 1107A and G1 11A TIFY mutations on the interaction of JAZ10.4 with other members of the JAZ family. Yeast strains were co-transformed with plasmids encoding an AD-fusion of JAZIO.4 (or the JAZlO.4I->A and JAZIO.4G->A mutants; lefi column) and a BD—fusion of one of the indicated (top row) twelve J AZ family members. Co-transforrned strains were plated on medium containing X-gal. The photograph was taken after 48 h of incubation of the plate at 30°C. Blue color formation is indicative of a positive protein-protein interaction. 143 (04.9. .m-._._.EOO<<0t.O 225.... 09.33. 32:..."— Button. 558“: 26.0 5:5 £5 5 .52. 955?:— 15 82.5239 he 55 ._.n ozah 3.5.5.395 144 ma mm~>ma w->ma au>m 4mg au>o 4mg au>o 4mg au>w 4mg m»;mg szma 52mg wh4ma mFAma mh4ma mh4ma wFAma .m-mv<0._.00<0<00._|_.0Oc mun—>32... au>cnmNm -Ewofis, -mmm au>m -VOPN<3 -mmm au>m IndP~o -voFNvmmm mam (4.19%., -wmm wno .vd.No dos—5:3 —.n 03:. Kwanzaa—am 147 REFERENCES Abe, H., Urao, T., Ito, T., Seki, M., Shinozaki, K., and Yamaguchi-Shinozaki, K. (2003). Arabidopsis AtMYCZ (bHLH) and AtMYB2 (MYB) function as transcriptional activators in abscisic acid signaling. Plant Cell 15: 63-78. Balbi, V., and Devoto, A. (2008). Jasmonate signalling network in Arabidopsis thaliana: crucial regulatory nodes and new physiological scenarios. New Phytol. 177: 301- 318. Barbazuk, W.B., Fu, Y., and McGinnis, K.M. (2008). Genome-wide analyses of alternative splicing in plants: Opportunities and challenges. Genome Res. 18: 1381-1392. Bove, J., Kim, C.Y., Gibson, C.A., and Assmann, S.M. (2008). Characterization of wound-responsive RNA-binding proteins and their splice variants in Arabidopsis. Plant Mol. Biol. 67: 71-88. Bracha-Drori K, Shichrur K, Katz A, Oliva M, Angelovici R, Yalovsky S, Ohad N. (2004). Detection of protein-protein interactions in plants using bimolecular fluorescence complementation. Plant J. 40: 419-27. Browse, J. (2005). Jasmonate: An oxylipin signal with many roles in plants. Vitam. Horm. 72: 431-456. Browse, J., and Howe, G.A. (2008). Update on jasmonate signaling: New weapons and a rapid response against insect attack. Plant Physiol. 146: 832-838. Chen, I.C., Huang, I.C., Liu, M.J., Wang, Z.G., Chung, S.S., and Hsieh, H.L. (2007). Glutathione S-transferase interacting with far-red insensitive219 is involved in phytochrome A-mediated signaling in Arabidopsis. Plant Physiol. 143: 1189-1202. Chico, J.M., Chini, A., Fonseca, S., and Solano, R. (2008). JAZ repressors set the rhythm in jasmonate signaling. Curr. Opin. Plant Biol. 11: 486-494. Chini, A., Fonseca, S., Fernandez, G., Adie, B., Chico, J.M., Lorenzo, 0., Garcia- Casado, C., Lopez-Vidriero, 1., Lozano, F.M., Ponce, M.R., Micol, J.L., and Solano, R. (2007). The JAZ family of repressors is the missing link in jasmonate signalling. Nature 448: 666-671. Chung, H.S., Koo, A.J.K., Gao, X., Jayany, S., Thines, B., Jones, A.D., and Howe, G.A. (2008). Regulation and function of Arabidopsis JASMONA TE ZIM-domain genes in response to wounding and herbivory. Plant Physiol. 146: 952-964. 148 Clough, S.J., and Bent, A.F. (1998). Floral dip: A simplified method for Agrobacterium-mediated transformation of Arabidopsis thaliana. Plant J. 16: 735-743. Devoto, A., Ellis, C., Magusin, A., Chang, H.S., Chilcott, C., Zhu, T., and Turner, J.G. (2005). Expression profiling reveals COIl to be a key regulator of genes involved in wound- and methyl jasmonate-induced secondary metabolism, defence, and hormone interactions. Plant Mol. Biol. 58: 497-513. Dombrecht, B., Xue, G.P., Sprague, S.J., Kirkegaard, J.A., Ross, J.J., Reid, J.B., Fitt, G.P., Sewelam, N., Schenk, P.M., Manners, J.M., and Kazan, K. (2007). MYC2 differentially modulates diverse jasmonate-dependent functions in Arabidopsis. Plant Cell 19: 2225-2245. Feys, B., Benedetti, C.E., Penfold, C.N., and Turner, J.G. (1994). Arabidopsis mutants selected for resistance to the phytotoxin coronatine are male sterile, insensitive to methyl jasmonate, and resistant to a bacterial pathogen. Plant Cell 6: 751-759. Glazebrook, J. (2005). Contrasting mechanisms of defense against biotrophic and necrotrophic pathogens. Annu. Rev. Phytopathol. 43: 205-227. Goossens, A., Hakkinen, S.T., Laakso, 1., Seppanen-Laakso, T., Biondi, S., De Sutter, V., Lammertyn, F., Nuutila, A.M., Soderlund, H., Zabeau, M., Inze, D., and Oksman-Caldentey, K.M. (2003). A functional genomics approach toward the understanding of secondary metabolism in plant cells. Proc. Natl. Acad. Sci. USA 100: 8595-8600. Griffiths, S., Dunford, R.P., Coupland, C., and Laurie, D.A. (2003). The evolution of CONSTANS-like gene families in barley, rice, and Arabidopsis. Plant Physiol. 131: 1855-1867. Herms, D.A., and Mattson, W.J. (1992). The dilemma of plants - To grow or defend. Q. Rev. Biol. 67: 283-335. Howe, G., and Jander, G. (2008). Plant immunity to insect herbivores. Annu. Rev. Plant Biol. 59: 41-66. Ishiguro, S., Kawai-Oda, A., Ueda, J., Nishida, I., and Okada, K. (2001). The DEFECTIVE IN ANTHER DEHISCIENCE gene encodes a novel phospholipase A1 catalyzing the initial step of jasmonic acid biosynthesis, which synchronizes pollen maturation, anther dehiscence, and flower opening in Arabidopsis. Plant Cell 13: 2191-2209. Izaguirre MM, Mazza CA, Biondini M, Baldwin IT, Ballaré CL. (2006). Remote sensing of future competitors: impacts on plant defenses. Proc. Natl. Acad. Sci. USA 103: 7170-7174. 149 Katsir, L., Chung, H.S., Koo, A.J.K., and Howe, G.A. (2008a). J asmonate signaling: a conserved mechanism of hormone sensing. Curr. Opin. Plant Biol. 11: 428-435. Katsir, L., Schilmiller, A.L., Staswick, P.E., He, S.Y., and Howe, G.A. (2008b). C011 is a critical component of a receptor for jasmonate and the bacterial virulence factor coronatine. Proc. Natl. Acad. Sci. USA 105: 7100-7105. Kim, J., Harter, K., and Theologis, A. (1997). Protein-protein interaction among the Aux/1AA proteins. Proc. Natl. Acad. Sci. USA 94: 11786-11791. Klemm, J.D., Schreiber, S.L., and Crabtree, GR. (1998). Dimerization as a regulatory mechanism in signal transduction. Annu. Rev. Immunol. 16: 569-592. Koo, A.J.K., Chung, H.S., Kobayashi, Y., and Howe, G.A. (2006). Identification of a peroxisomal acyl-activating enzyme involved in the biosynthesis of jasmonic acid in arabidopsis. J. Biol. Chem. 281: 33511-33520. Laurie-Berry, N., Joardar, V., Street, I.H., and Kunkel, B.N. (2006). The Arabidopsis thaliana JASMONA TE INSENSIT I VB 1 gene is required for suppression of salicylic acid-dependent defenses during infection by Pseudomonas syringae. Mol. Plant. Microbe Interact. 19: 789-800. Li, L., Zhao, Y., McCaig, B.C., Wingerd, B.A., Wang, J., Whalon, M.E., Pichersky, E., and Howe, G.A. (2004). The tomato homolog of CORONATINE- INSENSITIVEI is required for the maternal control of seed maturation, jasmonate-signaled defense responses, and glandular trichome development. Plant Cell 16: 126-143. Lorenzo, 0., Chico, J.M., Sanchez-Serrano, J.J., and Solano, R. (2004). JASMONATE-INSENSITIVEI encodes a MYC transcription factor essential to discriminate between different jasmonate-regulated defense responses in Arabidopsis. Plant Cell 16: 1938-1950. Mandaokar, A., Thines, B., Shin, B., Lange, R.M., Choi, G., Koo, Y.J., Yoo, Y.J., Choi, Y.D., Choi, G., and Browse, J. (2006). Transcriptional regulators of stamen development in Arabidopsis identified by transcriptional profiling. Plant J. 46: 984-1008. McConn, M., and Browse, J. (1996). The critical requirement for linolenic acid is pollen development, not photosynthesis, in an Arabidopsis mutant. Plant Cell 8: 403-416. Melotto, M., Mecey, C., Niu, Y., Chung, HS, Katsir, L., Yao, J., Zeng W., Thines, B., Staswick. P.E., Browse, J., Howe, GA, He SY. (2008). A critical role of two 150 positively charged amino acids in the Jas motif of Arabidopsis JAZ proteins in mediating coronatine- and jasmonoyl isoleucine-dependent interactions with the C011 F-box protein. Plant J. 55: 979-988 Mitra, S., and Baldwin, LT. (2008). Independently silencing two photosynthetic proteins in Nicotiana attenuata has different effects on herbivore resistance. Plant Physiol. 148: 1128-1138. Ouellet, F., Overvoorde, P.J., and Theologis, A. (2001). IAA17/AXR3: Biochemical insight into an auxin mutant phenotype. Plant Cell 13: 829-841. Park, J.H., Halitschke, R., Kim, H.B., Baldwin, I.T., Feldmann, K.A., and Feyereisen, R. (2002). A knock-out mutation in allene oxide synthase results in male sterility and defective wound signal transduction in Arabidopsis due to a block in jasmonic acid biosynthesis. Plant J. 31: 1-12. Pauwels, L., Morreel, K., De Witte, E., Lammertyn, F., Van Montagu, M., Boerjan, W., Inze, D., and Goossens, A. (2008). Mapping methyl jasmonate-mediated transcriptional reprogramming of metabolism and cell cycle progression in cultured Arabidopsis cells. Proc. Natl. Acad. Sci. USA 105: 1380-1385. Pozo, M.J., Van Loon, L.C., and Pieterse, C.M.J. (2004). Jasmonates - Signals in plant-microbe interactions. J. Plant Growth. Regul. 23: 211-222. Reddy, A.S.N. (2007). Alternative splicing of pre-messenger RNAs in plants in the genomic era. Annu. Rev. Plant Biol. 58: 267-294. Reymond, P., Bodenhausen, N., Van Poecke, R.M., Krishnamurthy, V., Dicke, M., and Farmer, E.E. (2004). A conserved transcript pattern in response to a specialist and a generalist herbivore. Plant Cell 16: 3132-3147. Sanders, P.M., Lee, P.Y., Biesgen, C., Boone, J.D., Beals, T.P., Weiler, E.W., and Goldberg, RB. (2000). The Arabidopsis DELAYED DEHISCENCEI gene encodes an enzyme in the jasmonic acid synthesis pathway. Plant Cell 12: 1041- 1061. Sasaki-Sekimoto, Y., Taki, N., Obayashi, T., Aono, M., Matsumoto, F., Sakurai, N., Suzuki, H., Hirai, M.Y., Noji, M., Saito, K., Masuda, T., Takamiya, K., Shibata, D., and Ohta, H. (2005). Coordinated activation of metabolic pathways for antioxidants and defence compounds by jasmonates and their roles in stress tolerance in Arabidopsis. Plant J. 44: 653-668. Sasaki, Y., Asamizu, E., Shibata, D., Nakamura, Y., Kaneko, T., Awai, K., Amagai, M., Kuwata, C., Tsugane, T., Masuda, T., Shimada, H., Takamiya, K., Ohta, H., and Tabata, S. (2001). Monitoring of methyl jasmonate-responsive genes in 151 Arabidopsis by cDNA macroarray: self-activation of jasmonic acid biosynthesis and crosstalk with other phytohormone signaling pathways. DNA Res. 8: 153-161. 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. USA 97 : 11655-11660. Schilmiller, A.L., Koo, A.J., and Howe, G.A. (2007). Functional diversification of acyl- CoA oxidases in jasmonic acid biosynthesis and action. Plant Physiol. 