p. E l’ '1‘ £3 '9 A. . mm) ‘r . v V a)“. ' w.» ruie‘aw um "um - v.1 . "5:23 ,. a; ”‘4‘” ‘53}; ' {$.56 ,- v . 5“ ,.. ’ 4d: 2%; Til-7‘09 “r. - 4 ' Wig} A J’ @7003 54504.14 2‘} £2 >9> :1:an In» Ctr-US Lang-LE .1153 E MS U is an Affirmative Action/Equal Opportunity Institution This is to certify that the dissertation entitled BIOCHEMICAL AND GENETIC ANALYSIS OF SYSTEMIC WOUND SIGNALING IN TOMATO (LYCOPERSICON ESCULENTUM) presented by Gyu In Lee has been accepted towards fulfillment of the requirements for Ph.D. degreein Plant Biology Date 8/15/03 042771 PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE l i 3 6/01 C'JCIRC/DateDquSS-p. 1 5 BIOCHEMICAL AND GENETIC ANALYSIS OF SYSTEMIC WOUND SIGNALING IN TOMATO (L YCOPERSICON ESCULENTUM) By Gyu In Lee A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Plant Biology 2003 ABSTRACT BIOCHEMICAL AND GENETIC ANALYSIS OF SYSTEMIC WOUND SIGNALING IN TOMATO (L YCOPERSICON ESCULENTUM) By Gyu In Lee Tomato plants activate the synthesis of defense proteins such as proteinase inhibitors (PIS) in response to insect attack. Systemic accumulation of PIS in wounded plants is mediated by a long-distance signaling pathway that transmits a mobile signal from the wound site to distal undamaged leaves. Previous studies have established that the polypeptide signal systemin, which is derived from a precursor protein called prosystemin, regulates systemic expression of PI genes via the jasmonic acid (J A) Si gnaling pathway. However, the precise role of J A and systemin in systemic wound signaling remains unclear. To address this question, two tomato mutants that are compromised in the systemic wound response were characterized. The suppressor of prosystemin-mediated responses] (sprI ) mutant was blocked in J A biosynthesis and subsequent PI gene expression in response to systemin. The systemin-insensitive phenotype of sprI plants indicates that the SprI gene product plays a role in linking the perception of systemin to the activation of J A biosynthesis. Reciprocal grafting experiments between wild-type and sprI plants demonstrated that this mutant is deficient in the production of the long-distance wound signal in damaged leaves, rather than the perception of that signal in neighboring undamaged leaves. This result indicates that systemin acts at or near the site of wounding to increase J A synthesis to a level that is required for the systemic response. The role of JA in systemic wound signaling was investigated using the J L-l mutant that is defective in both J A biosynthesis and wound-inducible expression of PI genes. Gas chromatography-mass spectrometry analysis Showed that J L-l plants are unable to produce J A from its precursor, 12-oxo-phytodienoic acid (OPDA), presumably because of a defect in fatty acid B-oxidation. It was found that OPDA is not an active signal for PI expression in tomato. Analysis of reciprocal grafis between wild- type and J L—l plants showed that J A biosynthesis is required for the generation of the systemic signal in wounded leaves. Taken together with the finding that J A perception is essential for systemic expression of P1 in undamaged leaves, these results indicate that JA or a derivative of J A is the long-distance Signal for systemic PI expression. J A biosynthesis is regulated by substrate availability. Two related cytochrome P4505, allene oxide synthase (A08) and hydroperoxide lyase (HPL), metabolize a common hydroperoxy fatty acid substrate to JA and volatile C6 aldehyde, respectively. To study the regulation of J A biosynthesis, genes encoding A08 and HPL were cloned from tomato and expressed in E. coli. Wounding induced the expression of the ADS gene in both wild-type and J A-deficient mutants, indicating that a JA-independent signaling pathway may contribute to systemic AOS expression. These results, together with those obtained from gene expression profiling in sprl plants, point to the existence of both systemin/JA-dependent and ~independent signaling pathways that regulate distinct sets of wound-inducible target genes in tomato plants. Acknowledgements My thesis could not have been completed without many people’s help and contributions. First of all, I would like to thank my adviser, Dr. Gregg Howe, for his support and guidance. I also thank my committee members, Drs. Hans Kende, Sheng Yang He, and Christoph Benning, for advice, encouragement and critical reviews. It was my honor to have Dr. Jan Zeevaart as an outside observer for my defense exam. The Howe lab members, Amy DeRocher, Aya Itoh, Lei Li, Ying-Tsu Loh, Sara Spencer, Dave Shaffer, Chuanyou Li, Mark Williams, Sara Birkeland, Liyan Liu, Sarah Norris, Tony Schihniller, Youfu Zhao, Hui Chen, Guanghui Liu, and Bonnie McCaig, have directly and indirectly contributed to this work. I appreciate their help and friendship. As the first year graduate students, Julie Zwiesler-Vollick, Carl Andre, and Wayne Riekhof, also participated in my research project during their lab rotations. Many results in this study were obtained from GC-MS analysis. I learned how to use GC-MS for JA measurement from Mr. Nate Riddle, Ms. Beverly Chamberlin, Dr. Doug Gage and Dr. Hans Weber. Dr. Benjamin Travis helped me to prepare internal standards. I appreciate their help. Thanks to all members working in Plant biology building, especially for Drs. Joong-Hoon Ahn, Kook-Ho Cho, Hyung Taeg Cho, Dongsu Choi, J eong-Hoe Kim, Yi Lee, Dong Ju Lee, Seung-Hwan Yang, and Mi Jung Suh, for support, advice, and help. Finally, I would like to thank our PRL editor, MS. Karen Bird, for critical reading of my dissertation draft. iv Table of Contents List of Tables .................................................................................................................... x List of Figures .................................................................................................................. xi Chapter 1 Systemic wound signaling in tomato ................................................................................ 1 Introduction ................................................................................................................ 2 Wound-induced gene expression ................................................................................ 3 Wound signals ............................................................................................................ 5 Oligogalacturonic acid ....................................................................................... 5 Systemin ............................................................................................................ 8 J asmonic acid ................................................................................................... 10 Physical Signals ............................................................................................... 1 5 Involvement of other plant hormones and signaling molecules in wound signaling. 15 Systemic wound signaling and the long-distance signal .......................................... 17 Systemic wound signaling in Arabidopsis ............................................................... 18 Wound response mutants of tomato ......................................................................... 19 Overview of thesis .................................................................................................... 21 References ................................................................................................................ 23 Chapter 2 The tomato mutant spr! is defective in systemin perception and the production of a systemic wound signal for defense gene expression ...................................................... 34 Introduction .............................................................................................................. 35 Results ...................................................................................................................... 37 sprl preferentially affects wound-induced systemic PI expression ................ 37 sprl plants are impaired in systemin-mediated Signaling ............................... 42 Effect of sprl on wound- and systemin-induced J A accumulation ................. 50 sprl plants are defective in the generation of a systemic wound signal for PI expression ........................................................................................................ 52 Discussion ................................................................................................................. 53 sprl defines a novel class of wound-response mutant .................................... 53 SprI-independent wound signaling ................................................................. 54 Role of systemin in wound signaling .............................................................. 56 Materials and Methods ............................................................................................. 57 References ................................................................................................................ 61 Appendix of Chapter 2 SprI does not encode SR160, the systemin receptor ...................................................... 65 Introduction .............................................................................................................. 66 Results ...................................................................................................................... 67 Cloning of SR160 from sprl plants ................................................................. 67 Analysis of linkage between Spr] and SR160 loci .......................................... 67 Expression of SR160 in sprl plants ................................................................. 75 Discussion ................................................................................................................. 75 Materials and Methods ............................................................................................. 78 References ................................................................................................................ 80 Chapter 3 vi The wound response mutant J L1 is defective in the conversion of 12-oxo-phytodienoic acid to jasmonic acid ...................................................................................................... 82 Introduction .............................................................................................................. 83 Results ...................................................................................................................... 85 J L1 plants are deficient in JA accumulation .................................................... 85 Exogenous OPDA does not activate the expression of PM] in J L1 plants ..... 85 J L1 plants fail to generate a systemic wound signal for PI—II expression ....... 86 Growth of J L1 seedlings is retarded ................................................................ 91 Discussion ................................................................................................................. 93 J L1 plants are unable to convert OPDA to JA ................................................ 93 OPDA is not a wound Signal for PI expression in tomato .............................. 97 Materials and Methods ............................................................................................. 99 References .............................................................................................................. 101 Chapter 4 Quantification of jasmonic acid and 12-oxo-phytodienoic acid in tomato plants by gas chromatography-mass spectrometry ............................................................................. 105 Introduction ............................................................................................................ 106 Results .................................................................................................................... 1 10 Preparation of internal standards and determination of GC-MS parameters 110 Extraction and quantification of endogenous J A and OPDA ........................ 113 Quantification of IA and OPDA from wound response mutants of tomato.. 118 Discussion ............................................................................................................... 121 Regulation of IA biosynthesis in tomato ....................................................... 122 vii Materials and Methods ........................................................................................... 123 References .............................................................................................................. 127 Chapter 5 Analysis of wound-induced root-to-leaf signaling in tomato plants ............................ 132 Introduction ............................................................................................................ 133 Results and Discussion ........................................................................................... 135 Materials and Methods ........................................................................................... 140 References .............................................................................................................. 142 Chapter 6 Cytochrome P450-dependent metabolism of oxylipins in tomato. Cloning and expression of allene oxide synthase and fatty acid hydroperoxide lyase ..................... 144 Introduction ............................................................................................................ 145 Results .................................................................................................................... 149 cDNA isolation and sequence analysis .......................................................... 149 Functional expression of LeA 0S and LeHPL in E. coli ................................ 156 Developmental expression of LeA OS and LeHPL ......................................... 159 Wound-inducible expression of LeA OS is mediated by a Def] -independent signaling pathway .......................................................................................... 162 Discussion ............................................................................................................... 167 Characterization of LeA 0S ............................................................................ 168 Characterization of LeHPL ............................................................................ 169 Wound-inducible expression of LeA OS is mediated by a Def] -independent signaling pathway .......................................................................................... 171 viii Materials and Methods ........................................................................................... 173 References .............................................................................................................. 1 83 Chapter 7 Conclusions and future perspectives ............................................................................ 189 References .............................................................................................................. 1 97 ix List of Tables Table 3-1. Wound response of wild-type and JLl plants ................................... 88 Table 4-1. Summary of IA levels in wound response mutants of tomato. . . . . . . . .......1 16 Table 4-2. Wound-inducible J A and OPDA levels in wound response mutants ........ 117 Table 6-1. Percent amino acid and nucleotide identity between different AOSS and HPLS ................................................................................ 153 List of Figures Figure 1-1. The current model of systemic wound signaling ................................. 7 Figure 1-2. The octadecanoid pathway ......................................................... 12 Figure 2-1. Spatial pattern of PH and PH! mRNA accumulation relative to the wound Site ...................................................................................... 38 Figure 2-2. Time course of wound-induced gene expression in wild-type (WT) and sprl plants .............................................................................. 41 Figure 2-3. PI-II accumulation in wild-type (WT) and sprl plants in response to exogenous signaling compounds .................................................. 43 Figure 2-4. Response of sprl plants to exogenous systemin, oligogalacturonides (OGA), and chitosan .................................................................. 44 Figure 2-5. Effect of exogenous systemin on gene expression in wild-type (WT) and sprl plants ............................................................................. 46 Figure 2-6. J asmonic acid (J A) accumulation in response to systemin and mechanical wounding .............................................................................. 49 Figure 2-7. Wound-inducible PI-II expression in grafts between wild-type (WT) and Figure 2-8. Genomic DNA sequence of SR1 60 isolated from L. esculentum ............ 68 Figure 2-9. Comparison of deduced amino acid sequences of SR1 60 and BRII . . .. . ....70 Figure 2-10. Linkage analysis of SR] 60 and SprI ........................................... 73 Figure 2-11. Expression of SR] 60 in wounded wild-type and sprl plants. . . . . . . . . ..76 Figure 3-1. Conversion of OPDA to JA ........................................................ 87 xi Figure 3-2. Accumulation of J A and OPDA in wild-type and JL1 plants in response to wounding ............................................................................. 89 Figure 3-3. Effect of exogenous J A and OPDA on the accumulation of PI in wild-type and JL1 plants .......................................................................... 90 Figure 3-4. Wound-inducible PI-II expression in grafts between J L1 and wild-type plants .................................................................................... 92 Figure 3-5. Comparison of root length between wild-type and JL1 seedlings .......... 94 Figure 4-1. Structure ofJA, OPDA and their internal standards. . . . ...109 Figure 4-2. Chromatogram and mass spectra ofMeJA and MeDHJA. . ..1 11 Figure 4—3. Chromatogram and mass spectra of MeOPDA and MeH4OPDA. . . ..1 12 Figure 4-4. Standard curves for the quantification of IA and OPDA.........114 Figure 4-5. Accumulation of J A and OPDA in wild-type and def] plants challenged with Spider mites .................................................................... 119 Figure 5-1. Systemic induction of PHI in leaves in response to wounding of roots..136 Figure 5-2. Systemic expression of wound-inducible genes in response to root damage ................................................................................. 138 Figure 6-1. The octadecanoid pathway ....................................................... 146 Figure 6-2. Comparison of cDNA-deduced protein sequences of plant AOS and HPL genes .................................................................................. 150 Figure 6-3. Southern-blot analysis of LeA OS and LeHPL ........................................... 155 Figure 64. Activity ofLeAOS and LeI-LPL expressed in E. colt .. 157 Figure 6-5. Expression of LeA 0S and LeHPL genes in different organs of tomato. . . 160 Figure 6-6. Accumulation of LeAOS protein in different organs of tomato. . . . . . . l 63 xii Figure 6-7. Accumulation of LeA 0S, LeHPL, and Inh-II mRNAs in tomato plants in response to herbivory ............................................................... 164 Figure 6-8. Analysis of wound-induced gene expression in wild-type and def] mutant plants ................................................................................ 165 Figure 7-1. Revised model of systemic wound signaling in tomato. . . . . . . . . . . ..192 xiii Chapter 1 Introduction: Systemic wound signaling in tomato Introduction Herbivore attack causes plants to induce several defense responses (Kessler and Baldwin, 2002; Walling, 2000). One of the most extensively studied defense responses is the wound-induced synthesis of defensive proteinase inhibitors (PIS) in tomato (Gatehouse, 2002; Ryan, 2000). Accumulation of PIS in response to herbivory and mechanical wounding is regulated at the level of gene expression (Nelson et al., 1983). Accumulated PI proteins in tomato adversely affect the grth of lepidopteran caterpillars by interfering with digestive enzymes in the gut of the attacking insect (Ryan, 1990). Wound-induced synthesis of PIS occurs not only in the vicinity of the wound site, but also in undamaged leaves throughout the plant (Green and Ryan, 1972). This systemic response implies the existence of a long-distance signal transduction pathway that connects local injury to the activation of gene expression in distal undamaged tissue. Detachment of the damaged leaf immediately after wounding was Shown to abolish systemic PI accumulation, indicating that the mobile Signal is synthesized in the damaged leaf and transported to distal undamaged leaves (Green and Ryan, 1972). Extensive research effort has been focused on the identification of the systemic wound signal. Chemical elicitors of PI expression have been isolated from tomato plants and characterized as candidates for the mobile signal (Bishop et al., 1981; Pearce et al., 1991). These studies have Shown that a polypeptide signal called systemin, and the plant hormone jasmonic acid (J A), play crucial roles in the systemic wound response (Farmer and Ryan, 1992; McGurl et al., 1992; Pearce et al., 1991). Our current understanding of systemic wound signaling is summarized in this chapter. Wound-induced gene expression Over 20 genes have been shown to be expressed systemically in response to wounding of tomato plants. Based on their expression pattern, these genes have been classified into two groups (Ryan, 2000). The first group is the ‘early genes’ that are induced rapidly and transiently upon wounding. Expression of early genes is detected within 30 min of wounding. Maximal expression occurs about 2 hr after wounding, after which mRNA levels quickly decrease. The group of early genes includes those encoding calcium calmodulin (Bergey and Ryan, 1999), transcription factors (Stankovic et al., 2000; Strassner et al., 2002), and enzymes involved in the biosynthesis of IA (Heitz et al., 1997; Sivasankar et al., 2000; Strassner et al., 2002). Because the levels of cytosolic calcium and endogenous J A rapidly increase in response to wounding and elicitor treatment (Moyen et al., 1998; Parchmann et al., 1997), the induction of genes encoding calcium calmodulin and JA biosynthetic enzymes may represent a positive feedback mechanism to amplify the wound response. The second group of wound response genes in tomato is referred to as ‘late genes’. Transcript levels of these genes increase slowly and steadily in response to wounding. The expression of late genes is detected about 2 hr after wounding, with the highest expression level occun'ing 8 to 12 hr after wounding. Examples from this group include genes encoding PIS, polyphenol oxidases, leucine aminopeptidase and LE RNase (Chao et al., 1999; Constabel et al., 1995; Lers et al., 1998). Similar to P15, polyphenol oxidases mediate plant defense against attacking insects by inhibiting insect digestive processes (Constable et al., 1995). LE RNase and proteolytic enzymes such as leucine aminopetidase (Chao et al., 1999; Schaller and Ryan 1996) may play a role in reprogramming plant metabolism to allow for massive synthesis of PIS and other defensive phytochemicals. Advances in cDNA microarray technology have enabled investigators to monitor global changes in gene expression in response to wounding. A microarray experiment using 230 tomato cDNAs supports the idea that wound—inducible genes in tomato can be classified according to their temporal expression pattern (Strassner et al., 2002). The sequential expression of these genes indicates that transcription factors encoded by early genes may activate the expression of late genes. Alternatively, proteins encoded by the early genes may amplify the signaling pathway leading to the expression of late genes, which results in maximal production of defensive chemicals. Larger-scale microarray experiments have been performed in Arabidopsis. Mechanical wounding significantly changed the expression level of 657 genes (Cheong et al., 2002). About 20% of these wound—inducible genes encode signal transduction-related proteins such as protein kinases, protein phosphatases, GTP-binding proteins, calcium binding proteins, phosphatidylinositol-related enzymes, JA biosynthetic enzymes and transcription factors. Although it is likely that these gene products are involved in wound signaling, the precise contribution of each of gene remains to be determined. Other wound-inducible genes encode enzymes involved in the phenylpropanoid pathway. Activation of these genes leads to the synthesis of defensive chemicals and the reinforcement of the cell wall (Constabel, 1999; F ranke et al., 2002; Peters and Constabel, 2002), which may enhance resistance to pest attack. Wound signals A Simple bioassay has been used extensively to identify systemic wound signals in tomato plants (Ryan, 1992). This assay involves supplementation of young plants through the cut stem with compounds purified fiom tomato leaf extract, followed by measurement of PI accumulation using an immunodiffirsion assay (Ryan, 1967). Two structurally distinct molecules were found using this bioassay: oligogalacturonic acid (OGA) derived from the cell wall (Bishop et al., 1981), and a polypeptide, systemin (Pearce et al., 1991), In addition, the plant hormone J A and its methyl ester (MeJ A) were identified as inducers of PI expression (Farmer and Ryan, 1990; Farmer and Ryan, 1992). Studies of these compounds have established a signaling cascade that regulates wound-induced expression of PI and other defense-related genes (Figure 1-1; Ryan 2000). Oligogalacturonic acid Previously it was reported that a pectin-enriched fraction purified from tomato leaf hydrolysate induced PI synthesis (Bishop et al., 1981). Further purification of the fraction showed that the active component was the oligosaccharide OGA, which is a hydrolysis product of pectin. Therefore, it was speculated that wounding results in the release of OGA from the cell wall. A cDNA encoding a polygalacturonase that degrades pectin to OGA was recently cloned from tomato (Bergey et al., 1999). The activity of this enzyme is up-regulated locally and systemically upon wounding. This finding supports the idea that OGA is an authentic signal for the activation of wound-inducible genes in tomato. However, genetic evidence to support this hypothesis is lacking. Although OGA is a potent elicitor of PI expression, the mobility of OGA in plant tissue is very limited. Radioactively labeled OGA applied to wound sites was not transported to unwounded tissue (Baydoun and Fry, 1985). Thus, it is unlikely that OGA is the long-distance signal of the systemic wound response. In addition to its effect on P1 expression, OGA treatment results in phosphorylation of a plasma membrane protein (Farmer et al., 1989), depolarization of the plasma membrane (Moyen and Johannes, 1996; Thain et al., 1995), a reactive oxygen burst (Stennis et al., 1998), and an increase in endogenous JA levels (Doares et al., 1995b). Among these rapid cellular responses, it was shown that activation of J A biosynthesis is essential for OGA-mediated expression of PI genes. OGA failed to induce PI accumulation in plants treated with inhibitors of J A biosynthesis (Doares et al., 1995a), or in plants that are compromised in IA biosynthesis (Howe et al., 1996). These results indicate that OGA activates PI expression through the J A signaling pathway. The oligosaccharide, chitosan, which is derived from fungal cell walls, also induces J A synthesis and PI accumulation in tomato plants (Walker-Simmons and Ryan, 1984). This observation is consistent with the fact that JA regulates defense response against some fungal pathogens of tomato (Diaz et al., 2002). These results indicate that plants use chitosan as a signal to trigger defense response. systemin wounding FEE \ ” I prosystemin * \ Local +\‘ / OGA response t jasmonic acid ++ ethylene \ .1. / SR160 \ \ I Systemic jasmonic acid response § + ethylene t \ PIS j Figure 1-1. The current model of systemic wound signaling. This schematic model is adapted from Ryan (2000). Wounding activates J A biosynthesis via systemin- and OGA-mediated signaling pathways in wounded tissue (local response). Systemin functions as a mobile signal to activate PI expression in distal unwounded tissue (systemic response). A recent study indicates that J A, not systemin, is the systemic wound signal (Li et al., 2002). SR160, the systemin receptor. Systemin The 18-amino-acid polypeptide systemin was purified from tomato leaf extracts on the basis of its ability to induce PI accumulation when supplied to young tomato plants through the cut stem (Pearce et al., 1991). Synthetic systemin is active in the frnol range, indicating that it is the most potent elicitor of PI accumulation ever found. Systemin is derived fiom the C-terminal region of a 200 amino acid precursor protein named prosystemin (McGurl et al., 1992). Tomato has a single copy of the prosystemin gene, which is expressed at low levels in aerial parts of unwounded plants. Expression of the prosystemin gene is elevated locally and systemically within cells of the vascular bundles in response to wounding (J acinto et al., 1997; Mch1 et al., 1992). Several lines of evidence indicate that systemin plays a critical role in systemic wound signaling. First, systemic PI expression was severely reduced in transgenic plants that express the prosystemin cDNA in antisense orientation (McGurl et al., 1992). Moreover, these transgenic plants were more susceptible to insect attack than wild-type plants (Orozco-Cardenas et al., 1993). Second, PI and other wound-inducible genes were constitutively expressed in transgenic plants (called 35S::prosys) that overexpress prosystemin from the cauliflower mosaic virus (CaMV) 35S promoter (McGurl et al., 1994). In a grafting experiment in which wild-type scions were grafted onto 35S: :prosys transgenic rootstocks, the wild-type scions accumulated significant amounts of Pls in the absence of wounding. This observation indicates that 35S:.°prosys transgenic rootstocks produce a systemic wound signal that is transmitted to the wild- type scions. Unlike OGA, systemin may be a mobile signal in vivo because radioactively-labeled systemin applied to wound sites moved via the phloem to unwounded tissue (Narvaez-Vasquez et al., 1995; Pearce et al., 1991). Therefore, systemin was regarded as a strong candidate for the systemic wound signal. Previous studies demonstrated that systemin requires J A biosynthesis for induction of wound-inducible genes. Endogenous levels of J A rapidly increase in response to systemin (Conconi et al., 1996; Doares et al., 1995b), followed by expression of PI genes. Systemin treatment failed to induce PI expression in tomato mutant plants that are impaired in J A biosynthesis (Howe et al., 1996). Also, inhibitors of J A biosynthesis abolished systemin-induced PI accumulation (Doares et al., 19953). These results indicate that systemin, like OGA, acts upstream of J A in the wound signaling pathway (Figure 1-1). In addition to up-regulating J A synthesis, exogenous systemin activates several rapid cellular events. These events include increased cytosolic calcium levels, inactivation of a proton-ATPase located on plasma membrane, activation of a wound- inducible mitogen-activated protein kinase (MAPK), and membrane depolarization (Felix and Boller, 1995; Moyen and Johannes, 1996; Moyen et al., 1998; Schaller and Oecking, 1999; Stratrnann and Ryan, 1997). Pharmacological studies showed that inhibition of the proton-ATPase and changes in calcium ion flux induced PI expression (Schaller and Frasson, 2001; Schaller and Oecking, 1999). Thus, calcium mobilization and modulation of proton-ATPase may be early causal events of the wound signaling pathway. Systemin specifically binds a plasma membrane protein called SR160 (Meindl et al., 1998; Scheer and Ryan, 1999). This 160 kDa protein is a leucine-rich repeat receptor kinase that shows high similarity to the brassinolide receptor, BRII , of Arabidopsis (Scheer and Ryan, 2002). Interestingly, the tomato mutant curl3 is insensitive to brassinolide as a result of a nonsense mutation in the SR160 gene (Montoya et al., 2002). This finding indicates that SR160 is a dual receptor for systemin and brassinolide. It is currently unclear how the same receptor regulates two distinct signaling pathways. Orthologues of the prosystemin gene have been found in members of the Solanaceae family, such as potato, black nightshade, and bell pepper (Constabel et al., 1998). In tobacco, two polypeptides were found to strongly induce PI expression. These polypeptides are processed from the same precursor protein, and do not show sequence similarity to tomato systemin (Pearce et al., 2001). Therefore, it is possible that other plant species use polypeptide signals to regulate the wound response. Jasmonic acid The cyclopentanone compound J A is synthesized from linolenic acid by the octadecanoid pathway (Figure 1-2; Wastemack and Hause, 2002). J A biosynthesis is initiated in chloroplasts, where phospholipases release linolenic acid from membrane lipids (Ishiguro et al., 2001). Lipoxygenase (LOX) converts linolenic acid to 13- hydroperoxylinolenic acid (13-HPOT), which is metabolized by various enzymes belonging to the CYP74 family of cytochrome P4508 (Howe and Schilmiller, 2002; Howe et al., 2000). Among these, allene oxide synthase (AOS) commits 13-HPOT to the biosynthesis of J A. AOS transforms l3-HPOT to an unstable epoxide intermediate (12, 13-epoxy1inolenic acid), which is subsequently transformed by allene oxide cyclase 10 to 12-oxo-phytodienoic acid (OPDA). The remaining reactions of the octadecanoid pathway take place in peroxisomes (Strassner et al., 2002). The double bond of OPDA is reduced by OPDA reductase (OPR3) to yield 3-oxo-2-(22-pentenyl)-cyclopentane-1- octanoic acid (OPC-8). Finally, three rounds of B-oxidation shorten the carboxylate side chain of OPC-8, yielding J A. It remains unknown how OPDA is transported from chloroplasts to peroxisomes. J A is subject to several metabolic transformations such as methylation, conjugation to amino acid or glucose, adenylation, hydroxylation and sulfonation (Gidda et al., 2003; Hause et al., 2000; Kramell et al., 1997; Seo et al., 2001; Staswick et al., 2002). Me] A, amino acid-conjugated J A, and glucose-conjugated J A have been shown to induce expression of PI (Farmer and Ryan, 1990; Kramell et al., 1997 ; Wastemack et al., 1998). These results indicate that some JA derivatives act as signals for gene expression. In consistent, defective adedylation of J A abolishes JA-inducible gene expression (Staswick et al., 2002). Alteration of the expression of J A- metabolizing enzymes in transgenic Arabidopsis has provided insight into the role of J A metabolism. For example, a Specific methyl transferase converts JA to volatile Me] A (See et al., 2001). Overexpression of this enzyme in Arabidopsis results in elevated endogenous Me] A levels and enhanced resistance to fungal pathogens. Thus, MeJ A seems to be an active Signal for defense gene expression. A wealth of evidence indicates that J A is an essential component of the wound signaling pathway. First, tomato plants produce high levels of PIS in response to JA and MeJA (Farmer and Ryan, 1990; Farmer and Ryan 1992). Second, wounding results in a rapid increase in J A biosynthesis (Parchmann et al., 1997), which is followed by 11 Chloroplast Peroxisome Membrane Lipid I Lipase I - _ . . . (m OOH a linolenic acnd OOH ‘ Lipoxygenase \ l 13-hydroperoxylinolenic acid COOH I Allene oxide synthase 0 I 12, 13-epoxylinolenic acid 0 Allene oxide cyclase 12-oxo-phytodienoic acid COOH I I V \ I OPDA reductase - UPC-8 COOH * * ,B-oxidation 0 v — Jasmonic acid kNI:COOH A / J Figure 1-2. The octadecanoid pathway. Linolenic acid is released from chloroplast membranes and converted to OPDA by the sequential action of three enzymes. OPDA is processed to J A in peroxisomes. PI expression. Finally, wound-induced PI accumulation is blocked in JA-deficient tomato mutants (Howe et al., 1996; Li et al., 2002). It has also been established that wounding rapidly induces systemic expression of several J A biosynthetic genes including LOX, AOS, ADC, and OPR3 (Heitz etal., 1997; Howe et al., 2000; Sivasankar et al., 2000; Stenzel et al., 2003; Strassner etal., 2002). JA biosynthetic enzymes such as ADC and OPR3 are highly enriched in vascular bundles, indicating that IA is mainly synthesized in these tissues (Stenzel et al., 2003; Strassner et al., 2002). Overexpression of A US and AOC did not result in increased J A levels or constitutive expression of wound-inducible genes (Laudert et al., 2000; Stenzel et al., 2003). These observations indicate that J A biosynthesis is not regulated by transcriptional control of the genes. Because application of linolenic acid to tomato results in P1 expression (Farmer and Ryan, 1992; Howe et al., 1996), JA biosynthesis is probably regulated by substrate availability. Despite extensive knowledge of the effects of J A on gene expression and defense-related processes, our understanding of the signaling events that couple the production of J A to the activation of target genes is still in its infancy. Our current understanding of this problem has been significantly enhanced by the identification and characterization of mutants that are insensitive to JA. For example, an Arabidopsis mutant coronatine insensitive] (coil ) is unable to express wound-inducible genes in response to J A and MeJA (Feys et al., 1994). The C011 gene encodes a protein containing an F-box motif and leucine-rich repeats that functions in ubiquitin-mediated protein degradation (Xie et al., 1998). As a component of ubiquitin ligase complex, F - box proteins act as receptors that recruit regulatory proteins as substrates for ubiquitin- 13 mediated degradation (Bai et al., 1996). Previous studies demonstrated that the C011 protein participates in the assembly of a ubiquitin ligase complex that presumably recruits regulatory proteins to the 26S proteosome for degradation (Feng et al., 2003; Xu et al., 2002). A recent study indicates that COIl mediates degradation of a histone deacetylase that modulates chromosome remodeling (Devoto et al., 2002). These results indicate that C011 is involved in the removal of negative regulators acting downstream of J A There is also evidence indicating that a MAPK cascade operates downstream of J A to regulate gene expression. JA failed to activate wound-inducible genes in Arabidopsis plants deficient in MAPK4 activity (Petersen et al., 2000). In addition, a pharmacological study indicated that protein dephosphorylation catalyzed by protein phosphatase 2C is required for the induction of J A-responsive genes in Arabidopsis (Rojo et al., 1998). Previous studies indicated that OPDA, in addition to serving as an intermediate in J A biosynthesis, is a signal for several plant developmental processes (Koch et al., 1999; Weiler et al., 1993; Weiler et al., 1994). The Arabidopsis opr3 mutant is unable to convert OPDA to J A due to the loss of OPDA reductase activity (Stintzi et al., 2001 ). Whereas J A-deficient and J A-insensitive mutants are susceptible to insect attack, opr3 plants were shown to be as resistant as wild-type plants. Moreover, wild-type and opr3 plants showed very Similar gene expression patterns in response to wounding, and OPDA treatment induced a subset of wound-inducible genes in opr3 plants. These results demonstrate that OPDA is an active signal for defense responses in Arabidopsis. 