Botanical Studies (2012) 53: 75-84.
physiology
Induction of tomato Jasmonate-Resistant 1-Like 1 gene expression can delay the colonization of Ralstonia solanacearum in transgenic tomato
Jhy-Gong WANG and Hsu-Liang HSIEH*
Institute of Plant Biology, College of Life Science, National Taiwan University, Taipei 106, Taiwan
(Received April 8, 2011; Accepted October 6, 2011)
ABSTRACT. The bacterial wilt in tomato caused by Ralstonia solanacearum infection is common and widespread, especially in hot and humid environments. Combating this disease is difficult due to unstable host resistance and the variation and diversity of the bacterial strains. Thus, the molecular mechanisms underlying tomato resistance against Ralstonia solanacearum remain unknown. Here, we isolated a homolog of tomato Solanum lycopersicum Jasmonate-Resistant 1 (SUAR1), named SI,JAR1-like 1 (SURL1), and generated trans­genic tomato lines harboring an inducible promoter-driven SURL1 construct. SlJRL1 shares 99% amino acid identity with SlJAR1. Intriguingly, SURL1 showed preferential expression in aerial parts and SlJAR1 in roots. DNA gel blot analysis revealed multiple copies of SURL1 in the tomato genome. Transgenic tomato contain­ing a single copy of the transgene SURL1 exhibited high levels of SURRL1 expression two days after dexam-ethasone (DEX) induction. Moreover, DEX-induced SURL1 expression could delay the symptoms of tomato bacterial wilt, and efficiently reduce the amount of Ralstonia in stems. The phytohormone jasmonic acid may play a role in this resistance response. This study of inducible SlJRL1 expression in transgenic tomato contrib­utes to the molecular understanding of tomato resistance against bacterial wilt.
Keywords: Jasmonic acid; Ralstonia solanacearum; SlJAR1; Tomato bacterial wilt-SlJAR1.
introduction
Ralstonia solanacearum, a soil-borne bacterium and one of the most devastating pathogens, causes a lethal dis­ease known as bacterial wilt in more than 200 plant spe­cies in tropical, subtropical and temperate regions of the world. Its hosts are ornamentals, weeds, and crops, includ­ing tomato, potato, banana and pepper (Hayward, 1991; Schell, 2000; Deslandes et al., 2002). At the early stage of infection, R. solanacearum attaches to and enters the lat­eral roots where wounding has taken place. Subsequently, the bacterium invades the root cortex, translocates to the xylem vessels, and spreads rapidly throughout the vascular system of infected hosts. The bacterium then colonizes the xylem tissues of plants and causes wilt symptoms and death (Schell, 2000; Gemn and Boucher, 2004). R. solan-acearum is later released from the roots or the collapsed stems into the soil for the infection of new hosts.
Because of the serious impact R. solanacearum infec­tion has on plants, several approaches have been used for control, including crop rotation (Adhikari and Basnyat, 1998), biocontrol agent application (Guo et al., 2004; Xue

*Corresponding author: E-mail: hlhsieh@ntu.edu.tw; Tel: +886-2-33662540; Fax: +886-2-23918940.
et al., 2009), and resistant cultivar breeding via traditional or genetic engineering strategies (Chan et al., 2005; Tho-quet et al., 1996; Wang et al., 2000). Crop rotation has a limited effect on controlling R. solanacearum due to the its wide range of hosts. Although biological control using bacteria has drawn much attention, its efficiency is limited because of environmental factors such as temperature, moisture, and the storage of bacteria (Guo et al., 2004). Al­though traditional breeding systems have produced some resistant breeds, their numbers are still limited and some undesirable traits have occurred. Genetic engineering has been used to generate plants with high resistance against R. solanacearum infection. Chan et al. (2005) reported that transgenic tomatoes harboring an Arabidopsis thionine gene THI2.1 driven by a fruit-inactive promoter displayed high resistance to this wilt bacteria.
In addition, previous studies suggested that plant hor­mones, including salicylic acid (SA) and ethylene (ET), play roles in wilt bacteria resistance. SA participates in R. solanacearum resistance in Arabidopsis in a gene-for-gene strategy (Deslandes et al., 2002; Deslandes et al., 2003) involving the bacterial PopP2 protein and Arabi-dopsis RRS1-R protein. The wilt bacteria first invade plant cells and release PopP2, a type III effector acting as an avirulence (Avr) protein, into the cytoplasm of plant cells. PopP2 then interacts with RRS1-R protein, a resistant
76
(R) protein containing a WRKY domain, to form a com­plex. The complex enters the nucleus to activate defense-related genes, thus leading to enhanced resistance against R. solanacearum (Deslandes et al., 2002; Deslandes et al., 2003). In the ET signaling pathway, Ethylene-Insensitive 2 (EIN2) is a positive regulator and degraded via 26S pro-teasome (Qiao et al., 2009; Stepanova and Alonso, 2009). Previous reports have shown the ein2-1 mutant with delayed wilt symptoms after inoculation with R. solan-acearum (Hirsch et al., 2002). Furthermore, some effec­tors of wilt bacteria can promote hormone production and influence plant defense responses (Hirsch et al., 2002). For example, inoculation with wilt bacteria induced the pro­duction of ET, leading to the induced expression of Eth­ylene Response Factor 1 (ERF1) and Pathogen Response Gene 4 (Valls et al., 2006). Therefore, phytohormone bio­synthesis and signaling play important roles in defending against R. solanacearum infection.
