. .3... 4.5.5....3 , 39:53.... 2,, R53». , out): 1.... a .9- i _ . , . m ,... it: . . £59.: a: . mi . ”s f. i 3.“ «.6»? ‘ .II I :r . ? III a JP ‘ 9,! n 1‘... 0* sfzfiaws‘w m: . ... I 151.133.. .3... i 4...! . 11..."... .3. K3..- 3., .3 333.55%. é- : : “swab: Y... inauguwu. «GMrliuir I 0., i. “was: 4 wt!- Ito-if! 5?!th 0.1.! I. . . . 533. . it... o i. 3...}. Q 3 it .1. slat}; iii: t. L . 5.1 .. I. «Full-ii!!! it 1) .1... ”Hill! :Jxajéfigfifi . A . tut-3:3 .. 1) T5 f This is to certify that the dissertation entitled innate Immunity in Arabidopsis thaliana: Induction and Suppression by Pseudomonas syringae I presented by l ;. William Robert Underwood has been accepted towards fulfillment of the requirements for the Ph.D. degree in Genetics 7" \s -rfiém WW Major Professor’s Signature 13-49-06» Date MSU is an Affinnative Action/Equal Opportunity Institution LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2/05 p:/ClRC/Date0ue.indd~p.1 .7. p INNATE IMMUNITY IN ARABIDOPSIS T HALIANA: INDUCTION AND SUPPRESSION BY PSEUDOMONAS S YRINGAE By William Robert Underwood A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Genetics Program 2006 ABSTRACT INNATE IMMUNITY IN ARABIDOPSIS T HALIANA : INDUCTION AND SUPPRESSION BY PSEUDOMONAS SYRINGAE By William Robert Underwood Innate immunity refers to the ability of eukaryotic organisms to detect the presence of potentially harmful microbes through recognition of highly conserved surface molecules, referred to as pathogen-associated molecular patterns (PAMPs), and to activate defense responses to prevent infection. Like mammals and insects, plants can perceive microbial PAMPs and activate defenses. Microbial pathogens such as bacteria cause significant agricultural losses each year and chemical treatments aimed at abating these losses represent a substantial input cost for agriculture and are damaging to the environment. Therefore, new strategies to control crop losses due to pathogens without the use of chemicals will be highly desirable. To develop such strategies, a thorough understanding of plant defenses and the mechanisms by which pathogens overcome them will be necessary. My research has contributed to our knowledge of both innate immunity in plants, and the mechanisms used by pathogenic bacteria to overcome innate immune responses. I performed genome-wide microarray analyses in collaboration with Dr. Roger Thilmony to determine global gene expression changes associated with the activation of innate immunity in Arabidopsis thaliana. This study provided the first look at the gene expression changes induced by perception of PAMPs present on live bacteria. We also investigated the specific effects of the Pseudomonas syringae pv. tomato DCBOOO (Pst DC3000) virulence factors coronatine (COR) and the type III secretion system (T'TSS) on modulating gene expression in Arabidopsis. My work also contributed to the discovery of bacterium-induced stomatal closure as a novel component of innate immunity in plants. This research was performed in collaboration with Dr. Maeli Melotto and revealed that the guard cells of stomata, pores on the surface of the plant leaf, can sense bacterial PAMPs and close stomata to restrict entry of bacteria into the leaf tissue. Prior to this discovery, stomata were generally regarded as passive ports of bacterial entry. We found that bacterium-induced closure of stomata required the plant hormone abscisic acid (ABA) and an intact ABA signaling pathway. Additionally, we discovered that the phytotoxin COR is required for Pst DC3000 to open stomata, explaining the primary virulence function of COR. Successful phytopathogens such as Pst DC3000 can overcome innate immunity and infect plants. The TTSS and effector proteins injected into the plant cells are crucial for the bacteria to overcome defenses and cause disease. I investigated the function of a single 'ITSS effector, the protein tyrosine phosphatase HopD2, by creating transgenic Arabidopsis lines expressing HopD2 or the catalytically-inactive HopD2cms. I found that HopD2 blocks defense responses associated with PAMP-induced innate immunity, but not the HR. This activity was dependent on an intact phosphatase catalytic site and, interestingly, HopD2C3788 had a dominant-negative effect on the function of the wild-type HopD2. Global gene expression profiling revealed that HopD2 does not affect gene expression levels in Arabidopsis, suggesting that HopD2 acts at a late stage, downstream or independent of PAMP-induced signaling and gene expression changes, to block activation of innate immunity. C0pyright by William Underwood 2006 ACKNOWLEDGEMENTS I would first like to thank my advisor, Dr. Sheng Yang He for the opportunity to perform my doctoral studies in his laboratory. Sheng Yang has been an excellent teacher and mentor and I have benefited greatly from his advice and support. Working with Sheng Yang has been a really wonderful learning experience. I would like to thank the members of my guidance committee: Dr. Rebecca Grumet, Dr. Jonathan Walton, and Dr. Mike Thomashow, for their support and advice throughout my progress as a graduate student. Their suggestions and insights have been very helpful. I would also like to thank all of my colleagues, past and present, in the He lab: Elena Bray-Speth, Wei-Ning Huang, Christy Mecey-Smith, Maeli Melotto, Francisco Uribe, Weiqing Zeng, Kinya Nomura, Young Nam Lee, Lori Imboden, Roger Thilmony, Paula Hauck, Julie