143: 812- 824. Schmidt, D.D., and Baldwin, LT. (2006). Transcriptional responses of Solanum m'grum to methyl jasmonate and competition: a glasshouse and field study. Func. Ecol. 20: 500-508. Shikata, M., Matsuda, Y., Ando, K., Nishii, A., Takemura, M., Yokota, A., and Kohchi, T. (2004). Characterization of Arabidopsis ZIM, a member of a novel plant-specific GATA factor gene family. J. Exp. Bot. 55: 631-639. Shoji, T., Ogawa, T., and Hashimoto, T. (2008). Jasmonate-induced nicotine formation in tobacco is mediated by tobacco C011 and JAZ genes. Plant Cell. Physiol. 49: 1003-1012. Sineshchekov, V.A., Loskovich, A.V., Riemann, M., and Nick, P. (2004). The jasmonate-free rice mutant hebiba is affected in the response of phyA 'lphyA " pools and protochlorophyllide biosynthesis to far-red light. Photochem. Photobiol. Sci. 3: 1058-1062. Stintzi, A., and Browse, J. (2000). The Arabidopsis male-sterile mutant, opr3, lacks the 12-oxophytodienoic acid reductase required for jasmonate synthesis. Proc. Natl. Acad. Sci. USA 97: 10625-10630. Suzuki, H., Reddy, M.S., Naoumkina, M., Aziz, N., May, G.D., Huhman, D.V., Sumner, L.W., Blount, J.W., Mendes, P., and Dixon, R.A. (2005). Methyl jasmonate and yeast elicitor induce differential transcriptional and metabolic re- programming in cell suspension cultures of the model legume Medicago truncatula. Planta 220: 696-707. Takahashi, F ., Yoshida, R., Ichimura, K., Mizoguchi, T., Seo, S., Yonezawa, M., Maruyama, K., Yamaguchi-Shinozaki, K., and Shinozaki, K. (2007). The mitogen-activated protein kinase cascade MKK3-MPK6 is an important part of the jasmonate signal transduction pathway in Arabidopsis. Plant Cell 19: 805-818. 152 Thines, B., Katsir, L., Melotto, M., Niu, Y., Mandaokar, A., Liu, G., Nomura, K., He, S.Y., Howe, G.A., and Browse, J. (2007). JAZ repressor proteins are targets of the SCFCO” complex during jasmonate signalling. Nature 448: 661-665. Ulmasov, T., Hagen, G., and Guilfoyle, T.J. (1999). Dimerization and DNA binding of auxin response factors. Plant J. 19: 309—319. Uppalapati, S.R., Ayoubi, P., Weng, H., Palmer, D.A., Mitchell, R.E., Jones, W., and Bender, CL. (2005). The phytotoxin coronatine and methyl jasmonate impact multiple phytohormone pathways in tomato. Plant J. 42: 201-217. Vanholme, B., Grunewald, W., Bateman, A., Kohchi, T., and Gheysen, G. (2007). The tify family previously known as ZIM. Trends Plant Sci. 12: 239-244. Wenkel, S., Turck, F., Singer, K., Gissot, L., Le Gourrie rec, J., Samach, A., and Coupland, G. (2006). CONSTANS and the CCAAT box binding complex share a functionally important domain and interact to regulate flowering of Arabidopsis. Plant Cell 18: 2971-2984. White, D.W.R. (2006). PEAPOD regulates lamina size and curvature in Arabidopsis. Proc. Natl. Acad. Sci. USA 103: 13238-13243. Xie, D.X., F eys, B.F., James, S., Nieto-Rostro, M., and Turner, J.G. (1998). C011: an Arabidopsis gene required for jasmonate-regulated defense and fertility. Science 280: 1091-1094. Yan, Y., Stolz, S., Chetelat, A., Reymond, P., Pagni, M., Dubugnon, L., and Farmer, E.E. (2007). A downstream mediator in the growth repression limb of the jasmonate pathway. Plant Cell 19: 2470-2483. 153 CHAPTER 4: ROLE OF JAS INTRON IN FUNCTIONAL DIVERSIFICATION OF JAZ PROTEINS 154 ABSTRACT JASMONATE ZIM-domain (JAZ) proteins are negative regulators of jasmonate (JA) signaling. Bioactive forms of JA activate the expression of JA response genes by targeting JAZ proteins for ubiquitin-mediated proteolysis via the SCFCOU/26S proteasome pathway. The Jas domain interacts directly with C01] in a hormone- dependent manner and is thus a key determinant of the stability and repressive function of JAZs. Here, we report that 9 of the 12 JAZ genes in Arabidopsis contain a conserved intron (the Jas intron) that separates the N-terrninal 20 amino acids and C-terminal 7 amino acids of the Jas domain. Retention of the Jas intron during splicing of JAZ pre- mRNAs results in the production of alternatively spliced transcripts that encode JAZ isoforms with a modified or truncated Jas domain. We demonstrate a role for Jas intron retention in generating splice variants of JAZZ, JAZ3, and JAZIO that have reduced affinity for C011. Conservation of the J as intron in angiosperms, S. maellendorflii, and P. patens suggests that this sequence has provided an enduring selective advantage during the evolution of land plants. We conclude that alternative splicing plays a central role in generating functional diversity of J AZ proteins through altered protein stability. 155 INTRODUCTION The lipid-derived hormone jasmonate (JA) is implicated in diverse aspects of plant growth and development. JA is particularly well known for regulating plant stress responses, including defense against insect herbivores and pathogens, and exposure to UV radiation and ozone (Glazerbrook, 2005; Howe and Jander, 2008; Conconi et al., 1996; Browse, 2005). In healthy plants, JA also plays an important role in cell division, carbon partitioning, photomorphogenesis, reproductive development, and trichome development (McConn and Browse, 1996; Creelman and Mullet, 1997; Li et al., 2004; Feys et al., 1994; Sineshchekov et al., 2004; Browse, 2005; Yoshida et al., 2009). JA exerts this wide range of effects through large-scale reprogramming of gene expression. Hundreds of JA response genes in many plant species have been identified and characterized (Schenk et al., 2000; Reymond et al., 2000; Sasaki, 2001; Goossens et al., 2003; Sasaki-sekimoto et al., 2005; Suzuki et al., 2004; Devoto et al., 2005; Uppalapati et al., 2005; Mandaokar et al., 2006; Schmidt and Baldwin, 2006; Reymond et al., 2004). The expression of many JA-response genes is positively regulated by the basic helix-loop-helix transcription factor MYC2/JASMONATE INSENSITIVE 1 (JINl) (Abe et al., 2003; Lorenzo etal., 2004; Laurie-Berry et al., 2006; Chini etal., 2007; Dombrecht et al., 2007; Takahashi et al., 2007). Current models of JA signaling indicate that, in at the presence of low JA levels, MYC2 is directly repressed by JASMONATE ZIM- domain (JAZ) proteins (Chini et al., 2007, 2009; Melotto et al., 2008; Chung and Howe, 2009). Increased production of JA in response to environmental or developmental cues leads to ubiquitin-mediated proteolysis of JAZ proteins, which releases MYC2 from repression and permits expression of early JA-response genes (Chini et al., 2007; Thines 156 et al., 2007; Chung and Howe, 2009). A key event in this signal transduction chain is the hormone-induced targeting of JAZ proteins to the E3 ubiqutin ligase complex, SCFCO“. JAZs physically interact with the F-box protein CORONATINE IN SENSITIVE 1 (C011) in a manner that depends on bioactive forms of JA, such as (3S,7R)-jasmonoyl-L- isoleucine (JA-Ile) (Xie et al., 1998; Thines et al., 2007; Katsir et al., 2008b; Melotto et al., 2008; Katsir et al., 2008b; Fonseca et al., 2009). The JA-Ile-induced JAZ-COIl interaction leads to ubiquitination and subsequent degradation of JAZs by the 26S proteasome. Direct evidence of J AZ ubiquitination was recently reported by Saracco et al. (2009). The bacterial toxin coronatine is a structural mimic of JA-Ile and a potent agonist of the COIl-JAZ receptor system (Feys et al., 1994; Melotto et al., 2008; Katsir et al., 2080b; Chung and Howe, 2009). A highly conserved 27-amino-acid sequence motif, referred to as the Jas (JASMONATE-ASSOCIATED) domain, is located near the C-terminal end of JAZ proteins. Yeast two-hybrid and in vitro pull-down experiments have shown that this sequence is necessary and sufficient for hormone-dependent binding of JAZ to C01] (Thines et al., 2007; Karsir et al., 2008b; Melotto et al., 2008; Chini et al., 2009; Chung and Howe, 2009). These studies also showed that full-length JAZ proteins fused to GUS or GFP/YFP are degraded in JA-treated cells, whereas JAZ proteins lacking the Jas domain (JAZAJas) are stabilized against JA-induced protein degradation. Significantly, expression of JAZAJas proteins in Arabidopsis produces various JA-insensitive phenotypes, including male sterility, resistance to JA-inhibited root growth, resistance to pathogen infection, and susceptibility to insect herbivores (Thines et al., 2007; Chini et al., 2007; Yan et al., 2007; Chung et al., 2008; Chung and Howe, 2009). The dominant 157 negative effect of JAZAJas proteins on JA signal output is consistent with a role for JAZs as negative regulators of JA signaling whose abundance is controlled by the ubiquitin/26S proteasome pathway. Alternative splicing (AS) of pre-mRNA provides an important mechanism for creating protein diversity and biological complexity in eukaryotic organisms. It has been shown that approximately 95% of all human genes undergo AS (Pan et al., 2008), and that many of the resulting proteins serve essential roles in physiology and development (Leeman and Gilmore, 2008; Irimia et al., 2009; Tazi et al., 2009). Cross-species conservation in the exon-intron organization within a gene family often reflects the biological importance of AS events (Kalyna et al., 2006; Iida et al., 2006; Li et al., 2006). Although it is now clear that AS contributes to the high complexity of plant transcriptomes, our understanding of the functional significance of AS in plants is still largely unexplored (Wang and Brendel, 2006; Reddy et al., 2007). There is evidence to indicate that AS is important for plant adaptation to stress (Barbazuk et al., 2008; Reddy et al., 2007). However, recent bioinformatics studies provided evidence that conserved splicing events across plant species are rare, thus suggesting a limited role for AS in generating functionally diversified proteins (Severing et al., 2009). The recent identification of functionally distinct splice variants of JAZIO (also known as JASl) has provided new insight into the role of AS in the regulation of JA signaling. Differential splicing of JAZ10 pre-mRNA generates three transcripts that encode proteins that differ in the sequence of the Jas domain. The JAZ10.) transcript (At5g13220.l) encodes a full-length protein harboring an intact Jas domain. This protein strongly interacts with C01] in a ligand-dependent manner and is readily degraded via 158 the 26S proteasome pathway in response to JA treatment (Chung and Howe, 2009). The JAZlO.3 protein (encoded by At5g13220.3) lacks seven amino acids from the C-terminal end of the of Jas motif (Yan et al., 2007). JAZlO.3 interacts weakly with COIl in a ligand-dependent manner and is degraded in planta in response to high concentrations of exogenous JA (Chung and Howe, 2009). Consistent with the intermediate stability of JAZlO.3, overexpression of this splice variant in Arabidopsis confers partial insensitivity to JA (Yan et al., 2007; Chung and Howe, 2009). A third JAZIO splice variant (JAZlO.4) lacks the entire Jas domain and, as predicted, does not interact with C011 and is highly resistant to J A-mediated degradation (Chung and Howe, 2009). Although these findings establish a role for AS in the production of functionally distinct JAZ isoforms, it is not known whether this phenomenon is specific for Arabidopsis JAZIO or whether it reflects a general mechanism to create diverse JAZ isoforms. Here, we show that the Jas domain of most JAZ genes from Physcomitrella patens (a moss), Selaginella moellendmflii (a vascular non-seed plant), and diverse seed- bearing plants is encoded by two exons separated by a highly conserved intron, which we refer to as the Jas intron. As is the case for JAZlO.3, retention of the Jas intron during processing of these JAZ pre-mRNAs is predicted to generate truncated JAZ proteins lacking the C-terminal 7 amino acids of the Jas domain. We present experimental evidence for the existence of these AS transcripts in Arabidopsis, and show that the corresponding protein products for JAZZ, JAZ3, and JAZIO have reduced capacity to interact with C01] in the presence of coronatine and JA-Ile. We conclude that AS events involving retention of the Jas intron play a central role in generating functional diversity of JAZ proteins through altered protein stability. 