14 Physical signals Mechanical wounding generates an electrical pulse (Wildon et al., 1992) and changes in xylem tension capable of producing a hydraulic signal (Malone and Alarcon, 1995; Malone et al., 1994). Because these physical signals are rapidly propagated long distance through the whole plant, they were proposed as candidates for the systemic signal for PI expression (Malone, 1996; Wilden et al., 1992). To date, however, there is no genetic evidence that an electrical signal is the causative agent of PI expression. In the case of hydraulic Signals, it was proposed that mass flow of xylem sap might deliver signaling molecules produced at the wound Site to undamaged leaves (Rhodes et al., 1999). However, this idea is not consistent with the observation that steam girdling of the petiole, which prevents phloem transport, but not xylem transport abolished systemic P1 expression in tomato (N elson et al., 1983). Thus, a causal role of physical signals in wound-induced systemic PI expression has not been established. Involvement of other plant hormones and signaling molecules in wound signaling Ethylene is required for wound-inducible gene expression in tomato. Wounding and systemin treatment activated ethylene biosynthesis. Inhibitors of ethylene perception, such as silver thiosulphate, were Shown to prevent the induction of J A biosynthesis and P1 expression. Also, wounding failed to stimulate PI synthesis in transgenic plants that are deficient in ethylene biosynthesis (O'Donnell et al., 1996). These results demonstrate that ethylene is a positive regulator of wound signaling in tomato. However, ethylene is not an independent signal for wound-inducible gene expression 15 because ethylene treatment alone does not induce P1 expression. It was proposed that ethylene and J A work together in a positive feedback loop to amplify wound-induced expression of PIS (O'Donnell et al., 1996). Abscisic acid (ABA) was proposed to be a positive regulator of wound signaling in tomato because ABA-deficient mutants fail to express P1 in response to wounding (Herde et al., 1996). It is not clear whether ABA alone is a signal for P1 expression because studies involving ABA treatment gave contradictory results (Birkenmeier and Ryan, 1998; Pena-Cortes et al., 1995). It is unlikely that ABA is the systemic wound Si gnal because ABA-inducible genes were not expressed systemically upon wounding (Birkenmeier and Ryan, 1998). Hydrogen peroxide accumulates systemically in response to wounding (Orozco- Cardenas and Ryan, 1999). Wounding, systemin, OGA, and JA did not induce PI accumulation when hydrogen peroxide synthesis was suppressed by inhibitors (Orozco- Cardenas et al., 2001). Interestingly, such inhibition did not affect expression of early wound response genes. Consistent with this, application of the hydrogen peroxide- generating enzyme glucose oxidase plus glucose to tomato plants resulted in the induction of late wound response genes such as PI, but not early wound response genes. Therefore, hydrogen peroxide appears to act downstream of J A to activate expression of late wound response genes. Other signals have been shown to inhibit wound signaling in tomato. For example, salicylic acid (SA) treatment suppressed J A biosynthesis and PI expression (Doares et al., 1995a). Likewise, nitric oxide (NO) was shown to decrease the level of hydrogen peroxide and PI accumulation (Orozco-Cardenas and Ryan, 2002). Because 16 SA and NO regulate defense responses to pathogenic bacteria (Dumer and Klessi g, 1999; McDowell and Dangl, 2000), these results indicate that defense responses against herbivores and bacterial pathogens are mediated by signaling pathways that antagonize one another (Kunkel and Brooks, 2003). Systemic wound signaling and the long-distance signal Based on previous studies, Ryan and coworkers proposed a model for systemic wound signaling in tomato (Orcozco-Cardenas et al., 2001; Ryan, 2000). According to this model, wounding activates the processing of prosystemin to systemin by unknown mechanism. Subsequently, systemin is loaded to the phloem and transported to distal undamaged leaves. Upon binding to its receptor, SR160, systemin activates JA biosynthesis in vascular bundle cells where prosystemin and JA-biosynthetic enzymes are located. J A then activates expression of early wound response genes in vascular bundles. In addition, J A induces release of OGA from cell wall by the activation of wound-inducible polygalacturonase, which triggers an oxidative burst. Hydrogen peroxide produced during the oxidative burst diffuses from vascular bundles to mesophyll cells, where P1 expression is activated. In this model, systemin acts as the long-distance signal for systemic gene expression. Recent studies using reciprocal grafting techniques have challenged the idea that systemin is the systemic signal for PI expression (Li et al., 2002). Analysis of wound- induced systemic P1 expression in grafts using J A—deficient and JA-insensitive mutants demonstrated that systemic P1 expression requires J A biosynthesis in wounded leaves but not in distal undamaged leaves. Conversely, J A perception is essential for 17 recognition of the systemic wound signal in undamaged leaves, but not for the generation of that signal in wounded leaves. The simplest interpretation of these results is that J A or a derivative of J A is the systemic wound signal. Because systemin induces PI expression by activation of JA biosynthesis (Doares et al., 1995a; Doares et al., 1995b; Howe et al., 1996), these experiments indicate that systemin is unlikely to act in unwounded tissue as the systemic Si gnal. Since prosystemin and J A biosynthetic enzymes are enriched in vascular bundle cells, systemin and JA might interact synergistically to amplify the production of the systemic wound signal along the vascular bundles (Ryan and Moura, 2002). Emerging evidence indicates that some systemic wound responses are mediated by a signaling pathway that operates independently of J A and systemin. From example, neither systemin nor J A activates expression of the wound-inducible glucosyl transferase gene in tomato (O'Donnell et al., 1998). It was also Shown that rapid systemic induction of a wound-inducible MAPK activity is unlikely to involve J A or systemin (Stratmann and Ryan, 1997). Moreover, destruction of phloem transport by steam girdling did not block the systemic induction of the kinase activity. This finding clearly demonstrates that a systemin- and J A-independent signal is involved in this particular systemic wound response. Systemic wound signaling in Arabidopsis Arabidopsis has been used as an alternative model system to study the wound responses (Berger, 2002; Leon et al., 2001). In this system, it is clear that JA plays a crucial role in wound-induced systemic expression of various defense-related genes. Wounding 18 induces JA biosynthesis in damaged leaves, which is followed by local and systemic gene expression 0(ubigsteltig et al., 1999; Laudert and Weiler, 1998; Reymond et al., 2000; Titarenko et al., 1997). Because these responses are impaired in J A-deficient and JA-insensitive mutants (Berger et al., 1996; McConn et al., 1997; Park et al., 2002; Reymond et al., 2000; Rojo et al., 1999), JA seems to be an essential component of the systemic wound signaling pathway. Interestingly, the JA-mediated signaling pathway appears to work independently of OGA in Arabidopsis, whereas J A is required for OGA-induced gene expression in tomato (Doares et al., 1995b; Rojo etal., 1999; Titarenko et al., 1997). Furthermore, expression of JA-responsive genes is inhibited in wounded Arabidopsis leaves by OGA and ethylene (Rojo et al., 1999). Therefore, although J A is required for systemic wound responses in both tomato and Arabidopsis, the regulatory mechanism of J A-mediated signaling pathway may depend on the plant species. This idea is consistent with the observation that systemin-mediated signaling is absent in Arabidopsis. Wound response mutants of tomato Mutants impaired in the wound response provide valuable tools to elucidate the systemic wound signaling pathway. J L1 and defenseless] (defI) were identified by screening an ethyl methane sulfonate (EMS)-mutagenized population for plants unable to produce PIS in response to wounding (Lightner et al., 1993). Genetic analysis Showed that these mutations are recessive. Since Me] A treatment induced PI accumulation in JL1 and def] plants, these mutants likely contain a defect in J A biosynthesis or an upstream step in the wound signaling pathway. Additional characterization of def] 19 plants showed that this mutant is deficient in wound-inducible J A accumulation, and has reduced AOC activity (Howe et al., 1996; Stenzel et al., 2003). Transgenic plants (35S::prosys) that overproduce prosystemin constitutively express defensive proteins such as PIS and polyphenol oxidase (PPO) in the absence of wounding (McGurl et al., 1994). To investigate the systemin-mediated signaling pathway, 35S::prosys plants were subject to EMS-mutagenesis, and mutant plants that fail to accumulate PPO and PIS were isolated (Howe and Ryan, 1999). Genetic analysis of these mutants showed four complementation groups, named suppressor of prosystemin-mediated responses (spr) 1-4. Interestingly, sprI plants showed a severe reduction in systemic but not local PI expression. Thus, Sprl appears to function specifically in the systemic response. In contrast, wounded spr2 plants did not accumulate PIS in either damaged or undamaged leaves. Recently, it was shown that Spr2 encodes an omega-3 fatty acid desaturase that is required for the synthesis of linolenic acid, which is the precursor of J A (Li et al., 2003). Tomato mutants that are insensitive to J A and Me] A were isolated by screening of EMS-mutagenized and fast-neutron mutagenized populations for plants that fail to express P1 in response to volatile MeJ A (Li et al., 2001). Genetic analysis showed that these lines define a Single gene named Jasmonic acid-insensitive! (Jail). As is predicted from the current model of wound signaling (Figure 1-1), wounding, systemin, and JA treatment are unable to induce PI expression in jail plants (Li et al., 2002). 20 Overview of thesis Systemic wound responses play important roles in plant defense against insect pests. Since the identification of wound-inducible PIS by Green and Ryan (1972), tomato has been used as a model system for the study of systemic wound signaling. A wealth of evidence indicates that systemin and J A are two essential signals required for the systemic wound response. Recent studies using wound response mutants indicate that JA or a derivative of J A acts as a long-distance Signal for the systemic P1 expression. In addition, other systemic signals appear to be involved in systemic responses that are induced rapidly upon wounding. The current thesis is focused on understanding the role of systemin and J A in the systemic wound response. Systemin-insensitive sprl plants were characterized to investigate how systemin and J A regulate systemic PI expression. Chapter 2 describes the characterization of sprl and discusses the role of systemin in the systemic wound response. In Chapter 3, the role of J A and its precursor OPDA in systemic wound signaling was examined by characterizing the J L1 wound response mutant. To study the regulation of J A biosynthesis during the wound response, endogenous levels of J A and OPDA were measured in wild-type and mutant plants by gas chromatography-mass spectrometry (GC-MS). These results are presented in Chapter 4, along with a discussion of the regulation of J A biosynthesis. Chapter 5 reports on JA-independent and J A-dependent signaling events that occur during wound-induced root-to-leaf Signaling. Chapter 6 presents a study aimed at characterizing two related cytochrome P450 enzymes, AOS and hydr0peroxide lyase, that metabolize hydroperoxy fatty acids 21 to J A and volatile C6 aldehyde, respectively. In the final chapter, a model of systemic wound signaling is proposed based on the results from thesis study. 22 References Bai, C., Sen, P., Hofmann, K., Ma, L., Goebl, M., Harper, J.W., and Elledge, SJ. (1996). SKPl connects cell cycle regulators to the ubiquitin proteolysis machinery through a novel motif, the F -box. Cell 86, 263—274. Baydoun, EA-H., and Fry, SC. (1985). The immobility of pectic substances in injured tomato leaves and its bearing on the identity of the wound hormone. Planta 165: 269-276. Berger, S. (2002). J asmonate-related mutants of Arabidopsis as tools for studying stress signaling. Planta 214: 497-504. Berger, 8., Bell, E., and Mullet, J .E. (1996). Two methyl jasmonate-insensitive mutants Show altered expression of AtVsp in response to methyl jasmonate and wounding. Plant Physiol. 11 1: 525-531. Bergey, DR, and Ryan, CA. (1999). Wound- and systemin-inducible calmodulin gene expression in tomato leaves. Plant Mol. Biol. 40:815-23. Bergey, D.R., Orozco-Cardenas, M., de Moura, BS, and Ryan, CA. (1999). A wound- and systemin-inducible polygalacturonase in tomato leaves. Proc. Natl. Acad. Sci. USA. 96, 1756-1760. Birkenmeier, GR, and Ryan, CA. (1998). Wound signaling in tomato plants. Evidence that ABA is not a primary signal for defense gene activation. Plant Physiol. 117: 687-693. Bishop, P., Makus, K.J., Pearce, G., and Ryan, GA. (1981). Proteinase inhibitor- inducing factor activity in tomato leaves resides in oligosaccharides enzymatically released from cell walls. Proc. Natl. Acad. Sci. USA. 78: 3536- 3540. Bowles, DJ. (1998). Signal transduction in the wound response of tomato plants. Phil. Trans. R. Soc. Lond. B. 353: 1495-1510. Chao, W.S., Gu, Y-Q., Pautot, V., Bray, EA, and Walling, LL. (1999). Leucine aminopeptidase RNAS, proteins, and activities increase in response to water deficit, salinity, and the wound Signals systemin, methyl jasmonate, and abscisic acid. Plant Physiol. 120: 979-992. Cheong, Y.H., Chang, H.S., Gupta, R., Wang, X., Zhu, T., and Luan, S. (2002). Transcriptional profiling reveals novel interactions between wounding, pathogen, abiotic stress, and hormonal responses in Arabidopsis. Plant Physiol. 129: 661-677. 23 Conconi, A., Miquel, M., Browse, J ., and Ryan, CA. (1996). Intracellular levels of free linolenic and linoleic acids increase in tomato leaves in response to wounding. Plant Physiol. 111: 797-803. Constabel, GP. (1999). A survey of herbivore-inducible defensive proteins and phytochemicals. in Inducible plant defenses against pathogens and herbivores: Biochemistry, ecology, and agriculture. Edited by Agrawal, A.A., Tuzun, S., and Bent, E. American Phytopathological Society Press, St. Paul, MN. Pp 137-166. Constabel, C.P., Bergey, DR, and Ryan, CA. (1995). Systemin activates synthesis of wound-inducible tomato leaf polyphenol oxidase via the octadecanoid defense signaling pathway. Proc. Natl. Acad. Sci. USA. 92: 407-411. Constabel, C.P., Yip, L., and Ryan, CA. (1998). Prosystemin from potato, black nightshade, and bell pepper: primary structure and biological activity of predicted systemin polypeptides. Plant Mol. Biol. 36: 55-62. Devoto, A., Nieto-Rostro, M., Xie, D., Ellis, C., Harmston, R., Patrick, E., Davis, J ., Sherratt, L., Coleman, M., and Turner, J .G. (2002). C011 links jasmonate signalling and fertility to the SCF ubiquitin-ligase complex in Arabidopsis. Plant J. 32: 457-466. Diaz, J ., ten Have, A., and van Kan, J .A.L. (2002). The role of ethylene and wound signaling in resistance of tomato to Botrytis cinerea. Plant Physiol. 129: 1341- l 35 1 . Doares, S.H., Narvaez-Vasquez, J ., Conconi, A., and Ryan, C.A. (1995a). Salicylic acid inhibits synthesis of proteinase inhibitors in tomato leaves induced by systemin and jasmonic acid. Plant Physiol. 108: 1741-1746. Doares, S.H., Syrovets, T., Weiler, E.M., and Ryan, C.A. (1995b). Oligogalacturonides and chitosan activate plant defensive genes through the octadecanoid pathway. Proc. Natl. Acad. Sci. USA. 92: 4095-4098. Dumer, J ., and Klessig, D. (1999). Nitric oxide as a signal in plants. Curr. Opin. Plant Biol. 2: 369-374. Farmer, E.E., Pearce, G., and Ryan, CA. (1989). In vitro phosphorylation of plant plasma membrane proteins in response to the proteinase inhibitor inducing factor (PIIF). Proc. Natl. Acd. Sci. USA. 86: 1539-1542. Farmer, BE, and Ryan, CA. (1990). Interplant communication: Air-borne methyl jasmonate induces synthesis of proteinase inhibitors in plant leaves. Proc. Natl. Acd. Sci. USA. 87: 7713-7716. 24 Farmer, BE, and Ryan, CA. (1992). Octadecanoid precursors of jasmonic acid activate the synthesis of wound-inducible proteinase inhibitors. Plant Cell 4: 129-134. Felix, G., and Boller, T. (1995). Systemin induces rapid ion fluxes and ethylene biosynthesis in Lycopersicon peruvianum cells. Plant J. 7: 381-3 89. Feng, S., Ma, L., Wang, X., Xie, D., Dinesh-Kumar, S.P., Wei, N., and Deng, X.W. (2003). The COP9 signalosome interacts physically with SCF CO" and modulates jasmonate responses. Plant Cell 15:1083-1094. Feys, B., Benedetti, C.E., Penfold, C.N., and Tmner, 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. Franke, R., Hemm, M.R., Denault, J .W., Ruegger, M.O., Humphreys, J.M., and Chapple, C. (2002). Changes in secondary metabolism and deposition of an unusual lignin in the ref8 mutant of Arabidopsis. Plant J. 30: 47-59. Gatehouse, J .A. (2002). Plant resistance towards insect herbivores: a dynamic interaction. New Phytol. 156: 145-169. Gidda, S.K., Miersch, 0., Levitin., A., Schmidt, J ., Wastemack, C., and Varin, L. (2003) Biochemical and molecular characterization of a hydroxyj asmonate sulfotransferase from Arabidopsis thaliana. J Biol Chem. 278: 17895-17900. Green, T.R., and Ryan, GA. (1972). Wound-induced proteinase inhibitor in plant leaves: a possible defense mechanism against insects. Science 175: 776-777. Hause, 13., Stenzel, I., Miersch, O., Maucher, H., Kramell, R., Ziegler, J ., and Wastemack, C. (2000). Tissue-specific oxylipin signature of tomato flowers: allene oxide cyclase is highly expressed in distinct flower organs and vascular bundles. Plant J. 24: 113-126. Heitz, T., Bergey, DR, and Ryan, CA. (1997). A gene encoding a chloroplast-targeted lipoxygenase in tomato leaves is transiently induced by wounding, systemin, and methyl jasmonate. Plant Physiol. 114: 1085-1093. Herde, 0., Atzom, R., Fisahn, J ., Wastemack, C., Willrnitzer, L., and Pena-Cortes, H. (1996). Localized wounding by heat initiates the accumulation of proteinase inhibitor 11 in abscisic acid-deficient plants by triggering jasmonic acid biosynthesis. Plant Physiol. 112: 853-860. Howe, G.A., Lee, G.I., Itoh, A., Li, L., and DeRocher, A. (2000). Cytochrome P450- dependent metabolism of oxylipins in tomato. Cloning and expression of allene oxide synthase and fatty acid hydroperoxide lyase. Plant Physiol. 123: 711-724. 25 Howe, G.A., Lightner, L., Browse, 1., and Ryan, CA. (1996). An octadecanoid pathway mutant (J L5) of tomato is compromised in signaling for defense against insect attack. Plant Cell 8: 2067-2077. Howe, G.A., and Ryan, CA. (1999). Suppressors of systemin Signaling identify genes in the tomato wound response pathway. Genetics 153: 1411-1421. Howe, G.A., and Schilmiller, AL. (2002). Oxylipin metabolism in response to stress. Curr. Opin. Plant Biol. 5: 230-236. Ishiguro, S., Kawai-Oda, A., Ueda, J ., Nishida, I., and Okada, K. (2001). The DEFECTIVE 1N ANT HER DEHISCENCEI 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. J acinto, T., McGurl, B., Franceschi, V., Delano-Freier, J., and Ryan, CA. (1997). Tomato prosystemin promoter confers wound-inducible vascular bundle-specific expression of the B-glucuronidase gene in transgenic tomato plants. Planta 203: 406-412. Kessler, A., and Baldwin, LT. (2002). Plant responses to insect herbivory: the emerging molecular analysis. Annu. Rev. Plant Biol. 53: 299-328. Koch, T., Krumm, T., Jung, V., Engelberth, J., and Boland, W. (1999). Differential induction of plant volatile biosynthesis in the lima bean by early and late intermediates of the octadecanoid-signaling pathway. Plant Physiol. 121: 153- l 62. Kramell, R., Miersch, 0., Hause, B., Ortel, B., Parthier, B., and Wastemack, C. (1997). Amino acid conjugates of jasmonic acid induce jasmonate-responsive gene expression in barley (Hordeum vulgare L.) leaves. FEBS Lett. 414: 197-202. Kubigsteltig, I., Laudert, D., and Weiler, E.W. (1999). Structure and regulation of the Arabidopsis thaliana allene oxide synthase gene. Planta 208: 463-471. Kunkel, B.N., and Brooks, D.M. (2002). Cross talk between signaling pathways in pathogen defense. Curr. Opin. Plant Biol. 5: 325-331. Laudert, D., Schaller, F., and Weiler, E. (2000). Transgenic Nicotiana tabacum and Arabidopsis thaliana plants overexpressing allene oxide synthase. Planta 211: 163-165. Laudert, D., and Weiler, E.W. (1998). Allene oxide synthase: a major control point in Arabidopsis thaliana octadecanoid signalling. Plant J. 15: 675-684. 26 Leon, J., Rojo, E., and Sénchez-Serrano, J.J. (2001). Wound signalling in plants. J. Exp. Bot. 5: 1-9. Lers, A., Khalchitski, A., Lomaniec, E., Burd, S., and Green, P.J. (1998). Senescence- induced RNases in tomato. Plant Mol. Biol. 36: 439-49. Li, L., Li, C., and Howe, GA. (2001). Genetic analysis of wound signaling in tomato. Evidence for a dual role of jasmonic acid in defense and female fertility. Plant Physiol. 127: 1414-1417. Li, L., Li, C., Lee, G.I., and Howe, GA. (2002). Distinct roles for jasmonate synthesis and action in the systemic wound response of tomato. Proc. Natl. Acad. Sci. USA. 99: 6416-6421. Li, C., Liu, G, Xu, C., Lee, G. 1., Bauer, P., Ling, H., Ganal, M., and Howe, GA. (2003). The tomato suppressor of prosystemin-mediated responses 2 (spr2) gene encodes a fatty acid desaturase required for the biosynthesis of jasmonic acid and the production of a systemic wound signal for defense gene expression. Plant Cell 15: 1646-1661. Li, C., Williams, M., Loh, Y.-T., Lee, G.I., and Howe, GA. (2002). Resistance of cultivated tomato to cell-content feeding herbivores is regulated by the octadecanoid signaling pathway. Plant Physiol. 130: 494-503. Lightner, J., Pearce, G., Ryan, C.A., and Browse, J. (1993). Isolation of signaling mutants of tomato (Lycopersicon esculentum). Mol. Gen. Genet. 241: 595-601. Malone, M. (1996). Rapid, long-distance signal transmission in higher plants. Adv. Bot. Res. 22: 163-228. Malone, M., and Alarcon, J. (1995). Only xylem-home factors can account for systemic wound Signaling in the tomato plant. Planta 196: 740-746. Malone, M., Palumbo, L., Boari, F., Monteleone, M., and Jones, H.G. (1994). The relationship between wound-induced proteinase-inhibitors and hydraulic signals in tomato seedlings. Plant Cell Environ. 17: 81-87. McConn, M., Creelrnan, RA., Bell, E., Mullet, J .E., and Browse, J. (1997). Jasmonate is essential for insect defense in Arabidopsis. Proc. Natl. Acad. Sci. USA. 94: 5473-5477. McDowell, J .M., and Dangl, J .L. (2000). Signal transduction in the plant immune response. Trends Biochem. Sci. 25: 79-82. Mch1, 3., Orozco-Cardenas, M., Pearce, G., and Ryan, CA. (1994). Over-expression of the prosystemin gene in transgenic tomato plants generates a systemic signal 27 that constitutively induces proteinase inhibitor synthesis. Proc. Natl. Acad. Sci. USA. 91: 9799-9802. McGurl, B., Pearce, G., Orozco-Cardenas, M., and Ryan, CA. (1992). Structure, expression, and antisense inhibition of the systemin precursor gene. Science 255: 1570-1573. Meindl, T., Boller, T., and Felix, G. (1998). The plant wound hormone systemin binds with the N-terminal part to its receptor but needs the C-terminal part to activate it. Plant Cell 10: 1561-1570. Montoya, T., Nomura, T., Farrar, K., Kaneta, T., Yokota, T., and Bishop, G]. (2002). Cloning the tomato curl3 gene highlights the putative dual role of the leucine- rich repeat receptor kinase tBRIl/SR16O in plant steroid hormone and peptide hormone Signaling. Plant Cell 14: 3163-3176. Moyen, C., Hammond-Kosack, K.E., Jones, J., Knight, M.R., and Johannes, E. (1998). Systemin triggers an increase of cytoplasmic calcium in tomato mesophyll cells: Ca2+ mobilization fiom intra- and extracellular compartments. Plant Cell Environ. 21: 1101-1111. Moyen, C., and Johannes, E. (1996). Systemin transiently depolarises the tomato mesophyll cell membrane and antagonises fusicoccin-induced extracellular acidification of mesophyll tissue. Plant Cell Environ. 19: 464-470. Narvaez-Vasquez, J., F lorin-Christensen, J ., and Ryan, CA. (1999). Positional specificity of a phospholipase A activity induced by wounding, systemin, and oligosaccharide elicitors in tomato leaves. Plant Cell 11: 2249-2260. Narvaez-Vasquez, J., Pearce, G., Orozco-Cardenas, M., Franceschi, V.R., and Ryan, CA. (1995). Autoradiographic and biochemical-evidence for the systemic translocation of systemin in tomato plants. Planta 195: 593-600. Nelson, C. E., Walker-Simmons, M., Makus, D., Zuroske, G., Graham, J., and Ryan, C. A. (1983). Regulation of synthesis and accumulation of proteinase inhibitors in leaves of wounded tomato plants. in Plant Resistance to Insects, Symposium Series 208, Edited by Hedin, P. (Am. Chem. Soc., Washington, DC), pp. 103- 122. O'Donnell, P.J., Calvert, C., Atzom, R., Wastemack, C., Leyser, H.M.O., and Bowles, DJ. (1996). Ethylene as a signal mediating the wound response of tomato plants. Science 274: 1914-1917. O'Donnell, P.J., Truesdale, M.R., Calvert, C.M., Dorans, A., Roberts, M.R., and Bowles, DJ. (1998). A novel tomato gene that rapidly responds to wound- and pathogen-related signals. Plant J. 14: 137-142. 28 Orozco-Cardenas, M., MchI, 8., and Ryan, CA. (1993). Expression of an antisense prosystemin gene in tomato plants reduces resistance toward Manduca sexta larvae Proc. Natl. Acad. Sci. USA. 90: 8273-8276. Orozco-Cardenas, M., Narvaez-Vasquez, J., and Ryan, CA. (2001). Hydrogen peroxide acts as a second messenger for the induction of defense genes in tomato plants in response to wounding, systemin, and methyl j asmonate. Plant Cell 13: 17 9-191. Orozco-Cardenas, M., and Ryan, CA. (2002). Nitric oxide negatively modulates wound Signaling in tomato plants. Plant Physiol. 130: 487-493. Orozco-Cardenas, M., and Ryan, CA. (1999). Hydrogen peroxide is generated systemically in plant leaves by wounding and systemin via the octadecanoid pathway. Proc. Natl. Acad. Sci. USA. 96: 6553-6557. Parchmann, S., Gundlach, H., and Mueller, M. (1997). Induction of 12-oxo- phytodienoic acid in wounded plants and elicited plant cell cultures. Plant Physiol. 115: 1057-1064. Park, J .H., Halitschke, R., Kim. H.B., Baldwin, I.T., F eldmann, K.A., F eyereisen, 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 j asmonic acid biosynthesis. Plant J. 31: 1-12. Pearce, G., Moura, D.S., Stratrnann, J ., and Ryan, CA (2001). Production of multiple plant hormones from a single polyprotein precursor. Nature 411: 817-820. Pearce, G., Strydom, D., Johnson, 8., and Ryan, CA. (1991). A polypeptide from tomato leaves induces wound-inducible proteinase inhibitor proteins. Science 253: 895—898. Pena-Cortes, H., Fisahn, J., and Willmitzer, L. (1995). Signals involved in wound- induced proteinase inhibitor 11 gene expression in tomato and potato plants. Proc. Natl. Acad. Sci. USA. 92: 4106-4113. Peters, D.J., and Constabel, GP. (2002). Molecular analysis of herbivore-induced condensed tannin synthesis: cloning and expression of dihydroflavonol reductase from trembling aspen (Populus tremuloides). Plant J. 32: 701-712. Petersen, M., Brodersen, P., Naested, H., Andreasson, E., Lindhart, U., Johansen, B., Nielsen, H.B., Lacy, M., Austin, M.J., Parker, J .E., Sharma, S.B., Klessig, D.F., Martienssen, R., Mattsson, O., Jensen, AB, and Mundy, J. (2000). Arabidopsis map kinase 4 negatively regulates systemic acquired resistance. Cell 103:1111- 1120. 29 Reymond, P., Weber, H., Damond, M., and Farmer, BE. (2000). Differential gene expression in response to mechanical wounding and insect feeding in Arabidopsis. Plant Cell 12: 707-720. Rhodes, J .D., Thain, J.F., and Wilden, DC. (1999). Evidence for physically distinct systemic Signalling pathways in the wounded tomato plant. Ann. Bot. 84: 109- 116. Rojo, E., Leon, J ., and Sanchez-Serrano, J .J . (1999). Cross-talk between wound signalling pathways determines local versus systemic gene expression in Arabidopsis thaliana. Plant J. 20: 135-142. Rojo, E., Titarenko, E., Leon, J ., Berger, 8., Vancanneyt, G., and Sanchez-Serrano, J. J. (1998). Reversible protein phosphorylation regulates jasmonic acid-dependent and -independent wound Signal transduction pathways in Arabidopsis thaliana. Plant J. 13: 153-165. Ryan, CA. (2000). The systemin signaling pathway: differential activation of plant defensive genes. Biochim. Biophys. Acta. 1477: 112-121. Ryan, C. A. (1992). The search for the proteinase inhibitor-inducing factor, PIIF. Plant M01. Biol. 19: 123-133. Ryan, C. A. (1990). Proteinase inhibitors in plants: genes for improving defenses against insects and pathogens. Annu. Rev. Phytopathol. 28: 425-449. Ryan, CA. (1967). Quantitative determination of soluble cellular proteins by radial diffusion in agar gels containing antibodies. Anal. Biochem. 19: 434-440. Ryan, C.A., and Moura, BS. (2002). Systemic wound Signaling in plants: a new perception. Proc. Natl. Acad. Sci. USA. 99: 6519-6520. Schaller, A, and Frasson, D. (2001). Induction of wound response gene expression in tomato leaves by ionophores. Planta 212:431-5. Schaller, A., and Oecking, C. (1999). Modulation of plasma membrane H+-ATPase activity differentially activates wound and pathogen defense responses in tomato plants. Plant Cell 11: 263-272. Schaller, A., and Ryan, CA. (1996). Molecular cloning of a tomato leaf cDNA encoding an aspartic protease, a systemic wound response protein. Plant Mol. Biol. 31:1073-1077. Scheer, J .M., and Ryan, CA (1999). A 160-kD systemin receptor on the surface of Lycopersicon peruvianum suspension-cultured cells. Plant Cell 11: 1525-1536. 30 Scheer, J .M., and Ryan, CA. (2002). The systemin receptor SR160 from Lycopersicon peruvianum is a member of the LRR receptor kinase family. Proc. Natl. Acad. Sci. USA. 99: 9585-9590. Seo, H., Song, J ., Cheong, J., Lee, Y., Lee, Y., Hwang, 1., Lee, J ., and Choi, Y. (2001). J asmonic acid carboxyl methyltransferase: a key enzyme for jasmonate- regulated plant responses. Proc. Natl. Acad. Sci. USA. 98: 4788-4793. Sec, 8., Okamoto, M., Seto, H., Ishizuka, K., Sano, H. and Ohashi, Y. (1995). Tobacco MAP kinase: a possible mediator in wound signal transduction pathways. Science 270: 1988-1992. Sivasankar, S., Sheldrick, B., and Rothstein, SJ. (2000). Expression of allene oxide synthase determines defense gene activation in tomato. Plant Physiol. 122: 1335-1342. Stankovic, B., Vian, A., Henry-Vian, C., and Davies, E. (2000). Molecular cloning and characterization of a tomato cDNA encoding a systemically wound-inducible bZIP DNA-binding protein. Planta 212: 60-66. Staswick, P.E., Tiryaki, I., and Rowe, ML. (2002). Jasmonate response locus JAR] and several related Arabidopsis genes encode enzymes of the firefly luciferase superfamily that show activity on jasmonic, salicylic, and indole-3-acetic acids in an assay for adenylation. Plant Cell 14: 1405-1415. Stennis, M.J., Chandra, S., Ryan, C.A., and Low, PS. (1998). Systemin potentiates the oxidative burst in cultured tomato cells. Plant Physiol. 117 : 1031-1036. Stenzel, 1., Hause, B., Maucher, H., Pitzschke, A., Miersch, 0., Ziegler, J ., Ryan, C., and Wastemack, C. (2003). Allene oxide cyclase dependence of the wound response and vascular bundle-specific generation of j asmonates in tomato- amplification in wound signalling. Plant J. 33: 577-589. Stintzi, A., Weber, H., Reymond, P., Browse, B., and Farmer, E. (2001). Plant defense in the absence of jasmonic acid: the role of cyclopentenones. Proc. Natl. Acad. Sci. USA. 98: 12837-12842. Strassner, J ., Schaller, F., Frick, U., Howe, G.A., Weiler, E., Amrhein, N., Macheroux, P., and Schaller, A. (2002). Characterization and cDNA-microarray expression analysis of 12-oxo-phytodienoate reductases reveals differential roles for octadecanoid biosynthesis in the local versus the systemic wound response. Plant J. 32: 585-601. Stratrnann, J .W., and Ryan, CA. (1997). Myelin basic protein kinase activity in tomato leaves is induced systemically by wounding and increases in response to 31 systemin and oligosaccharide elicitors. Proc. Natl. Acad. Sci. USA. 94: 11085- 1 1089. Thain, J.F., Gubb, I.R., and Wildon, DC. (1995). Depolarization of tomato leaf-cells by oligogalacturonide elicitors. Plant Cell Environ. 18: 211-214. Titarenko, E., Rojo, E., Leon, J ., and Sanchez-Serrano, J .J . (1997). J asmonic acid- dependent and -independent signaling pathways control wound-induced gene activation in Arabidopsis thaliana. Plant Physiol. 115: 817-826. Walker-Simmons, M.K., and Ryan, CA. (1984). Proteinase inhibitor synthesis in tomato leaves. Induction by chitosan oligomers and chemically modified chitosan and chitin. Plant Physiol. 76: 787-790. Walling, LL. (2000). The myriad plant responses to herbivores. J. Plant Growth Regul. 19: 195-216. Wastemack, C., and Hause, B. (2002). J asmonates and octadecanoids: signals in plant stress responses and development. Prog. Nucleic Acid Res. Mol. Biol. 72: 165- 221. Wastemack, C., Ortel, B., Miersch, 0., Kramell, R., Beale, M., Greulich, F., Feussner, I., Hause, B., Krumm, T., Boland, W., and Parthier, B. (1998). Diversity in octadecanoid-induced gene expression of tomato. J. of Plant Physiology 152: 345-352. Wildon, D.C., Thain, J.F., Minchin, P.E.H., Gubb, I.R., Reilly, A.J., Skipper, Y.D., Doherty, H.M., O’Donnell, P.J., and Bowles, DJ. (1992). Electrical Signaling and systemic proteinase-inhibitor induction in the wounded plant. Nature 360: 62-65. Weiler, E., Albrecht, T., Groth, B., Xia, Z.., Luxem, M., Liss, H., Andert, L., and Spengler, P. (1993) Evidence for the involvement of jasmonates and their octadecanoid precursors in the tendril coiling response of Bryonia dioica. Phytochemistry 32: 591-600. Weiler, E., Kutchan, T., Gorba, T., Brodschelm, W., Niesel, U., and Bublitz, F. (1994). The Pseudomonas phytotoxin coronatine mimics octadecanoid signalling molecules of higher plants. FEBS Lett. 345: 9-13. 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. 32 Xu, L., Liu, F., Lechner, E., Genschik, P., Crosby, W.L., Ma, H., Peng, W., Huang, D., and Xie, D. (2002). The SCFCO" ubiquitin-ligase complexes are required for jasmonate response in Arabidopsis. Plant Cell 14: 1919-1935. Ziegler, J ., Stenzel, I., Hause, B., Maucher, H., Miersch, 0., Hamberg, M., Grimm, M., Ganal, M., and Claus Wastemack, C. (2000). Molecular cloning of allene oxide cyclase: the enzyme establishing the stereochemistry of octadecanoids and jasmonates. J. Biol. Chem. 275: 19132-19138. 