In addition to SA and ET, another biotic-stressed hormone, jasmonic acid (JA), plays an important role in the control of biotic invaders, such as necrotrophic pathogens and herbivores (as reviewed in Katsir et al., 2008). Previous studies have reported that pathogen infection may induce the production of JA-isoleucine (JA-Ile), an activeform of JA, in Arabidopsis, thus leading to increased expression of JA-responsive genes. Moreover, low levels of JA-Ile in the jar1-1 mutant suppressed JA responses and reduced the resistance against microbial pathogens and herbivores (Staswick and Tiryaki, 2004; Browse, 2009). JA-Ile formation is mediated by jasmonate-resistant 1 (JAR1) protein, which conjugates JA and isoleucine (Staswick et al., 2002; Staswick and Tiryaki, 2004). The binding of JA-Ile and coronatine-insensitive 1 (COI1) F-box protein enhances interactions between COI1 and jasmonate ZIM-domain (JAZ) proteins, resulting in the degradation of JAZ proteins via an ubiquitin/26S proteasome-mediated process. This leads to a release of the transcription factor MYC2, the expression of downstream JA responsive genes, and physiological responses (Chini et al., 2007; Dombrecht et al., 2007; Thines et al., 2007; Yan et al., 2009). JAR1 plays an important role in the regulation of JA responses in Arabidopsis. JAR1 homologs have been found in other plant species, including tomato, tobacco, and rice (Kang et al., 2006; Wang et al., 2007; Riemann et al., 2008; Suza et al., 2010). Kang et al. (2006) reported that silencing the JAR1 homologous gene Nicotiana tabacum JAR4, via virus-induced gene silencing, decreased JA-Ile levels and increased plant susceptibility to Manduca sexta attack. JAR1 and its homologs may thus participate in the control of biotic invaders.
In this study, we isolated Solanum lycopersicum jas-monate-resistant 1-like 1 (SURL1) encoding a 577 amino-acid protein with 99.3% amino-acid identity with Solanum lycopersicum jasmonate-resistant 1 (SlJAR1) in tomato (Suza et al., 2010). Moreover, we generated transgenic tomato lines harboring a SlURL1 gene construct driven by a glucocorticoid-inducible promoter. The transcripts of SlURL1 were greatly increased in the roots of transgenic
Botanical Studies, Vol. 53, 2012
plants treated with dexamethasone (DEX). High levels of SlURL1 in transgenic tomato resulted in delayed coloniza­tion of R. solanacearum Pss4 and delayed the develop­ment of wilt symptoms after bacterial inoculation. Thus, JA interferes with the virulence of the wilt bacteria.
materials and methods
Plant materials and growth conditions
Tomato (Solanum lycopersicum Mill.) cultivar CL5915 was used for gene isolation and transformation. Seeds were kindly provided by AVRDC-The World Vegetable Center, Tainan, Taiwan. Seeds were surface sterilized with 1% NaOCl by standard procedures and then germinated on Murashige and Skoog (MS) basal medium supplemented with 1% sucrose with a 16-h photoperiod at 26°C.
SIJRL1 cloning from tomato cultivar CL5915
To isolate the tomato homolog of AtFIN219/JAR1 gene, we searched the tomato expressed sequence tag (EST) library (Computational Biology and Functional Genomics Laboratory, http://compbio.dfci.harvard.edu/ tgi/) using the sequence for the AtFIN219/JAR1 cDNA. The EST clones TC129298, TC124338 and TC124339 showed high similarity. These 3 EST clones can be assembled into a full-length clone that is predicted to encode a 577 amino acid protein, named SlJRL1. The cDNA fragment of SlJRL1 was amplified with use of the primers 5'- TCTAGAATGAAGATGATGGTGGAAAATATTG-3' and 5'-ACATTGGGACGACCGGTAAGATCT-3' with the cDNA template derived from the reverse transcription of total RNAs isolated from the roots of tomato cultivar CL5915.
DNA and RNA gel blot analyses of SIJRL1 in tomato line CL5915
To detect the copy number of SlJRL1 in CL5915 or transgenic tomato lines, genomic DNA was isolated from tomato seedlings and digested with various restriction enzymes at 37°C overnight (Hsieh et al., 1996). The resulting DNA was then separated on a 0.7% agarose gel and blotted onto a charged nylon membrane (Roche). The gene-specific fragment of SlJRL1 in CL5915 or the full-length fragment of hygromycin phosphotransferase (HptII) for transgenic tomato lines was used as a probe for DNA gel blot analysis. The probes were labeled with DIG-dUTP by PCR amplification with the primers 5'-TGTCTGGACAAATCGTTCCT-3' and 5'-TTGGACCAACACATCTAGGG-3' for SlJRL1 and 5ˊ-GTGCTTGACATTGGGGAGTT-3' and 5'- TTATGCTCCAGCGGTTGTAG-3' for the HptII fragment. To detect expression patterns, total RNA was isolated from various tissues of tomato CL5915 plants or from transgenic tomato lines induced by 5 μM dexamethasone (DEX), then subjected to RNA gel blot analysis using a SlJRL1-specific probe. The results were then detected with the LAS-3000 imaging system (FujiFilm).
WANG and HSIEH ― SIJRL1 expression delays bacterial wilt
77
Plasmid constructs and transformation
The cDNA fragment of SlURL1 was amplified with the primers 5'-TCTAGAATGAAGATGATGGT GGAAAATATTG-3' and 5'-ACATTGGGACGACCGG TAAGATCT-3' by RT-PCR with the template generated from the total RNA of tomato roots. The resulting cDNA fragment was cloned into yT&A vector (Yeastern Bio­tech. Co.) and named yT&A/SlJRL1 for sequence verifi­cation. SlJRL1 was then excised from yT&A/SlJRL1 as aXbaI fragment and sub-cloned into thepTA7002 vector (Aoyama and Chua, 1997) with a DEX-inducible pro­moter. The pTA7002/SlURL1 was then transformed into CL5915 via Agrobacterium-mediated transformation (Chan et al., 2005).