159 RESULTS Structural Organization of Arabidopsis JAZ Genes and Identification of the Jas Intron We used gene models at TAIR (The Arabidopsis Information and Resources; http://www.arabidopsisorg) to compare the structural organization of the 12 Arabidopsis JAZ genes. In general, genes encoding phylogenetically related proteins, such as JAZ3/JAZ4/JAZ9 and JAZS/JAZ6, shared a similar intron-exon organization (Fig. 4.1A; Fig. $4.1). This finding supports the hypothesis that expansion of the Arabidopsis JAZ family involved gene duplication events giving rise to paralogous gene pairs (Vanholme et al., 2007). One notable exception to this was JAZI and JAZ2 which, despite their overall amino acid sequence similarity, are encoded by structurally diverse genes that serve non-redundant roles in JA signaling (Grunewald et al., 2009). The structural organization of sequences encoding the Jas domain is remarkably similar for all Arabidopsis JAZ genes except of JAZ1, JAZ 7, and JAZ8. Specifically, we found that the Jas domain of nine JAZ proteins is encoded by two exons, the first (5’) of which specifies the N-terminal 18 to 20 amino acids (depending on the specific gene) of the domain (Fig. 4.1A and B). With the exception of JAZ5 and JAZ6, this short exon (55 to 61 nucleotides) does not code for amino acids outside the Jas domain. The second (3’) exon encodes the C-terminal 7 amino acids of the Jas domain and the remainder of the protein’s C-terminus (Fig. 4.1). In all nine JAZ genes that share this structure, the intron separating the two exons is located at the same position and phase (i.e., +2). We henceforth refer to this conserved intron as the J as intron. 160 A functional role for the bipartite organization of the Jas domain coding region was established for JAZlO.3, which is generated by an alternative splicing event involving retention of the Jas intron (Yan et al., 2007; Chung and Howe, 2009). In this case, retention of the 5’ end of Jas intron creates an in-frame stop codon immediately after the highly conserved KRKXR motif (Fig. 4.1B). Y2H studies showed that this naturally truncated JAZ interacts weakly with C011 and, as a consequence, is stabilized against JA-induced degradation in vivo (Chung and Howe, 2009). Our sequence analysis revealed that, like JAZIO, hypothetical retention of the Jas intron in JAZZ, JAZ6, and JAZI 1 would create an in-frame stop codon after the KRKXR motif (Fig. 4.2). Similarly, retention of the Jas intron in JAZ3 and JAZ4 is predicted to generate splice variants containing one and five amino acids, respectively, afler the KRKXR motif. Identification of Alternatively Spliced JAZ Transcripts with a Retained Jas Intron The TAIR database contains an EST-supported gene model for JAZ4 (Atlg—iSStiOQ. JAZ4.2) and JAZIO (At5g13220.2, JAZlO.2; At5g13220.3, JAZlO.3) that are indicative of Jas intron retention. To systemically determine whether the Jas intron is retained during splicing of other JAZ pre-mRNAs, we performed reverse transcriptase-mediated PCR (RT-PCR) with primers designed to specifically amplify transcripts in which 5’ end of the Jas intron is retained (i.e. the 5’ splicing donor site is not used). Using RNA from JA-treated seedlings as a template for RT-PCR, this experiment identified Jas intron- containing transcripts for JAZZ, JAZ3, JAZ4, JAZ6, JAZ9, JAZIO, and JAZIZ but not JAZ5 and JAZ/1 (Fig. 4.3A). Sequence analysis of cloned PCR products showed that these transcripts contain the Jas intron but not other intron sequences, therefore excluding 161 the possibility of genomic DNA contamination in the PCR reactions. Control experiments using reverse primers that span the two Jas domain-coding exons showed that JAZ transcripts lacking the Jas intron are also present in JA-treated seedlings (Fig. 438). Based on these results, we conclude that the Jas intron in most (at least 7) Arabidopsis JAZ genes can be retained during pre-mRNA processing. Retention of the Jas Intron Produces JAZ Isoforms with Altered Function The importance of the Jas domain as a JAZ destabilization element via a mechanism involving hormone-dependent COIl interaction is now well established (Thines et al., 2007; Katsir et al., 2008b; Melotto et al., 2008; Chung and Howe, 2009; Fonseca et al., 2009). Given the effect of Jas-intron retention on the function of JAZlO.3 (Yan et al., 2007; Chung and Howe, 2009), we tested the hypothesis that other JAZ isoforms produced by this intron retention event are altered in their capacity to interact with COI1. We first used the yeast two-hybrid (Y2H) system to assess all 12 full-length JAZ proteins for their ability to bind COIl in a ligand-dependent fashion. In the presence of COR, a potent mimic of JA-Ile (Feys et al., 1994; Staswick, 2008; Katsir et al., 2008), C011 interacted strongly with JAZ1, JAZZ, JAZ3, JAZ9, and JAZlZ, and moderately with JAZ4 and JAZlO.] (Fig. 4.4A). COR-dependent interaction of COIl with JAle was very weak but nevertheless detectable, whereas no interaction with the four remaining family members (JAZ5, JAZ6, JAZ7, and JAZ8) was observed in this system. These results are consistent with previous studies showing that COR stimulates COI1 binding to JAZI, JAZ3, JAZ9, and JAZ10.1 (Katsir et al., 2008b; Melotto et al., 2008; Chung and Howe, 2009; Fonseca et al., 2009). 162 We also assessed COIl-JAZ interactions in the presence of the two native stereoisomers of JA-Ile, namely (3R,7S)-JA-Ile and (3R,7R)-JA-Ile. In agreement with recent studies (Fonseca et al., 2009) showing that the (3R,7S) isomer is the active form of the endogenous hormone, we observed that (3R,7S)-JA-Ile but not (3R,7R)-JA-Ile stimulates COIl-JAZ binding (Fig. 4.4A and data not shown). At the concentration of (3R,7S)-JA-Ile used in this experiment, a positive COIl-JAZ interaction was observed only with those JAZs (i.e., JAZ1, JAZZ, JAZ3, and JAZ12) that interacted strongly with C01] in the presence of COR. Western blot analysis showed that all 12 JAZ proteins were produced in yeast cells used for this experiment (Fig. $4.2). These results suggest that members of the JAZ family differentially interact with C01], and provide a rational basis for selecting specific JAZs for further study. We chose JAZZ and JAZ3 for additional studies aimed at characterizing splice variants that are produced as a consequence of Jas intron retention. In accordance with established gene models in TAIR, we refer to proteins produced from the completely spliced transcripts as JAZZ.1 and JAZ3.1, and the corresponding splice variants encoded by Jas intron-containing transcripts as J AZ2.Z and JAZ3.4. Both COR and (3R,7S)-JA-Ile stimulated the interaction of COIl with JAZZ.1 and JAZ3.1 in a dose-dependent manner (Fig. 4.4B). COR appeared to be more active than (3R,7S)-JA-Ile over the range of concentrations tested. Parallel analysis of J A222 and J AZ3.4 showed that these truncated isoforms interact more weakly with COIl in comparison to the full-length JAZZ.1 and JAZBJ proteins. This effect was observed both in the presence of COR and (3R,7S)-JA- Ile. Notably, relatively high concentrations of COR stimulated COIl interaction with JAZZ.Z and JAZ3.4, whereas (3R,7S)-JA-Ile did not promote these interactions at any of 163 the concentrations tested. These findings provide evidence that retention of the Jas intron during pre-mRNA splicing represents a general mechanism for producing functional diversity among JAZ proteins. Alternative Splicing of JAZIO pre-mRN A Affects JA Signal Output Previous fiinctional analyses of JAZIO isoforms produced by alternative splicing relied on overexpression of cDNAs encoding specific splice variants (Yan et al., 2007; Chung and Howe, 2009). To further test the hypothesis that alternative splicing of JAZ pre- mRNA in vivo has an effect on JA signal output, we generated stable lines of Arabidopsis that express the full-length JAZIO genomic DNA (Fig. 4.1A) under the control of the Cauliflower Mosaic Virus (CaMV) 358 promoter. Fifteen of 20 independent Tl lines expressing the 35S-JAZIOG transgene produced progeny that were insensitive to JA- induced root growth inhibition. Subsequently, we compared the root growth phenotype of three representative T2 homozygous lines (JAZI 06-2, JAZI 06-4, and JAZI OG-5) to lines overexpressing one of the three specific JAZ10 splice variants, JAZlO.1, JAZlO.3, or JAZ10.4 (Chung and Howe, 2009). Root length measurements performed 9 days after germination showed that the level of JA insensitivity exhibited by all three 35S-JAZIOG lines was comparable to that of 35S—JAZIO.4 plants (Fig. 4.5A and B). This finding indicates that 35S-JAZI OG plants produce one or more dominant negative splice variants (i.e., JAZlO.3 and/or JAZ10.4) as a result of alternative splicing of JAZ/0 pre-RNA by the endogenous splicing machinery. In contrast to the male-sterile phenotype of 35S- JAZIO.4 plants (Chung and Howe, 2009), we failed to observe reproductive defects in any of the 50 independent 35S-JAZIOG lines tested, including homozygous lines (data 164 not shown). This result suggests that alternative splicing of JAZIO pre-RNA may occur in a tissue-specific manner. Conservation