33 Chapter 2 The tomato mutant spr1 is defective in systemin perception and the production of a systemic wound signal for defense gene expression A version of this chapter has been published. Gyu In Lee and Gregg A. Howe (2003) Plant J. 33: 567-576 34 Introduction Many plants respond to insect attack and wounding by modulating the expression of genes involved in various defense-related processes. The synthesis and deployment of wound-induced phytochemicals is regulated by signal transduction pathways that operate both locally at the site of wounding and systemically in undamaged leaves throughout the plant (Green and Ryan, 1972; Karban and Baldwin, 1997). Wound- inducible defensive proteinase inhibitors (PIS) in Solanaceous plant species provide an attractive model system in which to study the mechanism of long-distance wound signaling, and several ideas have been proposed regarding the identity of the systemic signal transmitted from wound sites (reviewed by Bowles, 1998; Leén et al., 2001; Malone, 1996; Ryan, 2000; Walling, 2000). Among the proposed intercellular signals for wound-induced PI gene expression are systemin, an 18-amino acid peptide that is produced from cleavage of a larger precursor protein called prosystemin, and the octadecanoid pathway-derived hormone jasmonic acid (J A) (Farmer and Ryan, 1992; Li et al., 2002a; McGurl et al., 1992; Pearce et al., 1991). A wealth of biochemical and genetic evidence indicates that systemin and J A work together in the same signaling pathway to activate expression of P1 and other defense-related genes (Li et al., 2001; Ryan, 2000). The systemin/J A signaling pathway is activated upon binding of systemin to a l60-kDa plasma membrane-bound receptor called SR160 (Meindl et al., 1998; Scheer and Ryan, 1999). This receptor was recently identified as a member of the leucine-rich repeat (LRR) receptor kinase family (Scheer and Ryan, 2002). Binding of systemin to the cell surface is associated with several rapid signaling events including increased 35 cytosolic Ca2+ levels, membrane depolarization, inhibition of a plasma membrane proton ATPase, and activation of a MAP kinase activity (Felix and Boller, 1995; Moyen and Johannes, 1996; Moyen et al., 1998; Schaller and Oecking, 1999; Stratmann and Ryan, 1997). The systemin signaling pathway leading to P1 expression is thought to cuhninate in activation of a phospholipase that releases linolenic acid, the metabolic precursor of J A, from membrane lipids (F arrner and Ryan, 1992; Narvaez-Vasquez et al., 1999). Chitosan oligomers and Oligogalacturonides (OGAS) derived fi‘om fungal and plant cell walls, respectively, also activate PI expression via the octadecanoid pathway (Bishop et al., 1981; Doares et al., 1995; Walker-Simmons and Ryan, 1984). The presence of wound-inducible polygalacturonase activity in tomato leaves (Bergey et al., 1999), together with the relative immobility of OGAS in the plant vascular system (Aldington and Fry, 1996), indicates that these compounds induce P1 expression at or near the site of wounding. J A synthesized in response to wounding, systemin, and OGAS acts in concert with ethylene (O’Donnell et al., 1996) and hydrogen peroxide (Orozco-Cardenas et al., 2001) to coordinate the induction of PI gene expression. Recent studies in Arabidopsis indicate that J A signaling depends upon assembly of ubiquitin-ligase complexes that presumably target transcriptional repressors of J A- responsive genes for proteolytic degradation (Xie et al., 1998; Xu et al., 2002). Mutants affected in the synthesis or perception of prosystemin and JA provide useful tools to understand the mechanism of systemic wound signaling. For example, antisense-mediated depletion of prosystemin expression in tomato plants abrogated wound-induced systemic accumulation of P13, indicating that this gene is essential for a normal systemic wound response (McGurl et al., 1992). To identify genes involved in 36 the systemin/J A signaling pathway using a forward genetic approach, we took advantage of a transgenic plant (called 355::prosys) that overexpresses prosystemin from the CaMV 358 promoter and, as a consequence, constitutively expresses PI and other defense-related genes in the absence of wounding (Constabel et al., 1995; Mch1 et al., 1994). An ethyl methane sulfonate (EMS)-mutagenized population derived from this line was screened for mutants that are impaired in 35S::prosys-mediated signaling (Howe and Ryan, 1999). Among several 'spr' suppressed in prosystemin-mediated responses mutations identified, five independent alleles were shown to define one locus called SprI . Evidence presented herein indicates that SprI is involved in a Signaling step that couples systemin perception to activation of the octadecanoid pathway. The results of grafting experiments further indicate that SprI function is required at or near the Site of wounding to amplify J A accumulation to a level sufficient to promote long- distance signaling. The existence of a wound response pathway that operates independently of SprI is also described. These results are discussed in the context of the role of systemin in the wound response of tomato plants. Results spr1 preferentially affects wound-induced systemic Pl expression Recessive mutations in Sprl were previously shown to suppress 35S::prosys-mediated expression of the well-characterized serine PI genes, PH and PM] (Howe and Ryan, 1999). Further characterization of this mutant was conducted using spr1/spr1 homozygous lines in which the 35S::prosys transgene was removed by outcrossing (see 37 Figure 2-1. Spatial pattern of PM and PM] mRNA accumulation relative to the wound site. A single wound was inflicted at the distal end of the terminal leaflet of the lower leaf of two-leaf-stage plants. Eight hours thereafter, various sections of the leaf blade and petiole were dissected for RNA isolation and analysis. Leaf sections from six plants were pooled for RNA isolation. (A) Schematic drawing illustrating the leaf and petiole sections that were harvested for RNA extraction. Panels in (B) and (C) Show the results obtained for analysis of wild-type (WT) and spr1-I plants, respectively. Lanes 1 through 6 rep resent analysis of RNA isolated from the corresponding tissue sections shown in panel in (A). Lanes 1', 2', and 3' represent analysis of RNA isolated from various tissue sections of unwounded plants: 1', pooled tissue from sections 1 and 6; 2', pooled tissue from sections 2 and 5; 3', pooled tissue from sections 3 and 4. RNA gel blots containing 5 pg total RNA were hybridized to probes for PM and PI—II. To facilitate the comparison between WT and spr1, RNA blots shown in panels in (B) and (C) were hybridized together in the same containers and exposed to autoradiographic film for the same time. Blots were hybridized to a probe for eIF 4A as a loading control. Note, however, that e1F4A mRNA abundance is greater in petiole tissue relative to leaf lamina. As an additional loading control, a picture of an ethidium bromide-stained gel of the total RNA (rRNA) is shown. The results shown are representative from three independent experiments. 38 (A) |1|2I3|4I5>I¢6>l (B) wild-type control wounded 1'2'3'12 34 5 6 n“ .. eIF4A w»vw~~~--um PI-l Pl-II rRNA yapunu-Eu rum-o sat-‘— fill. (C) an" control wounded 1’2’3’1 23456 PM K“ ...--. PM! .. eIF4A ~ua-..- I'RNA wgwwsswngr' 39 Materials and Methods). To determine the effect of spr1 on wound-induced local and systemic gene expression, RNA gel blot analysis was used to measure PH and PH] transcript levels in leaf tissue located at defined distances from a single wound inflicted at the distal end of the lower leaf (Figure 2-1A). As previously observed in wild-type (WT) tomato plants (Howe et al., 1996), P1 mRNA accumulation in the undamaged section (section 2) of the wounded leaf was significantly greater than that in the damaged section (section 1) of the same leaf (Figure 2-1B). A relatively strong systemic response was observed in the undamaged leaf (sections 5 and 6), whereas WT petioles (sections 3 and 4) showed little or no PI expression. In the case of spr1, PI mRNA accumulation in the damaged section of the wounded leaf was comparable to that in WT (Figure 2-1C). Mutant plants also showed PI expression in adjacent unwounded tissue (section 2), albeit at a level lower than in WT. More significantly, however, the steady— state level of P1 mRNA in the unwounded leaf (sections 5 and 6) of wounded spr1 plants was <10% of that observed in WT. This result indicates that spr1 impairs a signaling pathway that mediates or amplifies wound-induced systemic P1 expression, but contributes less to P1 expression near the wound site. This interpretation was supported by measurements of PI—II protein levels in damaged (local response) and undamaged (systemic response) leaves of wounded plants (data not shown). In six independent experiments involving at least six plants per genotype, the local response of spr1 plants ranged between 50 and 75% of the WT response. Systemic PI-II accumulation in spr1 plants ranged between 0 and 35% of that in WT plants, with the average response in the mutant being approximately 15% of WT levels. Analysis of plants homozygous for an independent allele of spr1 (spr1-2) gave very similar results; 40 Local Systemic wild-type spr1 wild-type spr1 012 48224012 481224 0124 812240124 81224 PH! .. fl 1“ an.“ ”a an». 0".” «mama-wuoou fl. I CD, ..... an. m. LoxD g”..- ”a--- ._,.,,,_ e _......., 3a :2; a; =.. . A081 " 4... g. Q WIPK Ms. M w wu,u~~~~~o fi!‘uuhwfi'fi~whu -- . o. a m..- \ .-- * 7 3L. v.. " . .. vaim'..‘5 eIF4Awwumuzwnmufiwun.a «wfiflflfifififl‘bmwu Figure 2-2. Time course of wound-induced gene expression in wild-type (WT) and spr1 plants. WT and spr1 plants (15-day-old) were wounded once on the lower leaf with a hemostat. Lower damaged (local response) and upper undamaged (systemic response) leaves were harvested at various times (hours) after wounding for RNA isolation and analysis as described in the legend for Figure 2-1. For each time point, Six plants were harvested and pooled for RNA extraction. 41 wound-induced local and systemic PI-II accumulation in spr1-2 plants was 51 and 13%, respectively, of WT levels. RNA gel blot analysis was used to determine the time course of local and systemic expression of various wound-responsive genes in spr1 plants. Two classes of genes that differ with respect to their timing of wound-induced expression have been described in tomato (Ryan, 2000). Transcripts of so called 'late'-response genes, including PI-I, PI—II, and cathepsin D inhibitor (CD1), accumulate to maximum levels 8-12 hr after wounding of WT plants (Figure 2-2). Consistent with the results shown in Figure 1, sprl plants were deficient in the magnitude but not the timing of induction of these genes. Genes whose expression is induced rapidly and transiently in response to wounding comprise a second class of 'early'-response genes. Included among this group are genes encoding signaling-related proteins such as lipoxygenase (LoxD), allene oxide synthasel (A081), and a putative mitogen-activated protein kinase (WIPK). Interestingly, wound-induced local and systemic expression of these early genes was not affected in spr1 plants (Figure 2). These results indicate that spr1 specifically affects the expression of late-response genes (i.e. PI genes). spr1 plants are impaired in systemin-mediated signaling To gain additional insight into the wound response phenotype of spr1 , the capacity of the mutant to respond to various PI-inducing compounds was determined (Figure 2-3). Consistent with the ability of spr1 to suppress 35S.'.°prosys-mediated P1 expression, spr1 plants did not accumulate PI-II in response to exogenous systemin or its bioactive precursor, prosystemin (Dombrowski et al., 1999). However, the mutant was responsive 42 Pl-ll (pg/ml leaf juice) CD 0 Con Sys PS Chit OGA LA JA Figure 2-3. PM] accumulation in wild-type (WT) and spr1 plants in response to exogenous signaling compounds. WT (filled bar) and spr1 (open bar) seedlings (lS-day-old) were excised at the base of the stem and supplied with 15 mM sodium phosphate buffer (Con), systemin (Sys, 15 nM), recombinant prosystemin (PS, 0.1 pg ml"), chitosan (Chit,250 pg ml 'I), oligogalacturonide (OGA, 250 pg ml '1), linolenic acid (LA, 5 mM), or jasmonic acid (JA, 100 nM). PI-II levels were measured 24 hr after treatment. Data points represent the mean and SD (n = 6). 43 160 (A) _L ..s N 1% o o 1 1 ..s o o 801 404 it; , i 0 0.01 0.1 1 10 100 Pl-ll (pg/ml leaf juice) M O) O O O systemin (pmollplant) (B) 160 140 4 1204 100 - 80 4 60 - 4o 4 20 - o - a r , . o 15 40 100 061A (ngml) PI-II (pg/ml leaf juice) *1 1 80 250 160 140 . 120 « 100 4 I (C) PI-II (pg/ml leaf juice) O 15 40 100 180 250 chitosan (pg/ml) Figure 2-4. Response of spr1 plants to exogenous systemin, oligogalacturonides (OGA), and chitosan. Two-leaf-stage wild-type (WT) (filled bars), defl (gray bars), and spr1 (open bars) plants were excised at the base of the stem and supplied with either phosphate buffer ('0') or buffered solution containing various concentrations of systemin (A), OGA (B), or chitosan (C). PI-II accumulation in leaves was measured 24 hr after treatment. Values indicate the mean and SD (n = 6). to octadecanoid signaling compounds (linolenic acid and J A) and to the polysaccharide elicitors OGA and chitosan. Because exogenous chitosan, OGA, and systemin activate P1 expression via the octadecanoid pathway (Doares et al., 1995), these findings indicated that spr1 affects systemin-mediated signaling at a point upstream of the octadecanoid pathway. To further test this hypothesis, the responsiveness of spr1 plants to a range of concentrations of systemin, OGA, and chitosan was compared to that of WT (Figure 2-4). Parallel analysis of the J A-deficient defenseless] (def?) that is impaired in P1 expression in response to systemin, OGA, and chitosan (Howe et al., 1996) was included as a control. WT plants showed a strong response to systemin concentrations of 1 pmol per plant and greater, whereas def] accumulated low levels of PI—II in response to high concentrations of systemin, as previously reported (Howe et al., 1996). spr1 plants failed to accumulate significant levels of PHI (<5% WT levels) in response to all concentrations of systemin tested. As expected, the spr1 mutant responded normally to a range of concentrations of OGA and chitosan, whereas def] plants were unresponsive to these elicitors (Figures 2-4B, C). These results support the idea that spr1 specifically affects the systemin branch of the wound-response pathway. RNA gel blot analysis was used to determine the effect of spr1 on systemin- mediated expression of early and late wound-response genes. Excision of seedlings at the base of the stem resulted in a gradual, low-level increase in PM] mRNA accumulation in leaves of both WT and spr1 plants (Figure 2-5A). WT plants accumulated high levels of PM] mRNA in response to exogenous systemin, whereas spr1 showed no response above that observed in the buffer control. This finding is consistent with the inability of spr1 plants to accumulate PI—II protein in response to 45 Figure 2-5. Effect of exogenous systemin on gene expression in wild-type (WT) and spr1 plants. (A) Excised seedlings (2-week-old) were incubated for 45 min in a solution containing 5 pmol systemin, and then transferred to water. At various times (hours) after the beginning of systemin treatment, leaf tissue was harvested for RNA isolation. Leaf tissue from six plants was pooled for each RNA isolation. RNA gel blots containing 5 pg total RNA were hybridized to probes for PI—II, LoxD, AOS], WIPK and, as a loading control, eIF4A. (B) Excised seedlings were pre-incubated in water for 4 h and then transferred either to 300 pl phosphate buffer (buffer) or the same volume phosphate buffer containing 5 pmol of systemin ('0'). Following uptake of this solution (approximately 45 min), seedlings were transferred to water. At various times (hours) thereafter, leaves were harvested for RNA isolation. 46 (A) buffer systemin wild-type spr1 wild-type spr1 051248051248{0512480511248 "m Pl'” :1 he. 4, au W “a . ‘1, 111:: .~,.... M, a” LoxD 'w. 3.. "'e- .- 'e- m [‘3' "2' r?- "- m ..m J 123. I!!! ('41 Inn 1531:. ‘ Linn , WIPK gm“ N r .101 M in!” . ‘~ .5”. .g....vdofi ALL. ""3“... .. . mglum-Mm... “..1. ..1-{.1 eIF4A ”up- uflflfiflnfl QCQCCQOOGCOU (8) buffer (after pretreatment) systemin (after pretreatment) wild-type spr1 wild-type spr1 051248051248 0.51248051248 PI-II fififl- ..fiu U- a». u LoxD .- ‘" “' "' A031 . ..... . '1‘ . ’ p ‘ . . J. WIPK " " '1 “ "f' "' j _"'".‘ 1;“. W -\o- eIF4A -----~flfl---~ uflflaauogubutui 47 systemin (Figure 2-4). Mock treatment (excision and incubation in buffer) of both WT and spr1 seedlings resulted in rapid and transient expression of three early wound- response genes: LoxD, AOS], and WIPK. In WT plants, systemin clearly enhanced the accumulation ofLoxD and AOS] mRNA, as previouSly reported (Heitz et al., 1997; Sivasankar et al., 2000). However, the expression of these genes in spr1 was not enhanced by systemin. In contrast to LoxD and AOS], the steady-state level of WIPK mRN A in both WT and spr1 plants was not affected by systemin treatment. The analysis of systemin-induced gene expression in WT and mutant plants was complicated by the fact that excision of seedlings at the base of the stem induced significant expression of early-response genes (Figure 2-5A). To determine the effect of systemin in the absence of this excision-induced effect, excised plants were pre- incubated in water for 4 hr (to allow mRNA levels to return to basal level), and then transferred to tubes containing either buffer or systemin (Figure 2-5B). A very low level of LoxD and AOS] expression was detected in buffer-treated WT and spr1 plants, presumably as a result of handling (i.e. touching) of plants during the transfer procedure. Transfer of pre-incubated WT plants to a systemin-containing solution resulted in a rapid and transient increase in the steady-state level of LoxD and AOS] mRNAs, and a more gradual, massive accumulation of PHI transcripts. The level of WIPK mRNA in WT plants was unaffected by systemin treatment, indicating that exogenous systemin stimulates expression of some (e. g. LoxD, AOS], and PM!) but not all (e. g. WIPK) wound-response genes. Treatment of pre-incubated spr1 plants with systemin did not increase the accumulation of LoxD, AOS] , PM], or WIPK mRNAs above the level observed in buffer-treated plants. In summary, these results show that 48 (A) 300 JA(pmolIgFW) o 8 8 8 8 8 0 1 3 Incubation time (hour) (3) ears“ 4'0“ 1L1 JA (pmol/g FW) 0 8 8 8 fi—r—I buffer systemin Treatment Figure 2-6. Jasmonic acid (JA) accumulation in response to systemin and mechanical wounding. (A) Leaves of 2-week-old wild-type (WT) (filled bar) or spr1 (open bar) plants were mechanically wounded with a hemostat. At various times after wounding (1 or 3 hr), wounded leaf tissue was harvested for J A extraction. J A was also extracted from leaves of unwounded plants ('0').(B) Two-week-old seedlings were excised at the base of the stem, pre-incubated in water for 4 hr, and then transferred to either buffer unwanted plant ('0') or a buffered solution containing systemin (5 pmol per plant). Leaves were harvested for JA extraction 2 hr 45 min after systemin application. The amounts of J A in plant extracts were quantified by GC-MS. Data represent the mean and SD of three independent experiments. 49 spr1 impairs systemin-mediated activation of both early- and late-response genes. However, the mutation does not affect the rapid and transient activation of early genes in response to excision or wounding, indicating the existence of an Spr]-independent wound signaling pathway. Effect of spr1 on wound— and systemin-induced JA accumulation Because wound- and systemin-induced PI gene expression is dependent upon the synthesis and subsequent action of J A, it was of interest to determine the capacity of spr1 plants to accumulate J A in response to wounding and systemin. J A levels in unwounded leaves of WT and mutant plants were 15.1 :t 1.5 pmol JA/g Fresh Weight (FW) and 14.5 i 4.6 pmol J A/ g FW, respectively (Figure 2-6A). In wounded WT plants, J A levels increased 15- and 7.5-fold, 1 and 3 hr after wounding, respectively. Wounding also increased J A accumulation in spr1 plants, albeit to levels that were significantly lower (P < 0.05) than WT levels. The amount of J A in wounded spr1 leaves throughout the time course was estimated to be approximately 57% of that in WT leaves. Exogenous systemin induced high levels of J A accumulation in WT, but had no effect in spr1 (Figure 2-6B). These findings indicate that Spr] is necessary for maximal levels of JA accumulation in response to wounding, but is strictly required for systemin- induced J A accumulation. 50 WT spr1 WT spr1 WT spr1 spr1 WT - + - + - + - + PHI to Q. o 3 eIF4A PHI 2) o 9. (I) 8 eIF4A S’- Figure 2-7. Wound-inducible PI-II expression in grafts between wild-type (WT) and spr1 plants. WT and spr1 plants were grafted in the four combinations indicated. The genotypes listed above and below the horizontal line correspond to the scion and rootstock, respectively. For each graft combination, plants were divided into a control (-) and experimental (+) group consisting of four grafted plants per group. For the experimental group, each leaflet on the rootstock was mechanically wounded with a hemostat. Eleven hours after wounding, leaf tissue was harvested separately from wounded rootstock leaves and undamaged scion leaves (scion) for RNA extraction. The control set of plants received no wounding, other than that inflicted by the grafting procedure itself. Levels of PHI mRNA were analyzed by RNA blot analysis, using an eIF4A cDNA probe as a loading control. The results shown are representative of three independent experiments. 51 spr1 plants are defective in the generation of a systemic wound signal for PI expression The deficiency of wound-induced systemic PI expression in spr1 plants could result from a defect in production of a long-distance wound signal or a defect in the perception of this signal in distal undamaged leaves. To address this question, wound- induced PI-II expression was analyzed in reciprocal grafts between WT and spr1 plants. F our-week-old plants were grafted such that both the rootstock (stock) and the scion contained at least two healthy leaves. After the graft junction healed, stock leaves were wounded and PH] mRN A levels were measured 11 hr after in both the damaged stock leaves (local response) and the undamaged scion leaves (systemic response). Wounding of WT stock leaves resulted in local and systemic accumulation of PHI transcripts to levels well above that observed in unwounded control plants that had also been grafted (Figure 2-7, lanes 1 and 2). This result demonstrates that wounding of WT stock leaves leads to the production of a graft-transmissible signal that is recognized in undamaged scion leaves. Consistent with the pattern of PHI expression in two-leaf-stage spr1 plants (Figure 2-1), wounded spr1 stock leaves showed a relatively strong local response and a weak (10% WT) systemic response (Figure 2-7, lanes 3 and 4). Analysis of spr1/WT hybrid grafts Showed that upon wounding of spr1 stock leaves, WT scions failed to activate PI-II expression to levels greater than that in spr1 scions that had been grafted to spr1 stock (Figure 2-7, lanes 5 and 6). In the reciprocal combination, however, spr1 scions were responsive to a signal emanating from wounded leaves of WT stock (Figure 2—7, lanes 7 and 8). Taken together, these findings indicate that spr1 impairs wound-induced systemic PI expression mainly by blocking the production of 52 the long-distance wound signal in damaged leaves, rather than the recognition of that signal in systemic, undamaged leaves. Discussion spr1 defines a novel class of wound-response mutant Forward genetic screens have identified two general classes of mutants that are defective in wound-induced systemic PI expression (Howe and Ryan, 1999; Li et al., 2001; Lightner et al., 1993). One group includes JA biosynthetic mutants (e.g. defl and spr2) that are unresponsive to upstream signals (e.g. systemin, OGA, and chitosan) that activate the octadecanoid pathway, but are responsive to exogenous J A. The second group includes JA-insensitive mutants (e.g. jail) that are responsive neither to upstream signals nor to J A. Here, we show that spr1 differs from existing wound-response mutants in that it is responsive to OGA and chitosan but muesponsive to systemin and its precursor, prosystemin. Consistent with the fact that OGA and chitosan activate PI expression through J A (Doares et al., 1995), spr1 plants were responsive to exogenous J A and its metabolic precursor, linolenic acid (Figure 2-3). These findings indicate that the octadecanoid pathway and downstream signaling steps leading to P1 expression are intact in sprl plants. The capacity of the mutant to accumulate significant levels of J A (approximately 57% WT levels) in response to wounding supports this idea. Considered collectively, the most straightforward interpretation of the results is that Spr] is involved in the perception of systemin or a subsequent systemin-specific signaling event necessary for activation of the octadecanoid pathway. Because systemin 53 perception occurs at the level of the plasma membrane (Meindl et al., 1998; Scheer and Ryan, 1999) and the initial steps of the octadecanoid pathway occur in the chloroplasts, it is possible that Spr] is involved in relaying a signal from the plasma membrane to the chloroplast. Included among the early signaling events induced by systemin are increased cytosolic Ca2+ levels, membrane depolarization, inhibition of a plasma membrane proton ATPase, activation of a MAP kinase activity, and activation of a phospholipase A2 activity (Felix and Boller, 1995; Moyen and Johannes, 1996; Moyen et al., 1998; Narvaez-Vasquez et al., 1999; Schaller and Oecking, 1999; Stratmann and Ryan, 1997). The insensitivity of sprl plants to both prosystemin and systemin indicates that the mutant is not defective in the synthesis or proteolytic processing of prosystemin. However, the data leave open the possibility that spr1 impairs the interaction of systemin with SR160, or signaling output of the activated receptor complex. Additional work is needed to distinguish these possibilities. Spr1-independent wound signaling It is noteworthy that sprl appears to impair wound-induced systemic PI expression much more than it affects local P1 expression. This finding indicates that the signaling pathway for systemic PI expression can be uncoupled from the signaling pathway that operates in tissue adjacent to the wound site. This aspect of spr1 is reminiscent of the wound-response phenotype of prosystemin antisense plants that are compromised in prosystemin production but nevertheless have the capacity to respond to exogenous prosystemin (Dombrowski et al., 1999; Mch1 et al., 1992). The robust local wound response of sprl and prosystemin antisense plants supports the hypothesis that multiple 54 signals generated at the wound site activate the octadecanoid pathway in wounded leaves (Doares et al., 1995; Farmer and Ryan, 1992; Ryan, 2000). This interpretation is consistent with the observation that spr1 plants respond normally to OGAS, and accumulate significant levels of J A in response to wounding (Figures 2-4 and 2-6). Whether OGAS are responsible for the entire pool of wound-induced J A in spr1 leaves or whether other mechanisms are involved in initiating the octadecanoid pathway remains to be determined. Genetic analysis of the wound response in tomato plants indicates that the bulk of wound-induced systemic PI expression requires the systemin/J A signaling pathway. However, we did observe that sprl plants exhibit a low but significant level of wound- induced systemic Pl expression. This residual signaling activity could reflect incomplete loss of Spr] function or, alternatively, a Spr] -independent pathway for systemic PI expression. Given the complete lack of systemin-induced gene expression in spr1 plants, however, the latter possibility seems more likely. The existence of a systemin-independent wound-response pathway is clearly supported by the observation that spr1 plants are not affected in wound- or cut-induced expression of early wound- response genes such as LoxD and AOS] . The fact that exogenous systemin enhances expression of LoxD and AOS] in WT plants (Figure 2-5) indicates that wounding and exogenous systemin may regulate the expression of these genes in somewhat different ways. For instance, it is possible that responses to systemin, when supplied through the transpiration stream, do not accurately reflect the wound-induced activity of systemin produced in vascular bundle cells of intact plants (Jacinto et al., 1997; Ryan, 2000). The expression pattern of WIPK, which was wound and cut inducible in both WT and sprl 55 plants, provides fiirther evidence for systemin-independent wound signaling. Unlike the LoxD and AOS] genes, exogenous systemin did not stimulate WIPK mRNA accumulation in WT (or sprl) plants. Several other wound-induced rapid systemic responses have been described in plants (e. g. O'Donnell et al., 1998; Seo et al., 1995; Stratmann and Ryan, 1997), and may involve physical (e. g. hydraulic) Si gnals propagated through the xylem (Malone, 1996) or the phloem (Rhodes et al., 1999). Role of systemin in wound signaling The signaling-related phenotypes of spr1 plants are fiilly consistent with a role for prosystemin in regulating wound-induced systemic PI expression through the octadecanoid pathway (Farmer and Ryan, 1992; Li et al., 2001; Mch1 et al., 1992). What is less clear, however, is how systemin and J A (and other signals) interact to effect wound signaling over long distances. Insight into this question was recently provided by means of grafting experiments showing that J A biosynthesis in rootstock leaves is essential for production of a long-distance Signal for P1 expression, whereas J A biosynthesis in undamaged leaves is not required for PI expression (Li et al., 2002a). Given that prosystemin works through J A, the most straightforward interpretation of these data is that systemin, acting in the rootstock portion of the graft, amplifies the synthesis of J A to levels that are required for long-distance signaling. Support for this model comes from the analysis of systemic P] expression in reciprocal grafts between WT and spr1 (Figure 2-7). These experiments Show that Spr] function is involved primarily in the production of the long-distance signal, rather than the recognition or processing of that signal in systemic undamaged leaves. The reduced level of wound- 56 induced J A in sprl plants further indicates that systemin may activate the synthesis of a specific pool of J A that is necessary for the systemic response, which is consistent with the lack of systemin-induced JA accumulation in spr1 (Figure 2-6). Along these lines, two hypotheses have recently been proposed to explain the role of systemin in long- distance P1 expression (Ryan and Moura, 2002). First, systemin may induce localized production of J A that subsequently exits the wounded leaflet and activates PI expression in distal leaves. Alternatively, systemin produced at the wound site may be translocated in the phloem where it activates JA synthesis in vascular tissues of leaves, petioles, and stems of the rootstock. The latter scenario indicates that a positive feedback loop between systemin and J A may amplify and propagate the systemic signal along the vascular system (Ryan, 2000; Ryan and Moura, 2002). Additional insight into the function of Spr] and other genes involved in systemic wound signaling may help to distinguish these hypotheses. Materials and Methods Plant material and treatments Tomato (Lycopersicon esculentum Mill cv Castlemart) seedlings were grown in Jiffy peat pots (Hummert International, Earth City, MO) in a grth chamber maintained under 17 hr of light (200 pE m 2 sec '1) at 28°C and 7 hr of dark at 18°C. Seed for def] was collected from a defl/defl homozygous line that was back-crossed four times using Castlemart as the recurrent parent. To simplify and standardize the genetic nomenclature, herein we refer to the previously described spr-l 593p and spr—l 9615 alleles (Howe and Ryan, 1999) as spr1-1 and sprl -2, respectively. Lines homozygous 57 for either of these two alleles were generated as described below. Unless otherwise indicated, sprl-1 was used for all experiments. Wounding and chemical elicitor experiments were performed with two-leaf- stage plants (14 to 16-day-old) as previously described (Howe etal., 1996). To assay the responsiveness of mutants to PI-inducing compounds, plants were excised at the base of the stem and placed in 0.5 ml microfuge tubes containing 300 pl of the inducing compound. When >75% of the elicitor solution had been imbibed (approximately 45 min), plants were transferred to glass vials containing 20 ml of water, and incubated in a Lucite box for 24 hr under continuous light. PI-II levels in leaves were measured by radial irnmunodiffusion assay (Ryan, 1967). Systemin, oligogalacturonic acid (OGA), and recombinant prosystemin (Dombrowski et al., 1999) were obtained from Dr. C .A. Ryan (Washington State University). Chitosan, J A, and linolenic acid were obtained from Sigma. All inducers except linolenic acid were diluted from stock solutions into sodium phosphate buffer (15 mm sodium phosphate, pH 6.5) prior to use. Linolenic acid was diluted into 15 mm sodium phosphate, pH 6.5, containing 0.05% (v/v) ethanol. Control experiments Showed that 0.05% ethanol in 15 mm soditun phosphate buffer did not affect the background level of PI expression (data not shown). Grafting experiments were performed as described by Li et al. (2002a). Genetic analysis sprl-l and spr1 -2 were originally isolated as recessive suppressors of 35S: :prosys- mediated responses O-Iowe and Ryan, 1999). Segregation of spr1-1 and spr1-2 from the 35S::prosys transgene, for the purpose of isolating homozygous spr1 alleles in an 58 otherwise WT genetic background, was achieved by crossing mutant lines (homozygous for the 35S::prosys transgene and either spr1-1 or spr1-2) to WT (cv Castlemart). Plants in the resulting F2 populations were tested for wound-induced PI-II accumulation with the assumption that spr1 homozygotes would display a Significantly reduced systemic response, as is the case for prosystemin antisense plants (McGurl etal., 1992). F2 plants showing a deficiency (<10% of WT levels) in the systemic response comprised approximately one-quarter of the population (data not shown) and were selected as putative spr1 homozygotes. A polymerase chain reaction (PCR)-based assay (Li and Howe, 2001) was used to test these plants for the presence of the 35S::prosys transgene. Individuals lacking the transgene were brought to the greenhouse for collection of F3 seed. To verify homozygosity of the spr1 allele, the deficiency in wound-induced systemic PI-II expression and insensitivity to exogenous systemin were confirmed in F3 seedlings. Southern blot analysis confirmed the absence of 35S::prosys. Selected spr1 homozygotes (either spr1-l 0r spr1-2) were back-crossed again to WT (cv Castlemart). Seedlings in the resulting F2 populations were scored for wound-induced systemic accumulation of PI-H, as well as for systemin-induced PI-II accumulation. The results were consistent with the expectation that spr1-l and sprl -2 impair both responses and behave as single recessive mutations. RNA gel blot analysis RNA was isolated fiom tomato leaves and analyzed by gel blot hybridization as described by Li et al. (2002b). Gels were run in duplicate, with one set stained with ethidium bromide to check for equal loading of the samples and intactness of the RNA. 59 DNA probes were isolated and radiolabeled with [32P-a] dCTP as described by Howe et al. (2000). The tomato EST clones, cLEC9C14 and cLET1D13, were used as probes for detection of AOS] (Sivasankar et al., 2000) and WIPK transcripts, respectively. Hybridization results were visualized by autoradiography using Kodak XAR-S film and, when appropriate, quantified using a Phosphorimager (Molecular Dynamics). Hybridization signals were normalized to the signal obtained using a cDNA probe (EST clone cLED1D24) for translation initiation factor eIF 4A. To directly compare transcript levels in WT and spr1 plants, blots containing RNA from both plant types were hybridized in the same container, washed under the same conditions, and exposed to film for the same length of time. Measurement of jasmonic acid J asmonic acid was extracted from leaves (10 g FW) of two-leaf-stage plants as previously described (Li et al., 2002b). Dihydrojasmonic acid (DHJA) was used as an internal standard for quantification of J A levels by gas chromatography-mass spectrometry (Li et al., 2002a). 60 References Aldington, S., and Fry, SC. (1996). Oligosaccharins. Adv. Bot. Res. 19: 1-101. Bergey, D.R., Orozco-Cardenas, M., de Moura, D.S., and Ryan, CA. (1999). A wound- and systemin-inducible polygalacturonase in tomato leaves. Proc. Natl. Acad. Sci. USA. 96: 1756-1760. Bishop, P., Makus, K.J., Pearce, G., and Ryan, CA. (1981). Proteinase inhibitor- inducing factor activity in tomato leaves resides in oligosaccharides enzymatically released from cell walls. Proc. Natl. Acad. Sci. USA. 78: 3536- 3540. Bowles, DJ. (1998). Signal transduction in the wound response of tomato plants. Phil. Trans. R. Soc. Lond. B. 353: 1495-1510. Constabel, C.P., Bergey, D.R., and Ryan, CA. (1995). Systemin activates synthesis of wound-inducible tomato leaf polyphenol oxidase via the octadecanoid defense signaling pathway. Proc. Natl. Acad. Sci. USA. 92: 407-411. Doares, S.H., Syrovets, T., Weiler, E.M., and Ryan, CA. (1995). Oligogalacturonides and chitosan activate plant defensive genes through the octadecanoid pathway. Proc. Natl. Acad. Sci. USA. 92: 4095-4098. Dombrowski, J .E., Pearce, G., and Ryan, CA. (1999). Proteinase inhibitor-inducing activity of the prohormone prosystemin resides exclusively in the C-terminal systemin domain. Proc. Natl. Acad. Sci. USA. 96: 12947-12952. Farmer, BE, and Ryan, CA. (1992). Octadecanoid precursors of jasmonic acid activate the synthesis of wound-inducible proteinase inhibitors. Plant Cell 4: 129-134. Felix, G., and Boller, T. (1995) Systemin induces rapid ion fluxes and ethylene biosynthesis in Lycopersicon peruvianum cells. Plant J. 7: 381-389. Green, T.R., and Ryan, CA. (1972). Wound-induced proteinase inhibitor in plant leaves: a possible defense mechanism against insects. Science 175: 776-777. Heitz, T., Bergey, D.R., and Ryan, CA. (1997). A gene encoding a chloroplast-targeted lipoxygenase in tomato leaves is transiently induced by wounding, systemin, and methyl jasmonate. Plant Physiol. 114: 1085-1093. Howe, G.A., Lee, G.I., Itoh, A., Li, L., and DeRocher, A. (2000) Cytochrome P450- dependent metabolism of oxylipins in tomato. Cloning and expression of allene oxide synthase and fatty acid hydroperoxide lyase. Plant Physiol. 123: 711-724. 61 Howe, G.A., Lightner, L., Browse, J ., and Ryan, CA. (1996) An octadecanoid pathway mutant (J L5) of tomato is compromised in signaling for defense against insect attack. Plant Cell 8: 2067-2077. Howe, GA. and Ryan, CA. (1999) Suppressors of systemin signaling identify genes in the tomato wound response pathway. Genetics 153: 1411-1421. J acinto, T., McGurl, B., Franceschi, V., Delano-Freier, J ., and Ryan, CA. (1997). Tomato prosystemin promoter confers wound-inducible vascular bundle- specific expression of the B-glucuronidase gene in transgenic tomato plants. Planta 203: 406-412. Karban, R., and Baldwin, LT. (1997). Induced Response to Herbivory. Chicago: Chicago University Press. LeOn, J ., Royo, E., and Sénchez-Serrano, J .J . (2001) Wound signalling in plants. J. Exp. Bot. 52: 1-9. Li, L., and Howe, GA. (2001) Alternative splicing of prosystemin pre-mRNA in tomato generates two active forms of the prosystemin wound signal. Plant Mol. Biol. 46: 409-419. Li, L., Li, C., and Howe, GA. (2001). Genetic analysis of wound signaling in tomato. Evidence for a dual role of jasmonic acid in defense and female fertility. Plant Physiol. 127: 1414-1417. Li, L., Li, C., Lee, G.I., and Howe, GA. (20023). Distinct roles for jasmonate synthesis and action in the systemic wound response of tomato. Proc. Natl. Acad. Sci. USA. 99: 6416-6421. Li, C., Williams, M., Loh, Y.-T., Lee, G.I., and Howe, G.A. (2002b). Resistance of cultivated tomato to cell-content feeding herbivores is regulated by the octadecanoid signaling pathway. Plant Physiol. 130: 494-503. Lightner, J ., Pearce, G., Ryan, C.A., and Browse, J. (1993). Isolation of signaling mutants of tomato (Lycopersicon esculentum). Mol. Gen. Genet. 241: 595-601. Malone, M. (1996). Rapid, long-distance signal transmission in higher plants. Adv. Bot. Res. 22: 163-228. McGurl, B., Orozco-Cardenas, M., Pearce, G., and Ryan, GA. (1994) Over-expression of the prosystemin gene in transgenic tomato plants generates a systemic signal that constitutively induces proteinase inhibitor synthesis. Proc. Natl. Acad. Sci. USA. 91: 9799-9802. 62 McGurl, B., Pearce, G., Orozco-Cardenas, M., and Ryan, C.A. (1992).Structure, expression, and antisense inhibition of the systemin precursor gene. Science 255: 1570-1573. Meindl, T ., Boller, T., and Felix, G. (1998). The plant wound hormone systemin binds with the N-terminal part to its receptor but needs the C-terminal part to activate it. Plant Cell 10: 1561-1570. Moyen, C., Hammond-Kosack, K.E., Jones, J ., Knight, M.R., and Johannes, E. (1998). Systemin triggers an increase of cytoplasmic calcium in tomato mesophyll cells: Ca2+ mobilization from intra- and extracellular compartments. Plant Cell Environ. 21: 1101-1111. Moyen, C ., and Johannes, E. (1996). Systemin transiently depolarises the tomato mesophyll cell membrane and antagonises fusicoccin-induced extracellular acidification of mesophyll tissue. Plant Cell Environ. 