Infection with R. solanacearum
Transgenic seeds of the T1 generation were screened on MS medium supplemented with 20 [ig/ml hygromycin for two weeks, and wild-type seeds were germinated on MS medium alone. The hygromycin-resistant T1 tomato plants and wild-type plants were transferred to 2-inch-diameter plastic pots containing sand, soil, rice husks, and compost (1:3:1:1) and incubated at 28°C under 16-h light/8-h dark for two weeks. The preparation of the inoculum R. solan-acearum strain Pss4 was described previously (Lin et al., 2008). Briefly, bacteria were grown on 523 medium (0.3 g/l MgSO4'7H2〇,2.0 g/l K2HPO4, 4.0 g/l yeast extract, 8.0 g/l casein hydrolysate, 10.0 g/l sucrose, 15g/l agar) supplemented with 50 mg/l TTC (2,3,5-triphenyltertrazo-lium chloride) at 28°C for 24 h, harvested and suspended in sterile water to adjust the concentration to 108 colony
forming units/ml. Roots of four-week-old tomato plants were damaged with a knife, and bacteria inoculum was poured into pots. Inoculated tomato plants were incubated at 28°C to develop wilt symptoms. For the induction of transgenic SlJRL1 expression, 5 μM DEX was added to each pot at the indicated time. As a control, 75 ml water was added to pots. Two biological experiments were per­formed, each experiment with three replicates, and each replicate involved ten plants.
Analysis of bacterial growth
To detect the propagation of bacteria, tissue 2 cm below the shoot apex of infected tomato plants was collected at different times and ground in sterile water. The extracts were then serially diluted with sterile water, and the ti-ter of bacteria was analyzed on SM1 medium (100 mg/l polymyxin B sulfate, 20 mg/l tyrothricin, 5 mg/l chloram-phenicol, 5 mg/l cycloheximide, 50 mg/l TTC, 5 mg/l crystal violet) at 28°C (Lin et al., 2008). Five plants were collected at each time, along with three bacterial growth replicates per plant.
results
Isolation of tomato JAR1-Iike gene, SIJRL1, from heat-tolerant tomato CL5915
Arabidopsis JAR1/FIN219 is a JA-conjugating enzyme responsible for producing JA-Ile (Staswick et al., 2002) and plays vital roles in far-red light and JA signaling inte­gration (Hsieh et al., 2000; Chen et al., 2007). To obtain tomato homologs of JAR1/FIN219, we searched the to-
Figure 1. Alignment of Solanum lycopersicum jasmonate-resistant 1 (SlJAR1) and SIJAR1-like 1 (SlJRL1) amino acid residues. The solid lines indicate adenylate-forming domains, boxI, boxII, and boxIII. Different amino acid residues between S1JAR1 and SlJRL1 are not shaded.
78
Botanical Studies, Vol. 53, 2012
mato EST library (Computational Biology and Functional Genomics Laboratory http://compbio.dfci.harvard.edu/tgi/) using the sequence for AtFIN219/JAR1 cDNA, and found three overlapping EST clones that could be assembled into a full-length cDNA fragment. The cDNA clone corresponding to the tomato JAR1-like gene, named SlJRL1, was eventually obtained by RT-PCR with cDNA templates derived from total RNA of the heat-tolerant tomato line CL5915. SlJRL1 encodes a 577 amino-acid polypeptide consisting of three adenylate-forming domains (Figure 1) that participate in ATP/AMP binding (Chang et al., 1997) and JA-Ile synthesis (Staswick and Tiryaki, 2004). Moreover, SlJRL1 shared more than 99% amino acid identity with SlJAR1, with only three amino-acid differences between sequences (Figure 1). To further determine the copy number of SlJRL1 in the tomato genome, we subjected genomic DNA digested with various restriction enzymes to DNA gel blot analysis using a gene-specific probe in the 3' end of SlJRL1 cDNA (Figure 2A). Both BamHI and HindIII restriction sites are absent from the coding region of SlJRL1 cDNA (Figure 2A). We found 6~8 bands under highly stringent conditions among different enzymedigested DNA samples (Figure 2B), which suggests that the tomato genome may contain multiple copies of SlJRL1 gene. An assay of tissue-specific expression patterns revealed SlJRL1 transcripts in all tissues examined, with the greatest abundance in leaves and the least in roots (Figure 2C).
Induction of SlJRL1 expression in transgenic tomato treated with DEX
To further understand the functions of SlURL1 in toma­to, we tried to introduce an overexpression or RNA inter­ference construct of SlURL1 into the heat-tolerant tomato line CL5915 by Agrobacterium-mediated transformation; however, we failed to obtain any putative transgenic toma­to plants (data not shown), implying that SlURL1 may have an essential role in tomato.We further generated transgenic tomato plants harboring the SlURL1 gene driven by a glu-cocorticoid-inducible promoter (Figure 3A) induced by the application of DEX (Aoyama and Chua, 1997). Transgenic tomatoes were screened on medium supplemented with hygromycin, and the insertion number of the transgenic SlURL1 gene was examined by DNA gel blot analysis with the full-length HPTII used as a probe. Only a single inser­tion was observed in the test lines (Figure 3B).
To understand the expression patterns of SlURL1 in transgenic tomatoes, the roots of the transgenic line 1

Figure 2. SURL1 has multiple copies in the tomato genome and high expression in aerial tissues. (A) Schematic diagram of the coding region of SlURL1 cDNA. The specific probe used for DNA gel blot analysis shown in (B) is indicated below the diagram. One EcoRI site is present in the coding region of SlURL1 Cdna; (B) DNA gel blot analysis of the SlURL1 gene in the tomato genome. Genomic DNA isolated from heat-tolerant tomato CL5915 was digested with various restriction enzymes shown above the blot and hybridized with a dig-labeled specific DNA probe corresponding to the C-terminal region of SlJRLl. The number shown at the left of the blot in­dicates 1 kb plus DNA ladder (Invitrogen); (C) RNA gel blot analysis of tissue-specific expression patterns of SURL1 in tomato. Total RNA isolated from different tissues of tomato CL5915 was analyzed for SlURL1 expression. Each lane was loaded with 10 [ig total RNA. 28S rRNA stained with methylene blue was a loading control. The numbers below the blot represent relative expression ratios normalized to the levels of 28S rRNA expression; the level of SlURL transcripts in roots was arbitrarily set to 1.