Zwiesler-Vollick, Sruti DebRoy, Qiaoling J in, Sarah Sarkar, Shuocheng Zhang,Ola Kolade, Yong Hoon Lee, Yong Bum Kwak, and Anne Plovanich- Jones. Our lab has always been a very positive work environment and I have been lucky to have great interactions with my colleagues. I would particularly like to thank Maeli Melotto and Roger Thilmony with whom I worked closely on the projects described in chapters 2 and 3. I really enjoyed collaborating with both Maeli and Roger and learned quite a lot from both of them. In addition to an great lab environment, the PRL has also been an excellent environment with many wonderful colleagues and friends. I would like to all of the members of the PRL for providing such an intellectually stimulating environment and for the collaborative and interactive spirit that is a hallmark of this institution. Finally, I would like to thank my parents for their constant love and support and all of my fi'iends in East Lansing. Especially my girlfriend Julia Letoutchaia for understanding my sometimes long and ofien erratic work hours and my friends Harrie van Brp, Hiroshi Maeda, Neil Adhikari, and the many other fi'iends who have attended our frequent graduate student “therapy” sessions at the Peanut Barrel. Without the support of my friends and family this would not have been possible. vi TABLE OF CONTENTS List of Tables .......................................................................................... x List of Figures ........................................................................................ xi Chapter 1 Introduction and Literature Review ................................................................ 1 Introduction ................................................................................... 2 Life cycles and virulence strategies of phytopathogenic bacteria ..................... 3 The Arabidopsis thaliana-Pseudomonas syringae model for studying plant disease ......................................................................................... 5 The bacterial pathogen Pseudomonas syringae .......................................... 6 Bacterial type III secretion .................................................................. 7 Type III effector proteins .......................................................... 9 Virulence functions of type III effectors ....................................... 11 Phytotoxins of plant pathogenic bacteria ................................................ 14 Plant defense responses .................................................................... 16 Plant basal defenses ............................................................... 17 Gene-for-gene resistance ......................................................... 21 Project summary and rationale ............................................................ 23 References ................................................................................... 27 Chapter 2 Genome-wide transcriptional analysis of the Arabidopsis thaliana interaction with the plant pathogen Pseudomonas syringae pv. tomato DC3000 and the human pathogen Escherichia coli 01571H7 .......................................................................... 36 Abstract ...................................................................................... 37 Introduction ................................................................................. 38 Experimental Procedures .................................................................. 41 Plant growth, bacterial strains and bacteria enumeration .................... 41 RNA isolation and labeling ...................................................... 42 Affymetrix GeneChip hybridization and data collection ..................... 43 Affymetrix GeneChip data analysis ............................................. 43 RNA isolation and blot analysis ................................................ 46 Results and Discussion .................................................................... 47 Rationale and experimental design ............................................. 47 PAMP-induced transcriptional changes ........................................ 50 Hierarchical cluster analysis or PAMP-induced genes ....................... 51 Transcription factors and signaling components are induced in response to PAMPs .............................................................................. 54 PAMPs induced genes that may be involved in secretion and cell wall modification ........................................................................ 56 vii F lagellin does not contribute uniquely to PAMP-mediated transcriptional changes ............................................................................. 61 Analysis of the TTSS- and COR-regulated genes ............................ 63 Impact of Pst pathogenesis on host metabolism .............................. 67 Coronatine activates J A responses and secondary metabolism ............. 71 Pst induces ABA/abiotic stress response genes ............................... 76 Defense response-related gene expression ..................................... 76 Expression of auxin- and cytokinin-related genes is altered during Pst infection ............................................................................ 78 Conclusion .................................................................................. 80 Acknowledgements ........................................................................ 