19: 464-470. Narvaez-Vasquez, J ., Florin-Christensen, J ., and Ryan, CA. (1999). Positional specificity of a phospholipase A2 activity induced by wounding, systemin, and oligosaccharide elicitors in tomato leaves. Plant Cell 11: 2249-2260. O'Donnell, P.J., Calvert, C., Atzom, R., Wastemack, C., Leyser, H.M.O., and Bowles, DJ. (1996). Ethylene as a signal mediating the wound response of tomato plants. Science 274: 1914-1917. O'Donnell, P.J., Truesdale, M.R., Calvert, C.M., Dorans, A., Roberts, M.R., and Bowles, DJ. (1998). A novel tomato gene that rapidly responds to wound- and pathogen-related signals. Plant J. 14: 137-142. Orozco-Cardenas, M., Narvaez-Vasquez, J ., and Ryan, CA. (2001). Hydrogen peroxide acts as a second messenger for the induction of defense genes in tomato plants in response to wounding, systemin, and methyl jasmonate. Plant Cell. 13: 179- 191. Pearce, G., Strydom, D., Johnson, 8., and Ryan, CA. (1991). A polypeptide from tomato leaves induces wound-inducible proteinase inhibitor proteins. Science 253: 895-898. Rhodes, J .D., Thain, J.F., and Wildon, DC. (1999). Evidence for physically distinct systemic signalling pathways in the wounded tomato plant. Ann. Bot. 84: 109- 116. Ryan, CA. (1967). Quantitative determination of soluble cellular proteins by radial diffusion in agar gels containing antibodies. Anal. Biochem. 19: 434-440. 63 Ryan, CA. (2000). The systemin signaling pathway: differential activation of plant defensive genes. Biochim. Biophys. Acta 1477: 112-121. Ryan, C.A., and Moura, BS. (2002). Systemic wound signaling in plants: a new perception. Proc. Natl. Acad. Sci. USA. 99: 6519-6520. Schaller, A., and Oecking, C. (1999). Modulation of plasma membrane HI-ATPase activity differentially activates wound and pathogen defense responses in tomato plants. Plant Cell, 11, 263-272. Scheer, J .M., and Ryan, CA. (1999). A 160-kD systemin receptor on the surface of Lycopersicon peruvianum suspension-cultured cells. Plant Cell 11: 1525-1536. Scheer, J .M., and Ryan, CA. (2002). The systemin receptor SR160 from Lycopersicon peruvianum is a member of the LRR receptor kinase family. Proc. Natl. Acad. Sci. USA. 99: 9585-9590. Sec, 8., Okamoto, M., Seto, H., Ishizuka, K., Sano, H., and Ohashi, Y. (1995). Tobacco MAP kinase: a possible mediator in wound signal transduction pathways. Science 270: 1988-1992. Sivasankar, S., Sheldrick, B., and Rothstein, SJ. (2000). Expression of allene oxide synthase determines defense gene activation in tomato. Plant Physiol. 122: 1 335-1342. Stratmann, J .W., and Ryan, CA. (1997). Myelin basic protein kinase activity in tomato leaves is induced systemically by wounding and increases in response to systemin and oligosaccharide elicitors. Proc. Natl. Acad. Sci. USA. 94: 11085- 11089. Walker-Simmons, M.K., and Ryan, CA. (1984). Proteinase inhibitor synthesis in tomato leaves. Induction by chitosan oligomers and chemically modified chitosan and chitin. Plant Physiol. 76: 787-790. Walling, LL. (2000). The myriad plant responses to herbivores. J. Plant Growth Regul. 19: 195-216. 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 ., Lechner, E., Genschik, P., Crosby, W.L., Ma, H., Peng, W., Huang, D. and Xie, D. (2002). The SCFCO" ubiquitin-ligase complexes are required for jasmonate response in Arabidopsis. Plant Cell 14: 1919-1935. 64 Appendix of Chapter 2 Spr1 does not encode SR160, the systemin receptor The cloning of SR160 from spr1 plants was performed by Mr. Carl Andre as part of his lab rotation project. The BC] mapping population used in this study was generated by Dr. Chuanyou Li. 65 introduction The previous chapter describes the putative role of the Suppressor of prosystemin- mediated responses] (Spr]) gene in systemin-mediated signaling. Systemin failed to induce synthesis of proteinase inhibitors (PIS) in spr1 plants, whereas other elicitors such as oligogalacturonic acid (OGA) and jasmonic acid (J A) activate normal PI accumulation in the mutant (Figure 2-3 and 24). Because both systemin and OGA induce expression of P1s by stimulation of J A biosynthesis (Doares et al., 1995a; Doares et al., 1995b; Howe et al., 1996), these results indicate that spr1 plants are compromised in the perception of systemin or a subsequent systemin-specific Signaling event necessary for activation of J A biosynthesis. Systemin perception is mediated by the systemin receptor, SR160, which is bound to the plasma membrane (Meindl et al., 1998; Scheer and Ryan, 1999). Using a photoaffinify labeling approaCh, SR160 was purified from cultured cells of the wild tomato species Lycopersicon peruvianum. This work led to the isolation of the corresponding cDNA encoding SR160. DNA sequence analysis showed that SR160 encodes a leucine-rich repeat (LRR) receptor kinase that shows high similarity to the BRII gene that encodes the brassinolide receptor of Arabidopsis (Scheer and Ryan, 2002). The systemin-insensitivity of spr1 plants indicates the possibility that Spr] encodes SR160. Experiments described here were designed to test this hypothesis. The results demonstrate that the sprl mutation does not affect SR160. Thus, it is likely that the Spr] gene product acts downstream of SR160 in the wound signaling pathway to stimulate J A biosynthesis. 66 Results Cloning of SR160 from spr1 plants To determine whether spr1 plants carry a mutation in the SR160 gene, the sequence of SR160 cloned from spr1 and its corresponding wild-type L. esculentum was compared. Because the SR160 cDNA was previously isolated from L. peruvianum (Scheer and Ryan, 2002), this sequence information was used to design two oligonucleotide primers to clone the SR160 from spr1 and wild-type L. esculentum. Genomic DNA from these plants was used as a template for PCR to clone SR160. Amplified products from these reactions were cloned and sequenced. The results demonstrated that SR160 genes from both sprl plants and wild-type (L. esculentum) contain an intronless 3624 bp open reading frame, 62 bp of 5’ untranslated region (UT R), and 122 bp 3’ UTR. Comparisons between SR160 from wild-type and spr1 plants Showed no sequence differences (Figure 2-8). Comparison of the deduced amino acid sequence of SR160 from L. esculentum to L. peruvianum showed sequence variation at 10 amino acid residues (Figure 2-9). Analysis of linkage between Spr1 and SR160 loci To further investigate whether Spr] and SR160 correspond to the same gene, linkage of the two loci was examined in a BC 1 mapping population that segregates for both the spr1 trait (i.e., systemin insensitivity) and a restriction fragment length polymorphism (RF LP) at the SR160 locus. The mapping population was generated fiom a cross 67 Figure 2-8. Genomic DNA sequence of SR160 isolated from L. esculentum. A PCR- strategy was used to clone SR160 from spr1 and wild-type L. esculentum. DNA sequence analysis Showed no sequence differences in SR160 between sprl plants and L. esculentum. Both genes contain an intronless ORF (3624 bp). The initiation and termination codons are boxed. Nucleotide sequences corresponding to the PCR primers are underlined. 1 AGAACTCAAGCTATAGATTCAAGAAAATCACCATTTAAGCTATAAAGTTTCAATCTTTGA 61 AQBIQAAAGCTCACAAAACTGTGTTTAACCAACATCCTTTGAGCTTAAACAAGCTTTTCT 121 TTGTTCTTCTTCTTATCTTTTTTCTTCCACCAGCTTCACCAGCAGCTTCTGTTAATGGTC 181 TTTATAAAGACTCCCAACAGCTTCTTTCCTTTAAAGCTGCACTCCCACCAACCCCAACTC 241 TGCTTCAGAACTGGTTGTCATCTACTGACCCTTGTAGTTTCACTGGTGTTTCATGCAAGA 301 ATTCTAGAGTTTCTTCTATAGATCTCAGTAACACTTTTTTAAGTGTGGATTTCAGTTTGG 361 TCACTTCTTATTTGCTTCCTCTTTCTAATTTGGAGTCTTTGGTGTTAAAGAATGCTAATC 421 TTAGTGGTTCTTTAACTTCTGCTGCAAAATCCCAATGTGGGGTTACTTTAGACTCCGTAG 481 ATCTAGCTGAGAACACAATTTCTGGACCTATTTCTGATATCTCTAGCTTTGGTGTTTGTT S41 CAAACCTTAAGTCTCTTAATCTTTCTAAGAATTTCTTGGACCCTCCTGGTAAAGAAATGC 601 TTAACGCTGCAACCTTTAGCCTCCAAGTTCTTGATCTTTCTTACAATAATATCTCAGGGT 661 TTAACTTGTTTCCATGGGTTTCATCTATGGGGTTTGTTGAACTTGAGTTCTTTTCTCTCA 721 AGGGTAACAAGCTAGCTGGAAGTATTCCTGAATTAGACTTCAAGAATTTGTCATATTTGG 781 ATCTTTCTGCAAATAATTTCTCAACTGTTTTTCCTTCATTCAAAGATTGCTCCAATTTGC 841 AGCACTTGGATTTGTCATCCAACAAGTTTTATGGTGATATTGGTTCTTCACTTTCTTCAT 901 GTGGGAAGCTCAGTTTTCTCAACCTTACCAATAACCAGTTTGTAGGTTTGGTCCCTAAGC 961 TACCAAGTGAAAGTCTACAGTATTTGTACTTAAGAGGGAATGATTTTCAGGGTGTGTACC 1021 CAAACCAACTTGCTGATTTGTGCAAAACTGTGGTGGAATTGGACTTGTCATACAATAATT 1081 TCTCAGGCATGGTTCCTGAGAGCCTTGGTGAATGTTCAAGTTTGGAACTTGTTGATATTT 1141 CCTACAATAATTTCTCTGGTAAGTTGCCTGTTGATACTCTCTCCAAGTTGAGTAATATTA 1201 AGACTATGGTCTTATCATTCAACAAATTTGTTGGTGGTTTGCCTGATTCTTTCTCTAATT 1261 TACTGAAATTGGAGACTTTGGATATGAGTTCTAATAATCTCACAGGGGTTATTCCATCTG 1321 GGATTTGCAAAGATCCTATGAATAACTTGAAAGTGCTGTACCTTCAGAATAACTTGTTTA 1381 AAGGCCCTATACCTGACAGTCTAAGCAACTGTTCACAGCTGGTGTCACTTGATCTTAGCT 1441 TTAATTACTTGACTGGGAGTATACCATCTAGTTTGGGGTCATTGTCAAAGCTAAAGGATC 1501 TCATCCTTTGGTTAAATCAGCTTTCAGGGGAAATCCCACAGGAGTTGATGTACTTGCAGG 1561 CTTTGGAGAATTTGATTCTTGATTTTAATGACTTAACTGGACCAATACCTGCAAGTCTTA 1621 GCAACTGTACCAAGTTGAATTGGATTTCATTGTCAAATAACCAATTGAGTGGTGAGATAC 1681 CGGCTTCTCTTGGGCGTTTGTCAAATCTAGCTATTCTTAAGCTTGGAAACAACTCAATCT 1741 CAGGGAATATACCTGCTGAATTGGGTAATTGCCAGAGCTTGATATGGTTGGATCTCAATA 1801 CTAATTTCCTGAATGGATCCATTCCGCCACCTTTGTTCAAGCAATCTGGCAATATTGCAG 1861 TGGCATTACTGACCGGGAAGCGATACGTGTATATCAAGAATGATGGGAGTAAGGAGTGCC 1921 ATGGAGCAGGGAATCTGCTGGAGTTTGGAGGGATTAGACAGGAACAGCTGGATAGAATCT 1981 CAACAAGGCATCCTTGCAATTTCACAAGAGTTTATAGAGGTATCACTCAGCCAACATTTA 2041 ACCACAATGGCTCTATGATATTTCTTGATTTATCTTATAATAAGTTGGAAGGTAGTATCC 2101 CAAAGGAATTAGGGGCAATGTACTATCTGTCTATATTGAATTTGGGGCATAATGATCTGT 2161 CTGGTATGATTCCTCAACAACTTGGAGGCTTGAAGAATGTTGCAATTCTTGATTTGTCAT 2221 ATAATAGGTTCAATGGCACGATCCCGAATTCCCTCACCAGTCTTACATTGCTTGGAGAGA 2281 TTGACCTGTCAAACAATAATCTCAGTGGAATGATTCCTGAATCTGCACCATTTGACACAT 2341 TCCCTGATTATAGGTTTGCGAATAATTCCCTCTGTGGGTATCCTCTCCCCATACCTTGTA 2401 GCTCGGGGCCGAAATCGGATGCAAATCAGCATCAGAAGTCTCACCGCAGACAAGCATCGT 2461 TGGCAGGGAGTGTGGCCATGGGTTTGTTATTTTCCCTCTTTTGTATCTTTGGTTTGATTA 2521 TTGTTGCCATAGAGACGAAGAAGAGGAGGAGGAAGAAGGAGGCTGCTCTTGAAGCTTATA 2581 TGGATGGTCATTCACATTCTGCAACTGCCAACAGTGCCTGGAAGTTTACGAGTGCTCGTG 2641 AGGCGTTAAGCATCAACCTTGCAGCATTTGAGAAGCCTCTCAGGAAGCTCACATTTGCTG 2701 ATCTTCTCGAAGCCACCAATGGTTTCCACAACGACAGTCTTGTAGGCTCTGGTGGTTTTG 68 2761 2821 2881 2941 3001 3061 3121 3181 3241 3301 3361 3421 3481 3541 3601 3661 3721 3781 GTGATGTCTACAAAGCTCAGTTGAAGGATGGGAGTGTTGTAGCTATTAAGAAATTGATAC ACGTCAGTGGACAGGGTGATCGAGAATTCACTGCTGAAATGGAAACCATAGGGAAGATCA AGCACCGCAACCTTGTCCCTCTTTTGGGCTACTGCAAAGTAGGGGAAGAAAGACTACTGG TTTATGAATACATGAAGTATGGAAGTCTTGAAGATGTCCTGCATGATCGGAAGAAAATTG GGATCAAGCTGAATTGGCCTGCAAGAAGGAAAATTGCCATTGGAGCTGCGAGAGGTTTGG CTTTCCTACACCATAACTGCATTCCACACATCATTCACCGGGACATGAAATCAAGTAATG TCTTGCTTGATGAAAATTTGGAAGCCAGAGTATCTGATTTCGGAATGGCAAGGTTAATGA GTGCTATGGACACTCATTTGAGTGTCAGCACTCTTGCCGGCACTCCAGGATACGTACCTC CTGAATATTACCAAAGCTTTAGATGTTCTACAAAAGGAGACGTTTATAGTTATGGTGTCG TATTACTTGAGCTTCTAACCGGCAAACAGCCAACAGATTCAGCTGATTTTGGTGACAACA ATCTTGTCGGATGGGTAAAGCTGCACGCTAAGGGAAAAATAACAGATGTCTTTGACCGGG AGCTATTGAAAGAGGATGCAAGCATTGAGATTGAACTTCTACAACACTTAAAGGTAGCTT GTGCTTGCTTAGATGATCGACATTGGAAACGTCCCACAATGATACAAGTTATGGCTATGT TTAAGGAGATTCAAGCAGGGTCAGGCATGGATTCGACATCGACAATCGGAGCTGATGATG TTAATTTTAGTGGAGTTGAAGGAGGGATAGAAATGGGGATAAATGGAAGTATAAAAGAAG GCAATGAGCTGAGCAAACACCrnfijzprcCACTAAATGAAGAGTTTATTGAAAGCTCACA AATTTTCCAAAATCATCATATGCAAAGTGTAATTTTTAGCCCCCAATTATTGTATGTACC ACTAGTTCCCATCCATAAAATCTTGTGT 69 Figure 2-9. Comparison of deduced amino acid sequences of SR160 and BRII. SR160 genes were isolated from sprl plants, cultivated tomato (L. esculentum), and the wild tomato species, L. peruvianum. Deduced amino acid sequences of these genes were aligned with that of Arabidopsis brassinolide receptor BRII, using the Clustal W program available at http://www.ch.embnet.org. Black boxes indicate conserved amino acid residues among all aligned sequences. spr1 VKAHKTVPNQHPLSLNKLFFVLLLlFFLPPASPAASVNGLYKDSQQLLEF esculentum Hi AHKTVFNQHPLSLNKLFFVLLLIFFLPPASPAASVNGLYKDSQQLLSi peruvianum Yfi-HRTVFNOHPLSLNRLFFVLLLTFFLPPASPAASVNGLIVDSOOLLQL Arabidopsis “ ‘ ' spr1 V-PLPPTPTLLQNHLSSTDPCSFTG‘CCKNSRVSSIDLSNTFLSVDF L; esculentum .UaLPPTPTLLQNWLSSTDPCSFTGVSCKNSRVSSIDLSNTFLSVDFQL- peruvianum H-HLPPTPTLLONHLSSTDPCSFTGVSCKNSPVSSIDLSNTFLSVDFQI Arabidopsis ‘ ‘ spr1 TSYLLPLSNLESLVLKNANLSGSLTSAAKSQCGVTLDSVDLAENTISGPI esculentum TSYLLPLSNLESLVLKNANLSGSLTSAAKSQCGVTLDSVDLAENTISGPi peruvianum Arabidopsis spr1lSSFGVCSNLKSLNLSKNFLDPPGKLNAATFSLQ\LDLSYNNISGE esculentum DISSFGVCSNLKSLNLSKNFLDPPGREMLNAATFSLQVLDLSYNNISGb peruvianum Arabidopsis spr1 esculentum peruvianum THFPHVSSMGFVELFPFCIPGNKLAGSIPELDFFNLSYLDLSANNFSTVP Arabidopsis “ 7 spr1 V3FKDCSNLQHLDLSSNKFYGDIGSSLSSCGKLSFLNLTNNQFVGLVPLL esculentum IHFKDCSNLQHLDLSSNKFYGDIGSSLSSCGKLSFLNLTNNQFVGLVPl peruvianum S3PKDCSNLOHLDLSSNKF{GDIGSSLSSCGRLSFLNLTNNQFVGLVPlL Arabidopsis v ‘ «ISSNQFVGPIP'L spr1 SLOYLYLRGNDFQGV PNQLADLCKTVVELDLSYNNFSGMVPBS esculentum H)ESLOYLYLRGNDFOGVYPNOLADLCKTVVELDLS{NNFSGMVPESLGT peruvianum ESLQ{LYLPGNDFOGVYPNOLADLCKTVVELDLSYNNFSGMVPECLEU Arabidopsis ‘ . spr1 esculentum peruvianum Arabidopsis spr1 lPSGICKDPMNNLKVLYLONNLFKGPIPDSl esculentum lPSGICKDPMNNLKVLYLONNLFKGPIPD”L” peruvianum Arabidopsis 70 spr1 esculentum peruvianum Arabidopsis spr1 esculentum peruvianum Arabidopsis spr1 esculentum peruvianum Arabidopsis spr1 esculentum peruvianum Arabidopsis spr1 esculentum peruvianum Arabidopsis spr1 esculentum peruvianum Arabidopsis spr1 esculentum peruvianum Arabidopsis spr1 esculentum peruvianum Arabidopsis spr1 esculentum peruvianum Arabidopsis spr1 esculentum peruvianum Arabidopsis spr1 esculentum peruvianum Arabidopsis VNLI LDFIJDLTC' HLI LEFHL‘LT HLllLlHD LSHCTI‘LI' SHCTITLI 71 l spr1 s4 ' Qfi'fiPLLM-KFLJHHI JCIPHl lHPDllKSSN \ LLDENLEHRVSDFG HPLl esculentum JGAARGLAFLHHNCI PHI IHRDMKSSNVLLDENLEIARVSDFGMARLMSE peruvianum J GAARGLAFLHHNC I PH I IHRDMKSSNVLLDENLEARVSDFGMARLMSR Arabidopsis ian ARG LAFLHHNC PH I IHRDlVlKSSNVLLDENLE'ARVSDFGMARLl‘v‘lE‘ A spr1 E3THLSVSTLAGTPGYVPPEYYQSFRCSTKGDVYSYGVVLLELLTGKQP? esculentum HDTHLSVSTLAGTPGYVPPEYYQSFRCSTKGDVYSYGVVLLELLTGKQPT peruvianum MDTHLSVSTLAGTPGYVPPEYYQSFRCSTKGDVYSYGVVLLELLTGKQPT Arabidopsis DTHLSVSTLAGTPGYVPPEYYQSFRCSTKGDVYSYGVVLLELLTGKfiPJ spr1 USADFGDNNLVGWVKLHAKGKITDVFDRELLKBDASIEIELLQHLKVAC; esculentum D8ADFGDNNLVGWVKLHAKGKITDVFDRELLKEDASIEIELLQHLKVAC¥ peruvianum D‘DDFGDNNLVGWVKLHAKGKITDVFDRELLKEDASIEIELLQHLKVAC~ Arabidopsis US'DFGDNNLVGWV’OHAKL'ISDVFDPELMKEDP‘ EIELLQHLKVA Ts spr1 ULDDRHWKRPTMlQVMAMFKEIQAGSGMDSTSTJG—ADDVNFSGVEGGJE esculentum ”LDDRHWKRPTMIQVMAMFKEIQAGSGMDSTSTIG—’DDVNFSGVEGGIE peruvianum - Arabidopsis 81:. spr1 HGINGSIKEGNELSKHJ esculentum EGINGSIKEGNELSKHL peruvianum JUJHGSTFEGNELC Arabidopsis 7 72 systemin-insensitive systemin-responsive (spr1/spr1) (Spr1/spr1) Figure 2-10. Linkage analysis of SR160 and Spr]. To test linkage between SR160 and Spr], a BCl mapping population was generated as described in the text. RFLP analysis was used to test for linkage between SR160 and SprI in this BCl population. XbaI digestion of genomic DNA generated a species- specific RFLP pattern at the SR160 locus (L. esculentum, 2 kb band; L. pennelliz', 12 kb band). To test whether the spr1 allele and the L. esculentum- specific RF LP cosegregate in the BCI population, genomic DNA was prepared from 26 BCl plants for RFLP analysis. The blot was probed with a 2 kb EcoRI fragment from the SR160 DNA. Asterisks show the cosegregation of the spr1 allele and L. esculentum-specific RFLP pattern. The cosegreation of the Spr] allele and L. pennellii-specific RFLP is noted with #. E, L. esculentum; S, the parental spr1 plant originated from L. esculentum; P, L. pennellii; F, the F1 plant. 73 between a homozygous spr1 plant (L. esculentum) and the wild tomato species L. pennellii (Spr]/Spr]), followed by a backcross of the resulting F 1 plant to the parental spr1 line (L. esculentum). To determine the genotype of individual plants within BCl population, each BCl plant was assayed for systemin-induced PI accumulation. Systemin-insensitive BCl plants do not accumulate P15 in response to systemin, and thus are homozygous for spr1. In contrast, systemin-responsive BCl plants must be heterozygous; they contain an spr1 allele inherited from L. esculentum and an Spr] allele transmitted fiom L. pennellii. If Spr] and SR160 correspond to the same gene or are tightly linked, all systemin-insensitive BC 1 plants are expected to be homozygous for the SR160 allele transmitted from L. esculentum (SR160m /SR160‘”‘). Conversely, systemin-responsive BCl plants should have one allele of SR160 inherited from L. pennellii (SR160M’) and another allele transmitted from L. esculentum (SR160“‘). To test this, the genotype (SR16Oe‘C /SR160‘”C or SR160” 312160“) of the SR160 locus for each BCl plant was determined using RFLP analysis that distinguishes the SR160” and SR160” alleles (Figure 2-10). Among 26 BC] plants examined, only 9 plants showed a match between the systemin-sensitivity phenotype and the expected RFLP pattern of SR160 (Figure 2- 10; lanes indicated with * or #). These results indicate that the Spr] and SR160 loci are assorting independently of one another. Therefore, Spr] does not encode the systemin receptor, SR160. 74 Expression of SR160 in spr1 plants Linkage analysis demonstrated that Spr] and SR 1 60 are different genes. However, this result does not rule out the possibility that the spr1 mutation may affect the expression of SR160. To test this possibility, RNA gel blot analysis was used to determine the steady state level of SR] 60 mRNA in spr1 and wild-type plants (Figure 2-11). Both wild-type and spr1 plants accumulated comparable levels of SR160 mRNA in leaves. Wounding did not significantly change the basal expression levels. Taken together with DNA sequence analysis ofSR160, these results indicate that spr1 plants are not compromised in SR160 or its expression. Discussion 1 tested the hypothesis that the systemin-insensitive phenotype of spr1 plants results from a defect in systemin receptor the SR160. Several lines of evidence disproved this hypothesis. First, linkage analysis demonstrated that Spr] does not encode SR160. Second, DNA sequence analysis showed that the SR160 gene in spr1 plants does not harbor a mutation in the coding region. Furthermore, spr1 plants showed normal expression of SR160. These results demonstrate that the spr1 mutation is not related to SR160. The Spr] gene product may, therefore, act downstream of SR160 to activate J A biosynthesis in systemin-mediated signaling. Alternatively, the Spr] product may assist in the binding of systemin to SR160. Precedence for this hypothesis comes from the observation that the Arabidopsis BRII associated receptor kinase 1 (BAKl) mediates brassinolide binding by forming a heterodimer with BRII (Nam and Li, 2002). 75 wild-type spr1 Hr 012 4 812240124 81224 . - ' q SR160 mu. m,“ «is .03». turn” “all“ 1.” fl eIF4A wwnmwwufiflummu” Figure 2-11. Expression of SR160 in wounded wild-type and spr1 plants. Two-leaf- stage spr1 and wild-type plants were wounded once on each leaf with a hemostat. Wounded leaves were harvested at indicated time (hour) after wounding for preparation of RNA. For each time point, six plants were harvested and pooled for RNA extraction. RNA gel-blot analysis was performed using a 2 kb EcoRI fragment of SR160 gene as a probe. A duplicated blot was probed with cDNA encoded eIF4A as a RNA loading control. 76 The tomato genome contains a single copy of SR1 60, which shows high similarity to Arabidopsis BRII gene that encodes the brassinolide receptor (Scheer and Ryan, 2002). Interestingly, a recent study indicates that SR160 acts as the brassinolide receptor in tomato (Montoya et al., 2002). It was found that brassinolide-insensitive curl3 plants have a nonsense mutation in SR160. This finding indicates that SR160 could be a dual ligand receptor of systemin and brassinolide. Additional insight into the dual role of SR160 could be obtained by investigating systemin binding and the systemic wound response in curl3 plants that lack a fimctional SR160. Because brassinolide regulates plant growth, curl3 plants display a dwarf phenotype (Montoya et al., 2002). If brassinolide and systemin compete for binding to SR160, it is expected that increased levels of systemin may interfere with brassinolide action. In mammals, the peptide hormone oxytocin and steroid hormone progesterone compete for the same receptor (Grazzini et al., 1998). Recent in vitro experiments indicate that brassinolide does not inhibit the binding of systemin to SR160 (Scheer and Ryan, 2002). Therefore, the two ligands seem not to compete for binding to SR160 in tomato plants. It remains to be determined how SR160 regulates two distinct signaling pathways. Systemin binding to the receptor activates several rapid responses in tomato cells, including increased cytosolic Ca2+ levels, depolarization of the plasma membrane, inactivation of a proton-ATPase associated with the plasma membrane, and activation of a mitogen-activated protein kinase (MAPK) activity (Felix and Boller, 1995; Moyen and Johannes, 1996; Moyen et al., 1998; Schaller and Oecking, 1999; Stratmann and Ryan, 1997). It is currently unclear how the activated SR160 promotes these effects and 77 how these changes are related to increased J A synthesis. Molecular cloning of Spr] could provide a clue to address these questions. Materials and Methods Plant material Plant growth conditions were the same as described in Chapter 2. The mapping population for linkage analysis was generated from a cross between L. esculentum carrying the homozygous spr1 mutation and wild-type L. pennellii. One of the resulting Fl plants was backcrossed to the parental homozygous spr1 plant to produce the BCI population. Individual BCl plants were tested for systemin-induced PI-II accumulation as described in Chapter 2. Briefly, two-leaf-stage plants were excised at the base of the stem and supplied with systemin dissolved in phosphate buffer (5 pmol systemin/plant). Accumulation of PI-II in leaves was measured 24 hr later using a radioimmunodiffusion assay. Each plant was transferred to a glass vial containing distilled water and incubated in the grth chamber to promote rooting. Plants were transferred to soil afier roots were established (approximately 7 days). For genotyping, DNA was isolated by leaf tissue of 6-week-old plants. To examine the expression of SR160, two-leaf-stage plants were wounded once on the lower leaf with a hemostat. Leaves were harvested at various times after wounding for extraction of total RNA. 78 Cloning of SR160 A PCR technique was used to clone SR160 from spr1 and wild-type L. esculentum. Two oligonucleotide primers were designed using the sequence of SR1 60 cloned from L. peruvianum: 5’-AGAACTCAAGCTATAGA-3’ and 5’- ACACAAGATTTTATGGATGGGA—3’. These primers correspond to the 5’ and 3’ UTRs, respectively. Genomic DNA was prepared from spr1 and wild-type L. esculentum to prepare template in PCR, following the method described in Chapter 6. PCR was performed with Pfu Turbo DNA polymerase (Stratagene, La Jolla, CA) following the manufacturer’s instruction. The amplified products were subcloned into EcoRV site of pBluescript SK (-) plasmid (Stratagene, La Jolla, CA) and sequenced (Genomics Technology Support Facility, Michigan State University, East Lansing, MI). Nucleic acid hybridization To find a RFLP for SR160, genomic DNA was prepared from L. esculentum and L. pennellii, and digested with several restriction enzymes. The reaction products were run on 0.8% agarose gel and transferred to I-beond membrane (Amersham, Piscataway, NJ). To prepare the radiolabeled probe, a 2 kb EcoRI fragment of SR160 was labeled with [32P-a] dCTP. The DNA gel-blot analysis was performed following the method describe in Chapter 6. The result showed that XbaI digestion generated an RFLP for SR] 60. Total RNA was extracted from leaves of spr1 and wild-type plants following the method described in Chapter 6. The RNA gel-blot was hybridized to the 2 kb EcoRI fragment of SR160, following the method described in Chapter 2. 79 References Doares, S.H., Narvaez-Vasquez, J., Conconi, A., and Ryan, C.A. (1995a). Salicylic acid inhibits synthesis of proteinase inhibitors in tomato leaves induced by systemin and jasmonic acid. Plant Physiol. 108:1741-1746. Doares, S.H., Syrovets, T., Weiler, E.M., and Ryan, C.A. (1995b). Oligogalacturonides and chitosan activate plant defensive genes through the octadecanoid pathway. Proc. Natl. Acad. Sci. USA. 92: 4095-4098. Felix, G., and Boller, T. (1995). Systemin induces rapid ion fluxes and ethylene biosynthesis in Lycopersicon peruvianum cells. Plant J. 7: 381-389. Grazzini, E., Guillon, G., Mouillac, B., and Zingg, H. (1998). Inhibition of oxytocin receptor function by direct binding of progesterone. Nature 392: 509-512. Howe, G.A., Lightner, L., Browse, J ., and Ryan, CA. (1996). An octadecanoid pathway mutant (JL5) of tomato is compromised in signaling for defense against insect attack. Plant Cell 8: 2067-2077. Howe, G.A., and Ryan, CA. (1999). Suppressors of systemin signaling identify genes in the tomato wound response pathway. Genetics 153: 1411-1421. Meindl, T., Boller, T., and Felix, G. (1998). The plant wound hormone systemin binds with the N-terminal part to its receptor but needs the C-terminal part to activate it. Plant Cell 10: 1561-1570. Montoya, T., Nomura, T., Farrar, K., Kaneta, T ., Yokota, T., and Bishop, GJ. (2002). Cloning the tomato curl3 gene highlights the putative dual role of the leucine- rich repeat receptor kinase tBRIl/SR16O in plant steroid hormone and peptide hormone signaling. Plant Cell 14: 3163-3176. Moyen, C., Hammond-Kosack, K.E., Jones, J ., Knight, M.R., and Johannes, E. (1998). Systemin triggers an increase of cytoplasmic calcium in tomato mesophyll cells: Ca2+ mobilization from intra- and extracellular compartments. Plant Cell Environ. 21: 1101-1111. Moyen, C., and Johannes, E. (1996). Systemin transiently depolarises the tomato mesophyll cell membrane and antagonises fusicoccin-induced extracellular acidification of mesophyll tissue. Plant Cell Environ. 19: 464-470. Nam, K.H., and Li, J. (2002). BRIl/BAKl , a Receptor Kinase Pair Mediating Brassinosteroid Signaling. Cell 110: 203-212. 80 Schaller, A., and Oecking, C. (1999). Modulation of plasma membrane H+-ATPase activity differentially activates wound and pathogen defense responses in tomato plants. Plant Cell 11: 263-272. Scheer, J .M. and Ryan, CA. (1999). A l60-kD systemin receptor on the surface of Lycopersicon peruvianum suspension-cultured cells. Plant Cell 11: 1525-1536. Scheer, J .M., and Ryan, CA. (2002). The systemin receptor SR160 from Lycopersicon peruvianum is a member of the LRR receptor kinase family. Proc. Natl. Acad. Sci. USA. 99: 9585-9590. Stratmann, J.W., and Ryan, CA. (1997). Myelin basic protein kinase activity in tomato leaves is induced systemically by wounding and increases in response to systemin and oligosaccharide elicitors. Proc. Natl. Acad. Sci. USA. 94: 11085- 11089. 81 Chapter 3 The wound response mutant JL1 is defective in the conversion of 12-oxo-phytodienoic acid to jasmonic acid 82 Introduction The plant hormone j asmonic acid (J A) is synthesized from linolenic acid via the octadecanoid pathway. The pathway is divided into two parts by the subcellular location of the enzymes involved. The first part of the octadecanoid pathway occurs in chloroplasts where the release of linolenic acid from plastid membranes leads to the synthesis of 12-oxo-phytodienoic acid (OPDA). The second part takes place in peroxisomes where OPDA, presumably transported from chloroplasts, is processed to J A (Figure 3-1; Turner et al., 2002; Wastemack and Hause, 2002). Several lines of evidence demonstrate that J A regulates the wound response of tomato plants. First, application of J A or its methyl ester Me] A) triggers the accumulation of wound-inducible proteins such as proteinase inhibitors (PIs; Farmer and Ryan, 1992; Farmer et al., 1992). Second, the endogenous level of J A increases rapidly upon wounding, which is followed by expression of PIs (Conconi et al., 1996). Third, the expression of PIs is abolished by both chemical inhibitors and mutations that block the octadecanoid pathway (Doares et al., 1995; Howe et al., 1996). Recent studies show that J A is a component of the long-distance signaling pathway that relays signals produced in the wound site to systemic expression of PIS in undamaged leaves (Li et al., 2002) Analysis of wound response mutants has expanded our knowledge of the role of J A in the wound response. The J L1 and J L5 (renamed defenseless] ; defI) mutant lines of tomato were isolated from an ethyl methanesulfonate (EMS)-mutagenized population by screening for plants that are unable to accumulate P15 in response to wounding (Li ghtner et al., 1993). Exogenous Me] A restored the production of P13 in both mutants, 83 indicating that they are defective in the biosynthesis of J A. Genetic analysis demonstrated that the mutations defined by J L1 and def] are non-allelic and recessive. Further study on def] plants showed that this mutant is deficient in the accumulation of J A upon wounding (Howe et al., 1996). The Defl gene appears to be involved in the regulation of allene oxide cyclase (AOC) that catalyzes the conversion of an unstable allene oxide to OPDA (Stenzel et al., 2003). The biochemical defect responsible for the wound response phenotype of J L1 remains to be established. Because JL1 plants synthesize P18 in response to exogenous Me] A, it is likely that the mutant is compromised either in the octadecanoid pathway or the activation of the pathway in response to wounding. To address this issue, the J L1 mutant was further characterized. Measurement of JA and OPDA levels in J L1 plants indicate that the mutant is compromised in the conversion of OPDA to J A. This hypothesis was supported by analysis of the response of J L1 plants to exogenous OPDA and J A. The retarded growth phenotype of JL1 seedlings further indicated that JL1 plants are affected in a B-oxidation step that is required for the conversion of OPDA to J A. The results of reciprocal grafting experiments indicate that J L1 plants are deficient in the production of the systemic wound signal, but are able to recognize the signal in unwounded leaves. Because J L1 plants accumulate OPDA but not J A, these results indicate that J A, not OPDA, is an essential component of the transmissible wound signal for PI expression. 84 Results JL1 plants are deficient in JA accumulation JL1 plants do not accumulate detectable levels of PHI protein in response to wounding (Table 3-1). To test whether the absence of PHI accumulation is because of a defect in J A synthesis, gas chromatography-mass spectrometry (GC-MS) was used to measure J A and OPDA levels in wild-type and J L1 plants. Wild-type plants accumulated 14 d: 1 pmol JA/g FW (fresh weight) in unwounded leaves (Figure 3-2). This basal level increased to 274 i 18 pmol J A/ g FW 1 hr after wounding. In contrast, unwounded J L1 plants produced significantly lower levels of JA (2 :1: 3 pmol/g FW) than wild-type (student’s t-test, P < 0.05). Furthermore, wounding did not elevate the level of J A (4 i 5 pmol J A/ g FW) in the mutant. These results indicated that J L1 plants are impaired in J A biosynthesis. Quantification of OPDA showed that JL1 plants accumulated OPDA to levels comparable to those in wild-type plants. Wild-type plants accumulated 597 i 41 pmol OPDA/ g FW in unwounded leaves, which increased to 752 :1: 19 pmol OPDA/ g F W 1 hr afier wounding. OPDA levels in JL1 plants were 413 i 205 pmol/g FW in unwounded leaves and 578 :t 134 pmol/g FW in wounded leaves. Therefore, J L1 plants synthesize OPDA but appear to be defective in its conversion to J A. Exogenous OPDA does not activate the expression of PM! In JL1 plants To further test the hypothesis that J L1 plants are unable to convert OPDA to J A, various amounts of OPDA and JA were applied to JL1 and wild-type plants through the cut stem, and the accumulation of PHI was measured 24 hr afier treatment. J L1 and wild- 85 type plants accumulated comparable amounts of PHI in response to J A (Figure 3-3). This finding is consistent with the previous observation that volatile MeJ A induced the synthesis of P15 in JL1 plants (Lightner et al., 1993). In wild-type plants, OPDA acted as a potent elicitor of PHI accumulation (Figure 3-3). In contrast, exogenous OPDA failed to induce PI- 11 production in J L1 plants. These results indicate that conversion of OPDA to J A is necessary for expression of PI—II in wild-type plants, and that the JL1 mutant is impaired in this metabolic process. JL1 plants fail to generate a systemic wound signal for PM! expression The inability of J L1 plants to synthesize HS in response to wounding (Table 3-1; Li ghtner et al., 1993) could result from a failure to produce or perceive a systemic wound signal. To address this question, reciprocal grafting experiments were performed with JL1 and wild-type plants. Mechanical wounds were inflicted to all leaves of the rootstock, and the expression of PM! was monitored in both the wounded rootstock and the undamaged scion leaves (Figure 3-4). Control experiments showed that the grafting procedure alone resulted in a moderate expression of PHI in both rootstock and scion leaves of grafted wild-type plants (Figure 3-4, lane 1). However, wounding of rootstock leaves resulted in a significant increase of PHI expression in the scion, indicating that the wounded rootstock produced a long-distance signal that was transmitted to the undamaged scion (Figure 3-4, lane 2). Grafted JL1 control plants did not show detectable local or systemic expression of PHI (Figure 3-4, lane 3), consistent with the absence of PHI in wounded J L1 seedlings (Table 3-1). In response to wounding, weak 86 OPDA O o | COOH L OPDA reductase OPC-8 3-oxo-2-(2Z-pentenyl)- cyclopentane-1-octanoic acid .0 3 0 * B-oxidation OPC'6 . OOOH 3-oxo-2-(2Z-pentenyl)- cyclopentane-1-hexanoic acid _ , l B—oxndatlon O OPC—4 . OOOH 3-oxo—2-(22-pentenyl)- cyclopentane-1-butyric acid 0 ‘ B-oxidation .. e “ COOH Figure 3-1. Conversion of OPDA to JA. OPDA is synthesized in chloroplasts and converted to J A in peroxisomes by consecutive reactions of OPDA reductase and three rounds of B-oxidation. 87 PI-ll levels (pg/ml Ieafjuice) genotype unwounded control wounded wild-type ND. 68.5 1: 14.1 JL1 ND. ND. Table 3-1. Wound response of wild-type and JL1 plants. Leaves of two-week-old wild-type and JL1 plants were wounded with a hemostat. The accumulation of PHI was measured in leaves harvested 24 hr after wounding. As a negative control, PI-II levels were also measured in leaves of unwounded plants. The data present mean and standard deviation from the measurement of three plants for each treatment. N.D. indicates that the accumulation of PHI was not detected. 88 (A) 350 300 a JA (pmol/g FW) _.. _. N N 0 0| O 01 O O O O O" O 1 [—‘—*—1 i .- O unwounded wounded (B) 900 800 1 OPDA (pmol/g FW) # 01 O) \l O O O O O O O O -‘ N (O o O o o o o m 1 4 O unwounded wounded Figure 3-2. Accumulation of JA and OPDA in wild-type and JL1 plants in response to wounding. Leaves of 2-week-old wild-type (open bar) and J Ll (filled bar) plants were wounded with a hemostat. Leaves were harvested for extraction of J A/OPDA 1 hr after wounding. JA/OPDA was also extracted from leaves of unwounded control plants. J A (A) and OPDA (B) levels were measured by GC-MS analysis. Data represent mean and SD of three independent experiments. 89 (A) (3) Figure 3-3. Effect of exogenous JA and OPDA on the accumulation of P1 in wild-type and JL-l plants. Two-week—old wild-type (open bar) and J L- 1 (filled bar) plants were supplied through the cut stem with phosphate buffer (pH 6.5) containing various amounts of J A (A) and OPDA (B). Plants were assayed for PI-II accumulation in leaves 24 hr after treatment. Data represent the mean and standard deviation of 6 plants. Pl-ll ( pglml Ieafjuice) Pl-ll (pg/ml leaf juice) .3.3 #05 CC .3 O O I‘LL 10 nmol JAlplant .3 .3 .3 .3 N h 0) oo 0 N «F O) O O O O O O O O 1 1 1 1 1 1 J A l—L O 0 9O 10 nmol OPDA/plant l expression of PHI was observed in the damaged J L1 rootstock but not in undamaged scion leaves (Figure 3-4, lane 4). In hybrid grafts between wild-type and J L1, wounding of JL1 rootstock leaves did not significantly increase the expression of PM] in the wild- type scion (Figure 3-4, lane 5 and 6). Thus, the wounded J L1 rootstock is defective in the production of the long-distance signal for systemic PI-II expression. In the reciprocal combination (Figure 3—4, lane 7 and 8), wounding of wild-type rootstock leaves resulted in strong induction of PHI expression in the JL1 scion. This result indicates that J L1 plants are able to respond to the long-distance signal that is produced and transmitted from the damaged wild-type rootstock. Taken together with the deficiency of J A in J L1 plants, these results indicate that the conversion of OPDA to J A is required for the generation of a systemic wound signal in damaged tissue. Growth of JL1 seedlings is retarded Defects of J L1 plants were further investigated. In addition to JA deficiency, J L1 plants displayed retarded growth at the seedling stage. To compare the grth rate between J L1 and wild-type plants, the root growth of seedlings was measured. To synchronize the seed germination, both J L1 and wild-type seeds were germinated on filter paper saturated with distilled water. Gerrninating seeds having a protruding radicle of the same size (about 2 mm) were then transferred to new filter paper saturated with distilled water. These seedlings were incubated for four days in darkness, and then root length was measured. The result shows that root length of J L1 seedlings was 42% of 91 scion rootstock wounding scion rootstock scion “-.fl-“-- eIF4A '°°t3t°°k C on up n on at b «I Figure 3-4. Wound-inducible PI-II expression in grafts between JL1 and wild-type plants. J L1(J L) and wild-type (WT) plants were grafted in the four combinations indicated. The genotypes listed above and below the horizontal line correspond to the scion and rootstock, respectively. For each graft combination, plants were divided into a negative unwounded control (-) and wounded (+) group consisting of four grafted plants per group. For wound treatment, each leaflet on the rootstock was mechanically injured with a hemostat. Eleven hours after wounding, leaf tissue was harvested separately from wounded rootstock leaves and undamaged scion leaves (scion) for RNA extraction. Levels of PHI mRNA were analyzed by RNA gel blot analysis, using an eIF4A cDNA probe as a loading control. 