WANG and HSIEH ― SlJRLl expression delays bacterial wilt
79
were collected before and after DEX treatments and used for RNA extraction, followed by RNA gel blot analyses. SlURL1 transcripts were detected in the roots of transgenic line 1 before DEX treatments (Figure 4A), indicating the basal levels of endogenous SlURL1 expression. Further, the expression of SlURL1 was enhanced and peaked at two days after 5 μM DEX treatment, then was greatly reduced at four days (Figure 4A). The transgenic tomato lines 11 and 16 also showed significant induction of SlURL1 expression with DEX treatment (Figure 4B). These trans-genic tomato lines were thus useful for further functional studies.
Induction of SlJRL1 expression could delay the colonization of R. solanacearum in transgenic tomato
Because Arabidopsis JAR1/FIN219 has a role in patho­gen-triggered defense responses (Staswick and Tiryaki,
Figure 4. SURL1 shows high levels of expression in transgenic tomatoes. (A) RNA gel blot analysis of SlURL1 expression at different times for transgenic tomato line 1 under dexametha-sone (DEX) treatment. Total RNA was isolated from the roots of transgenic tomato line 1 with induction by 5 μM DEX treatment, grown under white light, and used for RNA gel blot analysis. The probe used is a dig-labeled SlURL1 cDNA fragment. 28S rRNA was a loading control. The numbers below the blot rep­resent relative expression ratios normalized to the levels of 28S rRNA expression; the level of SlURL transcripts at time 0 was ar­bitrarily set to 1; (B) RNA gel blot analysis of SlURL1 expression in different lines of transgenic tomato under DEX treatment. To­tal RNA was isolated from the roots of different transgenic lines of tomatoes at 2 days after 5 μM DEX (+) or water (-) treatment and subjected to RNA gel blot analysis. WT: wild-type; 1, 11, and 16: different transgenic tomato lines. The numbers below the blot represent relative expression ratios normalized to the levels of 28S rRNA expression; the level of SlURL transcripts in various samples with water treatment (-) was arbitrarily set to 1.
2004; Browse, 2009), we wondered whether SlURL1 participates in the resistance responses against R. solan-acearum infection. Wild-type and transgenic tomatoes inoculated with R. solanacearum showed similar wilt dis­ease symptoms timing (Figure 5A). 40% showed wilt dis­ease two days after infection (dai); almost all plants were wilted at 6 dai (Figure 5A). A similar trend was observed in wild-type plants treated with DEX (Figure 5B), imply­ing that DEX treatment did not interfere with the virulence of R. solanacearum infection. In contrast, transgenic to­mato treated with DEX at 2 dai showed no wilt symptoms (Figure 5B). Furthermore, with DEX treatment, 40% of wild-type plants showed wilt disease at 3 dai compared with less than 10% for transgenic plants. The proportion of wild-type and transgenic plants with wilt disease sub­stantially differed at 4 dai, with fewer differences at later stages of infection (Figure 5B). Induction of SlURL1 ex­pression can thus delay the development of wilt symptoms in transgenic tomatoes after R. solanacearum infection.
Wilt symptoms in plants infected with R. solanacearum are caused by the colonization of R. solanacearum in the xylem tissues, thus interrupting the translocation of water to leaves (Chan et al., 2005). To further determine whether SlJRL1 is involved in the disruption of R. solanacearum colonization, we examined the titer of R. solanacearum in the stem 2 cm below the shoot tip. The titer of R. solan-
Figure 3. Transgenic tomato plants contain a single copy of the transgene SlURL1. (A) Schematic diagram of the SlURL1 expression construct driven by a glucocorticoid-inducible pro­moter. LB and RB are left and right border, respectively. 35S: cauliflower mosaic virus 35S promoter. GVG: artificial fusion protein of GAL4 DNA binding domain-VP16 transactivating domain-glucocorticoid receptor domain. E9: terminator of the pea ribulose bisphosphate carboxylase small subunit rbcs-E9. NOS: nopaline synthase promoter. HPTII: hygromycin phos-photransferase. 6xGAL4: six copies of the GAL4 upstream ac­tivating sequence. Nos-T: terminator of nopaline synthase. A3: terminator of pea rbc3-3A. The arrow indicates XhoI restriction site; (B) DNA gel blot analysis of the transgene SlURL1 in trans-genic tomatoes. The genomic DNAs of wild-type (WT) and transgenic tomato lines 1,11, and 16 were digested with XhoI overnight, subjected to DNA gel blot analysis, and hybridized with NPTII dig-labeled probe. The numbers on the left are 1 kb plus DNA ladder (Invitrogen).
80
Botanical Studies, Vol. 53, 2012
Previous reports showed that jar1 mutants and UAR1 -silenced plants contained reduced levels of JA-Ile (Stas-wick and Tiryaki, 2004; Kang et al., 2006; Wang et al., 2007; Suza et al., 2010). Arabidopsis jar1-1 and jar1-3 mutants were mutated in box I and box II motifs, respec­tively, and showed insensitive phenotypes to methyl JA (Staswick et al., 2002). Moreover, both jar1-1 and jar1-11 (a null mutant) appear to have equal levels of JA-Ile in un-wounded and wounded leaves, suggesting that the jar1-1 protein seems to have no conjugating activity (Staswick and Tiryaki, 2004; Suza and Staswick, 2007). In addition, UAR1 overexpression in jar1-1 could restore the JA-Ile to the wild-type level and the JA responses (Staswick and Tiryaki, 2004). All these data support the importance of three adenylate-forming boxes in JAR1 in mediating the conjugation of JA and Ile. At present, five homologous JAR1 proteins were annotated in four species, including Arabidopsis, tomato, tobacco, and rice (Staswick et al., 2002; Kang et al., 2006; Wang et al., 2007; Riemann et al., 2008; Suza et al., 2010). Except for Oryza sativa JAR1, other JAR1 homologs are involved in the generation of
Figure 5. Induction of SlURL1 expression results in delayed de­velopment of bacterial wilt symptoms in tomato. Wild-type and transgenic tomato lines were inoculated with Ralstonia and in­duced by water (A) or 5 [M DEX (B). The percentage of wilted tomato plants under white light was recorded. Arrows indicate the times of Ralstonia inoculation and DEX induction.