of the Jas Intron during Evolution of Land Plants To gain insight into the evolutionary significance of Jas intron retention as a mechanism for modulating JA signaling, we analyzed the intron-exon structure of JAZ genes in phylogenetically diverse taxa. Among the species in which JAZ genomic sequences were identified were Physcomitrella patens, Selaginella moellendorflii, Oryza sativa, Brachypodium distachyon, Populus trichocarpa, Vitis vim’fera, and Arabidopsis (see Methods). A total of 72 JAZ genes were identified from these seven species (Tab. 4.1). Although the overall amino acid sequence similarity between proteins from different species was low, multiple sequence alignments showed that the ZIM and Jas domain signature sequences are strongly conserved across species (Fig. 34.3). A phylogenetic tree constructed from the 72 proteins indicated that the family can be divided into four subgroups (Fig. 4.6). All P. patens and S. moellendorflii sequences clustered in subgroup A, which also contained JAZ proteins from the flowering plants. Arabidopsis JAZ3, JAZ4, JAZ9, which are encoded by highly similar genes (Fig. 4.1A), belong to this subgroup and thus appear to be related to the earliest evolved JAZ proteins. In contrast, members of subgroup D (e. g., Arabidopsis JAZ7/8) appear to have evolved more recently in eudicots. Remarkably, we found that the position and phase (+2) of the Jas intron is conserved in the vast majority (56 of 72; 78%) of JAZ genes from all species examined, including P. patens and S. moellendorflii (Tab. 4.1). Retention of the Jas intron during 165 pre-mRNA processing of 32 (57%) of these genes is predicted to produce an in-frame stop codon immediately after the conserved KRKXR motif (Tab. $41), as is the case for the alternative splicing event giving rise to Arabidopsis JAZlO.3. In many other cases, an in-frame stop codon near the 5’ end of the retained Jas intron is predicted to produce a JAZ variant containing less than 10 amino acids C-terrninal to the KRKXR motif (Tab. $4.1). EST sequence databases provide empirical evidence for the existence of Jas intron- containing transcripts in species other than Arabidopsis. For example, Jas intron- containing transcripts for JAZB3 and JAZB4 from B. distachyon (identification numbers Bradi4g31240.l and Bradi5gZ4410.l, respectively) were identified in the Brachypodium database (http://www.brachybaseorg:8075/). These observations suggest that the Jas intron existed in an ancestral JAZ gene prior to the divergence of the gene family. The persistence of the Jas intron through evolution further suggests an important biological role for this sequence. DISCUSSION The J as Domain is Encoded by DNA Having a Conserved Intron-Exon Structure We report here that the intron-exon organization of DNA encoding the Jas domain of most JAZ proteins is conserved in diverse land plants. In all 54 JAZ genes that share this organization (Tab. 84.1), the Jas domain is encoded by a 5’ exon that specifies the N- terminal 18 to 20 amino acids of the domain ending in the conserved KRKXR, and a 3’ exon specifying the C-terminal seven amino acids ending in the conserved PY (referred to as the FY motif; Fig. 4.18). These two exons are split by a sequence we refer to as the Jas intron. The presence of the Jas intron in JAZ orthologs from P. patens and S. 166 moellendorflii is noteworthy because JAZs and other members of the TIFY family appear to have arisen in early land plants (Vanholme et al., 2007; Chico et al., 2008; Katsir et al., 2008a). These observations suggest that the Jas intron arose early in the evolution of the TIFY family, and has been retained during the time since P. patens diverged from higher plants approximately 450 million years ago. Precise conservation of the position and phase (+2) of the Jas intron in all species strongly suggests an important biological role for this sequence (see below). The sequence of the N-terminal region of the Jas domain encoded by the 5’ exon resembles the CCT (C0, CO-like, T 0C1) domain that was first identified in the plant proteins TOCl and CONSTANS (CO), which serve important roles in controlling plant responses to environmental signals (Strayer et al., 2000; Robson et al., 2001). This sequence similarity is restricted to the N-terminal part of the CCT domain (Fig. 1.4). Although the C-terminal portion of the CCT domain has been implicated in protein- protein interaction, the biochemical function of the N-terminal CCT sequence resembling the Jas domain is not known (Wenkel et al., 2006). The presence of a CCT domain- containing C0 gene in Chlamydomonas reinhardtii (Griffiths et al., 2003; Serrano et al., 2009) indicates that the CCT domain predates the evolution of the TIFY family, which first appeared in lower plants (Vanholme et al., 2007). ZIM and ZIM-like (ZML) members of the TIFY family contain a CCT domain that, like the CCT domain in CO and CO-like proteins, is encoded by a single exon (Griffiths et al., 2003; Vanholme et al., 2007). These observations raise the speculative but interesting possibility that Jas domain evolved from an ancestral TIFY protein in which the CCT domain-coding region was interrupted by the J as intron. 167 Role of the Jas Intron in Functional Diversification of JAZ Proteins Initial insight into the phenomenon of JAZ alternative splicing came from the discovery that overexpression of a JAZ10 splice variant (JAZlO.3) lacking the C-terminal PY motif results in partial loss of JA responsiveness (Yan et al., 2007). Subsequent analyses showed that JAZlO.3 interacts weakly with COIl in vitro, and is stabilized against JA- dependent degradation in planta (Chung and Howe, 2009). Arabidopsis JAZlO.3 is encoded by a transcript (At5g13220.3) in which the 5' splice donor site of the Jas intron is not used. As a consequence, an in-frame stop codon (TAA) immediately following Argl 85 produces a truncated protein lacking the FY motif. Our results show that Jas intron retention is not specific for JAZIO but rather reflects a general mechanism for generating JAZ isoforms with an altered J as domain. At least two mechanisms can be proposed to explain how jas intron retention affects dominant negative effects on JA signal output. YZH studies performed with splice variants of JAZZ, JAZ3, and JAZ10 support the hypothesis that removal of the FY motif reduces the strength of hormone-dependent JAZ binding to C01] (Chung and Howe, 2009; this study). In vitro pull-down assays performed with JAZlO.1, JAZlO.3, and JAZlO.4 confirmed the differential, hormone-dependent interaction of these proteins with COIl (T Cooke, HS Chung, GA Howe, unpublished results). Functional analysis of additional splice variants will be useful to further test this hypothesis, as will structural elucidation of COIl-ligand-JAZ complexes. It is interesting to note that some JAZ genes in which the Jas intron is correctly spliced give rise to proteins that lack the PY motif (Fig. $4.3). The PEAPOD (PPD) members of the TIFY family also contain a Jas-like domain lacking the FY motif (Chung et al., 2009). Genetic analysis has shown that PPDs 168 regulate leaf size and shape in Arabidopsis (White, 2006). There is no evidence to indicate, however, that PPDs play a role in JA responses or function as substrates for SCFCO“. It is also possible that alteration of the Jas domain via intron retention affects protein subcellular distribution. A recent study provided evidence that the PY motif of JAZI is part of a larger nuclear localization signal (NLS) that includes basic amino acids within the Jas domain (Grunewald et al., 2009). Interestingly, NLSs having an overall basic character followed by a C-terminal R/K/HX2-5PY motif have been identified in yeast (Lee et al., 2006; Siiel and Chook, 2009). A dual role for the Jas domain as a C01]- interacting degron and an NLS (Grunewald et al., 2009) raises the possibility that modification of the domain via Jas intron retention affects both JAZ stability and localization. In the case of JAZlO.3, however, removal of the PY motif did not prevent nuclear targeting of the protein (Chung and Howe, 2009). Based on the recent findings of Grunewald et a1 (2009), additional studies are needed to determine whether splice variants produced by Jas intron retention are affected in their distribution within the nucleus. Regulation of JA Responses by Alternative Splicing Previous analysis of JAZ alternative splicing relied on the use of transgenic plants that overexpress cDNAs encoding individual splice variants (Yan et al., 2007; Chung and Howe, 2009). To determine whether these isoforms are produced in vivo from JAZIO pre- mRNA, we generated and analyzed 35S-JAZIOG transgenic plants that constitutively express a genomic copy of JAZIO. The JA-insensitive root growth phenotype of these 169 lines indicates that alternative splicing of JAZIO pre-mRNA results in production of splice variants that are resistant to JA-induced degradation. The fact that the strength of the 35S-JAZIOG root growth phenotype was more similar to 3SS—JAZIO.4 plants than to 35S-JAZIO.3 plants indicates that this effect likely involves JAZ10.4. This finding supports our previous proposal (Chung and Howe, 2009) that induced expression of JAZIO in response to wounding (Yan et al., 2007; Chung et al., 2008) or other inductive cues is sufficient to attenuate JA signaling through the production of stabilized JAZ10 isoforms. The identification herein of additional splice variants (JAZZ.Z and JAZ3.4) that weakly interact with C01] suggests that Jas intron retention is part of a general mechanism of negative feedback regulation of JA signaling. Gene models described at TAIR provide evidence that alternative splicing events other than Jas intron retention may be involved in functional diversification of the JAZ family. For example, EST sequence information supports the existence of transcripts encoding two splice isoforms of JAZ3 (JAZ3.Z and JAZ3.3) that lack a segment of the N—terminus but retain the ZIM and J as domains. Functional analysis of these isoforms may provide new insight into the role of the N-terminal region of JAZ3 and other “ancient” JAZs that comprise JAZ subgroup A. Alternative splicing plays a central role in expanding protein diversity in eukaryotic organisms. Our results indicate that the Jas intron has been maintained during the evolution because its retention during pre-mRNA splicing diversifies the cellular repertoire of JAZs. This conclusion provides a counter argument to the view that alternative splicing plays a limited role in generating functionally diversified proteins in plants (Severing et al., 2009). Indeed, our results are more in keeping with the view that 170 alternative splicing optimizes plant adaptation to environmental stress (Reddy, 2007; Barbazuk et al., 2008). Based on the results presented here and elsewhere (Yan et al., 2007; Chung and Howe, 2009), we propose that splice variants generated by Jas intron retention are more stable in the presence of J A and therefore act, in a dominant negative fashion, to attenuate JA-induced changes in gene expression. The high fitness cost associated with expression of JA-regulated defenses in the absence of stress (Baldwin, 1998) may explain why conditional expression of stabilized JAZs would be advantageous to the plant. Efficient mechanisms to attenuate JA responses may also serve to protect the cell from toxic effects associated with hyperactivation of the JA pathway. In addition to hormone desensitization, production of JAZ isoforms having a wide range of stabilities may increase versatility in J A responses. It is also possible that alternative splicing of JAZ genes is regulated in a tissue or cell type-specific manner. Several examples of tissue-specific mRNA splicing have been described in plants (Reddy, 2007; Fang et al., 2004; Zhang et al., 2009). Our finding that overexpression of the JAZIO gene results in strong JA insensitivity in roots but presumably not in flowers (as determined by fertility) suggests that JAZIO pre-mRNA may be differentially spliced in a tissue-specific manner. Additional studies are currently in progress to address this hypothesis. 