92 that of wild-type (Figure 3-5). This finding indicates that root growth of J L1 seedlings is significantly retarded relative to wild-type. However, growth retardation was not observed in adult J L1 plants that developed fully expanded leaves (data not shown). Discussion JL1 plants are unable to convert OPDA to JA The results of the present study indicate that the defective wound response of JL1 plants results from a deficiency in J A accumulation. Three lines of evidence indicate that JL1 plants have a defect in the conversion of OPDA to J A. First, J L1 leaves contained less than 2% of the level of J A observed in wild-type leaves (Figure 3-2A). This is the most severe J A deficiency among known wound response mutants of tomato (Howe et al., 1996; Lee and Howe, 2003; Li et al., 2002; Table 4-1, Chapter 4). Second, the OPDA level of J L1 plants was comparable to that of wild-type (Figure 3-2B). In contrast, other wound response mutants such as spr2 and def] are deficient in both of OPDA and J A (Table 4-2, Chapter 4; Stenzel et al., 2003). This result indicates that the mutation in J L1 affects a step downstream of OPDA in the octadecanoid pathway. Finally, exogenous J A, but not OPDA, induced the accumulation of PHI in J L1 plants (Figure 3-3). Because OPDA is a precursor of J A, these results indicate that J L1 plants are unable to metabolize OPDA to J A. 93 (D O root length (mm) —8 N 00 -h 01 O) \1 C) O O O O O O O wild-type JL1 Figure 3-5. Comparison of root length between wild-type and JL1 seedlings. Germinated seeds at the same stage of development (2 mm radicle) were transferred to 3-MM filter paper saturated with distilled water, and incubated at room temperature in darkness. Root length was measured after 4-day incubation. Data represent the mean and standard deviation of 10 plants. 94 The conversion of OPDA to IA is accomplished by consecutive reactions involving the transport of OPDA from chloroplasts to peroxisomes, reduction of the cyclopentenone ring of OPDA by OPDA reductase, and the removal of 6 carbon units from the carboxylate side chain by three rounds of B-oxidation (Figure 3-1; Wastemack and Hause, 2002). Virtually nothing is known about the mechanism involved in transport of OPDA to peroxisomes. Previous studies in Arabidopsis indicated that substrates for B-oxidation are imported into the peroxisome by an ATP-binding cassette transporter called anl/Ped3/COMATOSE, which is located on the peroxisomal membrane (Footitt et al., 2002; Hayashi et al., 2002; Zolman et al., 2001b). It is unlikely that the J L1 mutant is impaired in OPDA transport because exogenous OPDA failed to induce the accumulation of PM]. Rather, it is more likely that JL1 plants are defective in the conversion of OPDA to J A in peroxisomes. OPDA reductase genes (OPRs) comprise a small gene family in Arabidopsis and tomato. However, only OPR3 is involved in the biosynthesis of J A (Sanders et al., 2001; Schaller and Weiler 1997; Schaller et al., 2000; Strassner et al., 2002). The product of OPR3 is located in peroxisomes (Strassner et al., 2002), and the absence of OPR3 in Arabidopsis results in J A deficiency and male sterility (Sanders et al., 2001; Stintzi et al., 2001). Mapping experiments indicate that the mutation responsible for the wound response phenotype of J L1 does not correspond to the OPR3 locus of tomato (C. Li and GA. Howe, unpublished result) JL1 shows growth retardation at the seedling stage (Figure 3-5). A similar phenotype has not observed in the opr3 mutant of Arabidopsis or other J A-deficient mutants of tomato (Howe and Ryan, 1999; Howe et al., 1996; Sanders et al., 2001). The 95 defect in J A biosynthesis and retarded seedling growth of J L1 indicates that this mutant is compromised in B-oxidation. In addition to its role in J A biosynthesis, B-oxidation also mediates degradation of fatty acids by the sequential removal of two carbon units in the form of acetyl—coenzyme A (CoA; Graham and Eastrnond, 2002). In germinating seeds, storage lipids are metabolized by B—oxidation with subsequent synthesis of glucose. Thus, impeded seedling growth is often observed in mutants defective in [3- oxidation (Hayashi et al., 1998; Zolman et al., 2000). In such mutants, the defect in vegetative growth is typically alleviated after the transition to photoautotrophic growth (Hayashi et al., 1998; Hayashi et al., 2002; Zolman et al., 2001b). This phenotype was observed in JL1 plants; grth retardation was not observed in JL1 plants that had developed more than four leaves (data not shown). Therefore, it is less likely that the defective grth of JL1 seedlings results from a secondary mutation that adversely affects general aspects of plant development. Instead, the JL1 mutant appears to be compromised in B—oxidation, which inhibits both JA biosynthesis and seedling growth. fi-oxidation in plants occurs in peroxisomes and involves the sequential action of three enzymes: first, acyl—CoA oxidase; second, a multifunctional protein possessing L- 3-hydroxyacyl-COA hydrolyase, L-3-hydroxyacyl-dehydrogenase, D-3-hydroxyacyl- CoA epimerase, and 3, 2-enoyl-C0A isomerase activities; third, 3-ketoacyl-CoA thiolase (Eastmond et al., 2000; Germain et al., 2001; Richmond and Bleecker, 1999). Substrates for B-oxidation are first activated by acyl-CoA synthetase (F ulda et al., 2002). Enzymes involved in the B-oxidation cycle are encoded by small gene families in Arabidopsis, which results in the partial overlap of the substrate specificity among isozymes (Eastrnond and Graham, 2002; Eastmond et al., 2000; Germain et al., 2001; 96 Hayashi et al., 1998; Hayashi et al., 1999; Hooks et al., 1999; Richmond and Bleecker, 1999; Shockey et al., 2002). Additional work will be necessary to provide direct evidence for a defect in fi—oxidation in J L1 plants. OPDA is not a wound signal for PI expression in tomato Previous studies have shown that OPDA, in the absence of the conversion to J A, is a signaling molecule that regulates several aspects of plant development. For example, OPDA was more effective than J A in inducing tendril coiling of Cucurbitaceae plants (Weiler et al., 1993; Weiler et al., 1994). Furthermore, exogenous OPDA and J A treatments resulted in distinct patterns of volatile emission in lima bean (Koch et al., 1999). The study of the opr3 mutant demonstrated that OPDA is a signal for defense responses in Arabidopsis (Stintzi et al., 2001). In spite of being unable to convert OPDA to J A, opr3 plants were as resistant to insect and fungal pathogens as were wild- type plants. The resistant phenotype of opr3 plants contrasts the susceptible phenotype of other J A biosynthetic mutants of Arabidopsis that do not synthesize OPDA, or mutants that are insensitive to J A. cDNA microarray experiments firrther showed that wild-type and opr3 plants expressed defense-related genes in a similar manner upon wounding. These results indicate that OPDA functions as a defense signal in Arabidopsis. In contrast to the situation in Arabidopsis, the results presented in this chapter indicate that OPDA is not an active signal for the expression of wound-inducible PIs in tomato. J L1 plants failed to express PIs upon wounding or application of OPDA, whereas J A treatment restored PI-II accumulation in JL1 plants (Figure 3-3). Therefore, 97 the conversion of OPDA to J A is essential for PI expression in tomato. This finding is consistent with a previous study demonstrating that fl-oxidation is required for expression of P13 in tomato (Miersch and Wastemack, 2000). In this study, I examined PI-II expression in response to J A analogues containing a carboxylate side chain of differing length. PI-II expression was induced by analogues carrying an even-number of carbons in the side chain (i. e., OPC-8, OPC-6, and OPC-4; Figure 3-1). In contrast, J A analogues containing an odd-number of carbons in the side chain, which cannot be converted to J A by B-oxidation, were inactive. These results indicated that B-oxidation of OPC-8 is necessary for PM] expression in tomato. Increasing evidence indicates that J A biosynthesis is essential for the systemic wound response of tomato. Reciprocal grafting experiments using J A-deficient spr2 plants showed that J A biosynthesis is required for the production of the transmissible wound signal (Li et al., 2002). Because spr2 plants lack both OPDA and J A because of the loss of an co-3 fatty acid desaturase (Li et al., 2003; Table 4-2, Chapter 4), these results do not rule out the possibility that OPDA is a mobile signal for the systemic wound response. Grafting experiments performed with JL1 plants, which accumulate OPDA but not J A, showed that this mutant failed to generate a systemic wound signal in damaged leaves, but nevertheless was able to perceive that signal in unwounded leaves. These results indicate that J A rather than OPDA is the a mobile signal for systemic PI expression 98 Materials and Methods Plant material Seeds of J L1 were collected from a J L1 homozygous line that was back-crossed two times using Lycopersicon esculentum Mill cv Castlemart as the recurrent parent (Li ghtner et al., 1993). Wounding was performed on two-week-old plants as described in Chapter 2. Grafting experiments were performed following the method described by Li et al. (2002) Treatment of JA and OPDA (i)-J A and OPDA were purchased from Sigma (St. Louis, MO) and Cayman Chemical (Ann Arbor, MI), respectively. Two-week-old plants were excised at the base of the stem with a razor blade and immediately placed into 0.5 m1 plastic tubes containing various amounts of elicitors diluted in 300 11.1 of 15 mM sodium phosphate solution (pH 6.5). Because J A and OPDA were originally dissolved in ethanol, the mock control solution contained 0.1 % (vol/vol) ethanol. Excised plants were allowed to imbibe the solution during a 50 min incubation period, and then transferred to glass vials containing distilled water. PI-II levels in leaves are measured 24 hr after treatment as described in Chapter 2. RNA gel blot analysis RNA extraction from leaves and RNA hybridization experiments were performed as described in Chapter 2. 99 Root length measurement Seeds were surface sterilized as described (Zolman et al., 2000). To synchronize the seed germination, both J L1 and wild-type seeds were germinated on filter paper saturated with distilled water. Germinating seeds with a radicle of about 2 mm length were transferred to two layers of 3-MM filter paper soaked with distilled water. The transferred seedlings were incubated for four days in darkness at room temperature. Root growth was measured with a ruler using overhead projector to enlarge the image of seedlings. Analysis of JA and OPDA Quantification of J A and OPDA was performed by GC-MS analysis as described in Chapter 4. 100 References Conconi, A., Miquel, M., Browse, J ., and Ryan, C. (1996). Intracellular levels of free linolenic and linoleic acids increase in tomato leaves in response to wounding. Plant Physiol. 111: 797-803. Doares, S., Narvaez-Vasquez, J ., Conconi, A., and Ryan, C. (1995). Salicylic acid inhibits synthesis of Proteinase Inhibitors in tomato leaves induced by systemin and jasmonic acid. Plant Physiol. 108:1741-1746. Eastmond, P., Hooks, M., Williams, D., Lange, P., Bechtold, N., Sarrobert, C., Nussaume, L., and Graham, I. (2000). Promoter trapping of a novel medium- chain acyl-CoA oxidase, which is induced transcriptionally during Arabidopsis seed germination J. Biol. Chem. 275: 34375-343. Eastmond, P., and Graham, I. (2000). The multifunctional protein AtMF P2 is co- ordinately expressed with other genes of fatty acid beta-oxidation during seed germination in Arabidopsis thaliana (L.) Heynh. Biochem. Soc. Trans. 28:95-9. Farmer, E., Johnson, R., and Ryan, C. (1992). Regulation of proteinase inhibitor gene expression by methyl jasmonate and jasmonic acid. Plant Physiol. 98: 995-1002. Farmer, E., and Ryan, C. (1992). Octadecanoid precursors of jasmonic acid activate the synthesis of wound-inducible proteinase inhibitors. Plant Cell 4: 129-134. Footitt, S., Slocombe, S., Lamer, V., Kurup, S., Wu, Y., Larson, T., Graham, 1., Baker, A., and Holdsworth, M. (2002). Control of germination and lipid mobilization by COMA T OSE, the Arabidopsis homologue of human ALDP. EMBO J. 21: 2912-2922. Fulda, M., Shockey, J., Werber, M., Wolter, F., and Heinz, E. (2002). Two long-chain acyl-CoA synthetases from Arabidopsis thaliana involved in peroxisomal fatty acid beta-oxidation. Plant J. 32: 93-103. Germain, V., Rylott, E., Larson, T., Sherson, S., Bechtold, N, Carde, J., Bryce, J ., Graham, I., and Smith, S. (2001). Requirement for 3-ketoacyl-CoA thiolase-2 in peroxisome development, fatty acid beta-oxidation and breakdown of triacylglycerol in lipid bodies of Arabidopsis seedlings. Plant J. 28: 1-12. Graham, I., and Eastmond, P. (2002). Pathways of straight and branched chain fatty acid catabolism in higher plants. Prog. Lipid Res. 41: 156-181. Haider, G., Kislinger, T., and Kutchan, M. (1997). Barbiturate induced benzophenanthridine alkaloid formation proceeds by gene transcript 101 accumulation in the California poppy. Biochem. Biophys. Res. Commun. 241 : 606-610. Hayashi, H., De Bellis, L., Ciurli, A., Kondo, M., Hayashi, M., and Nishimura, M. (1999). A novel acyl-CoA oxidase that can oxidize short-chain acyl-CoA in plant peroxisomes. J. Biol. Chem. 274: 12715-21. Hayashi, M., Nito, K., Takei-Hoshi, R., Yagi, M., Kondo, M., Suenaga, A., Yamaya, T., and Nishimura, M. (2002). Ped3p is a peroxisomal ATP-binding cassette transporter that might supply substrates for fatty acid beta-oxidation. Plant Cell Physiol. 43: 1-11. Hayashi, M., Toriyama, K., Kondo, M., and Nishimura, M. (1998). 2, 4- dichlorophenoxybutyric acid-resistant mutants of Arabidopsis have defects in glyoxysomal fatty acid beta-oxidation. Plant Cell 10: 183-195. Hooks, M., Kellas, F., and Graham, I. (1999). Long-chain acyl-CoA oxidases of Arabidopsis. Plant J. 20: 1-13. Howe, G.A., and Ryan, CA. (1999). Suppressors of systemin signaling identify genes in the tomato wound response pathway. Genetics 153:1411-1421. Howe, G.A., Lightner, J ., Browse, J ., and Ryan, CA. (1996). An octadecanoid pathway mutant (J L5) of tomato is compromised in signaling for defense against insect attack. Plant Cell 822067-2077. Koch, T., Krumm, T., Jung, V., Engelberth, J ., and Boland, W (1999). Differential induction of plant volatile biosynthesis in the lima bean by early and late intermediates of the octadecanoid-signaling pathway. Plant Physiol. 121: 153- 162. Li, C., Liu, G, Xu, C., Lee, G. 1., Bauer, P., Ling, H., Ganal, M., and Howe, GA. (2003). The tomato suppressor of prosystemin-mediated responses 2 (spr2) gene encodes a fatty acid desaturase required for the biosynthesis of jasmonic acid and the production of a systemic wound signal for defense gene expression. Plant Cell 15: 1646-1661. Li, L., Li, C., Lee, G.I., and Howe, G (2002). Distinct roles for jasmonate synthesis and action in the systemic wound response of tomato. Proc Natl Acad Sci USA. 99: 6416-6421. Lightner, J ., Pearce, G., Ryan, C.A., and Browse, J. (1993). Isolation of signaling mutants of tomato (Lycopersicon esculentum). Mol. Gen. Genet. 241: 595-601. 102 Miersch, 0., and Wastemack, C. (2000). Octadecanoid and jasmonate signaling in tomato (Lycopersicon esculentum Mill.) leaves: endogenous jasmonates do not induce jasmonate biosynthesis. Biol. Chem. 381: 715-722. Ryan, CA. (1967). Quantitative determination of soluble cellular proteins by radial diffiision in agar gels containing antibodies. Anal. Biochem. 19, 434-440. Richmond, T., and Bleecker, A. (1999). A defect in beta-oxidation causes abnormal inflorescence development in Arabidopsis. Plant Cell 11: 1911-24. Sanders, P. M., Lee, P. Y., Biesgen, C., Boone, J. D., Beals, T. P., Weiler, E.W., and Goldberg, R. B. (2000). The Arabidopsis DELAYED DEHISCENCEI gene encodes an enzyme in the jasmonic acid synthesis pathway. Plant Cell 12:1041- 1061. Schaller, F ., and Weiler, E. (1997). Molecular cloning and characterization of 12- oxophytodienoate reductase, an enzyme of the octadecanoid signaling pathway from Arabidopsis thaliana. Structural and functional relationship to yeast old yellow enzyme. J. Biol. Chem. 272:28066—28072. Schaller, F ., Biesgen, C., Mfissig, C., Altrnann, T., and Weiler, E. (2000). 12- Oxophytodienoate reductase 3 (OPR3) is the isoenzyme involved in jasmonate biosynthesis. Planta 210: 979-984. Shockey, J ., Fulda, M., and Browse, J. (2002). Arabidopsis contains nine long-chain acyl-coenzyme a synthetase genes that participate in fatty acid and glycerolipid metabolism. Plant Physiol. 129: 1710-22. Stintzi, A., Weber, H., Reymond, P., Browse, B., and Farmer, E. (2001) Plant defense in the absence of jasmonic acid: the role of cyclopentenones. Proc. Natl. Acad. Sci. USA. 98: 12837-12842. Strassner, J ., Schaller, F., Frick, U., Howe, G.A., Weiler, E., Amrhein, N., Macheroux, P., and Schaller, A. (2002). Characterization and cDNA-microarray expression analysis of 12- oxophytodienoate reductases reveals differential roles for octadecanoid biosynthesis in the local versus the systemic wound response. Plant J. 32: 585-601. Turner, 1., Ellis, C., and Devoto, A. (2002). The jasmonate signal pathway. Plant Cell 14S: 153-164. Wastemack, C., and Hause, B. (2002). J asmonates and octadecanoids: signals in plant stress responses and development. Prog. Nucleic. Acid Res. Mol. Biol. 72: 165- 221. 103 Weiler, E., Kutchan, T., Gorba, T., Brodschelm, W., Niesel, U., and Bublitz, F. (1994). The Pseudomonas phytotoxin coronatine mimics octadecanoid signalling molecules of higher plants. F EBS Lett 345: 9-13. Weiler, E., Albrecht, T., Groth, B., Xia, Z.., Luxem, M., Liss, H., Andert, L., and Spengler, P. (1993). Evidence for the involvement of jasmonates and their octadecanoid precursors in the tendril coiling response of Bryonia dioica. Phytochemistry. 32: 591-600. Zolman, B., Monroe-Augustus, M., Thompson, B., Hawes, J ., Krukenberg, K., Matsuda, S., and Bartel, B. ( 2001a). chyI, an Arabidopsis mutant with impaired beta-oxidation, is defective in a peroxisomal beta-hydroxyisobutyryl-COA hydrolase. J. Biol. Chem. 276: 31037-31046. Zohnan, B., Silva, 1., and Bartel, B. (2001b). The Arabidopsis pxaI mutant is defective in an ATP-binding cassette transporter-like protein required for peroxisomal fatty acid beta- oxidation. Plant Physiol. 127: 1266-1278. Zolman, B., Yoder, A., and Bartel, B. (2000) Genetic analysis of indole-3-butyric acid responses in Arabidopsis thaliana reveals four mutant classes. Genetics 156:1323-1337. 104 Chapter 4 Quantification ofjasmonic acid and 12-oxo- phytodienoic acid in tomato plants by gas chromatography-mass spectrometry 105 Introduction J asmonic acid (J A) and its methyl ester (MeJ A) regulate plant defense responses against insects and pathogens (Cohen et al., 1993; Farmer and Ryan, 1992; Howe et al., 1996; McConn et al., 1997; Penninckx etal., 1996; Seo etal., 2001), and also play an important role in reproductive organ development in plants (McConn and Browse, 1996; Li et al., 2001; Sanders et al., 2000; Stintzi and Browse, 2000). JA is synthesized from linolenic acid by the octadecanoid pathway (Turner et al., 2002; Wastemack and Hause, 2002). The synthesis of J A is initiated in chloroplasts where linolenic acid is released from plastid membranes and converted to 12-oxo-phytodienoic acid (OPDA) by the consecutive action of lipoxygenase (LOX), allene oxide synthase (AOS), and allene oxide cyclase (AOC). The remaining biosynthetic steps occur in peroxisomes, where OPDA is processed to J A by OPDA reductase (OPR3) and three rounds of 5- oxidation. Interestingly, OPDA functions as a regulatory signal for defense and development, in the absence of its metabolic conversion to J A (Stelrnach et al., 1998; Stintzi et al., 2001; Weiler et al., 1993). These derivatives of linolenic acid, OPDA, J A and Me] A, are well-known examples of oxylipins that are linear or cyclic oxidation products derived from the catabolism of fatty acids (Howe and Schilmiller, 2002). Precise quantitative analysis of J A and OPDA is crucial for the study of the biosynthesis and function of these molecules (Creelrnan and Mullet, 1995). Quantification of J A and OPDA requires the use of appropriate internal standards and an analytical method for the detection. Because internal standards are used to correct for the loss of each oxylipin during extraction, it is important that the internal standard has the same or very similar physiochemical properties of the endogenous compound being 106 analyzed. Two types of internal standards have been used for quantification of JA and OPDA. The first type is isotope-labeled standards such as deuterium- and/or 13 C-labeled J A (Baldwin et al., 1994; Creehnan and Mullet, 1995; Creehnan et al., 1992; Kramell et al., 2000; Laudert and Weiler, 1998) and OPDA (Kramell et al., 2000; Parchmann et al., 1997; Stelrnach et al., 1998). Alternatively, 9, lO-dihydrojasmonic acid (DHJ A; Gundlach et al., 1992; Parchmann et al., 1997; Weber et al., 1997; Wilbert et al., 1998) and tetrahydro-OPDA (H4OPDA; Weber et al., 1997) have been used as internal standards for quantification of J A and OPDA, respectively (Figure 4-1). Two detection methods have been used to quantify J A and OPDA. The first method is an enzyme-linked immunosorbent assay (ELISA) that relies on a monoclonal antibody raised against MeJ A (Albrecht et al., 1993). To avoid possible cross-reaction of the antibody with J A-related molecules, the J A fraction was purified from plant extracts using hi gh-performance liquid chromatography (HPLC) and methylated prior to ELISA testing. Although this method has been used to measure endogenous J A levels in several plants including tomato (Albrecht et al., 1993; Conconi et al., 1996; Doares et al., 1995; Howe et al., 1996; Weiler et al., 1993), quantification of OPDA is net possible with this method. The second method, gas chromatography (GC)-mass spectrometry (MS), provides highly reliable identification and quantification of J A and related oxylipins. GC-MS analysis has been successfully used for the monitoring of J A levels in several plant species including soybean, broad bean, Arabidopsis, apple, tobacco, potato, and barley (Baldwin et al., 1994; Creelman et al., 1992; Fan et al., 1997; Kramell et al., 1997; McConn et al., 1997; Mueller and Brodschelm, 1994; Wang et al., 1999; Weber et al., 1997). In contrast to the ELISA method, GC-MS analysis 107 enables simultaneous quantification of J A, OPDA, and other oxylipins (Kramell et al., 2000; Vollenweider et al., 2000; Weber et al., 1997). Wound-induction of proteinase inhibitors (P13) in tomato has been widely used as a model system to study wound signaling in plants (Green and Ryan, 1972; Ryan, 2000). Several wound response mutants of tomato have been isolated with genetic screens for plants that lack accumulation of P18 in response to wounding (Howe and Ryan, 1999; Li et al., 2001; Lightner et al., 1993). Quantification of JA and OPDA in these mutants is an essential step in investigating the molecular and biochemical basis of these mutations. Previously, GC-MS analysis was used to measure the amount of J A and OPDA in tomato leaves (Parchmann et al., 1997). This published protocol relied on extensive purification of JA and OPDA prior to GC-MS analysis, which reduced the recovery of these compounds. The purpose of the experiments described in this chapter was to optimize and simplify this procedure using a solid-phase extraction method that was originally developed for oxylipin analysis in Arabidopsis (Weber et al., 1997). The results indicated that this modified method can be used to efficiently quantify J A and OPDA in tomato tissues. The method was used to measure the levels of J A and OPDA in various wound response mutants. The results obtained are discussed in the context of the regulation of JA synthesis. 108 O 7 _ C 3 (3R, 78)-JA (3R, 7R)-JA 0 (3R, 7S)-DHJA 98, 138-OPDA 98, 13R-OPDA O m... 98, 138-H4OPDA Figure 4-1. Structure of JA, OPDA and their internal standards. The octadecanoid pathway yields only the 3R, 7S-isomer of J A, which is derived from its precursor 9S, 138-OPDA (Laudert et al., 1997). During extraction and analysis by GC, 98, 13S-OPDA and 3R, 7S-JA epimerize to 98, l3R-OPDA, and 3R, 7R-J A, respectively (Mueller and Brodschelm, 1994). J A and OPDA were converted to DHJ A and H4OPDA, respectively, for use as internal standards. 109 Results Preparation of internal standards and determination of GC-MS parameters Because I A and OPDA are commercially available, these molecules were used to prepare the corresponding internal standards. J A has two chiral centers at the C-3 and C-7 positions (Figure 4-1), which yield four possible stereoisomers (Creelman and Mullet, 1997): 3R, 7S—JA; 3S, 7R-JA; 3R, 7R-JA; and 3S, 7S- JA [also called (+)-7-iso- J A, (-)-7-iso-JA, (-)-J A, and (+)-JA, respectively]. Chemically synthesized JA [(i)-J A] is a racemic mixture of these stereoisomers. OPDA also has two chiral centers at the C- 9 and C-13 positions. Commercially available OPDA [(i)-OPDA] consists of a mixture of 9S, 13R-OPDA and 9S, 13S-OPDA (Figure 4-1). The internal standards DHJA and H4OPDA were prepared from (i)-J A and (:t)-OPDA, respectively, by the reduction of double bonds (Weber et al., 1997 ; Wilbert et al., 1998). Endogenous DHJA and H4OPDA were not detected in the GC-MS analysis of tomato leaf extracts (data not shown), indicating that these compounds would be suitable as internal standards. (i)-J A, (i)-OPDA, and the corresponding internal standards were used to establish GC-MS conditions. To increase the sensitivity of detection (Fans et al., 1997), these molecules were methylated prior to injection into the GC. These methyl esters were readily separable by GC, and more determined by MS (positive-ion detection mode). Molecules were ionized in MS by the electron impact (EI) method to enhance the fragmentation of the molecules. Two temperature programs were used for the GC- MS analysis. The first program was optimized for quantification of J A (Figure 4-2). The 110 (A) 84055.: g 3.0 551 .83 22055 I“._ 31.0551 “ “ o ............. - .............. I. 310E5- g 8.0 E4 ~ 'g 6.0 E41 3 4.0 E4 - (3R. 7R) (3R. 78) ~20E4- 0M ' j 135' h 1610' ' ‘ 1615'T71'7To' ' ' Time (min) (B) 3 1.5 E51 83 5 J COOCH .51.OE5. 3 1: ; ‘ 153 153 a 0: ............ 1 I .............. 1:1 100 120 140 160 180 200 220 mlz 41 c 53054. §20 E4 3R, 7R 3R,7S N 1.0 E4 ( ) 1 ) 0 . .A'_.‘:\j+‘_vaA_'/AI '__.-I-\____‘. 15.5 16.0 16.5 17.0 Time (min) Figure 4-2. Mass spectrum and chromatogram of MeJA and MeDHJA. JA was extracted from wounded leaves of wild-type plants and methylated prior to GC-MS anaylsis. For quantification of endogenous JA levels, ions of m/z = 224 and 226 were monitored for MeJ A (A) and MeDHJA (B), respectively. Ions of m/z = 83, 151, 153 were used for the identification of each peak in the chromatogram. The configuration of each stereoisomer is indicated in the chromatogram. GC-MS was run with the temperature program optimized for J A quantification. 111 (A) (B) 238 3 3.0 1551 - 1.: 95 coocrlsl g 2.0 E5« 1: 1 0 E5 ’163 3 - x In ' 153 , \\ N -:-| MD 100 120 140 160 180 200 220 mlz 01.055 =80E4 § 6.0 E4 ‘ § 4.0 E4 .2 . I“ 2.0 E4. (98’ 13R) 4.22.: r k W f/YM‘ML’: 27.0 28. 0 Time (min) 240 g 3.0 E5 1 ~83 g 2.0 155 cooc113, C 24 3 1.0 E5- 153., 1,163 0 “ .............. L ................ I 100 120 140 160 180 200 220 240 mlz 3 1.0 55* GB 8.0 E4 (93’ 138) §6.0E4l .n N 4.0 E41 04 27.0 28.0 Time (min) Figure 4-3. Mass spectrum and chromatogram of MeOPDA and MeDHJA. OPDA and J A were extracted from wounded leaves of wild-type plants. For quantification of endogenous OPDA, ions of m/z = 238 and 240 were monitored for MeOPDA (A) and MeH4OPDA (B), respectively. The configuration of each stereoisomer is indicated in the chromatogram. Ions of m/z = 83, 95, 96, 153, 163 were used for identification of each peak. 112 second program has a longer running time, which is necessary to quantify J A and OPDA simultaneously in the same plant extract. The latter program permits the separation and detection of all peaks corresponding to MeJ A, MeOPDA (OPDA-methyl ester), MeDHJ A (DHJA-methyl ester), and MeH4OPDA (H4OPDA-methyl ester; Figure 4-3). For quantification of J A and OPDA, only a representative ion of each molecule was counted in the single-ion monitoring (SI) mode. Molecular ions were monitored for Me] A [mass-to-charge ratio (m/z) = 224] and MeDHJA (m/z = 226). Because the molecular ions of MeOPDA (m/z = 306) and MeH4OPDA (m/z = 310) were not sufficiently abundant for detection in SI mode, ions with m/z = 238 and 240 were monitored (Figure 4-3). Extraction and quantification of endogenous JA and OPDA To minimize unintended loss or modification of JA and OPDA during sample preparation, it was desirable to simplify the extensive purification procedure described by Parchmann et al. (1997). This was accomplished by a modification of the solid-phase extraction method originally developed for oxylipin analysis in Arabidopsis (Weber et al., 1997). Because this method does not require tedious purification steps such as preparatory HPLC, multiple samples could be prepared at the same time. Both J A and OPDA were extracted from tomato tissue with methanol and then partially purified on a C13 column. Known amounts of internal standards were added to the methanol extracts prior to the C13 column step. J A and OPDA recovered from the column were methylated with diazomethane, which converts endogenous J A and OPDA to their corresponding methyl esters. Therefore, the Me] A peak in the chromatogram represents 113 (A) 1o 9 _ y = 4.3392x _ 3 9 A R2 = 0.9926 n 2 I! 3 N N II a E H- O N 2 CB 0 T I I I o 0.5 1 1.5 2 2.5 [JA]I[DHJA] (B) 0.7 0.6 y = 0.0531x . R2 = 0.9532 area of mlz=238l area of mlz=240 o r I I T I 0 2 4 6 8 10 12 [OPDA]I[H40PDA] Figure 4-4. Standard curves for quantification of JA and OPDA. The standard curve of J A (A) was constructed using known amounts of J A and DHJA as described (Wilbert et al., 1998). Similarly, the standard curve of OPDA (B) was obtained using mixtures of OPDA and H4OPDA. The slope of each standard curve was used to calculate the endogenous level of J A and OPDA. 114 the sum total of endogenous J A and Me] A. For simplicity, the amount of endogenous J A was calculated from the Me] A peak in the chromatogram without further distinction between J A and Me] A. Endogenous MeOPDA was not detected in tomato leaves (data not shown). Two stereoisomers of JA (3R, 7R-J A and BK, 7S-JA) and OPDA (9S, 13R— OPDA and 9S, 138-OPDA) were detected in extracts from tomato leaves (Figure 4-2; Figure 4-3). The peak area of the two stereoisomers was added to calculate the level of J A or OPDA. The amount of endogenous J A and OPDA was calculated from the integrated peak areas using a standard curve constructed as follows. Various amounts of J A and DHJ A were mixed, methylated, and analyzed by GC-MS to obtain the standard curve for the quantification of endogenous JA in tomato (Figure 4-4). The area under each representative ion peak was measured fiom the chromatogram, and the ratios of these areas were plotted against the concentration ratios of J A and DHJ A, as described previously (Wilbert et al., 1998). Similarly, the standard curve for OPDA was obtained using mixtures of OPDA and H4OPDA (Figure 4-4). A new standard curve was constructed for each independent experiment. The resulting slope of the curve gave very consistent values among different experiments. The standard curve for J A was linear in the range of 0.5 to 25 nmol, with a correlation coefficient of 0.99. The standard curve for OPDA was linear in the range of 5 to 200 nmol with a correlation coefficient of 0.95. The recovery rate of internal standards was in the range of 60 to 90% in most measurements. Results obtained from experiments in which the recovery rate was less than 50% were discarded to avoid possible errors in the quantification. 115 pmol JA/g FW tissue genotype leaf (hour after wounding) flower 0 hr 1 hr 3 hr WT (cm) 12 1 1 (n=3) 262 1 41 (n=3) 151 :1: 26 (n=3) 1485 i: 290 (n=3) spr2 3 t 1 (n=3) 22 :l: 9 (n=3) 7 :t 1 (n=3) 265 :t 36 (n=3) WT (mt) 9 (n=2) 675 (n=2) ND. 555 :I: 58 (n=3) jai1-1 15 (n=3) 493 (n=2) ND. 85 :t 34 (n=3) WT (cm) 16 (n=1) 245 (n=1) 136 (n=1) N.D. jai1-3 15 (n=1) 130 (n=1) 53 (n=1) N.D. Table 4-1. Summary of JA levels in wound response mutants of tomato. J A levels were compared between wild-type (WT) and wound response mutants. Whereas spr2 and jail -3 plants originated in L. esculentum cv Castlemart (cm), jail-1 plants were originated in a semi-dwarf cultivar, L. esculentum cv Micro- Tom (mt). Therefore, J A levels of each mutant were compared to the appropriate wild-type cultivar. Wound-induced levels of J A were compared between WT and wound response mutants. Leaves of 2-week-old plants were wounded with a hemostat and harvested at the indicated time after wounding for J A extraction. As a negative control, leaves were harvested from unwounded plants (0 hr). J A levels in flowers were compared between wild- type and wound response mutants. Flowers were harvested from one month- old plants grown in the greenhouse. The number of independent measurements is indicated in parentheses. The mean of two independent measurements or the mean i standard deviation from three independent experiments is presented. N.D. indicates that the measurement was not performed. 116 unit: pmol/g FW oxylipin genotype 0 hr 1 hr 3 hr WT 15 (n=2) 241 (n=2) 120 (n=2) JA spr1 13 (n=2) 162 (n=2) 37 (n=2) spr2 5 (n=1) 11 (n=1) N.D. WT 644 (n=2) 741 (n=2) 1307 (n=2) OPDA spr1 301 (n=2) 499 (n=2) 556 (n=2) spr2 133 (n=1) 179 (n=1) ND. Table 4-2. Wound-inducible JA and OPDA levels in wound response mutants. Leaves of wild-type (WT), spr1, and spr2 plants were wounded and harvested for the extraction of oxylipins as described in the legend of Table 4- 1. Levels of J A and OPDA were simultaneously measured to examine the effect of these mutations on the biosynthesis of J A. Except for spr2, each value represents the mean of two independent measurements. ND, not determined. 117 Quantification of JA and OPDA from wound response mutants of tomato In wild-type plants (L. esculentum cv Castlemart), the JA level rapidly increased within one hour after wounding (Table 4-1). In contrast, wound response mutants showed reduced levels of J A. For example, the suppressor of prosystemin-mediated responsesZ (spr2; Howe and Ryan, 1999) mutant accumulated extremely low levels of J A over all time points. The OPDA level was also reduced in spr2 plants (Table 4-2). The semi-dwarf tomato cultivar, Micro-Tom, has been shown to be a usefirl model system for analysis of the wound response (Howe et al., 2000; Meissner et al., 1997). The basal level of J A in unwounded Micro-Tom leaves was comparable to that in Castlemart plants. However, Micro-Tom plants accumulated greater amounts of J A in response to wounding. These results indicate that the genetic background of the cultivar can affect the capacity for J A biosynthesis. Both wounding and exogenous Me] A failed to induce the accumulation of HS in jasmonic acid-insensitive] (jail) plants (Li et al., 2001). A deletion null allele, jail -I, was isolated in a genetic screen of fast-neutron mutagenized Micro-Tom plants. Within 1 hr after wounding, the J A level in jail -1 plants increased to 73% of the level observed in wild-type plants. A similar reduction of J A biosynthesis was found for the EMS- induced jai 1-3 allele (originally named spr5) that was isolated in the Castlemart background. Upon wounding, J A levels injai1-3 plants were approximately 50% of that observed in wild-type plants. These results indicate that insensitivity to J A prevents the maximal synthesis of J A in response to wounding. 118 (A) 60 50 - E 40 U) 3 30 ~ E 3 < 20 . _, 10 « o T control infested (B) 1400 1200 — I E 1000 ~ 1 0! =0 800 ~ E e 600 q 3 n_ 400 1 O 200 . o . control infested Figure 4-5. Accumulation of JA and OPDA in wild-type and defl plants challenged with spider mites. Sixteen-day-old wild-type (open bar) and def] (filled bar) plants were infested by spider mites as described in Materials and Methods. Two days after the challenge, leaflets showing visible symptom of damage (infested) were harvested for the quantification of J A (A) and OPDA (B). J A and OPDA were also quantified from leaves of untreated control plants (control). Values indicate the mean and standard deviation of three independent experiments. 119 The amount of JA in tomato flowers was higher than that in leaves (Table 4—1). Micro-Tom flowers appeared to contain less than half the amount of J A found in Castlemart flowers. Both spr2 and jail -I showed reduced JA levels in flowers. The J A level in spr2 flowers was only 18% of the level observed in Castlemart flowers, whereas jail -1 flowers accumulated less than 15 % of the level of JA observed in Micro-Tom flowers. Simultaneous monitoring of J A and OPDA levels showed that these oxylipins displayed different kinetics of accumulation (Table 4-2). The J A level increased transiently within 1 hr of wounding whereas the level of OPDA steadily increased up to 3 hr after wounding. This finding indicated that an increase in OPDA accumulation does not necessarily result in a proportional increase in J A. Consistent with a previous report (Lee and Howe, 2003), the basal level of J A in the suppressor of prosystemin—mediated responses] (spr1) mutant was normal. One hr afier wounding, this level increased to 65% of the level in wild-type. However, the basal level of OPDA was significantly reduced in spr1 plants compared to wild-type (Table 4-2). These results also indicate that the synthesis of OPDA is not proportional to that of J A. The wound-induced accumulation of OPDA was impaired in spr1 plants as well. Changes in the levels of J A and OPDA were also observed in tomato plants challenged with spider mites (Figure 4-5). This arachnid herbivore pierces epidermal cells and sucks out the cellular contents of the underlying mesophyll cells (Lange and Bronson, 1981). In infested leaves of wild-type plants, a moderate increase in JA was observed (Figure 4-5). Although the OPDA level increased upon challenge of wild-type 120 plants (student’s t-test, P < 0.05), the induction of J A and OPDA was not equivalent. The basal level of J A was comparable between wild-type and defenseless] (def?) plants, but the basal level of OPDA was significantly reduced in def] plants compared to wild- type. Neither J A nor OPDA levels increased in response to herbivore attack of defl plants (P < 0.05). This result is consistent with the previous observation that def] plants are defective in the induction of IA synthesis in response to wounding (Howe et al., 1996). Discussion Efficient quantification of J A and OPDA is important for the study of wound signaling because the induction of J A synthesis is one of the earliest events in the signaling cascade that results in expression of wound-inducible genes (Doares etal., 1995; Farmer and Ryan, 1992). Accordingly, this study was intended to establish a reliable and simple procedure to measure endogenous J A and OPDA in tomato tissues. Recent studies have shown the use of GC-MS to measure levels of J A and OPDA in flowers (Hause et al., 2000) and leaves (Stenzel et al., 2003; Strassner et al., 2002) of tomato plants. Despite minor differences in the methods used, the results are comparable to those presented here. A disadvantage of the method used in this study is that simultaneous quantification of endogenous J A and MeJ A is not possible owing to methylation of the plant extract prior to GC-MS analysis. Alternative derivatization methods (for example, silylation of the hydroxy group in the carboxylate side chains of J A and OPDA with trimethylsilyl reagents) may be helpful to circumvent this problem. 