acearum isolated from wild-type and transgenic tomatoes with water treatment remained largely the same, increased rapidly at 1 dai and peaked at 3 dai (Figure 6A). Similar results were observed in wild-type plants treated with DEX (Figure 6B). In contrast, transgenic tomatoes treated with DEX showed no R. solanacearum colonization at 1 dai; the titer of R. solanacearum was detected at 2 dai and peaked at 4 dai (Figure 6B). High levels of SlURL1 expres­sion can thus delay the colonization of R. solanacearum in xylem tissues and postpone the development of wilt symp­toms.
discussion
Here, we report on the isolation and characterization of The
tomato genome contains multiple copies of SlURL1 . SlURL1 shows high expression in aerial tissues, including stems, leaves and flowers, especially leaves. We gener­ated transgenic tomato lines containing a DEX-inducible SlURL1 construct, and DEX-induced SlURL1 expression could delay the symptoms of bacterial wilt disease. Thus, JA can interfere with the detrimental damage caused by Ralstonia infection.
Figure 6. Induction of SlURL1 expression results in a reduction of the Ralstonia titer in the stems of transgenic tomatoes. Wild type and transgenic tomato lines were inoculated with Ralstonia and induced by water (A) or 5 μM DEX (B). The titer of bacteria in the stems of wild type and transgenic tomatoes under white light was recorded. Arrows indicate the times of Ralstonia in­oculation and DEX induction.
WANG and HSIEH ― SlJRLl expression delays bacterial wilt
81
JA-Ile. Here, we identified SURL1 as a new homolog of the UAR1 gene in tomato. Three adenylate-forming boxes are conserved among 6 JAR1 homologs (Figure S1). Espe­cially in the Solanaceae family, no residues of adenylate-forming boxes are changed. Nicotiana tabacum JAR4 (NtJAR4) shows 86.9% amino-acid identity with NtJAR6, which suggests that they are involved in JA-Ile synthesis (Kang et al., 2006; Wang et al., 2007). In contrast, SlJRL1 and SlJAR1 share 99.3% identity. Thus, SlJRL1 may have conjugating activity for JA-Ile.
High levels of SlURL1 expression in transgenic toma­toes resulted in a delayed development of wilt symptoms after inoculation with R. solanacearum (Figure 5), which is similar to that observed in previous reports (Yang et al., 2007; Yang et al., 2008). LUAMP1 and LUAMP2 with antimicrobial activities were introduced into tobacco, thus leading to delayed development of wilt symptoms caused by R. solanacearum (Yang et al., 2007; Yang et al., 2008). In addition, the overexpression of Arabidopsis thionin 2.1 (AtTHI2.1) in tomato substantially reduced the population of R. solanacearum in stem tissues and further delayed the development of wilt symptoms (Chan at al., 2005). Simi-larly, overexpression of sweet pepper ferredoxin-I protein reduced colonization and wilt disease (Huang et al., 2007). Thus, ectopic expression of these genes inhibited the intru-sion and growth of bacteria and interfered with the upward movement of bacteria from root to stem, which further diminished the development of wilt symptoms (Wang et al., 2000; Wang and Lin, 2003; Chan at al., 2005). Here, our results suggest a similar mechanism against R. solan-acearum infection conferred by the DEX-induced SlURL1 in transgenic tomatoes, which led to the 1 dai abolishment of R. solanacearum population in transgenic tomatoes (Figure 6B). Transgenic tomatoes harboring high levels of SlURL1 expression can thus delay bacteria wilt disease and reduce the bacteria population in stems.
In addition to being antibacterial proteins, ET and SA are defense phytohormones against R. solanacearum infec­tion in plants (Deslandes et al., 2002; Hirsch et al., 2002; Deslandes et al., 2003; Valls et al., 2006; Qiao et al., 2009; Stepanova and Alonso, 2009). Tomato stress responsive factor 1 (TSRF1 ), an ET response factor, was upregulated by ET, SA, and R. solanacearum infection. TSRF1 protein directly bound to the GCC box, thus activating the expres­sion of PR genes to enhance resistance to R. solanacearum infection in tobacco (Zhang et al., 2004). Similar evidence was found with barley HvRAF and soybean GmERF3. HvRAF and GmERF3 belong to the ERF family, and di­rectly activate several PR genes to confer protection. The expression of HvRAF and GmERF3 was increased by ET and SA, and induced by JA (Jung et al., 2007; Zhang et al., 2009). Hirsch et al. (2002) reported that the jar1-1 mutant without conjugating activities for JA-Ile was more suscep­tible to R. solanacearum infection, implying that JAR1 may be involved in the defense response in Arabidopsis against R. solanacearum infection. The coi1 mutant loss-of-function in tomato resulted in the suppression of JA-
Figure S1. Alignment of SlJRL1-related proteins from differ-ent plant species. The solid lines indicate adenylate-forming domains. AtJAR1: Arabidopsis thaliana jasmonate resistance 1. OsJAR1: Oryza sativa JAR1. NtJAR4: Nicotiana tabacum JAR4. NtJAR6: Nicotiana tabacum JAR6. SlJAR1: Solanum ly-copersicum JAR1. SlJRL1: Solanum lycopersicum JAR1-Like 1.
responsive gene expression, including those involved in JA biosynthesis, signaling, and proteinase inhibitor proteins, leading to greater susceptibility to mites (Li et al., 2004). Moreover, the biocontrol agent Pythium oligandrum (PO) colonized the rhizosphere of tomato to suppress wilt dis­ease caused by R. solanacearum infection, but tomato plants became susceptible in the jai1-1 mutant (mutated in the COI1 gene) treated with PO homogenate (Hase et al., 2008). The PO homogenate also induced the expression of PR-6 via COI1-dependent signaling, but not SA signaling, and further promoted tomato resistance against R. solan-
82
Botanical Studies, Vol. 53, 2012
acearum infection (Hase et al., 2008). Thus, enhancing JA signaling provided protection against R. solanacearum infection. Furthermore, JA-Ile is the active form of the JA hormone and is required for plant defense against insect and pathogen invasion (see reviews in Browse, 2009). Our results indicated that a high SlURL1 expression may pro­mote resistance against R. solanacearum infection (Figures 4-6). Thus, SlURL1 reinforces protection against bacterial wilt disease via JA signaling in tomato.