171 MATERIALS AND METHODS Plant Material and Growth Conditions ArabidOpsis thaliana ecotype Columbia (Col-0) was used as the wild type for all experiments. Soil-grown plants were maintained in a growth chamber at 21°C under 16 h light (100 uE rn’2 s") and 8 h dark. Arabidopsis was transformed with Agrobacterium tumefaciens (strain C58Cl) using the floral dip method (Clough and Bent, 1998). Transformed lines (T1 generation) were selected on MS plates containing kanamycin (50 ug/mL) and vancomycin (100 ug/mL). F ifly independent Tl plants per genotype were transferred to soil for subsequent phenotypic analysis. Homozygous lines were identified by testing T3 progeny for resistance to kanamycin. Identification of Alternatively Spliced JAZ Transcripts JAZ transcripts containing or not containing the Jas intron were amplified by RT-PCR from RNA prepared from x-day-old wild-type seedlings treated with 100 uM MeJA for 2 h as previously described (Chung et al., 2008?). PCR reactions were performed with Taq DNA polymerase (Invitrogen) and the transcript-specific primer sets listed in Table S2. PCR products were cloned into pGEM-T easy vector (Promega) and sequenced with T7 and SP6 primers. Yeast Two-Hybrid Assays Cloning of 12 JAZ cDNAs and cloning into pB42AD vector is described in Chung and Howe (2009). cDNAs encoding JAZZZ and JAZ3.4 were subcloned into the Matchmaker pB42AD vector (Clonetech) to generate JAZ fusions with the B42 activation domain 172 (AD). Oligonucleotide primers used to generate these constructs are listed in Table 82. Yeast transformation, selection of transformants, and ligand-dependent yeast two-hybrid assays were conducted as described by Melotto et al. (2008) and Chung and Howe (2009). Coronatine (Sigma-Aldrich) or stereoisomers of JA-Ile ((3R,7S)-, (3S,7S)-, (3S,7R), and (3R,7R)-JA-Ile) was added to the medium before the yeast cells were plated. Stereoisomers of JA-Ile were preare described by Ogawa and Kobayashi (2008). Expression of the BD fusion JAZ proteins was detected by protein gel blot analysis using anti-hemagglutinin (HA; Covance) antibody. Construction of 35S-JAZMG Full-length JAZIO genomic DNA was amplified from DNA prepared from Arabidopsis rosette leaves. PCR reaction was performed with Pfu Turbo DNA polymerase (Stratagene) and the primers listed in Supplemntal Table 2. The amplified JAZIO genomic DNA was subcloned into pGEM-T Easy (Promega), followed by recloning into the BamHI site of pBI-TS (Schilmiler et al., 2007). Root Growth Inhibition Assay Seeds were surface sterilized with 30% (v/v) commercial bleach for 15 min and washed 10 times with sterile distilled water. Seeds of each genotype were placed on square plates (Fisher) containing MS medium that was supplemented with 50 uM MeJA. Plates were incubated at 4°C for 4 d in darkness and then incubated under normal growth conditions for the remainder of the experiment. Plates were oriented in a vertical position for 9 d prior to measurement of primary root length. 35S-JAZ]0.I, 35S-JAZIO.3, and 355- 173 JAZIO.4 transgenic lines, which exhibit different levels of JA sensitivity (Chung and Howe, 2009), were used as controls. Identification of JAZ Genes in Various Plant Species JAZ gene sequences were identified in publically available genome sequences of P. patens (Rensing et al., 2007), S. moellendorflii (htggfl/genomejgi- psforg/Selmol/Selmol.home.html), 0. sativa (Goff et al., 2002; ref), B. distachyon (http://www.brachybaseorg:8075/), P. trichocarpa (Tuskan et al., 2006), and V. vinifera (http://wwwncbi.nlm.nih.gov/genome/secfl3lastGen/BlastGen.cgi?taxid=Z9760). Using Arabidopsis JAZI as a query, databases (with default algorithm parameters) were searched with BLASTP and TBLASTN. Retrieved JAZ sequences for a particular species were then used to re-query the database for that species. Following removal of redundant sequences, all unique protein sequences were manually annotated for the presence of the ZIM and Jas domains. The intron-exon structure of all candidate genes was assessed with the genome browser in the respective database, as well as by ClustalW-based alignment of the genomic DNA sequence with predicted coding sequence of the gene. , A Hidden Markov Model gene prediction program (HMM-FGENESH) available at Soflberry (http://linuxl.softberry.com/berry.phtml) was used to confirm the structural organization of all genes. Sequence Alignment and Phylogenetic Analysis Full-length JAZ protein sequences were aligned and analyzed with the ClustalW program and Bioedit v.7.0.9 software (Thopmson et al., 1 994; 174 http://wwwmbioncsu.edu/BioEdit/bioedit.html). The resulting multiple sequence alignment was used for construction of a phylogenetic tree with the neighbor-joining (NJ) method and MEGA4.1 software (http://wwwmegasofiware.net/; Tamura et al., 2007). Node confidence was assessed with 1,000 bootstrap replicates. 175 i—JAZ1 . *- i—JAZZ 1.- H'l I 1 _—-—JA25 -—I- - - -——JAZO ~ H- - ——JAZ11 :I-fll—J—I I I i—JAZ12 El+—I.l I. JAZ10.1 'I-I-I—I + I JAZO ml : i—JAZ3 ' I————ml. _ i—JAZA I “II. I .__JAz7 2...; .__Jm H... — 0.2kb Jas Intron lneertlon Figure 4.1. Structural organization of Arabidopsis JAZ genes. A) Phylogenetic tree constructed by the neighbor-joining method from the amino acid sequence of JAZ proteins (lefi), and the intron-exon organization of the corresponding genes (right). Thick black and gray bars indicate coding regions and non-coding untranslated regions in exons, respectively. Introns are depicted by a thin black horizontal line or, in the case of the Jas intron, 3 blue line. Sequences encoding the TIFY motif and Jas domains are shown in red and blue, respectively. B) Web logo (Crooks et al., 2004) constructed for the Jas domain of 12 Arabidopsis JAZ proteins. The location of conserved Jas intron is indicated by the arrow. In 7 of the 9 Jas intron-containing genes, the N-terrninal l8 (JAZ9), l9 (JAZZ, JAZIO, JAZl 1, and JAZlZ)), or 20 (JAZB and JAZ4) amino acids of the Jas domain, ending in the conserved KRKXR motif, are encoded by a single exon (underlined). 176 +2 JAZZ —ITGCATTTTAATGAAAACATAATAT R F L E K R K D R +2 JAZ3 _5TACGCAACACTTCTTLZEAATACACCA R F L E K R K E R Y A L ’ T L +2 JAZ4 —GTACE€£GCTACAAGATTATTCACTTAT R F L E K R K E R Y ' ‘d TATTTAACCTTATCATACTTTTTGAAA R F F A K R K D R Y L T L S Y F L K +2 JAZB —GmAcrAAAcrcccAmcrcnnsr R F F A K R K D R ’ +2 JAZ! TTTGATTTTGTATTTTTTTTCTTTATA R F L E K R K E R F D F V F F F F I +2 JAZ“ _GIETGATTCTTCAACAATCCAAGGM a r r A K a K o R + +2 JAZ11 —GIZETTAGATATTCAGTGrrchrrrA R F F E K i! R H R ‘ +2 JA212 “TATGCAGCCTTTAAAATTACACTTCAT R r L E K a R o a v A A F K I T L H Figure 4.2. Retention of the Jas intron during pre-mRNA splicing is predicted to alter the Jas domain. Exon and Jas intron sequences are highlighted in black and gray, respectively. Only the last 26 nucleotides of the exon encoding N-terrninal portion of the Jas domain (ending in the KRKXR motif) are shown. In all cases, the Jas intron disrupts the exon in the +2 position. The predicted amino acid sequence is shown below the genomic DNA sequence. In the case of JAZZ, JAZ3, JAZ4, JAZ6, JAZIO, and JAZI 1, an in-frame stop codon (highlighted in dark gray) is found near the 5’ end of the Jas intron. 177 JAZZ JAZ3 JAZ4 JAZ5 JAZ6 JAZ9 JAZ10 JAZ11 JAZ12 Figure 4.3. PCR-based detection of Jas intron retention in various JAZ transcripts. RT-PCR was used to amplify JAZ transcripts in which the Jas intron was either spliced out (A) or retained (B). Gene-specific primers and RNA from MeJA-treated Arabidopsis seedlings was used as a template for reactions (see Methods). PCR-amplified products were separated by agarose gel electrophoresis and visualized by staining with ethidium bromide. 178 JAZ1 JA22 JAZ: JA24 JA25 JA26 JAZ7 JAZ8 JA29JA210. 1 JAZ11JA212 - 77W7x. 1'. - - , . 1' o — . . {' _ V ' J Mock *--' 7.17.:- 77171 .7. “rs-5" i:’.'m~.""“2’37’7~¥~, ‘3? '“ m *7 .TTA' :7. :7.-."'~*<:.:--',—a-.“7r,—-a: or“. - I - ' COR (9”) 0 1 5 10 30 100 , ' *Wr‘vmwmr“ ‘7 JA22.1 (\l ' . . .. JAZZZ ( 7777777) ALWA‘JAMAMA JAZ3.1( ,- » . 3'3; 4‘”) :25) . A. IA‘JA‘J‘U‘J Figure 4.4. Differential ligand-dependent interaction of JAZ proteins with C011. (A) Full-length JAZ proteins (expressed in pB42AD) were tested in the YZH system for interaction with C01] (expressed in pBILDA) in the presence of 50 uM coronatine (COR; top panel) or 100 uM (3R,7S)-JA-Ile (JA-Ile; bottom panel). (B) Full-length JAZZ (JAZZ.1) and JAZ3 (JAZ3.1) proteins, together with their respective splice variants (JAZZ.Z and JAZ3 .4) derived fi'om retention of the Jas intron (Fig. 2 and 3), were tested for C011 binding in YZH assays performed in the presence of the indicated concentration of COR (left panel) or (3R,7S)-JA-Ile (right panel). 