121 Regulation of JA biosynthesis in tomato The results summarized in Table 4-1 show that endogenous J A levels change in response to wounding, the particular tissue type under study, and the genetic background of the cultivar. These factors have been reported to regulate the activity of enzymes involved in IA biosynthesis (Hause et al., 2000; Lauder and Weiler, 1998; Strassner et al., 2002). In leaf tissue, J A levels seem to be regulated by substrate availability. Previous studies have shown that application of linolenic acid to leaves activates the expression of J A-inducible genes in tomato (Farmer and Ryan, 1992). Furthermore, linolenic acid and J A showed a similar accumulation pattern in tomato leaves in response to wounding (Conconi et al., 1996). Transgenic Arabidopsis plants that overexpress AOS did not have altered basal levels of J A in leaves, but showed stronger induction of J A synthesis in response to wounding (Laudert et al., 2000). These results indicate that the biosynthesis of J A in leaves is limited by the release of linolenic acid from membranes. The results of this study indicate that the conversion of OPDA to J A is another rate-limiting step in J A biosynthesis (Table 4-2; Figure 4-5). The level of OPDA was significantly higher than that of J A in unwounded control plants. Three hr after wounding, the level of OPDA continued to increase, whereas the J A content almost returned to the basal level. These results indicate that increased OPDA levels do not necessarily lead to a proportional increase in J A levels. The differential accumulation of OPDA and J A was also observed in Arabidopsis (Laudert and Weiler, 1998; Weber et al., 1997) and potato (Weber et al., 1997). 122 Wound response mutants showed reduced levels of J A and OPDA compared to wild-type plants (Table 4-1; Table 4-2; Figure 4-5). These observations provide clues to understanding the signaling defect in these mutants. For example, the product of Spr2 is expected to be involved in OPDA synthesis, because both J A and OPDA were severely reduced in spr2 plants (Table 4-2). This speculation was confirmed by the recent discovery that Spr2 encodes a fatty acid desaturase catalyzing the synthesis of linolenic acid (Li et al., 2003). Therefore, the deficiency in linolenic acid blocks the production of both OPDA and J A in spr2 plants. Both spr1 and def] plants accumulated less OPDA than did wild-type plants in response to wounding. These observations are consistent with recent findings that Spr] is a positive regulator of IA biosynthesis (Lee and Howe, 2003), and that the activity of AOC is reduced in def] plants (Stenzel et al., 2003). JA-insensitive jail -I and jail -3 plants showed reduced levels of J A in wounded leaves. These results indicate that J A biosynthesis may be regulated by a positive feedback mechanism. Materials and Methods Preparation of internal standards DHJ A and H4OPDA were prepared by PtOz-catalyzed hydrogenation of (:t)-J A (Sigma, St. Louis, MO) and OPDA (Cayman Chemical, Ann Arbor, MI) as described (Weber et al., 1997). The authenticity of the standards, as well as the absence of endogenous DHJA/H4OPDA in tomato leaf extracts, was verified by GC-MS. 123 Extraction of JA and OPDA from tomato tissues Plants containing two fully—expanded leaves and an emerging third leaf were wounded with a hemostat on each leaflet. Three to five g FW of leaves were used for quantification of J A. JA extraction was performed following the procedure described by Weber et al. (1997) with modifications. Harvested leaves were frozen in liquid nitrogen and ground to a fine powder with a chilled mortar and pestle. The tissue was dissolved in 28 ml methanol containing 500 ng DHJ A and 500 ng H4OPDA as internal standards and then homogenized with a Polytron for 1 min at 4°C. The homogenate was incubated for 2 hr at 4°C with shaking, diluted with 12 ml ice-cold water, and then centrifuged at 3,500 X g. The resulting supernatant was recovered and the pH adjusted to 8.0 with NH40H. This solution was centrifirged again at 3,500 x g, and the supernatant was passed through a tClg-SepPak cartridge (Waters Corporation, Milford, MA) that was preconditioned with 70% (vol/vol) methanol and collected in a new glass vials. The cartridge was washed with 7 ml of 75% (vol/vol) methanol. Eluates from both the sample and the wash steps were combined and adjusted to pH 4.0 with 10% (vol/vol) formic acid. This solution was diluted with 160 ml ice-cold water and then loaded on the same thg-SepPak column that was washed sequentially with methanol, diethylether, methanol, and water immediately after the elution with 75% (vol/vol) methanol. After washing the column with 7 ml of 15% (vol/vol) ethanol and 7 ml water, the J A fraction was eluted with 10 ml diethylether. The eluate was partially dried over anhydrous NaZSO4 and then dried completely under a stream of nitrogen gas at 35°C. The dried paste was dissolved in 0.5 ml methanol and subjected to methylation by the addition of diazomethane dissolved in 0.5 ml diethylether. This mixture was dried under nitrogen gas, resuspended in 20 pl hexane, and injected into the GC. Diazomethane was prepared 124 prior to use: 0.2 g N-nitroso-methylurea (Sigma, St. Louis, MO) was dissolved in a mixture of 2 ml ice-cold diethylether and 600 11.1 of 40 % (vol/vol) KOH solution. GC-MS analysis The amount of J A/MeJ A in leaf extracts was quantified by GC-MS by using a Hewlett- Packard GC 5890 equipped with a Hewlett-Packard 5970 mass detector. The GC was fitted with a DB-5 column (30 m X 0.25 mm id, J&W Scientific, Folsom, CA) and run with a temperature gradient of 100°C for l min, 100°C to 170°C at 5°C/min, 170°C for 2.5 min, and 170°C to 250°C at 20°C/min. GC-MS analysis was performed in the SI mode with monitoring of ions specific for Me] A (m/z = 224) and MeDHJ A (m/z = 226). Molecules were ionized by electron impact in MS. For quantification of JA/MeJ A, a standard curve was generated from samples in which MeJ A and MeDHJ A were mixed in known ratios. Because peaks corresponding to the 3R, 7S and 3R, 7R isomers of endogenous J A/MeJ A were detected, the areas of the two peaks were combined. For simultaneous analysis of J A and OPDA, a modified temperature program was used for GC: 100°C for 1 min, 100°C to 160°C at 20°C/min, 160°C to 238°C at 3°C/min, 238°C to 250°C at 30 °C/min. For the quantification, GC-MS was run in SI mode while monitoring ions for MeJ A (m/z = 224), MeDHJA (m/z = 226), MeOPDA (m/z = 23 8) and MeH4OPDA (m/z = 240). Retention times of these molecules were: 9.57 min for 3S, 7S—MeJA; 9.88 min for 3R, 7R-MeJA; 9.97 min for 3R, 7R- MeDHJA;10.51 min for 3R, 7S-MeJA; 10.54 min for 3R, 7S-MeDHJA; 26.93 min for 98, 13R-MeH4OPDA (also called cis-MeH4OPDA); 27.30 min for 9S, l3R-MeOPDA; 27.70 min for 9S, 138-MeH4OPDA (also called trans-MeH4OPDA); and 28.60 min for 125 9S, 13S-MeOPDA. Standard curves for OPDA were constructed using a mixture containing known amounts of OPDA and H4OPDA. The detection limit of Me] A (0.4 nmol) and MeOPDA (0.2 nmol) was determined by injection of known amounts of these molecules. Plant material Seeds of spr1-1, spr2, def] and jail -I were collected from each homozygous line afier at least two backcrosses using Castlemart cv (for spr1-1, spr2, and def!) or Micro-Tom cv (for jai 1-1) as the recurrent parent. Leaves were harvested from two-week-old plants for extraction of oxylipins. For the analysis of jail -3 plants, the second and the third youngest leaves were harvested from 5-leaf-stage plants. Wounding was inflicted on each leaflet with a hemostat. Flowers were harvested in both open and closed stages from plants grown in the greenhouse. Growth conditions are described in Chapter 2. Spider mite treatment Two-spotted spider mite (T etranychus urticae Koch) treatment was performed as described (Li et al., 2002). Spider mite eggs were hatched on leaves of lima bean (Phaseolus lunatus cv Fordhook) plants and allowed to grow to a population density of approximately 50 adult spider mites per fully expanded leaf. Infested leaves were cut and placed onto the upper surface of leaves of l6-day-old wild-type and def] plants. Two days later, tomato leaflets showing the visible symptom of damage were harvested for J A/OPDA extraction. 126 References Albrecht, T., Kehlen, A., Stahl, K., Knofel, H., Sembdner, G., and Weiler, E. (1993). Quantification of rapid, transient increases in jasmonic acid in wounded plants using a monoclonal-antibody. Planta 191: 86-94. Baldwin, 1., Schmelz, E., and Ohnmeiss, T. (1994). Wound-induced changes in root and shoot jasmonic acid pools correlate with induced nicotine synthesis in Nicotiana sylvestris spegazzini and comes. J. Chemical Ecology 20: 2139-2157. Cohen, Y., Gisi, U., and Nidennanm T. (1993). Local and systemic protection against Phytophthora infestans induced in potato and tomato plants by jasmonic acid and jasmonic methyl ester. Phytopathology 83: 1054—1062. Conconi, A., Miquel, M., Browse, J ., and Ryan, C. (1996). Intracellular levels of free linolenic and linoleic acids increase in tomato leaves in response to wounding. Plant Physiol. 111: 797-803. Creelman, R., and Mullet, J. (1997). Biosynthesis and action of jasmonates in plants. Annual Review. Plant Physiol. and Plant Mol. Biol. 48: 355-381. Creelman, R., and Mullet, J. (1995). J asmonic acid distribution and action in plants - regulation during development and response to biotic and abiotic stress. Proc. Natl. Acad. Sci. USA. 92: 4114-4119. Creehnan, R.., Tierney, M., and Mullet, J. (1992). J asmonic acid/methyl jasmonate accumulate in wounded soybean hypocotyls and modulate wound gene expression. Proc. Natl. Acad. Sci. USA. 89: 4938-4941. Doares, S., Narvaez-Vasquez, J ., Conconi, A., and Ryan, C. (1995). Salicylic acid inhibits synthesis of Proteinase Inhibitors in tomato leaves induced by systemin and jasmonic acid. Plant Physiol. 108:1741-1746. Fan, X., Mattheis, J ., Fellman, J ., and Patterson M. (1997). Changes in jasmonic acid concentration during early development of apple fruit. Physiol. Plantarum 101: 328-332. Farmer, E., and Ryan, C. (1992). Octadecanoid precursors of jasmonic acid activate the synthesis of wound-inducible proteinase inhibitors. Plant Cell 4: 129-134. Green, T., and Ryan, C. (1972). Wound-induced proteinase inhibitor in plant leaves - possible defense mechanism against insects. Science 175: 776-777. Gundlach, H., Mueller, M., Kutchan, T., and Zenk, M. (1992). Jasmonic acid is a signal transducer in elicitor-induced plant cell cultures. Proc. Natl. Acad. Sci. USA. 89: 2389-2393. 127 Hause, B., Stenzel, I., Miersch, 0., Maucher, H., Kramell, R., Ziegler, J ., and Wastemack, C. (2000). Tissue-specific oxylipin signature of tomato flowers: allene oxide cyclase is highly expressed in distinct flower organs and vascular bundles. Plant J. 24:113-126. Howe, G.A., Lightner, J ., Browse, J ., and Ryan, C. (1996). An octadecanoid pathway mutant (J L5) of tomato is compromised in signaling for defense against insect attack. Plant Cell 8: 2067-2077. Howe, G.A., and Ryan, C. (1999). Suppressors of systemin signaling identify genes in the tomato wound response pathway. Genetics 153: 1411-1421. Howe, G.A., and Schilmiller, A. (2002) Oxylipin metabolism in response to stress. Curr. Opin. Plant Biol. 5:230-236. Kramell, R.., Miersch, 0., Atzom, R., Parthier, B., and Wastemack, C. (2000). Octadecanoid-derived alteration of gene expression and the "oxylipin signature" in stressed barley leaves. Implications for different signaling pathways. Plant Physiol. 123: 177-188. Kramell, R., Miersch, 0., Hause, B., Ortel, B., Parthier, B., and Wastemack, C. (1997). Amino acid conjugates of jasmonic acid induce jasmonate-responsive gene expression in barley (Hordeum vulgare L.) leaves. FEBS Lett 414: 197-202. Laudert, D., Schaller, F., and Weiler, E. (2000) Transgenic Nicotiana tabacum and Arabidopsis thaliana plants overexpressing allene oxide synthase. Planta 211: 163-165. Lange, W., and Bronson, L. (1981) Insect pests of tomatoes. Annu. Rev. Entomol. 26: 345-371. Laudert, D., and Weiler, E (1998). Allene oxide synthase: a major control point in Arabidopsis thaliana octadecanoid signalling. Plant J. 15: 675-684. Lee, G.I., and Howe, GA. (2003). The tomato mutant spr1 is defective in systemin perception and the production of a systemic wound signal for defense gene expression. Plant J. 33: 567-576. Li, C., Liu, G, Xu, C., Lee, G.I., Bauer, P., Ling, H., Ganal, M., and Howe, GA. (2003). The tomato suppressor of prosystemin-mediated responses 2 (spr2) gene encodes a fatty acid desaturase required for the biosynthesis of j asmonic acid and the production of a systemic wound signal for defense gene expression. Plant Cell 15: 1646-1661. 128 Li, C., Williams, M., Loh, Y.-T., Lee, G.I., and Howe, GA. (2002). Resistance of cultivated tomato to cell-content feeding herbivores is regulated by the octadecanoid signaling pathway. Plant Physiol. 130: 494-503. Li, L., Li, C., and Howe, GA. (2001) Genetic analysis of wound signaling in tomato. Evidence for a dual role of j asmonic acid in defense and female fertility. Plant Physiol. 127:1414-1417. Lightner, J ., Pearce, C., Ryan, C., and Browse, J. (1993). Isolation of signaling mutants of tomato (Lycopersicon esculentum). Mol. Gen. Genet. 241:595-601. McConn, M., Creelman, R., Bell, E., Mullet, J ., and Browse J. (1997). J asmonate is essential for insect defense in Arabidopsis. Proc. Natl. Acad. Sci. USA. 94: 5473-5477. 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. Meissner, R., Jacobson, Y., Melamed, S., Levyatuv, S., Shalev, G., Ashri, A., Elkind, Y., and Levy, A. (1997). A new model system for tomato genetics. Plant J. 12: 1465-1472. Mueller, M., and Brodschelm, W. (1994). Quantification of jasmonic acid by capillary gas chromatography- negative chemical ionization-mass spectrometry. Anal. Biochem. 218: 425-435. Parchmann, S., Gundlach, H., and Mueller, M. (1997). Induction of 12-oxo- phytodienoic acid in wounded plants and elicited plant cell cultures. Plant Physiol. 115: 1057-1064. Penninckx, I., Eggerrnont, K., Terras, F., Thomma, B., Samblanx, G., Buchala, A., Metraux, J ., Manners, J ., and Broekaert. W. (1996). Pathogen-induced systemic activation of a plant defensin gene in Arabidopsis follows a salicylic acid- independent pathway. Plant Cell 8: 2309-2323. Ryan, GA. (2000) The systemin signaling pathway: differential activation of plant defensive genes. Biochim. Biophys. Acta 1477: 112-121. Sanders, P., Lee, P., Biesgen, C., Boone, J ., Beals, T., Weiler, E., and Goldberg, R. (2000). The Arabidopsis DELAYED DEHISCENCEI gene encodes an enzyme in the jasmonic acid synthesis pathway. Plant Cell 12: 1041-1061. Seo, H., Song, J ., Cheong, J ., Lee, Y, Lee, Y., Hwang, I, Lee, J ., and Choi, Y. (2001). J asmonic acid carboxyl methyltransferase: a key enzyme for jasmonate- regulated plant responses. Proc. Natl. Acad. Sci. USA. 98: 4788-4793. 129 Stelmach, B., Mueller, A., Hennig, P., Laudert, D., Andert, L., and Weiler, E. (1998). Quantitation of the octadecanoid lZ-oxo-phytodienoic acid, a signalling compound in plant mechanotransduction. Phytochemistry 47: 539-546. Stenzel, I., Hause, B., Maucher, H., Pitzschke, A., Miersch, 0., Ziegler, J ., Ryan, C., and Wastemack, C. (2003). Allene oxide cyclase dependence of the wound response and vascular bundle-specific generation of j asmonates in tomato - amplification in wound signalling. Plant J. 33:577-589. 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. Stintzi, A., Weber, H., Reymond, P., Browse, J ., and Farmer, E. (2001). Plant defense in the absence of jasmonic acid: the role of cyclopentenones. Proc. Natl. Acad. Sci. USA. 98: 12837-12842. Strassner, J ., Schaller, P., Frick, U., Howe, G., Weiler, E., Amrhein, N., Macheroux, P., and Schaller, A. ( 2002). Characterization and cDNA-microarray expression analysis of 12- oxo-phytodienoate reductases reveals differential roles for octadecanoid biosynthesis in the local versus the systemic wound response. Plant J. 32: 585-601. Turner, J ., Ellis, C., and Devoto, A. (2002). The jasmonate signal pathway. Plant Cell 148: 153-164. Vollenweider, S., Weber, H., Stolz, S., Chetelat, A., and Farmer E. (2000). Fatty acid ketodienes and fatty acid ketotrienes: Michael addition acceptors that accumulate in wounded and diseased Arabidopsis leaves. Plant J. 24: 467-476. Wang, C., Avdiushko, S., and Hildebrand, D. (1999). Overexpression of a cytoplasm- localized allene oxide synthase promotes the wound-induced accumulation of jasmonic acid in transgenic tobacco. Plant Mol. Biol. 40: 783-793. Wastemack, C., and Hause, B. (2002) Jasmonates and octadecanoids: signals in plant stress responses and development. Prog. Nucleic. Acid Res. Mol. Biol. 72: 165- 221. Wilbert, S., Ericsson, L., and Gordon, M. (1998). Quantification of jasmonic acid, methyl jasmonate, and salicylic acid in plants by capillary liquid chromatography electrospray tandem mass spectrometry. Anal. Biochem. 257: 186-194. 130 Weber, H., Vick, B., and Farmer, E. (1997). Dinor-oxo-phytodienoic acid: a new hexadecanoid signal in the j asmonate family. Proc. Natl. Acad. Sci. USA 94: 10473-10478. Weiler, E., Albrecht, T., Groth, B., Xia, Z., Luxem, M., Liss, H., Andert, L., and Spengler, P. (1993). Evidence for the involvement of jasmonates and their octadecanoid precursors in the tendril coiling response of Bryom'a dioica. Phytochemistry 32: 591-600. 131 Chapter 5 Analysis of wound-induced root-to-leaf signaling in tomato plants 132 Introduction Tomato plants express many genes systemically in response to wounding. Previous studies have shown that the polypeptide systemin and its precursor protein prosystemin play crucial roles in the systemic wound response in tomato (Ryan, 2000). Exogenous systemin induces the accumulation of defense proteins such as Proteinase Inhibitors (PIs; Pearce et al., 1991). Systemic accumulation of PIs in response to wounding is severely reduced in transgenic plants that constitutively express a prosystemin cDNA in the antisense orientation (McGurl et al., 1992). Systemin activates the biosynthesis of jasmonic acid (J A), which results in the expression of PIs in tomato (Doares et al., 1995a; Howe et al., 1996). These observations, together with the mobility of systemin applied at wound sites in the phloem (Pearce et al., 1991), led to the hypothesis that systemin is a mobile signal that couples local injury to systemic gene expression (Ryan, 2000). Results from recent studies do not support the proposed role of systemin as the long-distance signal for systemic gene expression. Grafting experiments performed with various wound response mutants of tomato indicate that J A, not systemin, is the long- distance signal for systemic PI expression (Chapter 2; Li et al., 2002). Moreover, it was shown that the systemin-insensitive suppressor of prosystemin-mediated responses] (spr1) mutant is compromised in the generation but not the perception of the systemic signal for PI expression (Chapter 2). This finding indicates that systemin is not the long-distance signal. J A-deficient tomato mutants such as J L1 (Chapter 3), defenseless] (def! ; Howe et al., 1996; Li et al., 2002), and spr2 (Li et al., 2002) are also defective in the generation of the systemic wound signal. However, these plants are able to perceive 133 the systemic wound signal transmitted from wild-type rootstock leaves. In contrast, the perception of the systemic wound signal was blocked in jasmonic acid-insensitive] (jail) plants that fail to express P1 in response to JA (Li et al., 2002). Therefore, JA or a derivative of J A appears to be synthesized in damaged leaves and transported to undamaged tissue to induce expression of P13. Increasing evidence indicates that not all systemic wound responses in tomato are regulated by J A and systemin. For example, wound-induced activation of a mitogen- activated protein kinase (MAPK) is mediated by a rapid, J A-independent signal (Stratmann and Ryan, 1997). It was also reported that wound-induced systemic expression of a glucosyl transferase gene occurs independently of J A and systemin (O’Donnell et al., 1998). Finally, the observation that a subset of wound-inducible genes is systemically expressed in spr1 plants (Figure 2-2 and 2-5, Chapter 2) points to the existence of a systemin-independent signaling pathway. Similar to systemic wound signaling in leaves, a long-distance signaling pathway appears to transmit a wound signal between roots and leaves. For example, mechanical wounding to roots induced PI expression in leaves of potato plants (Dammann et al., 1997). Conversely, leaf damage activates expression of a fatty acid desaturase gene in roots of Arabidopsis (N ishiuchi et al., 1997 ; 1999). At present, very little is known about the signaling pathway that mediates this root-to-shoot systemic wound response. The purpose of this study was to investigate the genetic basis of wound-induced root-to-leaf signaling in tomato. This question was addressed by analyzing the systemic expression of various wound-inducible genes in wild-type and mutant plants that are defective in the systemin/JA-dependent signaling pathway. The 134 results show that a JA-dependent signaling regulates the systemic expression of a subset of wound-inducible genes during root-to-leaf signaling, whereas a JA-independent pathway is involved in the controlling the systemic expression of a different set of genes. Results and Discussion In potato plants, damage to roots can induce gene expression in leaves (Dammann et al., 1997). To determine whether this is also the case in tomato plants, PI-H levels in leaves were measured in two-leaf-stage tomato plants afier wounding of roots. Mechanical crushing of roots with a hemostat resulted in accumulation of PHI in undamaged leaves (Figure 5-1). Although the level of PI accumulation was less than that observed in wounded leaves, this finding indicates that a systemic signal travels between roots and aerial parts of the plant. To investigate whether root-to-shoot signaling depends on systemin and J A, systemic accumulation of PHI was examined in systemin-insensitive spr1 and J A- deficient def] plants. The results showed that root-damaged spr1 plants accumulated normal levels of PHI in leaves. Thus, systemin action appears not to be required for this systemic response. This observation contrasts the role of systemin in a leaf-to-leaf systemic signaling (Chapter 2). A role for systemin in leaf-to-leaf but not root-to-leaf systemic signaling is consistent with the previous report that prosystemin is expressed in leaves but not roots (McGurl et al., 1992). In contrast to spr1 plants, root damage failed to induce systemic PI-II accumulation in def]. The lack of response in def] plants could result from a defect in 135 180 g 160 _ .5 140 - i: 120 1 19 2 100 1 J g 80 - g 60 j ; 4o 5' 2o 1 j 0 __.__L..L__ , 1L unwounded damaged leaf undamgged leaf systemic leaf-to-leaf root-to-leaf Figure 5-1. Systemic induction of PHI in leaves in response to wounding of roots. The accumulation of PHI in leaves was compared among wild- type (open bar), spr1 (gray bar), and def] (defI leaves did not accumulate detectable amounts of PHI) plants. Mechanical wounding was inflicted to roots of two-leaf-stage plants, and leaves were harvested 24 hr later for measurement of PHI (systemic). As a negative control, PI-II was measured in leaves harvested from unwounded plants (unwounded). For comparison to leaf-to-leaf systemic signaling, a single wound was inflicted on a lower leaf of two-leaf-stage plants, and PI-II levels were measured separately in damaged lower leaves (damaged leaf) and undamaged upper leaves (undamaged leaf) 24 hr later. Data represent the mean and standard deviation of six plants. 136 the production of the systemic wound signal in wounded roots, or from a defect in the recognition of that signal in unwounded leaves. Previous grafiing experiments showed that leaf-to-leaf wound signaling requires J A biosynthesis in wounded leaves, not in undamaged leaves (Chapter 3; Li et al., 2002). Because PI synthesis in leaves is mediated by the action of J A (Doares et al., 1995a; Farmer and Ryan, 1992), these results indicate that JA or a derivative of J A is synthesized in wounded leaves and transported to undamaged leaves (Li et al., 2002). Therefore, the lack of PI production in def] leaves in response to root damage indicates that def] roots are defective in wound-inducible J A biosynthesis and the generation of a systemic wound signal. RNA gel-blot analysis was used to determine whether root damage results in activation of gene expression in unwounded leaves. In these experiment, cDNA probes for two groups of wound-inducible genes were employed (Figure 5-2). The first group of these genes is the so-called ‘early’ wound response genes, which shows rapid and transient induction in response to wounding of leaves (Ryan, 2000; Figure 2-2). Early genes include lipoxygenase D (Lox D) involved in J A biosynthesis and a wound- inducible MAPK (WIPK) homologue (Seo et al., 1999). The second group is ‘late’ wound response genes, which are expressed slowly but steadily upon wounding (Ryan, 2002; Figure 2-2, Chapter 2). P13 are representative of this group. Root damage activated strong expression of PHI in wild-type leaves, similar to the pattern observed in response to leaf wounding or stem excision (Figure 2-5, Chapter 2). Wounding of wild-type roots also resulted in rapid and transient systemic expression of LoxD and WIPK. The general pattern of systemic expression of early and late wound response genes in spr1 plants was similar to wild-type plants, although differences in 137 WT spr1 iai1 spr2 hr 012481224012481224 012481224012481224 PI'” v: t 5. Q ’ LoxD "' 2* WIPK ‘ -' L‘ - ; , 1.4L; 1 - ‘ . eIF4A ..CQ. racmgytfioc Figure 5-2. Systemic expression of wound-inducible genes in response to root damage. Mechanical wounding was applied to roots of two-week-old plants with a hemostat. Leaves were harvested from six plants at each time point and pooled for RNA extraction. To facilitate direct comparison of transcript levels in wild-type (WT) and mutant plants, blots containing RNA from the four genotypes were hybridized in the same container and washed under the same conditions. Then X-ray film was exposed to the blots for the same length of time. 138 the magnitude of the response were noted. For example, the expression of PM] in spr1 leaves was significantly lower than in wild-type plants. Because the level of PHI protein in leaves of root-damaged spr1 plants was similar to that of wild-type plants (Figure 5-1), it appears that mRNA levels do not accurately reflect PI protein levels. These phenomena have previously been observed in wounded tomato leaves (Howe et al,1996) Systemic PI-II expression was abolished in leaves of J A-insensitive jail plants, indicating that a functional J A-dependent signaling pathway is required for this systemic wound response. Consistent with the absence of PHI in def] plants (Figure 5- 1), J A-deficient spr2 plants also failed to express PM] in leaves upon root damage. This result demonstrates that J A biosynthesis is required for systemic P] expression. Interestingly, a wound-inducible MAPK homologue (WIPK) was uniformly expressed in all genotypes in response to root damage. Therefore, expression of the WIPK appears to be mediated by a J A-independent signaling pathway. J A-independent wound signaling was previously proposed to account for the expression of an LE RNase gene, a glucosyl transferase gene, and activation of a MAPK activity in tomato leaves (Figure 6-8, Chapter 6; O’Donnell et al., 1998; Stratmann and Ryan, 1997). Unlike WIPK, LoxD expression in jail plants was less abundant than in spr1, spr2, and wild-type plants. This result indicates that the J A-dependent signaling pathway is required for maximal induction of LoxD. The residual expression of LoxD is unlikely due to ‘leakiness’ of the jai 1 mutation because PI-II expression was not detected in jail plants. Furthermore, this allele of jail (jail-1) corresponds to a fast- neutron-induced deletion in the Jail gene (Li et al., 2002). The expression of LoxD in 139 wild-type plants therefore appears to be regulated by both J A-independent and J A- dependent signaling pathways. This conclusion is consistent with the previous finding that LoxD expression is regulated by a systemin-independent pathway (Figure 2-5, Chapter 2). In Arabidopsis, a J A-independent pathway regulates local wound response through the action of oligogalacturonic acid (OGA) released from cell walls (Rojo et al., 1999). It is unlikely that JA-independent signaling in tomato is mediated by OGA because I A biosynthesis is required for OGA-induced gene expression (Doares et al., 1995b). Previous studies indicated that the JA-independent signal could be rapidly transmitted through the xylem (Malone et al., 1995; Rhodes et al., 1999; Stratmann and Ryan, 1997). It may be informative to examine the expression of WIPK using steam- girdled plants in which the xylem but not the phloem is intact. This experiment would provide insight into the question of whether the JA-independent signal for WIPK expression is phloem- or xylem-bome. Materials and Methods Plant material Seeds of spr1-1, spr2, def], and jai 1-1 were collected from each homozygous line after at least two backcrosses, using Castlemart cv as the recurrent parent. As a result, jai I -1 plants used in this study displayed normal growth compared to Castlemart cv. Plants were grown in peat pots as described in Chapter 2. Root damage was applied on two- leaf—stage plants by pinching the peat pot 7 times with a hemostat. The PI-II level was measured as described in Chapter 2. 140 RNA gel-blot analysis Tomato EST clones cLET1D13 and cLED1D24 were used as probes to detect WIPK and eIF 4A mRNA, respectively. Total RNA was harvested and hybridized following the methods described in Chapter 2. 141 References Dammann, C., Rojo, E., and Sanchez-Serrano, J. (1997). Abscisic acid and jasmonic acid activate wound-inducible genes in potato through separate, organ-specific signal transduction pathways. Plant J. 11: 773-782. Doares, S., Narvaez-Vasquez, J ., Conconi, A., and Ryan, C. (1995a). Salicylic acid inhibits synthesis of Proteinase Inhibitors in tomato leaves induced by systemin and jasmonic acid. Plant Physiol. 108: 1741-1746. Doares, S., Syrovets, T., Weiler, E.M., and Ryan, C. (1995b). Oligogalacturonides and chitosan activate plant defensive genes through the octadecanoid pathway. Proc. Natl. Acad. Sci. USA. 92: 4095-4098. Farmer, BE, and Ryan, CA. (1992). Octadecanoid precursors of jasmonic acid activate the synthesis of wound-inducible proteinase inhibitors. Plant Cell 4: 129-134. Hause, B., Stenzel, I., Miersch, O., Maucher, H., Kramell, R., Ziegler, J ., and Wastemack, C. (2000) Tissue-specific oxylipin signature of tomato flowers: allene oxide cyclase is highly expressed in distinct flower organs and vascular bundles. Plant J. 24:113-126. Li, L., Li, C., Lee, G.I., and Howe, G (2002). Distinct roles for jasmonate synthesis and action in the systemic wound response of tomato. Proc Natl Acad Sci USA. 99: 6416-6421. Malone, M. and Alarcon, J. (1995). Only xylem-bome factors can account for systemic wound signaling in the tomato plant. Planta 196: 740-746. McGurl, B., Pearce, G., Orozco-Cardenas, M. and Ryan, CA. (1992). Structure, expression, and antisense inhibition of the systemin precursor gene. Science 255: 1570-1573. Nishiuchi, T., Kodama, H., Yanagisawa, S., and Iba, K. (1999). Wound-induced expression of the FAD7 gene is mediated by different regulatory domains of its promoter in leaves/stems and roots. Plant Physiol. 121: 1239-1246. Nishiuchi, T., Hamada, T., Kodama, H., and Iba, K. (1997). Wounding changes the spatial expression pattern of the Arabidopsis plastid omega-3 fatty acid desaturase gene (FAD7) through different signal transduction pathways. Plant Cell 9: 1701-1712. O'Donnell, P., Turesdale, M., Calver, C., Dorans, A., Roberts, M., and Bowles, D. (1998). A novel tomato gene that rapidly responds to wound- and pathogen- related signals. Plant J .14: 137-142. 142 Pearce, G., Strydom, D., Johnson, S., and Ryan, CA. (1991). A polypeptide from tomato leaves induces wound-inducible proteinase inhibitor proteins. Science 253: 895-898. Rhodes, J ., Thain, J ., and Wildon, D. (1999). Evidence for physically distinct systemic signalling pathways in the wounded tomato plant. Annals Botany 84: 109-116. Rojo, E., Leon, J, and Sanchez-Serrano, J. (1999). Cross-talk between wound signalling pathways determines local versus systemic gene expression in Arabidopsis thaliana. Plant J. 20:135-142. Ryan, C. (2000). The systemin signaling pathway: differential activation of plant defensive genes. Biochem, Biophys. Acta 1477:112-121. Seo, S., Sano, H., and Ohashi, Y. (1999). Jasmonate-based wound signal transduction requires activation of WIPK, a tobacco mitogen-activated protein kinase. Plant Cell 11: 289-298. Strassner, J ., Schaller, F ., Frick, U., Howe, G, Weiler, E., Amrhein, N., Macheroux, P., and Schaller, A. (2002). Characterization and cDNA-microarray expression analysis of 12- oxo-phytodienoate reductases reveals differential roles for octadecanoid biosynthesis in the local versus the systemic wound response. Plant J. 32: 585-601. Stratmann, J .W., and Ryan, GA. (1997) Myelin basic protein kinase activity in tomato leaves is induced systemically by wounding and increases in response to systemin and oligosaccharide elicitors. Proc. Natl. Acad. Sci. USA. 94: 11085- 11089. 143 Chapter 6 Cytochrome P450-dependent metabolism of oxylipins in tomato. Cloning and expression of allene oxide synthase and fatty acid hydroperoxide lyase A version of this chapter has been published. Gregg A. Howe, Gyu In Lee, Aya Itoh, Lei Li, and Amy E. DeRocher (2000). Plant Physiol. 123:711-724. Data presented in Figure 6-4 were obtained by Dr. Aya Itoh. Data presented in Figure 6-6 were obtained by Dr. Amy DeRocher. Data presented in Figure 6-7 were obtained by Dr. Lei Li. Data presented in Figure 6-8 were obtained by Dr. Gregg Howe. 144 Introduction Fatty acid hydroperoxides produced by lB-lipoxygenases are important intermediates in the oxylipin pathway of fatty acid oxygenation in plants. In one branch of oxylipin metabolism often referred to as the octadecanoid pathway, allene oxide synthase (AOS) commits l3S-hydroperoxy—9(Z),11(E),15(Z)-octadecatrienoic acid (13-HPOT) to the formation of j asmonic acid (J A) and related cyclopenta(e)nones (Creehnan and Mullet, 1997; Figure 6-1). Products of the AOS pathway are essential signals for plant defense against pest attack (Staswick and Lehman, 1999), mechanical responses (Weiler et al., 1993), and some developmental processes (McConn and Browse, 1996). An alternative pathway for l3-HPOT metabolism is initiated by fatty acid hydroperoxide lyase (HPL; Figure 6-1). Short chain aldehyde products of HPL, together with their corresponding reduced alcohols, are important volatile constituents of the characteristic odor of fi'uits, vegetables, and green leaves (Gardner, 1991; Hatanaka, 1993). C6 aldehydes produced by HPL are also reported to act as phytoalexins against protozoa, bacteria, and fungi (for review, see Blée, 1998), and may be signals for gene regulation (Bate and Rothstein, 1998). The C12 oxo-acid product of HPL is the precursor of the previously identified ‘wound signal’ known as traumatin (Zimmerman and Coudron, 1979). 13— HPOT is metabolized by other plant enzymes including lipoxygenase (Salch et al., 1995), peroxygenase (Blée et al., 1993), and divinyl ether synthase (Grechkin et al., 1995; Hamberg, 1998), and may be subject to degradation by non-specific alkyl hydroperoxide reductases (Baier and Dietz, 1999). AOS and HPL comprise an unusual class of cytochrome (Cyt) P4503 that is specialized for the rearrangement of fatty acid hydroperoxides. Unlike typical P450 145 COOH Linolenic acid 1 Lipoxygenase coon _ \ Allene Oxide Synthase 13(9)::Por Hydroporoxide Lyase (CYP74A) (CYP74B) coon <_X\ :\\_/\ Cfv‘coon + or-Ic’\=/\ o - CI-IO 12-oxo-(Z)-9 dodecenoic acid (3Z)-hexenal Allene oxide 1 l 1 COOH COOH O 12-oxo-(E)-9 dodecenoic acid (2E)-hexenal 12-oxo-phytodienoic acid (Traumatin) l Jasmonic acid Figure 6-1. The octadecanoid pathway. Cyt P450-dependent metabolism of 13-HPOT. AOS (CYP74A) commits 13-HPOT to the production of J A and related cyclopenta(e) nones. In the absence of allene oxide cyclase (AOC), the epoxide product of AOS undergoes spontaneous hydrolysis to a- and 'y-ketols and racemic l2-OPDA. HPL (CYP74B) cleaves l3-HPOT to produce C6 and C12 products that are further metabolized as shown. 146 monoxygenases, AOS and HPL demonstrate low affinity for carbon monoxide and do not require 02 or NADPH-dependent Cyt P450 reductase for their activity (Song and Brash, 1991; Shibata et al., 1995a, 1995b). Identification of cDNA sequences encoding AOS and HPL has provided additional insight into the relationship between these two enzymes, and their divergence from classical P4508 (Song et al., 1993; Pan et al., 1995; Laudert et al., 1996; Matsui et al., 1996; Bate et al., 1998). Based on the amino acid sequence identity between AOS and HPL (approximately 38%), the two enzymes are classified as subfamilies CYP74A and CYP74B, respectively, within the CYP74 family of P4508 (Nelson, 1999). The importance of oxylipins as signals for plant stress responses has prompted interest in understanding the mechanisms by which their synthesis is regulated (Creelman and Mullet, 1997; Farmer et al., 1998). JA accumulation, for example, is stimulated by mechanical wounding and herbivory(Cree1man et al., 1992; Blechert et al., 1995; Conconi et al., 1996), pathogen attack (Penninckx et al., 1996), treatment with elicitors (Gundlach et al., 1992; Doares et al., 1995), and water or nutrient deprivation (Creehnan and Mullet, 1995; Lehmann et al., 1995). Similarly, mechanical injury and some plant-pathogen interactions lead to the production of HPL products (Hatanaka et al., 1987; Gardner, 1991; Croft et al., 1993). Formation of AOS- and HPL-derived oxylipins is controlled in large part by the availability of hydroperoxide substrates that are generated from lipase/acyl hydrolase-mediated release of fatty acids from membrane lipids, followed by lipoxygenase-catalyzed conversion to 9- and 13-hydroperoxides (Galliard et al., 1977; Hatanaka, 1993; Mueller et al., 1993; Narvaez-Vasquez et al., 1999). Nonenzymatic lipid peroxidation, such as that associated with the initial stages 147 of plant-pest interactions, may also contribute to the pool of hydroperoxides available to AOS and HPL (Gardner, 1989; Harnmond-Kosack and Jones, 1996). In addition to substrate availability, fatty acid hydroperoxide metabolism may also be influenced by the spatial and temporal expression of enzymes that utilize these substrates. For example, the localization of both AOS and HPL to the chloroplast (Vick and Zimmerman, 1987; Song et al., 1993; Blée and Joyard, 1996; Laudert et al., 1996; Bate et al., 1998; Froehlich et al., 1999) indicates that these enzymes utilize a common pool of hydroperoxide substrates. Recent studies indicate that AOS expression is positively regulated by wounding, as well as by terminal products of the AOS pathway (Laudert and Weiler, 1998). These results, together with transgenic studies showing that AOS is a rate-limiting step in J A biosynthesis (Harms et al., 1995), indicate that up- regulation of AOS activity during the wound response may provide a mechanism to amplify the octadecanoid signaling pathway. On the other hand, others have shown that exogenous methyl J A stimulates oxylipin metabolism through the HPL pathway, and thus may shift oxylipin metabolism away from J A biosynthesis (Avdiushko et al., 1995; Kohlmann et al., 1999). The aim of the present work was to gain an understanding of the molecular basis of Cyt P450-dependent metabolism of fatty acid hydroperoxides in tomato (Lycopersicon esculentum). Owing to the wealth of knowledge of plant-pest interactions in tomato, this system is likely to provide a good model for assessing the role of oxylipins in plant defense. The importance of HPL-derived volatiles in determining the flavor and aroma of fruits and vegetables provides additional incentive for investigating oxylipin metabolism in tomato (Kazeniac and Hall, 1970; Buttery and Ling, 1993). 