81 References ................................................................................... 83 Chapter 3 A novel function for plant stomata in innate immunity against bacteria ..................... 94 Abstract ...................................................................................... 95 Introduction ................................................................................. 96 Experimental Procedures .................................................................. 98 Plant Material ...................................................................... 98 Chemicals ........................................................................... 99 Assessment of response of stomata to treatments ............................. 99 Detection of nitric oxide in guard cells ....................................... 100 Bacterial growth assay .......................................................... 100 Results ...................................................................................... 102 Plant and human pathogenic bacteria induce closure of stomata ......... 102 Involvement of PAMPs in triggering stomatal closure ..................... 106 ABA synthesis and signaling components are required for bacteria- and PAMP-induced stomatal closure ............................................... 108 Pst DC3000 overcomes stomatal closure by production of the phytotoxin coronatine ......................................................................... 1 13 Biological relevance of stomatal closure in bacterial infection ............ 118 Discussion ................................................................................. 122 References ................................................................................. 128 Chapter 4 Characterization of the virulence function of HopD2, a type III effector tyrosine phosphatase of Pst DC3000 that suppresses basal defenses in Arabidopsis thaliana. . .. 131 Abstract .................................................................................... 132 Introduction ................................................................................ l 3 3 Experimental Procedures ................................................................ 136 Bacterial strains and media ..................................................... 136 Plant growth and bacteria enumeration ....................................... 137 Generation of transgenic plants ................................................ 138 Site-directed mutagenesis of HopD2 .......................................... 138 Production of antibody ........ ' .................................................. 139 DEX and estradiol induction of transgene expression ...................... 140 viii Microarray analysis .............................................................. 140 Lipid profiling analysis ......................................................... 142 Callose assay ..................................................................... 143 In-gel kinase assays .............................................................. 144 Immunoblotting and HopD2 protein fractionation .......................... 145 Results ...................................................................................... 147 Arabidopsis thaliana transgenic plants expressing HopD2 exhibit chlorosis and necrosis upon induction ........................................ 147 HopD2 transgenic plants allow enhanced multiplication of the Pst DC3000 hrpA‘ mutant ........................................................... 150 HopD2 transgenic plants show reduced callose deposition in response to hrpA' .............................................................................. 1 53 Development of the HR is not effected in HopD2 transgenic plants. . ...157 Activation of the Arabidopsis MAPKs AtMPK3 and AtMPK6 is enhanced in HopD2 transgenic plants .................................................... 159 HopD2 is a non-transmembrane protein tyrosine phosphatase and is present in both the soluble and membrane fractions of Arabidopsis thaliana protein extracts ........................................................ 162 Microarray analysis of HopD2 function using transgenic plants and bacterial mutants ................................................................. 165 Lipid profiling analysis of HopD2 transgenic plants ........................ 169 Discussion ................................................................................. 172 Acknowledgements ....................................................................... 176 References ................................................................................. 1 77 Chapter 5 Conclusions and Future Perspectives ............................................................ 183 References ................................................................................. 192 Appendix A Supplementary material for Chapter 2 .......................................................... 194 ix LIST OF TABLES Table 2-1. Arabidopsis genes reproducibly regulated by bacterial inoculation ............. 49 Table 2-2. Selected PAMP-induced Arabidopsis genes ....................................... 59 Table 4-1. Microarray analysis of HopD2 transgenic plants ................................. 167 Table 4-2. Genes differentially regulated by transgenic expression of HopD2 and HopD2C3788 ......................................................................................... 