179 WT JAZ101 JAZTOJ JAZ704 JAZ10G-2JAZ1OG-4JA210G5 .5 01 10 Root length (mm) WT JAZ10.1 JAZ10.3 JAZ10.4 JAZ1OG-2 JAZ1OG-4 JAZ1OG-5 Figure 4.5. Overexpression of JAZIO genomic DNA attenuates JA signaling in Arabidopsis. (A) Wild-type (WT) and transgenic seedlings of the indicated genotype were grown for 9 d on MS medium containing 50 uM MeJA. JAZIOJ, JAZlO.3, JAZIO.4 refer to control transgenic lines overexpressing cDNAs that encode the JAZlO.1, JAZlO.3, and JAZIO.4 splice variants, respectively. The JA-mediated root growth inhibition phenotype of these lines was compared to that of three independent transgenic lines (JAZIOG-Z, JAZI 06-4, and JAZIOG-5) overexpressing the JAZIO genomic DNA. (B) Quantification of JA-induced root growth inhibition of seedlings shown in A). Data show the mean :I: SD (n = 11 seedlings per genotype). 180 Figure 4.6. Phylogenetic tree of JAZ proteins from higher and lower in land plants. Unrooted neighbor-joining tree of predicted JAZ proteins from A. thaliana (black), P. patens (red), S. moellendorfiii (orange), 0. sativa (blue), B. aystachyon (light blue), P. trichocarpa (green), and V. vinifera (purple). Multiple sequence alignment was performed with ClutalW. Four JAZ subclades are indicated as A-D. The Arabidopsis JAZ nomenclature is as previously described (Thines et al., 2007; Chini et al., 2007). The nomenclature for all other JAZs is based on the phylogeneic classification within subclades A-D. 181 A B C D 0 91 64 0 3 2 12mm 334 mmuzaumnmzznngifé asnnuunnn1291u ....m_...,.J 1mmmmn1? sum 7.7.. 7.7.. AAAAAAAZA Z7.— 2 7.. 7.7.7.. 7.. AAAfingflAAanqnfinA “A“ JJJJJ JAJJA AAAAAAAAAA11Annn56nnnn12nnnnn1n“AT-Ban“ JJJJMJJJnmwwmmmJJN mmm m mJmmwpuwwmumfiuwnMJJmouJJuMJJuJuJJJrzerJJ mmWWJWWHMJJOO PPpWPPSSnams SMSSPPOBOOOBOBOOJ mBJJHMWJmeo mmmJJMWW z E? 182 33.25 MSSSNENAKA QAPEKSDFTR RCSLLSRYLK EKGSFGNIDL GLYRKPDSSL ALPGKFDPPG 31126 14er ------ QAPEKSNFSQ RCSLLSRYLK EKGSPGNINM GLARKSD--L ELAGKFDLKG m5 KQNAMHKAGH sx ------------ csps'rs sccm--nv ADLSESQP-- GSSQLTIFFG m6 QQNVIKKVET SETRPPKLIQ Krsrcnas'rs TEDKAIYIDL SEPAKVAPES GNSQLTIFFG m5 GKVLVYNEFP VDKAKEIMEV AKQAKPVTEI NIQTPINDEN NNNKSSMVLP DLNEPTDNNH JAZG GKVMVFNEFP Bumsmzv AREANHVAVD smsuosan NLDKSNWIP DLNEPTSSG- JAZS LTKEQQQQQE QNQIVERIAR RASLHRFFAK RKDRAVARAP YQVNQNAGHH RYPPKPEIVT ms -NNEDQETGQ QHQVVERIAR RASLHRFFAK RRDRAVARAP YQVNQHGSH- ~LPPKPEMVA m5 GQPLEAGQSS QRPPDNAIGQ rmrrtsnco KDDIMKIEEG OSSKDLDLRL" JAZ6 PS-IKSGQSS QHIATPPKPK AHNHMPMBVD K ------ KEG QSSKNLELKL“ Exon1 Exon2 Exon3 Exon4 Supplemental Figure 4.1. Amino acid sequence alignment of ArabidOpsis JAZS and JAZ6. Both genes contain three introns and four exons. Amino acids encoded by each exon are indicated with different colors. 183 AD-JAZ proteins 123456789 101112 Supplemental Figure 4.2. Western blot analysis of JAZ proteins expressed in the yeast strains used for the Y2H experiment shown in Figure 4.4A. Total protein extracted from yeast cells expressing activation domain (AD)-JAZ fusion proteins (JAZl-JAZlZ) and the DNA binding domain (BD)-C011 fusion protein was analyzed by western blot analysis. D-JAZ proteins were detected with an anti-HA antibody. 184 Supplemental Figure 4.3. Amino acid sequence alignment of the ZIM and Jas domains of JAZ proteins from various land plants. JAZI] contains an internal duplication and is predicted to encode a protein with two ZIM domains and two Jas domains. The ZIM and Jas domains located at the C-terminal end of JAle were used for the alignment. See Table 81 for additional information. 185 rJA:h< ;r.l,\:[ir 7r.lJ,:l-7 :IJATDH P'JAILQ p 311' I?" 186 4 ITAKEAA'E~-— . 3.7!;1.“ .._ :l-- 13-- TIK‘Si-K— » --”Lgucn2- LVSKJ'I- ~ -:vvaz—-rt —ZTTIK—-FY vaa‘ -—— ‘ ‘HVTHI — - ~- J .......... loauafl||—-» yvarannnrv Ira - 33 l. J. A A y. l l PpJAZA4 “Hm - SmJAZAQ -*I—-l-- OsJAZ3 I———--J-—I - AtJAZ3 I I If”! I PtJAZM r m- — I VVJAZA4 I "m- "- I Supplemental Figure 4.4. Structural organization of selected JAZ genes in subclade A. Thick black and gray bars indicate coding regions and non-coding untranslated regions in exons, respectively. Introns are depicted by a thin black horizontal line or, in the case of the Jas intron, a blue line. Sequences encoding the TIFY motif and Jas domains are colored red and blue, respectively. The 3’ end of the genes (boxed), including the last 4 exons and three introns, show the highest level of structural conservation. 187 Table 4.1. Identification of JAZ genes and the conserved Jas intron in various plant species Species Number of JAZ genes Number ofJAZ genes with the Jas intron (phase) P. patens 9 7 (+2) S. moellendorflii 9 9 (+2) 0. sativa 15 9 (+2) B. dystachyon 4 3 (+2) A. thaliana 12 9 (+2) P. trichocarpa 14 12 (+2) V. vinifera 9 7 (+2) Total 72 56 188 Z .mm<>4a<<“: NN >m<¥w>._mmommxwdmOJwImmSaqo 60E... PNSamO oummmmeomo $58 «$60 .mmxmmcammhqm > mimggmmxmmcfimbm 0<>u_._. m xom¥omoxm—x>._umm_m<¥m<0>oz > mfimw5§m<0>oz 02>“..F ~On_._o > mmmZOmen—mmxfiumcamlxwmzvaa Gw>n=._. o mz>_>hOxOxm¥O.EmC_w<~—m<0a._0 0<>n=._. m m_>mv..0wmxmxw4um04w<¥m<0a..< 00>...2h v m8 w< w<n=2h F >m_..>.5moxmv_w..um<._m_>v.v.>moxmv_w..n_m<._wu_._. waoxx>mo¥m¥w4um<4w wmocwmwmxmmoqumjmzxm’xzaan. wot... m4mxmw04umm4mmxm ommoomxmxmwOJummemxmghd 09E; m >awhm>mommmwdmfimu_._. vmw0mommmw..n_m<._w >amhmm:. mmoxmxo._um >a0xm>moxmxofimjwn=m Nmoxmxw4umjw<¥m<0a4w > >a0xm>moxmxw4um<4wu=._. Ni. toe—5: .0 «33% 38...; E tout—.02 8.39:— N<—. he 3: ._.v «Ea... 13582.3.5 9 8 l .moxmfimmmzqmém<_mw> . . .w4h4>moxm¥ «>mmxmxw4umngmk0Q46 .44h<>mm¥m¥m._umm0xm¥w4mmIqmzmm—um04m>¥m>>m._0 .mzxmxcqumOJwfim—Saim .mmuxmxwgn—m Imwzmmfiago .moxmquumIgwngQED .moxmxmqumIAm/xxméaqo . . .ZOOm>mw¥mxw4mm04m¥hE4mwx .mmxmxwgum5w<¥m< . . .IOmx>mmxmv_m4mKSw<¥w—< z>->->->->-z z>->->- >->->->->->-zzz>-z>-z >n_w._.<0_m..mmx0._n_w1._wmx2m<¥0> >a >a >n_w>z_>mw¥mv_m._umam>m§mmxmxmqmm<4m<¥mjm4< >Q<¥wtm0¥mxm._u_ml._ma<¥mgmoxm¥m4um14wn__momm¥w..u_m0._m1114.540 >Q<¥<_>mOxmxmdmmzawzmmgmmo >QKIOO¢IxKXIqumOJw<¥mSQAD >Q.=I>>m_2mmm¥w>n_m04w>¥m>>n30 >a00m<<4mOxmxmfium04m<¥m<>mgo >ame.mI¥m—¥O.Emci_w>_mzxmxwgumcmmngQED >at§moxmxwwmQn_._.xa<¥<04moxm¥m._u_mrqm<¥m<0n_..o >n_0...mhqmwxmxwmmmOJmXPKPAwa >Q< >a<>mm>mm¥mxw4mm<4w<¥m< 02>..E. Own—u; 00mm; 0<>n=h O<>n=._. 00E: 0<>n=H 00>“.E. 00>n=h ODE—2h 00>H=H 00>): 00>S=¥ 00>..E. 00E: 00E: 0®>n=h 0<>n=k 00>uz. 00:2. 00>n=h 02>u_._.r 0<>n=._. 00E; 5N3: 0N3. mN<... VNS. mNs. NN<... ..NS: 3225 mmNSabm NmN<2vm FmNsfim m PNmw¥mv.w._mm<..m<¥m<0a>< .mmxmxwgumfim< .mmxmquumjm<¥m<0n.>< .KMXKXMJumSm<¥m< . . .u4¥I>moxxxwm>m04w<¥mmmew .moxmquumIAwémSnfio .KGXKXMAumOJWXKmsQAD .mmVEXMJLKCvaEm/xaao .mommXqumOAmOmm’x—nfiw . . . .Ou2>mommv.mmumOAwImmfinfim .>mwv.mv.mun.mr..mo .>m0¥mxmu.um_14mo ...w><>>mw¥m¥m..um<..m<¥m<0n_>< . . .._.><>n_mwv.mv_w._n.m< .mwxmqumm5m<¥m<0n_>< .mmxmxmqmmsmwgméngm ...xu<<>mommxw4um04mImmSnfio .mImmxwmmmOJmmmm<_a>o .mwxmv. Mgumcgmxmmgnfio . . .uioummxmxw4m15m<¥mm >->->->->->-zz>->->->->->->->->->-> z>>>->- >m<¥mz>moxmxw4umIAmZm—msamw >am< >Qm< >n.<._.w_2>m wxmxwmum5m<¥m< >02.V..Cmmv.mv.m..n_m< >v.m.v.Z¥u¢0¥¥¥Mt>m04w<¥mWrw0w >n.am>wh4mmexwgumGwamm/xiao >n.._.n.mxm>4momm¥¥uum04wcmm5n...w >am¥m>qmommxwmumOAWImmSQJm. Daub/‘03I¥m¥¥4um04wm¥2Q40>a >n.w_amw¥mxmumm14mo >Qmoxm¥mumm14mo >th02>mw¥m¥w4umjw<¥m< >awhcgmm¥m¥mJumSw<¥m< >Q<mmxm¥m..u.m<..w<¥m<0n_>w >n. wghéwxmquum5w<¥m<0a>m >n.Z¥Z>Jmomm¥wqumOJmIm—mfiafiu >Q¥HI>uEImmxmmumOmemmSnso >n.m._.m>..mmxmxmqmmOAmxmm—lximo >Q2m >n.m._.<0.m.xmxcmuwcqwxxiwgaz 0<>...=._. 0<>m_._. 0<>n=... 0<>n=._. 0<>n=._. 00E... 0<>n=h OZ>u:... 02C... G0>u=h 00E; 02>n:... 02>...=... O<>n=._. 0<>u=._. 0<>n=h 0<>n=._. O<>n._._. 0.4:... 06mm... 00...... 02>n._._. 00E... 02>n._._. mmN> v> m> N> _.> v rmwxmxmn_um1._m¥mm<._a._o .moxmmemmIAw<¥m<_n3m >-ZZ>- >Q>>mh4mwmeMAmmIvafi—mjala >Qwhmwhmn=k 00>..E. m0N<...>> wON<_.>> BON<6>> omN<...>> 355:3 3. «Ban. _5=oEo=—n=m w 25¢»st Nmmommumw commons: wmvmvmsmr I92 :92: mag 55 flatomcmb N2... he Eon—tn 85>":— 9 EEE mg :55; muatomcmb N3. .5“ MEET—Q 0833. a .m-O3 IN> NFNS. :NS. 2N5. 9N3. 0N5. nNS. vNi, m N S. NN$.. mums—k anEtm 0293. .808...“ 3.261 905...". p.220“. >05... £5 5 tom: £25.:— uc an: N6 «3:. .3550395 193 REFERENCES Abe, H., Urao, T., Ito, T., Seki, M., Shinozaki, K. and Yamaguchi-Shinozaki, K. (2003). Arabidopsis AtMYCZ (bHLH) and AtMYBZ (MYB) function as transcriptional activators in abscisic acid signaling. Plant Cell 15: 63-78. Reusing, S.A., Lang, D., Zimmer, A.D., Terry, A. et al. (2008). The Physcomitrella genome reveals evolutionary insights into the conquest of land by plants. Science. 319: 64-69. Baldwin, LT. (1998). Jasmonate-induced responses are costly but benefit plants under attack in native populations. Proc Natl Acad Sci USA. 95: 8113-8118. Barbazuk, W.B., Fu, Y., and McGinnis, K.M. (2008). Genome-wide analyses of alternative splicing in plants: Opportunities and challenges. Genome Res. 18, 1381-1392. Browse, J. (2009). Jasmonate Passes Muster: A Receptor and Targets for the Defense hormone. Annu. Rev. Plant Biol. 60: 183-205. Chico, J.M., Chini, A., Fonseca, S., Solano, R. (2008). JAZ repressors set the rhythm in jasmonate signaling. Curr. Opin. Plant Biol. 11: 486-494. Chini, A., Fonseca, S., Fernandez, G., Adie, B., Chico, J.M., Lorenzo, 0., Garcia- Casado, G., Lopez-Vidriero, 1., Lozano, F.M., Ponce, M.R., Micol, J.L., and Solano, R. (2007). The JAZ family of repressors is the missing link in jasmonate signalling. Nature 448: 666-671. Chini, A., Fonseca, S., Chico, J., Fernandez-Calm, P., Solano, R. (2009) The ZIM domain mediates homo- and heteromeric interactions between Arabidopsis JAZ proteins. Plant J. (in press) Chung, H.S., Koo, A.J.K., Gao, X., Jayany, S., Thines, B., Jones, A.D., Howe, G.A. (2008). Regulation and function of Arabidopsis JASMONA TE ZIM-domain genes in response to wounding and herbivory. Plant Physiol. 146: 952-964. Chung, H.S., Howe, G.A. (2009). A critical role for the TIFY motif in repression of jasmonate signaling by a stabilized splice variant of the JASMONATE ZIM- domain protein JAZIO in Arabidopsis. Plant Cell 21: 131-145. Conconi, A., Smerdon, M. J., Howe, G. A. and Ryan, C. A., (1996). The octadecanoid signalling pathway in plants mediates a response to ultraviolet radiation. Nature 383: 826-829. 194 Creelman, R.A., Mullet, J.E. (1997). Oligosaccharins, brassinolides, and jasmonates: nontraditional regulators of plant growth, development, and gene expression. Plant Cell 9: 1211-1223. Devoto, A., Ellis, C., Magusin, A., Chang, H.S., Chilcott, C., Zhu, T.,and Turner, J.G. (2005). Expression profiling reveals C011 to be a keyregulator of genes involved in wound- and methyl jasmonate-inducedsecondary metabolism, defence, and hormone interactions. Plant Mol. Biol. 58: 497—513. Dombrecht, B., Xue, G.P., Sprague, S.J., Kirkegaard, J.A., Ross, J.J., Reid, J.B., Fitt, G.P., Sewelam, N., Schenk, P.M., Manners, J.M., et al. (2007). MYC2 differentially modulates diverse jasmonate-dependent functions in Arabidopsis. Plant Cell 19: 2225—2245. Fang, Y., Hearn, S., Spector, D.L. (2004). Tissue-specific expression and dynamic organization of SR splicing factors in Arabidopsis. Mol. Biol. Cell. 15: 2664- 2673. Feys, B., Benedetti, C.E., Penfold, C.N., Turner, J.G. (1994). Arabidopsis mutants selected for resistance to the phytotoxin coronatine are male sterile, insensitive to methyl j asmonate, and resistant to a bacterial pathogen. Plant Cell 6: 751—759. F onseca, S., Chini, A., Hamberg, M., Adie, B., Porzel, A., Kramell, R., Miersch, 0., Wastemack, C., Solano R. (2009). (+)-7-iso-Jasmonoyl-L-isoleucine is the endogenous bioactive jasmonate. Nat. Chem. Biol. 5: 344-350. Glazebrook, J. (2005). Contrasting mechanisms of defense against biotrophic and necrotrophic pathogens. Annu. Rev. Phytopathol. 