Of note, our results and other evidence, such as PO-treated plants, show a delay in bacterial wilt disease with infection (Figures 5 and 6; Hase et al., 2008). Compared with our results, PO treatment delayed wilt symptoms for about one day. PO-treated tomato may confer more resis­tance against R. solanacearum infection due to the puri­fied elicitms (POD-1 and POD-2) from the PO cell wall proteins and PO homogenate, which increase ET induce ET- and JA-related PR genes expression to promote pro­tection against pathogens (Takenaka et al., 2003; Takenaka et al., 2006; Takenaka et al., 2008). Moreover, JA and ET synergistically participate in the plant defense responses to many pathogens (for reviews, see Kunkel and Brooks, 2002; Van Loon et al., 2006). However, null or silenced JAR1-homolog mutants still contain certain levels of JA-Ile, suggesting that another component for JA-Ile biosyn­thesis in plants regulating JA signaling may exist (Staswick and Tiryaki, 2004; Suza and Staswick, 2007; Kang et al., 2006; Wang et al., 2007; Suza et al., 2010). This specula-tion is supported by the finding of two isoforms, NaJAR4 and NaJAR6, for the biosynthesis of JA-Ile in Nicotiana attenuata (Wang et al., 2007). In addition, we found mul-tiple copies of SlJAR1 homologs in tomato (Figure 2A). Although SlJRL1 and SlJAR1 share 99% amino acid identity, SlURL1 transcripts are abundant in stems, leaves, and flowers, but much less so in roots (Figure 2C). How-ever, SlUAR1 expression is abundant in roots and opened flowers, but not in leaves (Suza et al., 2010). Both genes exhibit tissue-specific expression patterns, which might be determined by specific transcription factors binding to the cis-elements present in the promoter regions of individual genes. This suggests that SlJRL1 and SlJAR1 may have different functions in JA signaling. In addition, because of the multiple copies of SlUAR1 homologs in tomato, an increase of only SlURL1 expression may not confer enough JA-Ile levels to trigger bacterial wilt disease resistance. Alternatively, SlURL1 induction may restrict the availability of the substrate in the aboveground tissues because SlURL1 showed less expression in roots, leading to delayed and reduced bacterial wilt disease symptoms. Therefore, ex-amining the effect of SlUAR1 induction on the infection or colonization of Ralstonia in tomato is of interest, as are the unique molecular mechanisms in SlUAR1 and SlURL1 that underly bacterial wilt disease resistance in tomato.
Acknowledgements. We thank Chiu-Ping Cheng, Insti­tute of Plant Biology, National Taiwan University, Taiwan, for the kind gift of the bacterial strain Ralstonia solan-acearum. This work was supported by a grant from Coun-
cil of Agriculture Executive Yuan (95Agriscience-6.1.3-crop-Z1(4)-2), Taiwan.
literature cited
Adhikari, T.B. and R.C. Basnyat. 1998. Effect of crop rotation and cultivar resistance on bacterial wilt of tomato in Nepal. Can. J. Plant Pathol. 20: 283-287.
Aoyama, T. and N.-H. Chua. 1997. A glucocorticoid-mediated transcriptional induction system in transgenic plants. Plant J. 11: 605-612.
Browse, J. 2009. Jasmonate Passes Muster: A receptor and tar­gets for the defense hormone. Annu. Rev. Plant Biol. 60: 183-205.
Chan, Y.-L., V. Prasad, Sanjaya, K.H. Chen, P.C. Liu, M.-T. Chan, and C.-P. Cheng. 2005. Transgenic tomato plants expressing an Arabidopsis thionin (Thi2.1) driven by fruit-inactive promoter battle against phytopathogenic attack. Planta 221: 386-393.
Chang, K.-H., H. Xiang, and D. Dunaway-Mariano. 1997. Acyl-adenylate motif of the acyl-adenylate/thioester-forming enzyme superfamily: A site-directed muta-genesis study with the Pseudomonas sp. strain CBS3 4-chlorobenzoate:coenzyme a ligase. Biochemistry 36: 15650-15659.
Chen, I.C., I.C. Huang, M.J. Liu, Z.G. Wang, S.S. Chung, and
H.L. Hsieh. 2007. Glutathione S-transferase interacting with far-red insensitive 219 is involved in phytochrome A-mediated signaling in Arabidopsis. Plant Physiol. 143:
1189-1202.
Chini, A., S. Fonseca, G. Fernandez, B. Adie, J.M. Chico, O. Lorenzo, G. Garcia-Casado, I. Lopez-Vidriero, F.M. Loza-
no, M.R. Ponce, J.L. Micol, and R. Solano. 2007. The JAZ
family of repressors is the missing link in jasmonate signal­ling. Nature 448: 666-671.
Deslandes, L., J. Olivier, F. Theulieres, J. Hirsch, D.X. Feng, P. Bittner-Eddy, J. Beynon, and Y. Marco. 2002. Resistance to Ralstonia solanacearum in Arabidopsis thaliana is con­ferred by the recessive RRS1-R gene, a member of a novel family of resistance genes. Proc. Natl. Acad. Sci. USA 99: 2404-2409.
Deslandes, L., J. Olivier, N. Peeters, D.X. Feng, M. Khoun-lotham, C. Boucher, I. Somssich, S. Genin, and Y. Marco. 2003. Physical interaction between RRS1-R, a protein conferring resistance to bacterial wilt, and PopP2, a type III effector targeted to the plant nucleus. Proc. Natl. Acad. Sci. USA 100: 8024-8029.
Dombrecht, B., G.P. Xue, S.J. Sprague, J.A. Kirkegaard, J.J.
Ross, J.B. Reid, G.P. Fitt, N. Sewelam, P.M. Schenk, J.M.