148 Toward this goal, we report here the isolation of cDNAs that encode functional members of the CYP74A (AOS) and CYP74B (HPL) subfamilies of P450 enzymes in tomato. The results of expression studies in Escherichia coli indicate a role for AOS and HPL in the commitment of l3-HPOT to the J A and C6 aldehyde/traumatin pathways, respectively. We also report findings relevant to the developmental and defense-related expression of these two genes in planta. The significance of these results for understanding the regulation of fatty acid hydroperoxide metabolism is discussed. Results cDNA isolation and sequence analysis An AOS-encoding cDNA from Arabidopsis (Laudert et al., 1996) was used to screen for related sequences in a tomato cDNA library. The longest clone obtained (designated as LeA 05') contained a 1,533-bp open reading frame, a 57-bp 5'-untranslated region (UTR), and a 11 l-bp 3'-UTR excluding the poly-(A) tail. The open reading frame was predicted to encode a 510-amino acid protein having a calculated molecular mass of 57,202 D. The presence of an in-frame stop codon (UAA) 30 nucleotides upstream of the putative AUG start codon indicated that LeA 0S contained the full-length coding sequence. The deduced amino acid sequence of LeAOS was approximately 61% identical to AOS from flax (Song et al., 1993), guayule (Pan et al., 1995), and Arabidopsis (Laudert et al., 1996) (Table 6-1). Thus, LeAOS is classified as a new member of the CYP74A subfamily of Cyt P450s. The N-terminal region of LeA US 149 Figure 6-2. Comparison of cDNA-deduced protein sequences of plant AOS and HPL genes. LeAOS and LeHPL sequences were aligned, using the ClustalW 1.7 program available at http://mbcr.bcm.tmc.edu/searchlauncher. AOS sequences were from flax (LuAOS; Song et al., 1993; accession no. U00428), guayule (PaAOS; Pan et al., 1995; accession no. X78166), and Arabidopsis (AtAOS; Laudert et al., 1996; accession no. Y12636). HPL sequences were from bell pepper (CaHPL; Matsui et al., 1996; accession no. U51674) and Arabidopsis (AtHPL; Bate et al., 1998; accession no. AF087932). Black boxes indicate amino acid residues that are conserved between all seven CYP74 members. Subfamily—specific substitutions are indicated with an asterisk. The three subfamily-specific motifs discussed in the text are underlined by the black bars. The " symbol denotes the T -* (I/V) change within the I helix that is a hallmark of CYP74 enzymes. The conserved Cys within the heme-binding domain is marked by a # symbol. The boxed residue (Pro—43) at the N terminus of LeAOS denotes the site where the His-tag was added in the pQE-AOS expression construct. 150 LuAOS AtAOS PaAOS LeAOS LeHPL CaHPL AtHPL LuAOS AtAOS PaAOS LeAOS LeHPL CaHPL AtHPL LuAOS AtAOS PaAOS LeAOS LeHPL CaHPL AtHPL LuAOS AtAOS PaAOS LeAOS LeHPL CaHPL AtHPL LuAOS AtAOS PaAOS LeAOS LeHPL CaHPL AtHPL LuAOS AtAOS PaAOS LeAOS LeHPL CaHPL AtHPL LuAOS AtAOS PaAOS LeAOS LeHPL CaHPL AtHPL LuAOS AtAOS PaAOS LeAOS LeHPL CaHPL AtHPL MASSALNNLVAVNPNTLSPSPKSTPLPNTFSNLRRVSAFRPIKASLFGDSPIKIPGITSQPPPSSDETTL MA ------- SISTPFPISLHPKT--VRSKPLKFRVL--TRPIKASGSETP----DLTVATRTGSKD---L ------------------------------------------------------------ MDPSSK---- MAL----TLSFSLPLP-SLHQK—--IPSKYSTF ------ RPIIVSLS ------- DKSTIEITQEIK---L ------------------------------------------ MNSAP--—----——-LSTPAPVT----L ----------------------------------- M---IPIMSSAP-—--—----—LSTATPIS-—--L ----------------------- MLLRTMAATSPRP---PPSTSLTS----------QQPPSPPSQ---L ~||TFWFQG-PDK 138 120 75 114 84 88 101 208 190 145 184 154 158 171 277 259 214 253 224 228 241 345 327 282 321 293 297 311 414 395 351 389 362 366 378 483 465 420 459 432 436 448 LLFNLIKNRRDYVIPEFSSS LFFLLKSSRNRIFPEFQAT LLFFMLKNSSNRVIPQFETT PFLLSSRRDHVIPEFHET FSQDILKRGSKTWVPTLLKE FSLDILKRSSKTWVPTLVKE FAMETLKRSSKVWLQELRSN * t i Q Q FTDLCEVVEYDLATKGKAAFNDPAEQ" YSELFDSLEKELSLKGKADFGGSSDGT‘E YTELFEGLEAELAKNGKAAFNDVGEQ" YTELFETLDKEMEEKGTVGFNSGSDQ"r LDTMFTTFEADLSKSNTASLLPALQKF * LDTLFGTFESDLSKSKSASLLPALQKF L LNIFWGTIESEISKNGAASYIFPLQRCI* SRAFFGVKP-IDTPLGKDAPSLIS ‘FYGTNP-ADTKLKADAPGLIT RAYFNSNP-EETKLGTSAPTLIS ‘RSLFGVNP-VETKLGTDGPALIG LP-KEVEEAT LP-RVIEEPLI LP-WFLQEPL LP-KFLDDVL YFQSVATPVMEQAE-KLGVPCl. FFYTNSANLFIEAE-KLGISCDHX QFVKNEAKEVLSRAQTEFQLTEQ‘L IKFVKSEAKEVLTRAQTDFQLTE! LKWIGVA-GENLHTQ LKSIAKA—GVEIHTR ;_‘_ a e a a e .,;ge5: ‘ -- ; ~ ' SSLGASITLTSLKRSTF fifyfiégfih L 1 ' ' t “ SPLGSSVNFSSLRKASF NHQCJQKU ' SPLGAAVTLTFLKRASI YTI‘A KL “ SALGSSITITSLKKA-- -------- SLTSVKKAS- -------- SLTSVKKAC- -------- SIKAVVKAK- -‘-“-- . “WTITGDS 151 displayed features of a typical chloroplast targeting peptide including an enrichment of hydroxylated amino acids. Conclusive evidence that LeA OS is localized to the chloroplast was recently obtained (F roehlich et al., 1999). These findings indicate that LeAOS is more similar to AOS from flax and Arabidopsis, which also reside in the chloroplast (Song et al., 1993; Harms et al., 1995; Laudert et al., 1996), than it is to the cytosolic AOS from guayule (Pan et al., 1995). Values on the upper right diagonal of the matrix indicate the percentage of amino acid identity between different family members. Values on the lower left of the matrix indicate the percentage of nucleotide identity. Percentage identity within the open reading frame of each pair of sequences was calculated using DNA Star software (Clustal method). AOS sequences were from flax (Linum usitatissimum), guayule (Parthenium argentatum), Arabidopsis, and tomato. HPL sequences were from pepper, Arabidopsis, and tomato. GenBank accession nos. for the sequences are given in Figure 6-2. A cDNA encoding HPL from bell pepper (Capsicum annum) was used to screen a tomato cDNA library for related sequences. Among the 15 positive clones identified, the longest cDNA (designated LeHPL) contained a 1,431-bp open reading frame that was predicted to specify a 476-amino acid protein with a molecular mass of 53,542 D. LeHPL contained a 169-bp 5'-UTR, and a 210-bp 3'-UTR excluding poly-(A) residues. The presence of an in-frame stop codon (U GA) 93 nucleotides upstream of the putative initiator AUG codon indicated that LeHPL encoded the entire protein. This was confirmed by DNA sequence analysis of RACE products derived from the 5' end of LeHPL transcripts (‘Materials and Methods’). The deduced amino acid sequence of 152 LuA OS AtAOS PaA OS LeA OS LeHPL CaHPL AtHPL LuAOS 59.9 65.0 61.1 36.1 36.0 34.9 AtAOS 55.7 60.5 62.2 38.2 37.2 36.3 PaAOS 56.2 55.6 60.1 36.5 36.1 34.8 LeAOS 53.1 57.5 55.9 35.6 35.3 35.9 LeHPL 36.3 38.7 38.0 36.0 87.6 55.3 CaHPL 36.3 38.5 38.7 36.5 85.5 53.2 AtHPL 37.8 38.9 36.6 36.3 51.2 49.1 Table 6—1. Percent amino acid and nucleotide identity between different AOSs and HPLs. Values on the upper right diagonal of the matrix indicate the percentage of amino acid identity between different family members. Values on the lower left of the matrix indicate the percentage of nucleotide identity. Percentage identity within the open reading frame of each pair of sequences was calculated using DNA Star software (Clustal method). AOS sequences were from flax (Linum usitatissimum), guayule (Parthenium argentatum), Arabidopsis, and tomato. HPL sequences were from pepper, Arabidopsis, and tomato. GenBank accession nos. for the sequences are given in Figure 6- 2. 153 LeHPL was 88% and 55% identical to the published sequence of HPL from bell pepper (Matsui et al., 1996) and Arabidopsis (Bate et al., 1998), respectively (Table 6-1). This establishes LeHPL as a new member of the CYP74B subfamily of Cyt P4508. Unlike HPL from Arabidopsis (Bate et al., 1998), LeHPL does not appear to contain a typical chloroplast targeting sequence at the N terminus of the protein (Figure 6-2). Comparisons between the primary structures of the seven known CYP74 members (four AOSs and three HPLs) showed 182 positions (38%) that were conserved in all members of both subfamilies (Figure 6-2). Many conserved residues were clustered at the N- and C-termini, and may be important for functions common to HPL and AOS (e. g. heme or substrate binding). We also noted subfamily-specific amino acid differences that might play a role in distinguishing AOS function fiom that of HPL. Specifically, there were 39 positions at which an amino acid was invariant among all AOSs, and was substituted to a different residue in all HPLS (Figure 6-2). One HPL- specific motif was PPxF P, which represents a variation of the N-terminal PPGP tetrapeptide that is important for stability and catalysis in many P4508 (Szczesna- Skorpa et al., 1993). A hallmark of all CYP74 enzymes, including LeAOS and LeHPL, is a T—)(V/I) substitution within the I helix that, in most P4508, participates in 0; binding (Song et al., 1993). Sequences surrounding this site show a subfamily-specific character, with the AOS and HPL consensus sequences being KI(L/F)F and (S/T)IFL, respectively. Several other subfamily-specific signatures were located near the C- terrninal heme-binding domain. The most striking of these was an eight-amino acid insertion in AOS sequences relative to HPL sequences, at the extreme C-terrninal end of the protein (Figure 6-2). 154 BEXBg BEXBg BEXBg ‘u h\ll" ' ‘ 12.2 - X 11 . ‘. 1 M1 . 2‘ ‘ w“ “(in 7.1 - l 5.1 — G ”I“ , 3.1 _ * v“ a " ‘ LeAOS LeHPL LeHPL 5’UTR Figure 6-3. Southern-blot analysis of LeAOS and LeHPL. Genomic DNA from tomato was digested with restriction enzymes BamHI (B), EcoRV (E), XbaI (X), or BglII (Bg). DNA blots were hybridized to labeled probes derived from the open reading frame of LeAOS (left), LeHPL (middle), or the LeHPL 5'-UTR (right). Blots were hybridized in 5x SSPE at 65°C and washed in 0.5)( SSPE at the same temperature, as described in "Materials and Methods." Molecular mass standards (in kb) are indicated on the left. 155 Genomic DNA-blot analysis using the LeA 0S cDNA as a probe showed a simple hybridization pattern for each of the restriction enzymes tested (Figure 6-3). This result indicates that LeA OS is derived from a single copy gene. However, detection of additional hybridizing bands (data not shown) under conditions of reduced stringency leaves open the possibility of related sequences in the genome. The results of hybridization analysis using a LeHPL cDNA probe showed a more complex pattern of weakly and strongly hybridizing bands (Figure 6-3). Use of a probe derived from the 5'- UTR of LeHPL reduced the complexity of the hybridization pattern as expected, but nevertheless still detected two or more bands for each restriction digest tested (Figure 6- 3, right panel). These results indicate that LeHPL is one member of a family of highly related genes that may be clustered as tandem repeats in one region of the genome. Functional expression of LeAOS and LeHPL in E. coli To confirm that LeA 0S and LeHPL encode the expected P450 enzymes, the cDNAs were subcloned into the pQE-30 expression vector to yield pQE-AOS and pQE-HPL, respectively, and transformed into an appropriate E. coli host. Bacterial cultures induced to express the constructs accumulated high levels of the recombinant proteins as determined by SDS-PAGE of bacterial lysates (data not shown). Crude lysates from cells expressing either pQE-AOS or pQE-HPL efficiently degraded l3-HPOT (Figure 6-4A) but did not metabolize 9-hydroperoxide derivatives of linolenic or linoleic acid (data not shown). Recombinant LeAOS and LeHPL metabolized the C20 hydroperoxide 15S-hydroperoxy-11(Z),l3(E),l7(Z)-eicosatrienoic acid at a rate comparable to that observed with 13-HPOT (Figure 6-4A), indicating that both enzymes can accommodate 156 > m — 13-HPOD — 134-WOO 5 15 « m 13-HP0T .5 15 . n— 1341901 3 _ 15”,“ {g- — 15-HPET o a 12‘ '5. 12‘ 0’ a: g 9 . g 9 . g .5 S 6 . g 6 . 3 3 . g 3 . 0 0 _ Figure 6-4. Activity of LeAOS and LeHPL expressed in E. coli. Total lysates of E. coli cells expressing pQE-AOS, pQE-HPL, or the empty vector (pQE- 30) were tested for their ability to metabolize C13 (l3-HPOD and l3-HPOT) and C20 (15-HPET) fatty acid hydroperoxides. Activity was measured either directly as a decrease in absorbance of the substrate at A 234 (A) or indirectly as the production of aldehydes using a NADH-coupled assay (B). Error bars represent the mean and SD of activity determined from three enzyme preparations of each culture. 157 a range of fatty acid hydroperoxide substrates. However, the two enzymes differed in their ability to metabolize 13S-hydroperoxy-9(Z), ll(E)-octadecadienoic acid (13- HPOD), a common C13 hydroperoxide derived from linoleic acid. Whereas LeAOS utilized l3-HPOD at about one-half the rate observed with 13-HPOT, the rate of breakdown of 13-HPOD by LeHPL was less than 5% of that observed for 13-HPOT. Similar results were obtained using a coupled enzyme assay (Vick, 1991) to measure aldehyde production in the in vitro reactions. As expected from the known products of AOS and HPL (Figure 6-1), aldehyde production was associated with reactions catalyzed by LeHPL but not by LeAOS (Figure 6-4B). This assay also confirmed the strong preference of LeHPL for l3-HPOT over 13-HPOD. Gas chromatography-mass spectometry (GC-MS) was used to identify the trimethylsilyl (TMS) derivatives of metabolites produced upon incubation of 13-HPOT with lysates from bacteria that expressed either pQE-AOS, pQE-HPL, or the pQE-30 mock control. In the case of pQE-AOS, three prominent peaks (A, B, and C) that were not present among the products of the mock reaction were observed. The relative abundance of these compounds, as estimated by integration of the GC peak areas, was 22% (peak A), 100% (peak B), and 13% (peak C) (values normalized to peak B). The retention time (11 min 9 8), molecular ion ([M]+' at m/z 364), and fragmentation pattern of peak A were identical to that of an authentic 12-oxo-phytodienoic (12-OPDA) standard (Cayman Chemical, Ann Arbor, MI). Peak B eluted at 12 min 37 8 and gave the following mass spectrum as m/z (% relative intensity, ion structure): 526 (18%, [M]+'), 511 (13%, [M - CH3]+'), 457 (100%, [M - C5H9]"), 367 (8%), 221 (5%), 179 (4%), 147 (12%), and 73 (28%, TMS+). This fragmentation pattern was consistent with 158 identification of the compound as the tri-TMS derivative of the a-ketol compound 12- oxo-l3-hydroxy-9(Z), 15(Z)-octadecadienoic acid (enolization of the 12-oxo group provided an additional hydroxyl for derivatization). The major fragment at m/z = 457 ([M - C5H9]+') indicated that this compound represented the a-ketol rather than the y- ketol (12-oxo-9-hydroxy-10[E],15(Z)-octadecadienoic acid). Peak C eluted with a retention time of 13 min 10 s and produced a mass spectrum identical to that of peak B, indicating probable double bond isomerization during derivatization or GC analysis. a- Ketol and 12-OPDA, together with minor amounts of 'y-ketol, are known to arise by spontaneous hydrolysis of the unstable epoxide product of AOS (Song and Brash, 1991) (Figure 6-1). The products of the pQE-HPL-catalyzed reaction were analyzed by GC- MS as the oxime, TMS derivatives. The major product eluted with a retention time of 8 min 59 s. A molecular ion, m/z 371 [M‘], was observed for the di-TMS derivative of this product, and the fragmentation pattern was identical to that of an authentic 12-oxo- trans-lO-dodecenoic acid standard, the expected product of HPL (Figure 6-1). These results confirmed the identity of LeAOS and LeHPL as fimctional members of the CYP74A and CYP74B subfamilies of P450 enzymes, respectively. Developmental expression of LeAOS and LeHPL RNA-blot analysis was used to investigate the distribution of LeHPL and LeA 0S mRNA in different organs of tomato (Figure 6-5). LeHPL transcripts accumulated to high levels in developing flowers, and decreased during flower maturation. LeHPL mRNA levels were also relatively high in leaf tissue, with greater expression detected in younger leaves compared to older leaves from the same plant (see Figure 6-7). Very 159 RSLBUFOFIFGFRF LeA OS ”4 u ' MP6 119-In u. . .. LeHPL 1 . .9 W LeHPL eIF4A _ " . Figure 6-5. Expression of LeAOS and LeHPL genes in different organs of tomato. Total RNA was extracted from roots (R), stems (S), leaves (L), developing flower buds (B), mature unopened flowers (UF), mature opened flowers (OF), small (<0.5 cm) immature green fruit (IF), mature green fruit (GF), or mature red fruit (RF). Ten- microgram samples of RNA were subjected to RNA-blot analysis. Specific transcripts were detected by hybridization of blots to probes corresponding to full-length LeAOS, full-length LeHPL, the 5'-UTR of LeHPL, or an eIF4A probe used as a loading control. Also shown is a photograph of an ethidium bromide-stained gel of the RNA used for the experiment (EtBr). 160 low levels of LeHPL mRNA were detected in stems and immature green fi'uit, whereas roots and mature green and red fruit lacked detectable transcripts. Hybridization probes derived from either the full-length LeHPL cDNA or the LeHPL 5'-UTR showed a similar organ-specific expression pattern (Figure 6-5). This result showed that LeHPL transcripts detected by RNA-blot analysis are derived from a single LeHPL gene, or highly related LeHPL genes that have a similar developmental expression pattern. LeA 0S mRNA was broadly distributed among all organs examined (Figure 6-5). LeA 0S transcript levels were relatively low in fruit, and appeared to decrease during fruit development. Polyclonal antibodies raised against recombinant LeAOS, but not preimmune serum fi'om the same rabbit, reacted with a polypeptide in the membrane fraction of extracts prepared from different organs (Figure 6-6). The estimated mass of the cross-reacting polypeptide as judged by SDS-PAGE was 55 kD, which is consistent with that expected for the LeA 0S gene product. Furthermore, the distribution of this polypeptide in different organs correlated with the distribution of LeA 0S mRN A. Taken together, these results indicate that LeA OS is expressed in all tomato organs with the possible exception of ripe fruit. A second cross-reacting polypeptide of slightly lower Mr was often observed in immunoblot experiments, particularly in extracts derived from flowers (Figure 6-6). This band could represent a polypeptide that shares common epitopes with LeA OS, or a post-translationally modified form of LeA OS. 161 Wound-inducible expression of LeAOS is mediated by a Defl-independent signaling pathway The importance of oxylipin metabolism for wound-inducible defense gene expression in tomato prompted us to examine the effect of wounding on LeAOS and LeHPL gene expression. Damage inflicted to tomato leaves by Manduca sexta larvae resulted in a modest (approximately 2-fold) increase in LeHPL mRNA accumulation (Figure 6-7). Wound-induced accumulation of LeHPL mRN A was more apparent in the lower damaged leaf than it was in the younger undamaged leaf. This is likely to reflect the higher constitutive expression of LeHPL in younger leaves. Wound-induced accumulation of LeA 0S transcripts was much more apparent, and thus became the focus of additional experiments. The time course and amplitude of LeA 0S expression differed in several ways from that of the well-characterized proteinase inhibitor II (Inh-II) gene (Figure 6-7). Whereas the maximum level of induction of LeA OS in local and systemic leaves was approximately 9- and 5-fold, respectively, Inh-II mRNA levels in these tissues increased by at least 60-fold. Wound-inducible accumulation of LeA 0S mRNA was also more transient than that of Inh-II. These results indicated that the mechanism controlling wound-inducible expression of LeA 0S might be different from that regulating Inh-II gene expression. 162 BRSPCL BRSPCL 81.0 - 51.2 - 33.6 - LeAOS Immune pre-lmmune Figure 6-6. Accumulation of LeAOS protein in different organs of tomato. Fifteen-microgram samples of membrane protein prepared from young flower buds (B), roots (R), stems (S), petioles (P), cotyledons (C), and leaves (L) were separated by SDS-PAGE. Protein was transferred to Immobilon-P membranes and probed with either antiserum raised against LeAOS (left) or an equivalent amount of preimmune serum (right). The numbers on the left of the figure indicate the position of Mr standards. 163 Local Systemic 000.5124 81224 C00.5124 81224 Inh-II '0!!! LeAOS o g - ..- LeHPL .. f a etc-C...- eIF4A --D--.--- vac-D---- Figure 6-7. Accumulation of LeAOS, LeHPL, and Inh-II mRNAs in tomato plants in response to herbivory. Tobacco hornworm larvae (third instar) were placed onto the lower leaf of 3-week-old cv Micro-Tom plants and allowed to feed for 5 to 10 min. During this period, approximately 5% to 10% of the area of the attacked leaf was consumed by the larvae. Leaf tissue was harvested for extraction of total RNA immediately after removal of the larvae (0 point) or at the times indicated (in hr). RNA was prepared separately from the lower damaged leaf (Local response) and from the third leaf (counted from the base of the plant) (Systemic response). RNA was also prepared from a set of control plants that received no damage (C). Duplicate RNA blots containing 5 pg of RNA per sample were hybridized to cDNA probes for proteinase inhibitor 11 (Inh-II), LeAOS, LeHPL, and eIF4A as a loading control. 164 wild-type def1 Hours: 012 4 81224 012 4 81224 Inh-Il a Q Q. . 30’ n a... AOS "fl““fi‘ «wise-911,11 v ‘ " r . LE .. s?" “‘6'?“ fl “ 11.. ' 1 .. 1‘ '1 m ”a. ..W eIF4A fi'fiflflfifiwflflflfl Ii i. M" ..- Figure 6-8. Analysis of wound-induced gene expression in wild- type and def] mutant plants. F ifteen-day-old wild-type (cv Castlemart) and def? mutant seedlings were mechanically wounded at the distal end of the terminal leaflet of the lower leaf. Undamaged tissue on the same leaflet was harvested for RNA extraction at the indicated times after wounding. RNA blots were hybridized to cDNA probes for proteinase inhibitor [1 (Inh-Il), cathepsin D inhibitor (CD1), T ornLoxD (LoxD), LeA OS (A US), LE RNase (LE), and eIF 4A as a loading control. 165 To further test this idea, we examined LeA OS expression in the tomato defenseless] (defI) mutant that is deficient in the octadecanoid-based signaling pathway that mediates the expression of Inh-II and other defense-related genes (Howe et al., 1996). Previous characterization of def] showed that it is deficient in J A accumulation in response to wounding and other elicitors (Howe et al., 1996). Genetic mapping studies have shown that the def] phenotype does not result fiom a defect in the LeA OS gene (A. Itoh and G.A. Howe, unpublished data). Moreover, direct measurements of 12- OPDA levels in def] and wild-type plants indicated that the mutant has both AOS and allene oxide cyclase activity (Stelmach, E. Weiler, G.A. Howe, unpublished data). Taken together, the available evidence indicates that the Def] gene product plays a role in the regulation of a late step in the biosynthesis of JA, or in the further metabolism of J A (e. g. transport or stability). A dramatic aspect of the def] phenotype is the lack of wound-induced accumulation of defensive proteinase inhibitor genes such as Inh-II and cathepsin D inhibitor (CD1) (Figure 6-8). In contrast to this, the pattern of wound- inducible LeA OS mRNA accumulation in defl plants was comparable to that in wild- type plants. This effect was not specific for LeAOS, as other transcripts, including those for lipoxygenase (LoxD) (Heitz et al., 1997) and a senescence-induced RNase (LE) (Lers et al., 1998), were also induced by wounding in both mutant and wild-type plants. These results demonstrate that wound-induced expression of LeA OS, LoxD, and LE mRNA is Def] independent. 166 Discussion Fatty acid hydroperoxides derived from lipoxygenase are precursors for an array of oxylipins that function in diverse aspects of plant growth and development. In this paper we report the isolation and characterization of tomato cDNAs encoding AOS and HPL, two similar P450 enzymes that commit 13-HPOT to different branches of oxylipin metabolism. The LeA OS and LeHPL proteins are 36% identical at the amino acid level, and are classified as members of the CYP74A and CYP74B subfamilies of Cyt P4508, respectively. Identification of AOS and HPL genes in tomato brings the total number of reported AOS and HPL sequences to seven. In comparing the primary structures of these, we noted 39 positions at which all AOS members contained one common amino acid and all HPLS contained a different residue. The significance of these subfamily- specific substitutions will become more or less apparent as additional CYP74 genes are identified. Subfamily-specific substitutions might reflect differences in the catalytic properties or substrate specificity of the two classes of enzymes. The facile expression of recombinant forms of AOS and HPL in E. coli should facilitate studies aimed at understanding the structure-function relationship that defines the catalytic identity of these unusual P4508. A major difference between the predicted amino acid sequence of LeA OS and LeHPL was the presence of a typical chloroplast targeting sequence at the N-terrninus of LeAOS (Figure 6-2). Previous studies indicate that chloroplast targeting peptides are present on AOS from flax and Arabidopsis (Song et al., 1993; Harms et al., 1995; Laudert et al., 1996), as well as HPL from Arabidopsis (Bate et al., 1998). A plastid location for AOS and HPL is consistent with biochemical studies demonstrating that 167 AOS and HPL activity is associated with chloroplasts (Vick and Zimmerman, 1987; Gardner et al., 1991; Blée and J oyard, 1996; Zhuang et al., 1996). Recently, we have shown that LeAOS is imported into chloroplasts where it specifically targets to the inner membrane of the chloroplast envelope (Froehlich et al., 1999). This finding indicates that LeAOS obtains its hydroperoxide substrates from one or both of the plastid-localized lipoxygenases (TomLoxC and TomLoxD) that have been described in tomato (Heitz et al., 1997). In contrast to LeAOS, the deduced N-terminus of LeHPL lacked a typical transit peptide. That the N—terminal sequence of LeHPL is very similar to that of bell pepper HPL (Figure 6-2) indicates that these proteins share a similar subcellular location. Given the preponderance of evidence indicating that HPL activity is associated with the chloroplast, additional experiments aimed at determining the subcellular location of LeHPL are clearly warranted. Characterization of LeAOS Expression of LeA OS in E. coli showed that the open reading frame encodes an authentic CYP74A enzyme (LeAOS) that metabolizes 13- but not 9-hydroperoxides of linoleic and linolenic acids. LeAOS expression was detected in all organs of the plant except mature red fruit. Similar expression patterns were observed for AOS in Arabidopsis and flax (Harms et al., 1998; Laudert and Weiler, 1998). Accumulation of LeAOS mRNA and protein in flowers is consistent with previous studies in Arabidopsis showing that AOS promoter activity is high in flowers, particularly in pollen sacs and pollen grains (Kubigsteltig et al., 1999). It is presently not known whether AOS-derived products are required for pollen development in tomato, as they are in Arabidopsis 168 (McConn and Browse, 1996). Detection of LeA OS mRNA and protein in leaves supports previous reports of AOS activity (Caldelari and Farmer, 1997) and inducible J A synthesis in tomato leaves (Pena-Cortes et al., 1993; Doares et al., 1995; Conconi et al., 1996). The relatively high accumulation of LeAOS mRNA in stems (Figure 6-5) compared to that in leaves indicates that LeAOS is expressed in vascular bundles, as was reported to be the case in wounded leaves of Arabidopsis (Kubigsteltig et al., 1999). LeA OS mRN A and protein were also detected in tomato roots. Given that root development appears normal in J A-deficient mutants of Arabidopsis (McConn and Browse, 1996), this result indicates that AOS-derived oxylipins serve a non- developmental role in roots. LeA OS mRNA expression in green fi'uit, while being relatively low, is consistent with previous studies showing increased levels of cis-JA during the early stages of tomato fruit ripening (Fan et al., 1998). However, our results do not exclude the possibility that J A synthesis in tomato fruit involves a different AOS-encoding gene that is undetectable by high stringency nucleic acid hybridization. Characterization of LeHPL Expression of LeHPL in E. coli confirmed that this cDNA encodes a functional member of the CYP74B subfamily of enzymes. LeHPL was similar to LeAOS in its ability to use 13- but not 9-hydroperoxides of C18 fatty acids. However, LeHPL was clearly distinguishable from LeAOS in its strong preference for l3-HPOT over 13-HPOD. This feature is shared by HPL isolated from other sources, including bell pepper (Shibata et al., 1995b), tea leaves (Matsui et al., 1991), and Arabidopsis (Bate et al., 1998). F auconnier et al. (1997) reported the purification from tomato leaves of an HPL that, 169 like recombinant LeHPL, did not utilize 9-hydroperoxides and showed a strong preference for l3-HPOT over l3-HPOD. However, the purified enzyme displayed a molecular mass (73 kD) much greater than that predicted for LeHPL (53.5 kD). Additional experiments are needed to clarify the relationship between LeHPL and this purified form of tomato leaf HPL. LeHPL mRNA was most abundant in developing flowers. This finding indicates that HPL-derived products might have a role in the production of floral scent. Relatively high levels of LeHPL mRNA were also detected in leaves. The overlapping expression pattern of LeA OS and LeHPL in leaves is consistent with the idea that these two enzymes compete for the same pool of substrate (Avdiushko et al., 1995; Blée and Joyard, 1996; Blée, 1998). However, additional studies aimed at determining the subcellular and tissue-specific location of both enzymes are needed to substantiate this hypothesis. The paucity of LeHPL mRNA accumulation in mature green and red fruit was surprising since cis-3-hexenal, derived from the action of HPL on l3-HPOT, is a prominent volatile component of the aroma and flavor of tomato fruit (Buttery and Ling, 1993). A possible explanation for these results is that LeHPL plays only a minor, if any, role in the production of C6 volatiles during tomato fruit ripening. Additional insight into the contribution of LeHPL to fruit aroma and flavor might be gained by altering LeHPL expression in transgenic plants. 170 Wound-inducible expression of LeAOS is mediated by a Defl-independent signaling pathway Damage inflicted to tomato leaves by hornworm larvae triggered the accumulation of LeA OS mRNA both in the damaged leaf and in the upper undamaged leaves of the plant (Figure 6-7). Similar changes in LeA OS expression occurred in plants subjected to mechanical wounding (Figure 6-8; data not shown). The transient accumulation of LeA OS transcripts in these experiments was a consequence of the limited damage inflicted to the plant (e. g. 5-10 min of feeding by the insect). It is likely that sustained feeding by herbivores, such as that occurring in natural and agricultural ecosystems, would result in much greater increases in LeAOS expression. Increased expression of AOS, and possibly other octadecanoid pathway enzymes, could serve to amplify the J A signaling cascade a8 a means of enhancing the induced resistance response. Wound- inducible increases in AOS mRNA, protein, and activity have been documented in Arabidopsis (Laudert and Weiler, 1998; Kubigsteltig et al., 1999) and flax (Harms et al., 1998). LeHPL transcript levels also appeared to increase in response to insect attack (Figure 6-7) and mechanical wounding (data not shown), but only by about 2- fold relative to unwounded controls. Additional studies are needed to determine whether the expression of LeHPL and LeA OS in tomato leaves is affected by other defense signals, or by interactions with pathogens. The wound-inducible expression pattern of LeA OS differed in several respects from that of proteinase inhibitor genes. First, the time course of LeA OS mRNA accumulation was more transient than that of the inhibitor genes. Second, the amplitude of LeA OS mRNA accumulation in both damaged and systemic leaves was 10- to 20- 171 fold lower than that of Inh-II mRNAs. Finally, wound-inducible expression of LeAOS was observed in the def] mutant, while that of the Inh-II and CD] genes was not. Two additional genes, LoxD and LE, were also wound inducible in the def] background. These results demonstrate the existence of two classes of genes whose requirements for wound induction in tomato can be defined as being either Def] -dependent or Def] - independent. Given the involvement of Def] in wound-inducible J A accumulation (Howe et al., 1996), we indicate that endogenous J A is a signal for Def] -dependent, but not Def] -independent, wound responses. It is noteworthy that some genes exhibiting Def] -independent expression, such as LeA OS and LoxD, are inducible by exogenous J A (Heitz et al., 1997; G.I. Lee and G.A. Howe, unpublished data). This raises the possibility that genes whose expression is altered by exogenous J A might not be under the control of the J A signaling pathway as it operates in planta. Alternatively, LeAOS and LoxD may be controlled by both J A-dependent and -independent wound-response pathways. This interpretation is consistent with the observation that wound-induced accumulation of LeA OS and LoxD mRNA in def] plants was slightly less than that in wild-type plants (Figure 6-8). Regulation of AOS and LOX activities by both J A- dependent and -independent signaling pathways might allow amplification or increased sensitization of wound responsiveness under different conditions. This notion is consistent with other studies showing that wound and defense responses in tomato involve multiple signaling pathways (Chao et al., 1999; Ryan, 2000). In Arabidopsis, J A-dependent and -independent wound responses have been shown to be differentially regulated by Ca2+/calmodulin, as well as by reversible protein phosphorylation events (Titarenko et al., 1997; Leon et al., 1998; Rojo et al., 1998). Thorough analysis of 172 mutants such as def] or those that are suppressed in the action of systemin (Howe and Ryan, 1999) may provide further insight into the role of oxylipins in wound and defense signaling pathways in tomato. Materials and Methods Plant Material and Growth Conditions cv Micro-Tom seed (Lycopersicon esculentum cv Micro-Tom) was obtained from Dr. Avraham Levy (Weizmann Institute, Rehovot, Israel). Seed for the tomato def] mutant was collected from a defl/def] homozygous line that had been back-crossed four times to L. esculentum cv Castlemart, the wild-type parent of def]. Seedlings were grown in Jiffy peat pots (Hummert International, St. Louis) in a growth chamber maintained under 17 hr of light (300 .115 m2 s") at 28°C and 7 hr ofdark at 18°C. Flowers and fruit were collected from plants maintained in a greenhouse. cDNA Cloning and Sequencing A 1.1-kb A7101 fragment derived from the coding region of an Arabidopsis AOS cDNA (Laudert et al., 1996) was labeled with [or-32P]dCTP and used to screen a tomato cDNA library constructed from tomato plants that overexpress the prosystemin gene, as described by Heitz et al. (1997). Duplicate filters were hybridized at 42°C in a solution containing 5X SSPE, 50% (v/v) forrnamide, 5x Denhardt's reagent, 0.5% (w/v) SDS, and 50 11 g/mL denatured salmon sperm DNA. Filters were washed at 42°C in a solution containing 5X SSPE and 0.5% (w/v) SDS, followed by an additional wash at 65°C. Four 173 positive clones were obtained among approximately 4 x 105 plaque-forming units screened. Following excision of the cDNA from the phagrnid, DNA sequence analysis showed that all clones were identical with the exception of minor differences in the length of the 5' end. The longest cDNA insert, designated LeAOS, was sequenced completely on both strands using a primer walking approach. The sequence of the LeAOS cDNA was deposited to GenBank (accession no. AF 23037 1). A bell pepper (Capsicum annum) cDNA encoding HPL was isolated using a reverse-transcription PCR (RT-PCR) kit (Life Technologies/Gibco-BRL, Cleveland) and total RNA isolated from bell pepper as a template. Two gene-specific primers were designed from the cDNA sequence reported by Matsui et al. (1996) (GenBank accession no. U51674). The sequence of the forward and reverse primers used for RT-PCR was 5'-(GTG-GAT- CCA-TTC-ATA-AAA-CAA-CAA-CTA-C)-3' and 5'-(GTG-AAT-TCA-GCA-ACC- TTT-AGT-ACC-TAC-C)-3', respectively. An amplified 1,477-bp product was subcloned into pBluescript SK(-) (Stratagene, La J olla, CA) and sequenced to confirm its identity to the published sequence (Matsui et al., 1996). This clone was used to screen a tomato leaf cDNA library as described above but with the following modifications. Filters were washed at 65°C in a solution containing 5X SSPE and 0.5% (w/v) SDS. Fifteen positive plaques were identified among 3 X 105 plaque-forming units screened. DNA sequence analysis of eight cDNA inserts showed that all cDNAs were identical except for minor differences in the length of the 5' end and the number of poly- (A) residues at the 3' end. The longest clone, designated LeHPL, was subcloned into smaller fragments and sequenced in its entirety on both strands. The sequence of the LeHPL cDNA was deposited