43: 205—227. Goff, S.A., Ricke, D., Lan, T.R., Presting, G. et al. (2002). A draft sequence of the rice genome (Oryza sativa L. ssp. japonica). Science. 296: 92-100. Griffiths, S., Dunford, R.P., Coupland, G., Laurie, D.A. (2003). The evolution of CONSTANS-like gene families in barley, rice, and Arabidopsis. Plant Physiol. 131: 1855-1867. Grunewald, W., Vanholme, B., Pauwels, L., Plovie, E., Inzé, D., Gheysen, G., Goossens, A. (2009). Expression of the Arabidopsis jasmonate signalling repressor JAZ1/TIFY10A is stimulated by auxin. EMBO Rep. 10: 923-928. Howe, G.A., Jander, G. (2008). Plant immunity to insect herbivores. Annu Rev Plant Biol. 59: 41-66. Iida, K., Go, M. (2006). Survey of conserved alternative splicing events of mRNAs encoding SR proteins in land plants. Mol. Biol Evol. 23: 1085-1094. 195 Irimia, M., Rukov, J.L., Roy, S.W., Vinther, J., Garcia-Fernandez, J. (2009). Quantitative regulation of alternative splicing in evolution and development. Bioessays. 1: 40-50. Kalyna, M., Lopato, S., Voronin, V., Barta, A. (2006). Evolutionary conservation and regulation of particular alternative splicing events in plant SR proteins. Nucleic Acids Res. 34: 4395-4405. Katsir, L., Chung, H.S., Koo, A.J.K., Howe, G.A. (2008a). Jasmonate signaling: a conserved mechanism of hormone sensing. Curr. Opin. Plant Biol. 11: 428-435. Katsir, L., Schilmiller, A.L., Staswick, P.E., He, S.Y., Howe, G.A. (2008b). C011 is a critical component of a receptor for jasmonate and the bacterial virulence factor coronatine. Proc. Natl. Acad. Sci. USA 105: 7100-7105. Laurie-Berry, N., Joardar, V., Street, I.H., Kunkel, B.N. (2006). The Arabidopsis thaliana JASMONATE INSENSITIVE 1 gene is required for suppression of salicylic acid-dependent defenses during infection by Pseudomonas syringae. Mol Plant Microbe Interact. 19: 789-800. Leeman, J.R., Gilmore, TD. (2008). Alternative splicing in the NF-kappaB signaling pathway. Gene. 423: 97-107. Li, L., Zhao, Y., McCaig, B.C., Wingerd, B.A., Wang, J., Whalon, M.E., Pichersky, E., Howe, G.A. (2004). The tomato homolog of CORONATINE- INSENSITIVEI is required for the maternal control of seed maturation, jasmonate-signaled defense responses, and glandular trichome development. Plant Cell 16: 126-143. Lorenzo, 0., Chico, J. M., Sanchez-Serrano, J. J. and Solano, R. (2004). JASMONATE-INSENSITIVEI encodes a MYC transcription factor essential to discriminate between different jasmonate-regulated defense responses in Arabidopsis. Plant Cell 16: 1938-1950. Mandaokar, A., Thines, B., Shin, B., Lange, B. M., Choi, G., Koo, Y. J., Yoo, Y. J., Choi, Y. D. and Browse, J. (2006). Transcriptional regulators of stamen development in Arabidopsis identified by transcriptional profiling. Plant J. 46: 984-1008. Melotto, M., Mecey, C., Niu, Y., Chung, H.S., Katsir, L., Yao, J., Zeng, W., Thines, B., Staswick, P.E., Browse, J., Howe, G.A., He, S.Y. (2008a). A critical role of two positively charged amino acids in the Jas motif of Arabidopsis JAZ proteins in mediating coronatine- and jasmonoyl isoleucine-dependent interactions with the C011 F -box protein. Plant J. 55: 979-988. 196 McConn, M., Browse, J. (1996). The Critical Requirement for Linolenic Acid Is Pollen Development, Not Photosynthesis, in an Arabidopsis Mutant. Plant Cell. 8: 403- 416. Ogawa, N. and Kobayashi, Y. (2008). Strategy for synthesis of the isoleucine conjugate of epi-jasmonic acid. Tetrahedron Lett. 49: 7124—7127. Pan, Q., Shai, 0., Lee, L.J., Frey, B.J., Blencowe, B.J. (2008). Deep surveying of alternative splicing complexity in the human transcriptome by high-throughput sequencing. Nat Genet. 40: 1413-1415. Reddy, A.S.N. (2007). Alternative splicing of pre-messenger RNAs in plants in the genomic era. Annu. Rev. Plant Biol. 58: 267-294. Reymond, P., Weber, H., Damond, M., Farmer, E.E. (2000). Differential gene expression in response to mechanical wounding and insect feeding in Arabidopsis. Plant Cell 12: 707—720. Reymond, P., Bodenhausen, N., Van Poecke, R.M., Krishnamurthy, V., Dicke, M., Farmer, E.E. (2004). A conserved transcript pattern in response to a specialist and a generalist herbivore. Plant Cell 16: 3132—31347. Robson, F., Costa, M.M., Hepworth, S.R., Vizir, 1., Pifleiro, M., Reeves, P.H., Putterill, J., Coupland, G. (2001). Functional importance of conserved domains in the flowering-time gene CONSTANS demonstrated by analysis of mutant alleles and transgenic plants. Plant J. 28: 619-631. Sasaki, Y., Asamizu, E., Shibata, D., Nakamura, Y., Kaneko, T., Awai, K., Amagai, M., Kuwata, C., Tsugane, T., Masuda, T., et al. (2001). Monitoring of methyl jasmonate-responsive genes in Arabidopsis by cDNA macroarray: selfactivation of jasmonic acid biosynthesis and crosstalk with other phytohormone signaling pathways. DNA Res 8: 153—161. Sasaki-Sekimoto, Y., et al. (2005). Coordinated activation of metabolic pathways for antioxidants and defence compounds by jasmonates and their roles in stress tolerance in Arabidopsis. Plant J. 44: 653—668. Schenk, P.M., Kazan, K., Wilson, 1., 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. USA 97 : 11655—11660. Schilmiller, A.L., Koo, A.J., and Howe, G.A. (2007). Functional diversification of acyl- CoA oxidases in jasmonic acid biosynthesis and action. Plant Physiol. 143: 812- 824. 197 Schmidt, S., Baldwin, LT. (2006). Systemin in Solanum nigrum. The tomato- homologous polypeptide does not mediate direct defense responses. Plant Physiol. 142: 1751-1758. Serrano, G., Herrera-Palau, R., Romero, J .M., Serrano, A., Coupland, G., Valverde, F. (2009). Chlamydomonas CONSTANS and the evolution of plant photoperiodic signaling. Curr Biol. 19: 359-368. Severing, E.I., van Dijk, A.D., Stickema, W.J., van Ham, RC. (2009). Comparative analysis indicates that alternative splicing in plants has a limited role in functionalexpansion of the proteome. BMC Genomics. 10: 154. Sineshchekov, V.A., Loskovich, A.V., Riemann, M., Nick, P. (2004). The jasmonate- free rice mutant hebiba is affected in the response of phyA'/phyA" pools and protochlorophyllide biosynthesis to far-red light. Photochem. Photobiol. Sci. 3: 1058-1062. Staswick, P.E. (2008). JAZing up jasmonate signaling. Trends Plant Sci. 13: 66-71. Strayer, C., Oyama, T., Schultz, T.F., Raman, R., Somers, D.E., Mas, P., Panda, S., Kreps, J.A., Kay, S.A. (2000) Cloning of the Arabidopsis clock gene TOCI, an autoregulatory response regulator homolog. Science 289: 768-771. Suzuki, H., Reddy, M.S., Naoumkina, M., Aziz, N., May, G.D.,Huhman, D.V., Sumner, L.W., Blount, J.W., Mendes, P., and Dixon,R.A. (2005). Methyl jasmonate and yeast elicitor induce differentialtranscriptional and metabolic re- programming in cell suspensioncultures of the model legume Medicago truncatula. Planta 220: 696—707. Takahashi, F., Yoshida, R., Ichimura, K., Mizoguchi, T., Seo, S., Yonezawa, M., Maruyama, K., Yamaguchi-Shinozaki, K., Shinozaki, K. (2007). The mitogen- activated protein kinase cascade MKK3-MPK6 is an important part of the jasmonate signal transduction pathway in Arabidopsis. Plant Cell. 19: 805-818. Tazi, J., Bakkour, N., Stamm, S. (2009). Alternative splicing and disease. Biochem. Biophys. Acta. 1792: 14-26. Thines, B., Katsir, L., Melotto, M., Niu, Y., Mandaokar, A., Liu, G., Nomura, K., He, S.Y., Howe, G.A., and Browse, J. (2007). JAZ repressor proteins are targets of the SCFCO“ complex during jasmonate signalling. Nature 448: 661-665. Uppalapati, S.R., Ayoubi, P., Weng, H., Palmer, D.A., Mitchell, R.E., Jones, W., and Bender, C.L. (2005). The phytotoxin coronatine and methyl jasmonate impact multiple phytohormone pathways in tomato. Plant J. 42: 201—217. 198 Vanholme, B., Grunewald, W., Bateman, Kohchi, T., Gheysen, G. (2007). The tify family previously known as ZIM. Trends Plant Sci. 12: 239-244. Wang, B.B., Brendel, V. (2006). Genomewide comparative analysis of alternative splicing in plants. Proc Natl Acad Sci U S A. 103: 7175-7180. Wenkel, S., Turck, F ., Singer, K., Gissot, L., Le Gourrierec, J., Samach, A., Coupland, G. (2006). CONSTANS and the CCAAT box binding complex share a functionally important domain and interact to regulate flowering of Arabidopsis. Plant Cell 18: 2971-2984. White, D.W.R. (2006). PEAPOD regulates lamina size and curvature in Arabidopsis. Proc. Natl. Acad. Sci. USA 103: 13238-13243. Yan, Y., Stolz, S., Chetelat, A., Reymond, P., Pagni, M., Dubugnon, L., and Farmer, E.E. (2007). A downstream mediator in the growth repression limb of the jasmonate pathway. Plant Cell 19: 2470-2483. Yoshida, Y., Sano, R., Wada, T., Takabayashi, J., Okada, K. (2009). Jasmonic acid control of GLABRA3 links inducible defense and trichome patterning in Arabidopsis. Development. 136: 1039-1048. Zhang, X.N., Mount, S.M. (2009). Two Alternatively Spliced Isoforms of the Arabidopsis SR45 Protein Have Distinct Roles during Normal Plant Development. Plant Physiol. 150: 1450-1458. 