Manners, and K. Kazan. 2007. MYC2 differentially modu­lates diverse jasmonate-dependent functions in Arabidopsis. Plant Cell 19: 2225-2245.
Genin, S. and C. Boucher. 2004. Lessons learned from the ge­nome analysis of Ralstonia solanacearum. Annu. Rev. Phy-topathol. 42: 107-134.
WANG and HSIEH ― SlJRLl expression delays bacterial wilt
83
Guo, J.-H., H.-Y. Qi, Y.-H. Guo, H.-L. Ge, L.-Y. Gong, L.-X. Zhang, and P.-H. Sun. 2004. Biocontrol of tomato wilt by plant growth-promoting rhizobacteria. Biol. Control 29: 66­72.
Hase, S., S. Takahashi, S. Takenaka, K. Nakaho, T. Arie, S. Seo, Y. Ohashi, and H. Takahashi. 2008. Involvement of jasmon-ic acid signalling in bacterial wilt disease resistance induced by biocontrol agent Pythium oligandrum in tomato. Plant Pathol. 57: 870-876.
Hayward, A.C. 1991. Biology and epidemiology of bacterial wilt caused by Pseudomonas solanacearum. Annu. Rev. Phyto-
pathol. 29: 65-87.
Hirsch, J., L. Deslandes, D.X. Feng, C. Balague, and Y. Marco. 2002. Delayed symptom development in ein2-1, an Arabi-dopsis ethylene-insensitive mutant, in response to bacterial wilt caused by Ralstonia solanacearum. Phytopathology
92: 1142-1148.
Hsieh, H.-L., C.-G. Tong, C. Thomas, and S.J. Roux. 1996.
Light-modulated abundance of an mRNA encoding a calmodulin-regulated, chromatin-associated NTPase in pea.
Plant Mol. Biol. 30: 135-147.
Hsieh, H.-L., H. Okamoto, M. Wang, L.-H. Ang, M. Matsui, H.
Goodman, and X.W. Deng. 2000. FIN219, an auxin-regu­lated gene, defines a link between phytochrome A and the downstream regulator COP1 in light control of Arabidopsis
development. Gene Dev. 14: 1958-1970.
Huang, H.-E., C.-A. Liu, M.-J. Lee, C.-G. Kuo, H.-M. Chen, M.-J. Ger, Y.-C. Tsai, Y.-R. Chen, M.-K. Lin, and T.-Y.
Feng. 2007. Resistance enhancement of transgenic tomato to bacterial pathogens by the heterologous expression of sweet pepper ferredoxin-I protein. Phytopathology 97: 900­906.
Jung, J., S. Won, S. Suh, H. Kim, R. Wing, Y. Jeong, I. Hwang,
and M. Kim. 2007. The barley ERF-type transcription fac­tor HvRAF confers enhanced pathogen resistance and salt tolerance in Arabidopsis. Planta 225: 575-588.
Kang, J.-H., L. Wang, A. Giri, and I.T. Baldwin. 2006. Silenc­ing threonine deaminase and JAR4 in nicotiana attenuata impairs jasmonic acid-isoleucine-mediated defenses against
manduca sexta. Plant Cell 18: 3303-3320.
Katsir, L., H.S. Chung, A.J.K. Koo, and G.A. Howe. 2008. Jas-
monate signaling: a conserved mechanism of hormone sens­ing. Curr. Opin. Plant Biol. 11: 428-435.
Kunkel, B.N. and D.M. Brooks. 2002. Cross talk between sig­naling pathways in pathogen defense. Curr. Opin. Plant
Biol. 5: 325-331.
Li, L., Y. Zhao, B.C. McCaig, B.A. Wingerd, J. Wang, M.E.
Whalon, E. Pichersky, and G.A. Howe. 2004. The tomato
homolog of CORONATINE-INSENSITIVE1 is required for
the maternal control of seed maturation, jasmonate-signaled defense responses, and glandular trichome development.
Plant Cell 16: 126-143.
Lin, Y.-M., I.-C. Chou, J.-F. Wang, F.-I. Ho, Y.-J. Chu, P.-C. Huang, D.-K. Lu, H.-L. Shen, M. Elbaz, S.-M. Huang, and
C.-P. Cheng. 2008. Transposon mutagenesis reveals differ-
ential pathogenesis of ralstonia solanacearum on tomato and
Arabidopsis. Mol. Plant-Microbe Interact. 21: 1261-1270.
Qiao, H., K.N. Chang, J. Yazaki, and J.R. Ecker. 2009. Interplay
between ethylene, ETP1/ETP2 F-box proteins, and degra­dation of EIN2 triggers ethylene responses in Arabidopsis. Gene Dev. 23: 512-521.
Riemann, M., M. Riemann, and M. Takano. 2008. Rice JAS-MONATE RESISTANT 1 is involved in phytochrome and
jasmonate signalling. Plant Cell Environ. 31: 783-792.
Schell, M.A. 2000. Control of virulence and pathogenicity genes of ralstonia solanacearum by an elaborate sensory network.
Annu. Rev. Phytopathol. 38: 263-292.
Staswick, P.E. and I. Tiryaki. 2004. The oxylipin signal jasmonic acid is activated by an enzyme that conjugates it to isoleu-cine in Arabidopsis. Plant Cell 16: 2117-2127.
Staswick, P.E., I. Tiryaki, and M.L. Rowe. 2002. Jasmonate re­sponse locus JAR1 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.
Stepanova, A.N. and J.M. Alonso. 2009. Ethylene signaling and response: where different regulatory modules meet. Curr.
Opin. Plant Biol. 12: 548-555.
Suza, W. and P. Staswick. 2008. The role of JAR1 in jasmonoyl-l-isoleucine production during Arabidopsis wound response
Planta 227: 1221-1232.
Suza, W., M. Rowe, M. Hamberg, and P. Staswick. 2010. A to­mato enzyme synthesizes (+)-7-iso-jasmonoyl-l-isoleucine in wounded leaves. Planta 231: 717-728.