to GenBank (accession no. AF 230372). A RACE 174 procedure (Life Technologies/Gibco-BRL, Cleveland) was used to obtain additional sequence information at the 5' end of LeHPL. First strand cDNA was synthesized from total RNA prepared from either tomato leaves or flowers as a template. The sequence of the gene-specific primer used for this reaction was 5'-(ACT-TCC-TTG-GCT-TCA- TTT-T)-3'. PCR amplification of the dC-tailed cDNA was performed using the manufacturer's abridged anchor primer and a gene-specific primer having the sequence 5'-(AGC-GCC-GAG-GAT-AGT-GAG-GGA-GAA)-3'. PCR products were re- amplified using the manufacturer's abridged universal amplification primer and a nested gene-specific primer having the sequence: 5'-(TGG-AGT-GCA-GGA-AGA-AGA- GAA-G)-3'. Amplified PCR products were subcloned into pBluescript SK(-). DNA sequencing of 5' RACE products derived from both leaf and flower mRNA confirmed the structure of the 5'-UTR of LeHPL, including the presence of the in-frame stop codon upstream of the initiator Met. Expression of LeA 0S and LeHPL in Escherichia coli A PCR strategy was employed to subclone the LeA OS cDNA into the E. coli expression vector pQE-30 (Qiagen USA, Valencia, CA). Forward and reverse primers were designed to contain BglII and Pstl restriction sites, respectively. The sequence of the forward primer was 5'-(GCT-AGA-TCT-CCT-ATA-AAA-TTA-TCT-ACC-AGG)-3' and that of the reverse primer 5'-(GTT-CTG—CAG-CCG-ATA-GTG-ACA-GTG-TAG- ACC)-3'. Using the LeA OS cDNA as a template, the PCR-amplified product was cut with Bng and Pstl and cloned into BamHI and Pstl sites of pQE-30. The resulting expression vector was called pQE-AOS. This strategy removed the first 42 amino acids 175 from the N terminus of LeA OS, and added the sequence MRGSHHHHHHGS to Pro 43 of LeA OS (Figure 6-2). A similar strategy was used to construct a vector for expression of LeHPL. Forward and reverse primers were designed to contain BamHI and SstI sites, respectively. The sequence of the forward primer was 5'-(CGG-GAT-CCC-CGA-TAA- TGA-ATT-CTG-CTC)-3' and that of the reverse primer 5'-(GCG-AGC-TCT-CAT- AAG-TCA-GAA-CAG)-3'. PCR products obtained using the LeHPL cDNA as a template were digested with BamHI and SstI, and cloned into BamHI and SacI sites of pQE-30 to give pQE-HPL. This strategy added the sequence MRGSHHI-II-II-IHGSPI to the deduced initiator Met of LeHPL. Expression constructs pQE-AOS and pQE-HPL were transformed into E. coli strain M15. Bacteria grown under standard conditions (37°C in Luria-Bertani medium) and induced with isopropylthio-fl-galactoside produced recombinant protein that was associated with inclusion bodies (data not shown). Induction of cultures using the following procedure significantly enhanced the recovery of active enzyme in the soluble fi'action of lysed cells. Bacterial cultures (50 mL) were grown in Terrific Broth medium at 37°C to logarithmic phase (A600 of 0.6), at which time the culture was induced by the addition of isopropylthio-B-galactoside to a final concentration of 0.5 mM. Cultures were incubated for an additional 8 hr at 26°C with gentle shaking (150 rpm). Bacteria were harvested by centrifugation and resuspended in 5 mL of a solution containing 50 mM potassium phosphate (pH 7.5) and 5% (v/v) glycerol. Following one freeze-thaw cycle, cells were broken by sonication and centrifuged at 10,000g for 20 min. SDS- PAGE analysis of supernatant protein from induced culture extracts showed the 176 presence of the recombinant protein, migrating with the expected molecular mass (data not shown). Enzyme Assays and Preparation of Fatty Acid Hydroperoxides The hydroperoxide-degrading activity of recombinant LeAOS and LeHPL was measured spectrophotometrically using two methods described by Vick (1991). One assay, which does not distinguish between AOS and HPL activity, involved monitoring the decrease in A 2 34 that results from disruption of the conjugated diene bond in the substrate. The second method was specific for HPL and involved an NADH-coupled assay for detection of aldehyde reaction products. The protein content of cell extracts was determined by the Bradford assay. F atty acid hydroperoxide substrates (9- and 13- substituted) were prepared using soybean lipoxygenase (Sigma-Aldrich, St. Louis) or com seed lipoxygenase as described (Vick, 1991). Fatty acids for these reactions were obtained from Nu-Chek-Prep, Inc (Elysian, MN). Substrate specificity results were confirmed using purified hydroperoxides (9-HPOD, 9-HPOT, 13-HPOD, 13-HPOT) purchased fiom Cayman Chemical. Identification of Metabolites Two micromoles of 13-HPOT, dissolved in 30 mL of 50 mM potassium phosphate (pH 7.0), was mixed with 1 mg of soluble protein (enzyme source) obtained from E. coli cells expressing either pQE-AOS, pQE-HPL, or the pQE-30 vector control. The reaction was allowed to proceed for 10 min at room temperature and then stopped by acidification to pH 4.0 with l M citrate. Products were extracted twice with diethyl 177 ether and dried under N2 gas. TMS derivatives of pQE-AOS reaction products were prepared by treatment of the extract with 30 1.1L of BSTFA (bis[TMS] trifluoroacetamide/trimethylchlorosilane) (99:1, v/v) (Supelco, Bellefonte, PA) and 10 11L of pyridine for 1 hr at 80°C. Oxirne TMS derivatives of pQE-HPL products were prepared by first reacting enzyme products with hydroxylarnine hydrochloride at 80°C for 1 hr, followed by treatment with BSTF A and pyridine as described above. One to 2 11L of the derivatized compounds was used for GC-MS analyses, which were carried out on an AX 505H double focusing mass spectrometer (J EOL, Peabody, MA) equipped with a 5890 gas chromatograph (Hewlett-Packard, Palo Alto, CA). GC separations employed a DB-l methyl silicone capillary column (30 m X 0.25 mm id.) (J &W Scientific, Folsom, CA) interfaced directly to the ion source via a heated transfer block. The temperature program was initiated at 50°C and ramped to 225°C at 20°C min". The temperature was then increased to 270°C at 2°C min". The ion source was operated at 70 eV with the scan rate of the instrument set to approximately 1 spectrum 8". Antibody Production and Western-Blot Analysis A 0.5-L culture of E. coli was induced for the expression of pQE-AOS as described above, except that induced cells were grown at 37 °C for 4 hr in Luria-Bertani medium. Cells were harvested by centrifugation and resuspended in 1/ 10 volume of lysis buffer (50 mM sodium-phosphate, pH 8.0, 10 mM imidazole, 300 mM NaCl, and 0.25% [v/v] Emulgen 911 [Kayo Corporation, Tokyo]). Following one freeze-thaw cycle, cells were broken by two passes through a French press calibrated at 17,000 pounds per square inch. Insoluble inclusion bodies containing LeAOS were recovered by centrifugation 178 for 20 min at 8,000g. The pellet was washed twice with 40 mL of lysis buffer and recovered by centrifugation. Washed pellets were solubilized at 4°C in 35 mL of a solution containing 6 M guanidine HCl, 100 mM sodium-phosphate, pH 8.0, and 10 mM Tris (tris [hydroxymethyl]aminomethane) HCl, pH 8.0. The mixture was sonicated for 10 min at 4°C to facilitate solubilization, and then centrifuged at 12,000g for 20 min. Recombinant LeAOS in the supernatant was purified by nickel affinity chromatography as described by the manufacturer (Qiagen). His-tagged LeAOS, which eluted from the nickel column at pH 4.5, was dialyzed twice for 4 hr against 100 volumes of 50 mM Tris, and 0.2% (v/v) Emulgen 911. Approximately 0.5 mg of protein was solubilized in Laemmli sample buffer and further purified by preparative SDS-PAGE (12% [w/v] gel). Acrylarnide gel slices containing His-tagged LeAOS were stained with Coomassie Brilliant Blue and macerated through a syringe as described (Harlowe and Lane, 1988). For the initial immunization, 100 11g of antigen in the mashed gel slice was mixed with F reund's complete adjuvant and injected at multiple subcutaneous and intramuscular sites of a New Zealand white rabbit. Four boosts, each consisting of 50 ug of antigen mixed with Freund's incomplete adjuvant, were administered over the course of a 90-d immunization schedule. Protein extracts for westem—blot analysis were prepared from fresh plant tissue that was extracted with a mortar and pestle at 4°C in a buffer containing 50 mM sodium-phosphate, pH 7.0. The buffer to tissue ratio (w/w) was about 2: 1. Crude cellular debris was removed by centrifugation at 2,000g for 10 min. The membrane fraction of the resulting supernatant was recovered by centrifugation at 100,000g for 15 min at 2°C. Pelleted membranes were washed with l M NaCl, and recovered by 179 centrifugation as described above. Membrane material equivalent to 15 ug of total protein was solubilized in Laemmli sample buffer, boiled for 5 min, and separated by SDS-PAGE (10% [w/v] gels). Separated proteins were electrophoretically transferred to Immobilon-P membranes (Millipore, Bedford, MA) in a solution consisting of 25 mM Tris, 192 mM Gly, and 20% (v/v) methanol, using standard procedures (Harlowe and Lane, 1988). Membranes were probed with anti-LeAOS antibodies used at a 1:2,000 dilution in a Tris-buffered saline solution containing 1% (w/v) bovine serum albumin as a blocking agent and 0.05% (w/v) Tween 20 to reduce non-specific binding. Anti gen- antibody complexes were detected with the use of an alkaline phosphatase-conjugated second antibody as described by the manufacturer (Kirkegaard and Perry, Gaithersburg, MD). Southern-Blot Analysis Genomic DNA from young leaves of cv Micro-Tom plants was purified as described by Rogers and Bendich (1985). Ten-micrograrn aliquots of DNA were digested with restriction enzymes, electrophoresed on a 0.8% (w/v) agarose gel, and blotted to Duralon-UV membranes (Stratagene) as indicated by the manufacturer. Blots were pre- hybridized at 65°C in a solution containing 5)( SSPE, 5x Denhardt's solution, 100 ug/mL denatured salmon sperm DNA, and 0.5% (w/v) SDS. Blots were hybridized at 65°C and washed at the same temperature in a solution containing 0.5x SSPE and 0.5% (w/v) SDS. DNA probes were prepared using a T7 Quickprime Kit (Pharmacia Biotech, Piscataway, NJ). The following cDNA fragments were labeled for use as probes: a 1.7- kb EcoRI-Xhol fragment containing full-length LeA OS; a 1.4-kb EcoRI-HindIH 180 fragment containing the coding region of LeHPL; and a 0.2-kb EcoRI-EcoRI fragment containing the LeHPL 5'-UTR. Wounding Experiments Manducta sexta larvae were reared on artificial diet as described by the vendor (Carolina Biological Supply, Burlington, NC) fiom which the eggs were purchased. One larva (third instar) was placed on the terminal leaflet of the oldest leaf of a 3-week- old cv Micro-Tom plant that contained three fully expanded leaves. Larvae were allowed to feed on the leaf for 5 to 10 min, during which time 5% to 10% of the area of the leaf was consumed. Plants were sampled for RNA analysis at different times after the challenge. Leaf tissue from six to eight plants per time point was pooled prior to RNA extraction. Mechanical wounding of plants was performed using a hemostat as described previously (Howe et al., 1996). RNA Gel Blot Analysis Total RNA was isolated from tomato tissue and analyzed by RNA-blot hybridization as previously described (Howe et al., 1996), except that Duralon-UV membranes were used in place of nitrocellulose. All gels were run in duplicate, with one set stained with ethidium bromide to ensure equal loading of the samples and intactness of the RNA. Hybridization signals were visualized by autoradiography using Kodak XAR-S fihn, or were measured using a Phosphorimager (Molecular Dynamics). These signals were normalized to signals obtained using a probe for translation initiation factor eIF4A mRN A (Taylor et al., 1993). Hybridization and subsequent washing of eIF4A-probed 181 blots was performed at 60°C in 2X SSPE. DNA probes were prepared as described above. 182 References Avdiushko, S., Croft, K.P.C., Brown, G.C., Jackson D.M., Hamilton-Kemp, TR, and Hildebrand, D. (1995). Effect of volatile methyl jasmonate on the oxylipin pathway in tobacco, cucumber, and Arabidopsis. Plant Physiol.109: 1227-1230. Baier, M., and Dietz, K-J. (1999). Alkyl hydroperoxide reductases: the way out of the oxidative breakdown of lipids in chloroplasts. Trends Plant. Sci. 4: 166-168. Bate, N.J., and Rothstein, SJ. (1998). C6-Volatiles derived from the lipoxygenase pathway induce a subset of defense-related genes. Plant J. 16: 561-569. Bate, N.J, Sivasankar, S., Moxon, C., Riley, J .M.C., Thompson, J .E., and Rothstein, SJ. (1998). Molecular characterization of an Arabidopsis gene encoding hydroperoxide lyase, a cytochrome P-450 that is wound inducible. Plant Physiol. 117: 1393-1400. Blechert, S., Brodschelm, W., Holder, S., Karnmerer, L., Kutchan, T.M., Mueller, M.J., Xia, Z-Q., and Zenk, M.H. (1995). The octadecanoic pathway: signal molecules for the regulation of secondary pathways. Proc. Natl. Acad. Sci. USA. 92: 4099- 4105. Blée, E. (1998). Phytooxylipins and plant defense reactions. Prog. Lipid Res. 37: 33-72. Blée, E, and J oyard, J. (1996). Envelope membranes from spinach chloroplasts are a site of metabolism of fatty acid hydroperoxides. Plant Physiol. 110: 445-454. Blée, E., Wilcox, A.L., Mamett, L.J., and Schuber, F. (1993). Mechanism of reaction of fatty acid hydroperoxides with soybean peroxygenase. J. Biol. Chem. 268: 1708- 1715 Buttery, R.G., Ling, LC. (1993). Volatile components of tomato fi'uits and plant parts. In R. Teranishi, R.G. Buttery, H. Sugisawa, eds, Bioactive Volatile Compounds in Plants. American Chemical Society, Washington, DC, pp 23-34 Caldelari, D., and Farmer EB. (1997). A rapid assay for the coupled cell free generation of oxylipins. Phytochemistry 47: 599-604. Chao, W.S., Gu, Y-Q, Pautot, V., Bray, E.A., and Walling, LL. (1999). Leucine aminopeptidase RNAs, proteins, and activities increase in response to water deficit, salinity, and the wound signals systemin, methyl jasmonate, and abscisic acid. Plant Physiol. 120: 979-992. 183 Conconi, A., Miquel, M., Browse, J .A., and Ryan, GA. (1996). Intracellular levels of free linolenic and linoleic acids increase in tomato leaves in response to wounding. Plant Physiol. 111: 797-803. Creelman, R.A., and Mullet, J .E. (1995). J asmonic acid distribution and action in plants: regulation during development and response to biotic and abiotic stress. Proc. Natl. Acad. Sci. USA. 92: 4114-4119. Creehnan, R.A., and Mullet, J .E. (1997). Biosynthesis and action of jasmonates in plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 48: 355-381. Creelman, R.A., Tierney, ML, and Mullet, J .E. (1992). J asmonic acid/methyl jasmonate accumulate in wounded soybean hypocotyls and modulate wound gene expression. Proc. Natl. Acad. Sci. USA. 89: 4938-4941. Croft, K.P.C., Juttner, F., and Slusarenko, A.J. (1993). Volatile products of the lipoxygenase pathway evolved from Phaseolus vulgaris (L.) leaves inoculated with Pseudomonas syringae pv phaseolicola. Plant Physiol. 101: 13-24. Doares, S.H., Syrovets, T., Weiler, E.W., and Ryan, CA. (1995). Oligogalacturonides and chitosan activate plant defense genes through the octadecanoid pathway. Proc. Natl. Acad. Sci. USA. 92: 4095-4098. Fan, X., Mattheis, J .P., Fellman, J .K. (1998) A role for jasmonates in climacteric fi'uit ripening. Planta 204: 444-449. Farmer, E.E., Weber, H., and Vollenweider, S. (1998). Fatty acid signaling in Arabidopsis. Planta 206: 167-174. Fauconnier, M-L., Perez, A.G., Sanz, C., and Marlier, M. (1997). Purification and characterization of tomato leaf (Lycopersicon esculentum Mill.) hydroperoxide lyase. J Agric Food Chem 45: 4232-4236. F roehlich, J.E., DeRocher, A., and Howe, G.A. (1999) Allene oxide synthase is imported and targeted to the inner envelope membrane of pea and tomato chloroplasts (abstract no. 317). Plant Biology '99 1999: 85. Galliard, T., Matthew, J .A., Wright, A.J., and Fishwick, M.J. (1977). The enzymatic breakdown of lipids to volatile and non-volatile carbonyl fragments in disrupted tomato fruits. J Sci Food Agric. 28: 863-868. Gardner, H.W. (1989). Oxygen radical chemistry of polyunsaturated fatty acids. Free Radic Biol Med. 7: 65-86 Gardner, H.W. (1991). Recent investigations into the lipoxygenase pathway of plants. Biochim. Biophys. Acta 1084: 221-239. 184 Gardner, H.W., Weisleder, D., and Plattner, RD. (1991) Hydroperoxide lyase and other hydroperoxide-metabolizing activity in tissues of soybean, Glycine max. Plant Physiol. 97: 1059-1072. Grechkin, A.N., F azliev, RN, and Mukhtarova, LS. (1995) The lipoxygenase pathway in garlic (Allium sativum L.) bulbs: detection of the divinyl ether oxylipins. F EBS Lett 371: 159-162. Gundlach, H., Muller, M.J., Kutchan, T., and Zenk, M.H. (1992) Jasmonic acid is a signal transducer in elicitor-induced plant cell cultures. Proc. Natl. Acad. Sci. USA. 89: 2389-2393. Hamberg, M. (1998). A pathway for biosynthesis of divinyl ether fatty acids in green leaves. Lipids 33: 1061-1071. Hammond-Kosack, K.E., and Jones, J .D.G. (1996). Resistance gene-dependent plant defense responses. Plant Cell 8: 1773-1791. Harlowe, E., and Lane, D. (1988) Antibodies: A Laboratory Manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY Harms, K., Atzom, R., Brash, A., Kuhn, H., Wastemack, C., Willmitzer, L., and Pena- Cortés, H. (1995) Expression of a flax allene oxide synthase cDNA leads to increased endogenous jasmonic acid (J A) levels in transgenic potato plants but not to a corresponding activation of J A-responding genes. Plant Cell 7: 1645-1654. Harms, K., Ramirez, 1., and Pena-Cortes, H. (1998). Inhibition of wound-inducible accumulation of allene oxide synthase transcripts in flax leaves by aspirin and salicylic acid. Plant Physiol. 118: 1057-1065. Hatanaka, A. (1993). The biogeneration of green odor by green leaves. Phytochemistry 34: 1201-1218. Hatanaka, A., Kajiwara, T., and Sekiya, J. (1987) Biosynthetic pathway for C6-aldehyde formation from linolenic acid in green leaves. Chem. Phys. Lipids 44: 341-361. Heitz, T., Bergey, D.R., and Ryan, CA. (1997) A gene encoding a chloroplast-targeted lipoxygenase in tomato leaves is transiently induced by wounding, systemin, and methyl jasmonate. Plant Physiol. 114: 1085-1093. Howe, G.A., Lightner, J ., Browse, J ., Ryan, CA. (1996). An octadecanoid pathway mutant (J L5) of tomato is compromised in signaling for defense against insect attack. Plant Cell 8: 2067-2077. 185 Howe, G.A., and Ryan, CA. (1999) Suppressors of systemin signaling identify genes in the tomato wound response pathway. Genetics 153: 1411-1421. Kazeniac, S.J., Hall, RM. (1970). Flavor chemistry of tomato volatiles. J Food Sci. 35: 519-530. Kohlrnann, M., Bachmann, A., Weichert, H., Kolbe, A., Balkenhohl, T., Wastemack, C., and F eussner, I. (1999). Formation of lipoxygenase-pathway—derived aldehydes in barley leaves upon methyl jasmonate treatment. Eur. J. Biochem. 260: 885-895. Kubigsteltig, I., Laudert, D., and Weiler, E.W. (1999). Structure and regulation of the Arabidopsis thaliana allene oxide synthase gene. Planta 208: 463-471. Laudert, D., Pfannschmidt, U., Lottspeich, P., Hollander-Czytko, H., and Weiler, E.W. (1996). Cloning, molecular and functional characterization of Arabidopsis thaliana allene oxide synthase (CYP 74), the first enzyme of the octadecanoid pathway to jasmonates. Plant Mol Biol. 31: 323-335. Laudert, D., and Weiler, E.W. (1998). Allene oxide synthase: a major control point in Arabidopsis thaliana octadecanoid signaling. Plant J. 15: 675-684. Lehmann, J ., Atzom, R., Briickner, C., Reinbothe, S., Leopold, J ., Wastemack, C., Parthier B (1995) Accumulation of jasmonate, abscisic acid, specific transcripts and proteins in osmotically stressed barley leaf segments. Planta 197: 156-162. Leon, J ., Rojo, E., Titarenko, E., and Sénchez-Serrano, J .J . (1998). J asmonic acid- dependent and -independent wound signal transduction pathways are differentially regulated by Ca2+/calmodulin in Arabidopsis thaliana. Mol. Gen. Genet. 258: 412- 419. Lers, A., Khalchitski, A., Lomaniec, E., Burd, S., and Green, P.J. (1998). Senescence- induced RNases in tomato. Plant Mol. Biol. 36: 439-449. Matsui, K., Shibutani, M., Hase, T., Kajiwara, T. (1996). Bell pepper fruit fatty acid hydroperoxide lyase is a cytochrome P450 (CYP74B). FEBS Lett 394: 21-24. Matsui, K., Toyota, H., Kajiwara, T., Kakuno, T., and Hatanaka, A. (1991) A fatty acid hydroperoxide cleaving enzyme, hydroperoxide lyase, from tea leaves. Phytochemistry 30: 2109-21 13. 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. 186 Mueller, M.J., Brodschelm, W., Spannagl, E., and Zenk, M.H. (1993). Signaling in the elicitation process is mediated through the octadecanoid pathway leading to jasmonic acid. Proc. Natl. Acad. Sci. USA. 90: 7490-7494. Narvaez-Vasquez, J ., Florin-Christensen, J ., and Ryan, CA. (1999). Positional specificity of a phospholipase A2 activity induced by wounding, systemin, and oligosaccharide elicitors in tomato leaves. Plant Cell 11: 2249-2260. Nelson, DR. (1999). Cytochrome P450 and the individuality of species. Arch. Biochem. Biophys. 369: 1-10. Pan, 2., Durst, F., Wreck-Reichhart, D., Gardner, H.W., Camara, B., Cornish, K., and Backhaus, R.A. (1995). The major protein of guayule rubber particles is a cytochrome P450. J. Biol. Chem. 15: 8487-8494. Pena-Cortes, H., Albert, T., Prat, S., Weiler, E.W., and Willmitzer, L. (1993). Aspirin prevents wound-induced gene expression in tomato leaves by blocking jasmonic acid biosynthesis. Planta 191: 123-128. Penninckx, I.A.M.A., Eggerrnont, K., Terras, F .RG., Thomma, B.P.H.J., Sarnblanx, G.W.D., Buchala, A., Metraux, J-P., Manners, J .M., and Broekaert, W.F. (1996) Pathogen-induced systemic activation of a plant defensin gene in Arabidopsis follows a salicylic acid-independent pathway. Plant Cell 8: 2309-2323. Rogers, SO, and Bendich, A.J. (1985). Extraction of DNA from milligram amounts of flesh, herbarium and mummified plant tissue. Plant Mol. Biol. 5: 69-76. Rojo, E., Titarenko, E., Leon, J., Berger, S., Vancanneyt, G., and Sénchez-Serrano, J .J . (1998). Reversible protein phosphorylation regulates jasmonic acid-dependent and -independent wound signal transduction pathways in Arabidopsis thaliana. Plant J. 13: 153-165. Ryan, CA. (2000). The systemin signaling pathway: differential activation of plant defensive genes. Biochem. Biophys. Acta 1477: 112-121. Salch, Y.P., Grove, M.J., Takamura, H., and Gardner, H.W. (1995). Characterization of a C-5,13-cleaving enzyme of 13(S)-hydroperoxide of linolenic acid by soybean seed. Plant Physiol. 108: 1211-1218. Shibata, Y., Matsui, K., Kajiwara, T., and Hatanaka, A. (1995a) Fatty acid hydroperoxide lyase is a heme protein. Biochem. Biophys. Res. Commun. 207: 438-443. Shibata, Y., Matsui, K., Kajiwara, T., and Hatanaka, A. (1995b) Purification and properties of fatty acid hydroperoxide lyase flom green bell pepper fluits. Plant Cell Physiol. 36: 147-156. 187 Song, W-C., and Brash, AR. (1991). Purification of an allene oxide synthase and identification of the enzyme as a cytochrome P-450. Science 252: 781-784. Song, W-C., Funk., CD, and Brash, AR. (1993). Molecular cloning of an allene oxide synthase: a cytochrome P450 specialized for the metabolism of fatty acid hydroperoxides. Proc. Natl. Acad. Sci. USA. 90: 8519-8523. Staswick, PE, and Lehman, CC. (1999). J asmonic acid-signaled responses in plants. In A.A. Agrawal, S. Tazun, and E., Bent, eds, Induced Plant Defenses against Pathogens and Herbivores. American Phytopathological Society Press, St. Paul, pp 117-136. Szczesna-Skorpa, E., Straul, P., and Kemper, B. (1993). Deletion of a conserved tetrapeptide, PPGP, in P450 2C2 results in loss of enzymatic activity without a change in its cellular location. Arch. Biochem. Biophys. 304: 170-175. Taylor, C.B., Bariola, P.A., Delcardayré, S.B., Raines, R.T., and Green, P.J. (1993) RN S2: a senescence-associated RNase of Arabidopsis that diverged flom the S- RNase before speciation. Proc. Natl. Acad. Sci. USA. 90: 51 18-5122. Titarenko, E., Rojo, E., Leon, J., Sénchez-Serrano, J .J . (1997) J asmonic acid-dependent and -independent signaling pathways control wound-induced gene activation in Arabidopsis thaliana. Plant Physiol. 113: 817-826. Vick, BA. (1991). A spectrophotometric assay for hydroperoxide lyase. Lipids 26: 315- 320. Vick, RA., and Zimmerman, DC. (1987) Pathways of fatty acid hydroperoxide metabolism in spinach leaf chloroplasts. Plant Physiol. 85: 1073-1078. Weiler, E.W., Albrecht, T., Groth, B., Xia, Z-Q., Luxem, M., Lib, H., Andert, L., and Spengler, P. (1993). Evidence for the involvement of jasmonates and their octadecanoid precursors in the tendril coiling response of Bryonia dioica. Phytochemistry 32: 591-600. Zhuang, H., Hamilton-Kemp, T.R., Andersen, R.A., and Hildebrand, DP. (1996) The impact of alteration of polyunsaturated fatty acid levels on CG-aldehyde formation of Arabidopsis thaliana leaves. Plant Physiol. 111: 805-812. Zimmerman, D2, and Coudron, CA. (1979). Identification of traumatin, a wound hormone, as 12-oxo-trans-dodecenoic acid. Plant Physiol. 63: 536-541. 188 Chapter 7 Conclusions and future perspectives 189 The systemic wound response of tomato is an important defense mechanism against insect herbivores. The discovery of systemin provided a breakthrough in our understanding of how local injury induces systemic expression of defense genes such as Proteinase Inhibitors (PIs; McGurl et al., 1992: Pearce et al., 1991). Manipulation of prosystemin expression in transgenic plants demonstrated that this polypeptide signal is an essential regulator of systemic PI expression (McGurl et al., 1992; Mch1 et al., 1994). Subsequent studies showed that systemin induces PI expression by activating the octadecanoid pathway for jasmonic acid (J A) biosynthesis (Doares et al., 1995; Howe et al., 1996). At the outset of this thesis research, a wealth of circumstantial evidence indicated that systemin might function as a long-distance signal for PI expression (Ryan and Pearce, 1998). To investigate this hypothesis, I characterized tomato mutants that are defective in the systemic wound response. Systemin-insensitive spr1 plants were previously isolated as suppressor of prosystemin-mediated responses (spr), and were shown to be deficient in wound-induced systemic PI expression (Howe and Ryan, 1999). As described in Chapter 2, systemin failed to induce J A biosynthesis and subsequent expression of PIs in spr1 plants (Figure 2-5; Figure 2-6). In contrast, application of oligogalacturonic acid (OGA), which is regarded as a local wound signal, resulted in normal PI expression in spr1 plants (Figure 2-3; Figure 2-4). These results indicate that the spr1 mutation defines the systemin-dependent pathway of systemic wound signaling, and further indicate that the Spr] gene product plays a role in coupling systemin perception at the plasma membrane to the initiation of J A biosynthesis in the chloroplast (Figure 7-1). 190 Reciprocal grafting experiments between wild-type and spr1 plants were used to investigate how systemin regulates systemic PI expression (Figure 2-7). Whereas spr1 plants were impaired in the generation of the systemic wound signal in wounded leaves, the mutant was able to perceive that signal in undamaged leaves. Because Spr] is required for the recognition of systemin, these results indicate that systemin is likely not the long-distance signal for PI expression. Systemin appears to act at or near the site of wounding (i.e. in rootstock tissues) to increase J A synthesis to a level that is required for the systemic response. Consistent with this interpretation, grafting experiments using J A-insensitive and J A-deficient mutants indicated that J A or a derivative of J A is the systemic wound signal (Li et al., 2002b). These experiments demonstrated that J A biosynthesis is required for the generation of the systemic wound signal, and that J A perception is essential for recognition of this signal in unwounded leaves. This finding supports the idea that systemin amplifies the systemic wound signal in damaged tissue by stimulating J A biosynthesis. The role of J A in systemic wound signaling was investigated further by characterization of J A-deficient J L1 plants. Gas chromatography-mass spectrometry (GC-MS) analysis showed that J L1 plants are compromised in the conversion of 12- oxo-phytodienoic acid (OPDA) to J A (Figure 3-2). OPDA application experiments showed that this octadecanoid pathway intermediate is not an active signal for PI expression in tomato (Figure 3-3). Reciprocal grafting experiments with wild-type and JL1 plants confirmed that JA biosynthesis is required for the generation of the long- distance signal for PI expression (Figure 3-4). 191 \ / p 1 Systemic \ ,’ response -+-> PI<----JA A I iai1 {I‘— JL1 Local 1 response + OPDA def1 —-n l spr2—-| : spr1 I 4 OGA/V systemin Figure 7-1. Revised model of systemic wound signaling in tomato. Wounding activates systemin/JA-dependent (broken arrows) and J A-independent signaling (solid arrows) pathways that regulate distinct sets of wound response genes in undamaged leaves (systemic response). See text for details. 192 It was indicated that two related cytochrome P4508, allene oxide synthase (AOS) and hydroperoxide lyase (HPL), regulate oxylipin levels in plants by competing for a common substrate, 13-hydroperoxy linolenic acid (l3-HPOT; Bate et al., 1998; Blee and J oyard, 1996; Lauder et al., 1996). AOS and HPL metabolize l3-HPOT to J A and volatile C6 aldehyde, respectively. To investigate the regulation of J A biosynthesis, genes encoding these two enzymes were cloned flom tomato. Expression of these genes in E. coli confirmed that AOS and HPL efficiently metabolize the same substrate, 13-HPOT, but do not metabolize other hydroperoxy fatty acid substrates (Figure 6-4). Both genes are expressed in leaves, which supports the idea that AOS and HPL compete for the same substrate in wounded tissue (Figure 6-5). Previous studies indicated that a J A-independent signaling pathway mediates some systemic wound responses in tomato (O'Donnell et al., 1998; Stratmann and Ryan, 1997). This idea is consistent with the observation that expression of AOS was up- regulated by wounding in both wild-type and J A-deficient defenseless] (defI) plants (Figure 6-7). In addition, early wound response genes such as lipoxygenaseD (LoxD) were systemically expressed upon wounding of spr1 plants, which indicates that a systemin-independent signaling pathway is involved in this systemic response (Figure 2-2; Figure 2-5). Wounding activated the systemic expression of a mitogen-activated protein kinase (MAPK) gene homologue (WIPK) in J A-deficient and J A-insensitive mutants. This observation indicates the existence of a J A-independent signaling. Expression of LoxD was determined to be under regulation by both J A-independent and -dependent signaling pathways. 193 Based on the findings flom this thesis study, a model of systemic wound signaling in tomato is proposed (Figure 7-1). In this model, wounding activates two distinct signaling pathways for systemic gene expression. One is the JA-dependent signaling pathway, which requires the action of Spr], Spr2, Def], and Jail. fi-oxidation is involved in this pathway, which is compromised in JL1 plants. This pathway regulates systemic expression of late genes such as PIs. J A or a derivative of J A that is transported by the phloem may be the long-distance signal in this pathway. Systemin promotes the synthesis of J A at or near the site of wounding. The other systemic wound signaling pathway operates independently flom J A and systemin, and is poorly understood. A subset of early wound response genes is regulated by this signaling pathway. The J A-independent pathway also regulates the systemic wound induction of a MAPK. In this case, the long-distance signal appears to be xylem-home (Stratmann and Ryan, 1997). Further biochemical studies are needed to determine whether J A or one of its derivatives is the long-distance signal of systemic PI expression. Because the systemic wound signal for PI expression moves though the phloem (Nelson et al., 1983), the relative level of the endogenous signal in phloem is expected to rapidly increase in response to wounding. Measurement of JA and related oxylipins in phloem sap will be useful to address this question. In addition, application of radiolabeled J A to the wound site may be helpful to monitor the movement of IA in vivo. As an initial step to identifying the Spr] gene by map-based cloning, a mapping population was generated by crossing the spr1 mutant to the wild tomato species Lycopersicon pennellii (Appendix of Chapter 2). Bulk segregant analysis and amplified 194 fragment length polymorphism (AF LP) analysis would be useful to find molecular markers closely linked to the Spr] locus. A similar approach was used to map the chromosomal location of Def] and Spr2 (Li et al., 2003; Li et al., 2002a). Molecular cloning of the Spr] gene will provide insight into how the binding of systemin to the SR160 receptor on the plasma membrane activates J A biosynthesis in chloroplasts (Scheer and Ryan, 2002). Map-based cloning of the gene defined by the J L1 mutant line is in progress in the Howe lab (C. Li and G. Howe, unpublished results). This gene product is presumably involved in B-oxidation, which is required for both degradation of fatty acids and conversion of OPDA to JA. The cloning of this gene will extend our understanding of how the conversion of OPDA to J A regulates endogenous J A levels. OPDA itself is not an active signal for wound-induction of PHI, which is contrary to the finding that OPDA regulates expression of defense genes in Arabidopsis (Stintzi et al., 2001). Microarray experiments with JL1 plants will be helpful to investigate the putative role of OPDA as a signal for gene expression in tomato plants. In vitro import assays have established that AOS is targeted to the inner envelope membrane of chloroplasts, whereas HPL is targeted to the outer envelope membrane (Froehlich et al., 2001). This finding fiirther supports the possibility that the two enzymes compete for the same substrate located in the envelope membrane. To further address this hypothesis, it will be useful to determine whether the two enzymes co-localize in the same cell and the same chloroplast. Microarray experiments using J A-deficient and J A-insensitive mutants would be helpful to identify wound-inducible genes that are regulated by the J A-independent 195 signaling pathway. It may also be informative to examine the expression of WIPK using stearn-girdled tomato plants in which the xylem but not the phloem is intact. This experiment would determine whether the JA-independent signal for WIPK expression is phloem-bome or xylem-bome. 196 References Bate, N.J., and Rothstein, SJ. (1998). C6-Volatiles derived flom the lipoxygenase pathway induce a subset of defense-related genes. Plant J. 16: 561-569. Blee, E., and Joyard, J. (1996). Envelope membranes flom spinach chloroplasts are a site of metabolism of fatty acid hydroperoxides. Plant Physiol. 110: 445-454. Doares, S., Narvaez-Vasquez, J ., Conconi, A., and Ryan, C. A. (1995). Salicylic acid inhibits synthesis of proteinase inhibitors in tomato leaves induced by systemin and jasmonic acid. Plant Physiol. 108:1741-1746. Froehlich, J ., Itoh, A., and Howe, G.A. (2001). Tomato allene oxide synthase and fatty acid hydroperoxide lyase, two cytochrome P4508 involved in oxylipin metabolism, are targeted to different membranes of chloroplast envelope. Plant Physiol. 125:306-17. Howe, G.A., Lightner, L., Browse, J ., and Ryan, CA. (1996). An octadecanoid pathway mutant (J L5) of tomato is compromised in signaling for defense against insect attack. Plant Cell 8: 2067-2077. Howe, G.A., and Ryan, CA. (1999). Suppressors of systemin signaling identify genes in the tomato wound response pathway. Genetics 153: 1411-1421. Laudert, D., Pfannschmidt, U., Lottspeich, F ., Hollander-Czytko, H., and Weiler E.W. (1996). Cloning, molecular and functional characterization of Arabidopsis thaliana allene oxide synthase (CYP 74), the first enzyme of the octadecanoid pathway to jasmonates. Plant Mol. Biol. 31: 323-335. Li, C., Liu, G, Xu, C., Lee, G.I., Bauer, P., Ling, H., Ganal, M., and Howe, G.A. (2003). The tomato suppressor of prosystemin-mediated responses 2 (spr2) gene encodes a fatty acid desaturase required for the biosynthesis of jasmonic acid and the production of a systemic wound signal for defense gene expression. Plant Cell 15: 1646-1661. Li, C., Williams, M., Loh, Y.-T., Lee, G.I., and Howe, G.A. (2002a). Resistance of cultivated tomato to cell-content feeding herbivores is regulated by the octadecanoid signaling pathway. Plant Physiol. 130: 494-503. Li, L., Li, C., Lee, G.I., and Howe, G.A. (2002b). Distinct roles for jasmonate synthesis and action in the systemic wound response of tomato. Proc. Natl. Acad. Sci. USA. 99: 6416-6421. 197 McGurl, B., Orozco-Cardenas, M., Pearce, G., and Ryan, CA. (1994). Over-expression of the prosystemin gene in transgenic tomato plants generates a systemic signal that constitutively induces proteinase inhibitor synthesis. Proc. Natl. Acad. Sci. USA. 91: 9799-9802. McGurl, B., Pearce, G., Orozco-Cardenas, M., and Ryan, CA. (1992). Structure, expression, and antisense inhibition of the systemin precursor gene. Science. 255: 1570-1573. Nelson, C. E., Walker-Simmons, M., Makus, D., Zuroske, G., Graham, J., and Ryan, C. A. (1983). Regulation of synthesis and accumulation of proteinase inhibitors in leaves of wounded tomato plants. in Plant Resistance to Insects, Symposium ' Series 208, ed. Hedin, P. (Am. Chem. Soc., Washington, DC), pp. 103-122. O'Donnell, P.J., Truesdale, M.R., Calvert, C.M., Dorans, A., Roberts, M.R., and Bowles, DJ. (1998). A novel tomato gene that rapidly responds to wound- and pathogen-related signals. Plant J. 14: 137-142. Pearce, G., Strydom, D., Johnson, S., and Ryan, CA. (1991). A polypeptide flom tomato leaves induces wound-inducible proteinase inhibitor proteins. Science 253: 895-898. Ryan, CA. (2000). The systemin signaling pathway: differential activation of plant defensive genes. Biochim. Biophys. Acta 1477: 112-121. Ryan, C.A., and Pearce, G. (1998), Systemin: a polypeptide signal for plant defensive genes. Annu. Rev. Cell. Dev. Biol.14:l-17. Scheer, J .M,. and Ryan, GA. (2002). The systemin receptor SR160 flom Lycopersicon peruvianum is a member of the LRR receptor kinase family. Proc. Natl. Acad. Sci. USA. 99: 9585-9590. Stintzi, A., Weber, H., Reymond, P., Browse, B., and Farmer, EB (2001). Plant defense in the absence of jasmonic acid: the role of cyclopentenones. Proc. Natl. Acad. Sci. USA. 98: 12837-12842. Stratmann, J .W., and Ryan, CA. (1997). Myelin basic protein kinase activity in tomato leaves is induced systemically by wounding and increases in response to systemin and oligosaccharide elicitors. Proc. Natl. Acad. Sci. USA. 94: 11085- 1 1089. 198 / IIIIIIIIIIIIIIIIIIIIIIIIIIIIIII llllll HI” 11111 1293 02469 2273