199 CHAPTER 5: SUMMARY AND FUTURE PERSPECTIVES 200 At the time this thesis research was initiated, little was known about the molecular mechanism of JAZ action in the regulation of IA responses. The study described in Chapter 2 was conceived with the goal of characterizing the expression pattern of Arabidopsis JAZ genes in response to mechanical wounding and herbivory and to test the model that JAZ proteins are destabilized by SCFCOIl-mediated ubiquitination in response to JA. Most JAZ genes exhibit increased expression in response to both mechanical wounding and feeding by S. exigua larvae (Fig. 2.1). The rapid expression of JAZ genes in response to mechanical wounding is dependent on C011 (Fig. 2.2) and is correlated with accumulation of JA and JA-Ile in damaged leaves (Fig. 2.3). This result is consistent with the hypothesis that JA-Ile is a wound signal that triggers degradation of JAZ repressors and subsequent transcription of early response genes. CHX-induced accumulation of JAZ transcripts is dependent on COIl (Fig. 2.5), which agrees with the idea that JAZ proteins are destabilized by SCFCOH-mediated ubiquitination (Chini et al., 2007; Thines et al., 2007). Perturbation of IA signaling by overexpression of JAZIA3A decreased plant resistance to S. exigua feeding (Fig. 2.7). This finding indicates that JAZ proteins play an important role in regulating the expression of plant processes that confer resistance to insect herbivores. Alternative splicing of pre-mRNAs provides a general mechanism to increase protein diversity from a limited number of loci and, ultimately, contributes to biological complexity. The work presented in Chapters 3 and 4 provides several independent lines of evidence that JAZ proteins are functionally diversified by alternative splicing. Chapter 3 describes the characterization of three JAZ10 splice variants differing in the sequence of the Jas domain. JAZlO.] contains an intact Jas domain that interacts with C011 in a 201 ligand-dependent manner. JAZIO.4, which lacks the entire J as domain, fails to bind C01 1, whereas J AZ10.3, which contains a partially truncated Jas domain, interacts weakly with COIl (Fig. 3.4). Analysis of JAZIO-YFP reporter lines provided direct evidence that JAZlO.] is degraded via the 26S proteasome pathway in response to JA treatment, and that JAZIO.4 is highly resistant to JA-induced destruction (Fig. 3.7). Consistent with C011-interaction studies, high concentrations of JA were required to induce turnover of JAZlO.3-YFP in viva. Consequently, overexpression of JAZlO.3 and JAZ10.4 conferred weak and strong JA-insensitive phenotypes, respectively, whereas 35S-JAZ10.1 plants that produce the full-length protein did not exhibit altered sensitivity to JA (Fig. 3.6). Interestingly, constitutive expression of JAZ10 pre-mRNA confered severe insensitivity to JA-induced root growth inhibition (Fig. 4.5). This finding indicates that one or more stable JAZIO variants (e.g., JAZ10.4) are produced by the endogenous splicing machinery and affect JA signal output. Comparative analysis of the structural organization of JAZ genes revealed that the position and phase of the intron involved in generating the JAZlO.3-encoding transcript (At5g13220.3) is conserved in 56 of 72 JAZ genes identified in diverse plant species (Tab. 4.1). Among the 56 JAZ genes that harbor the Jas intron (Tab. 4.1), 40 of these are predicted to encode proteins that terminate immediately or shortly after the KRKXR motif if the intron is retained during mRNA splicing (Tab. S1). The presence of the Jas intron in JAZ orthologs from P. patens suggests that this sequence arose early in the evolution of the JAZ family and has been retained in higher plants. Presumably, truncated JAZ isoforms generated by Jas intron retention participate in cellular processes that increase plant fitness. The reduced C011-binding capacity of JAZ2.2, JA23.4, and 202 JAZlO.3 provides evidence that Jas intron retention plays an essential role in functional diversification of JAZ proteins through increased protein stability and increased repression of JA signaling. The ZIM domain is the defining feature of the TIFY protein family that includes JAZ, PPD, and ZIM/ZML proteins (Vanholme et al., 2007). Experiments described in Chapter 3 demonstrate that the ZIM domain mediates homo- and heteromeric protein- protein interactions among JAZ proteins. Site-directed mutagenesis studies showed that the TIFY motif within the ZIM domain is a key determinant of JAZ3 and JAZ10 homo- and heteromerization (Fig. 3.2). Mutations in the TIFY motif that disrupt JAZ10.4 interaction with other JAZs also abrogate the dominant-negative effects of JAZIO.4 on JA signal output (Fig. 3.8), suggesting that a functional ZIM domain is critical for JAZ- mediated repression of JA signaling. Future Perspectives The research described in this thesis has led us to a better understanding of the molecular mechanism by which JAZ proteins regulate JA responses. However, many interesting questions remain to be answered. Differential binding of JAZ proteins to C011 (Fig. 4.4) suggests that individual members of the family have different stabilities in JA-stimulated cells. Based on current evidence that the Jas domain is sufficient for JAZ interaction with C011, it is likely that sequence variation in the Jas domain is a key determinant of JAZ stability. Site-directed mutagenesis studies (Melotto et al. 2008) are beginning to provide insight into regions of the domain that are required for C011 interaction. X-ray 203 crystallography studies are needed to provide more precise information about the structural basis of hormone-induced formation of C011-JAZ complexes. Modification of the Jas domain through alternative splicing further increases the diversity JAZs that differentially interact with C011. The involvement of alternative splicing in functional diversification of JAZ proteins increases our understanding of the complexity and dynamic nature of IA responses. Several studies have shown that differential pre-mRNA processing in plants is regulated by environmental cues and by cell type-specific factors (Lopato et al., 1999; Fang et al., 2004; Reddy et al., 2007; Barbazuk et al., 2008). Future studies of JAZ alternative splicing will benefit from quantitative analysis of the abundance of alternative transcripts, as well the development of approaches to detect specific protein variants in specific tissue and cell types. This information will ultimately lead to a better understanding of how the expression of diverse J A responses is temporally and spatially regulated. The ability of JAZ proteins to interact with one another (Chapter 3) raises the possibility that a network of JAZ-JAZ complexes differentially interacts with downstream transcription factors (TFs). Given the role of JA in such a wide range of plant biological processes, it is possible that different JAZ-JAZ complexes regulate distinct J A responses by interacting with different TFs that control specific subsets of JA response genes. It will be interesting to identify direct targets of each JAZ using ChIP experiments. Detailed phenotypic analysis of single and high-orderjaz null mutants will provide additional information about the biological relevance of JAZ-JAZ interaction in regulating various JA responses. 204 Although several TFs involved in IA signaling have been identified, MYC2 is the only TF known to be directly targeted by JAZs (Chini et al., 2007; Chung and Howe, 2009; Chini et al., 2009). Given that JA signaling in myc2/jin1 mutants is only partially impaired, other TFs must also be targeted by the SCFCOH/JAZ pathway. Establishment of functional links between JAZs and other TFs implicated in JA signaling, as well as identification of novel transcriptional regulator(s) whose function is controlled JAZs, are important directions for future research. The outpouring of genome sequence information from diverse plant species provides new opportunities to study the evolutionary roots of JA synthesis and signaling. In addition to containing several JAZ genes, P. patens also contains six F -box C011-like (F CL) genes (Chini et al., 2008). These observations, together with the absence of JAZ and F CL genes in photosynthetic aquatic eukaryotes, suggest that the JA signaling pathway arose during the evolution of early land plants. Although P. patens contains genes encoding plastidic enzymes of the octadecanoid pathway (e.g., allene oxide synthase/cyclase), it does not appear to produce JA or JA-Ile (Browse, 2009; Feussner, 2008). It is tempting to speculate that an ancient SCFCO'l/JAZ receptor system in P. patens recognizes an alternative oxylipin signal such as OPDA or an amino acid conjugated form of OPDA to regulate stress responses. Elucidation of the function and mechanism of action of JAZ and F CL genes in P. patens will provide important insight into evolution of the J A signaling pathway. 205 REFERENCES Barbazuk, W.B., Fu, Y., and McGinnis, K.M. (2008) Genome-wide analyses of alternative splicing in plants: Opportunities and challenges. Genome Res. 18, 1381-1392 Browse, J. (2009) Jasmonate Passes Muster: A Receptor and Targets for the Defense hormone. Annu. Rev. Plant Biol. 60: 183-205. Chini, A., Fonseca, S., Fernandez, G., Adie, B., Chico, J.M., Lorenzo, 0., Garcia- Casado, G., Lopez-Vidriero, 1., Lozano, F.M., Ponce, M.R., Micol, J.L., and Solano, R. (2007) The JAZ family of repressors is the missing link in jasmonate signalling. Nature 448: 666-671. Chini, A., Fonseca, S., Chico, J., Fernandez-Calvo, P., Solano, R. (2009) The ZIM domain mediates homo- and heteromeric interactions between Arabidopsis JAZ proteins. Plant J (in press) Chung, H.S., Koo, A.J.K., Gao, X., Jayany, S., Thines, B., Jones, A.D., Howe, G.A. (2008). Regulation and function of Arabidopsis JASMONA TE ZIM-domain genes in response to wounding and herbivory. Plant Physiol. 146: 952-964. Chung, H.S., Howe, G.A. (2009) A critical role for the TIFY motif in repression of jasmonate signaling by a stabilized splice variant of the JASMONATE ZIM- domain protein JAZ10 in Arabidopsis. Plant Cell 21: 131-145. Fang Y, Hearn S, Spector DL. (2004) Tissue-specific expression and dynamic organization of SR splicing factors in Arabidopsis. Mol Biol Cell. 15: 2664-2673. Feussner I. (2008) Oxylipins, evolutionary conserved signals in plants, mosses, fungi and bacteria. 18’” Int. Symp. Plant Lipids. Abstr. 036. http://www.isp12008- bordeaux.cnrs.fr/ Lopato S, Kalyna M, Dorner S, Kobayashi R, Krainer AR, Barta A. (1999) atSRp30, one of two SF2/ASF-like proteins from Arabidopsis thaliana, regulates splicing of specific plant genes. Genes Dev. 13: 987-1001. Melotto, M., Mecey, C., Niu, Y., Chung, H.S., Katsir, L., Yao, J., Zeng, W., Thines, B., Staswick, P.E., Browse, J., Howe, G.A., He, S.Y. (2008a). A critical role of two positively charged amino acids in the Jas motif of Arabidopsis JAZ proteins in mediating coronatine- and jasmonoyl isoleucine-dependent interactions with the C011 F-box protein. Plant J. 55: 979-988. Reddy, A.S.N. (2007) Alternative splicing of pre-messenger RNAs in plants in the genomic era. Annu. Rev. Plant Biol. 58: 267-294. 206 Thines, B., Katsir, L., Melotto, M., Niu, Y., Mandaokar, A., Liu, G., Nomura, K., He, S.Y., Howe, G.A., and Browse, J. (2007) JAZ repressor proteins are targets of the SCFCO” complex during jasmonate signalling. Nature 448: 661-665. Vanholme, B., Grunewald, W., Bateman, Kohchi, T., Gheysen, G. (2007) The tify family previously known as ZIM. Trends Plant Sci. 12: 239-244. 207 ”'111111111111111111111111ll“ 063 2495