Takenaka, S., Z. Nishio, and Y. Nakamura. 2003. Induction of defense reactions in sugar beet and wheat by treatment with cell wall protein fractions from the mycoparasite pythium oligandrum. Phytopathology 93: 1228-1232.
Takenaka, S., Y. Nakamura, T. Kono, H. Sekiguchi, A. Masu-naka, and H. Takahashi. 2006. Novel elicitin-like proteins isolated from the cell wall of the biocontrol agent Pythium oligandrum induce defence-related genes in sugar beet.
Mol. Plant Pathol. 7: 325-339.
Thines, B., L. Katsir, M. Melotto, Y. Niu, A. Mandaokar, G. Liu, K. Nomura, S.Y. He, G.A. Howe, and J. Browse. 2007. JAZ
repressor proteins are targets of the SCFCOI1 complex during jasmonate signalling. Nature 448: 661-665.
Thoquet, P., J. Olivier, C. Sperisen, P. Rogowsky, H. Laterrot, and N. Grimsley. 1996. Quantitative trait loci determining resistance to bacterial wilt in tomato cultivar hawaii7996.
Mol. Plant-Microbe Interact. 9: 826-836.
Valls, M., S. Genin, and C. Boucher. 2006. Integrated regulation of the type III secretion system and other virulence determi­nants in Ralstonia solanacearum. PLoS Pathog. 2: 798-807.
Van Loon, L.C., M. Rep, and C.M.J. Pieterse. 2006. Significance of inducible defense-related proteins in infected plants.
Annu. Rev. Phytopathol. 44: 135-162.
Wang, J.-F., J. Olivier, P. Thoquet, B. Mangin, L. Sauviac, and N.H. Grimsley. 2000. Resistance of tomato line Hawaii7996
84
Botanical Studies, Vol. 53, 2012
to Ralstonia solanacearum pss4 in Taiwan is controlled mainly by a major strain-specific locus. Mol. Plant-Microbe Interact. 13: 6-13.
Wang, L., R. Halitschke, J.-H. Kang, A. Berg, F. Harnisch, and I. Baldwin. 2007. Independently silencing two JAR family members impairs levels of trypsin proteinase inhibitors but
not nicotine. Planta 226: 159-167.
Xue, Q.-Y., Y. Chen, S.-M. Li, L.-F. Chen, G.-C. Ding, D.-W.
Guo, and J.-H. Guo. 2009. Evaluation of the strains of Acinetobacter and Enterobacter as potential biocontrol agents against Ralstonia wilt of tomato. Biol. Control 48:
252-258.
Yan, J., C. Zhang, M. Gu, Z. Bai, W. Zhang, T. Qi, Z. Cheng, W. Peng, H. Luo, F. Nan, Z. Wang, and D. Xie. 2009. The Arabidopsis CORONATINE INSENSITIVE1 protein is a
jasmonate receptor. Plant Cell 21: 2220-2236.
Yang, X., Y. Xiao, X. Wang, and Y. Pei. 2007. Expression of a novel small antimicrobial protein from the seeds of Moth-erwort (Leonurus japonicus) confers disease resistance in tobacco. Appl. Environ. Microb. 73: 939-946.
Yang, X., X. Wang, X. Li, B. Zhang, Y. Xiao, D. Li, C. Xie, and Y. Pei. 2008. Characterization and expression of an nsLTPs-like antimicrobial protein gene from motherwor. Plant Cell
Rep. 27: 759-766.
Zhang, G., M. Chen, L. Li, Z. Xu, X. Chen, J. Guo, and Y. Ma.
2009. Overexpression of the soybean GmERF3 gene, an AP2/ERF type transcription factor for increased tolerances to salt, drought, and diseases in transgenic tobacco. J. Exp.
Bot. 60: 3781-3796.
Zhang, H., D. Zhang, J. Chen, Y. Yang, Z. Huang, D. Huang, X.-C. Wang, and R. Huang. 2004. Tomato stress-responsive factor TSRF1 interacts with ethylene responsive element GCC box and regulates pathogen resistance to Ralstonia solanacearum. Plant Mol. Biol. 55: 825-834.
轉殖番茄中Uasmonate-Resistant 1-like 1 (SURL1)基因的誘導
表現能夠延遲番茄青枯菌的菌落
王志恭 謝旭亮
國立台灣大學植物科學研究所
番茄青枯病是一種非常普遍和廣泛的疾病,主要是由青枯菌引起,特別是在炎熱和潮濕的環境下
更爲嚴重;於自然環境中,因爲寄主抗性的不穩定,以及青枯菌種類與變異性極大,至目前爲止,並
無良好的防治方式;因此,番茄抵禦青枯菌的分子機制的瞭解仍然有限。本篇報導由番茄釣取到番茄
SUAR1同源性基因—SURL1 '並且建立可誘導SlURL1基因表現的番茄轉殖株。比對兩基因之蛋白質序
列竟然高達
99 %相同性;先前發現SUAR1於地下根部之基因表現量最高,相反地,SURL1於地下部表
現量最低,反而是地上部有較高的基因表現量;進一歩利用南方氏默點法分析,發現於
CL5915番茄品
系之基因體含有多個
SURL1基因。進一歩建構誘導啓動子驅動SlURL1之載體,將其轉殖進番茄基因體
中,
利用DEX處理轉殖株,其誘導SlURL1表現量於第二天即達到高峰;並利用此誘導方式,進行青枯
菌感染,經
DEX誘導之轉殖番茄可延緩萎凋病徵發生;並分析體內之青枯菌族群,發現轉殖番茄經由
DEX誘導後,亦可減緩於莖頂處青枯菌累積之數量;由此可知,植物茉莉酸荷爾蒙具有參與番茄防禦
青枯菌感染之功能。
因此'本篇SURL1轉殖番茄之結果,可提供抵抗番茄青枯病之分子機制的瞭解。
關鍵詞:番茄青枯病;青枯菌;番茄SUAR1同源性基因- SURL1 ;茉莉酸荷爾蒙;番茄轉殖株。