168 Table A-1. Known and predicted kinases induced by PAMP perception .................. 196 LIST OF FIGURES Images in this dissertation are presented in color. Fig. 2-1. Expression profile of 347 PAMP-induced Arabidopsis genes ..................... 53 Fig. 2-2. Expression profile of the 2439 Arabidopsis genes reproducibly regulated by 1 x106 and/or 5x107 Pst bacteria/mL ............................................................... 66 Fig. 2-3. COR toxin and type III effectors target different aspects of plant metabolism..69 Fig. 2-4. Pst coordinately regulates hormone and stress response genes .................... 75 Fig. 3-1. Pst DC3000 triggers closure of Arabidopsis stomata .............................. 104 Fig. 3-2. Arabidopsis responses to E. coli 0157:H7 .......................................... 105 Fig. 3-3. PAMPs and E. coli trigger stomatal closure ........................................ 107 Fig. 3-4. Stomatal responses in Arabidopsis ABA mutants .................................. 110 Fig. 3-5. Stomatal responses of Arabidopsis ABA response mutants 0st], gorkI, and rbohD/F ............................................................................................. 1 l 1 Fig. 3-6. Involvement of nitric oxide (N O) in PAMP-induced stomatal closure ......... 112 Fig. 3-7. Stomatal responses to cor' (DC3118) and hrcC mutants of Pst DC3000 ....... 115 Fig. 3-8. Purified COR blocks both ABA- and PAMP-induced closure of stomata......116 Fig. 3-9. COR does not effect NO production in guard cells in response to ABA or PAMPs .............................................................................................. 1 17 Fig. 3-10. Movement of GFP-labeled Pst DC3000 and cor’ Pst DC3118 through stomata of Arabidopsis Col-0 epidermal peels ........................................................... 120 Fig. 3-11. Multiplication of Pst DC3118 in wild-type Arabidopsis and stomatal closure mutants .............................................................................................. 121 Fig. 4-1. Leaf phenotypes of transgenic Arabidopsis expressing HopD2 .................. 149 Fig 4-2. Bacterial multiplication in HopD2 transgenic plants ............................... 152 Fig 4-3. Papilla-associated callose deposition in HopD2 transgenic plants ............... 155 xi Fig. 4-4. Transgenic expression of HopD2 blocks flg22-induced resistance to Pst DC3000 ............................................................................................. 156 Fig. 4-5. Onset of HR in C01 g1] and HopD2 transgenic plants ........................ 158 Fig. 4-6. Effect of HopD2 on activation of Arabidopsis MAPKs AtMPK3 and AtMPK6 ............................................................................................. 161 Fig. 4-7. Analysis of HopD2 localization and alignment of catalytic domain ............ 164 Fig. 4-8. Phosphoinositide (PI) lipid content in HopD2 transgenic plants ................. 171 Fig. A-l. RNA blot analysis of differentially expressed genes identified by microarray analysis .............................................................................................. 198 Fig. A-2. RNA blot analysis of the PAMP-responsive gene F lagellin-Induced Receptor Kinase 1 (FRKI) ................................................................................... 199 Fig. A-3. In planta multiplication of bacterial flagellin mutants ............................ 200 Fig. A-4. COR toxin induces genes involved in secondary metabolism and TTSS effectors repress genes involved in photosynthesis ........................................... 201 Fig. A-5. Pst regulates the gene expression of enzymes in the tryptophan, glucosinolate and related biosynthetic pathways ............................................................... 203 Fig. A-6. Bacterial inoculation alters auxin-related gene expression ....................... 205 Fig. A-7. Pst alters host cytokinin—related gene expression .................................. 208 xii Chapter 1 Introduction and Literature Review Phytopathogenic bacterial virulence and plant defense Introduction Plants constitute the primary source of energy and food for most terrestrial species, including humans. By converting the energy of the sun into chemical energy stored in the bonds of carbohydrates through photosynthesis, plants provide a readily exploitable source of energy for animals, insects, and other organisms including pathogenic microbes. Collectively, pathogens such as bacteria, fungi, nematodes, protozoa, and viruses infect essentially all species of land plants and, of particular importance, all varieties of crop plants. Crop loss due to disease represents a serious threat to global security of food, fiber, and animal feed (Strange and Scott, 2005). The percentage of crop losses worldwide caused by plant pathogens, insects, and weeds was estimated to be 31-42% of potential production capacity and to account for at least $500 billion in damages (Agrios, 2005; Oerke et al., 1994). In the United States, crop losses due to plant pathogens was estimated at $9.1 billion while worldwide losses as a result of plant disease reduce overall productivity by 12% (Food and Agriculture Organization, 1993) Set against a backdrop of a steadily increasing human population and stagnating increases in agricultural production, it is clear that novel, cost effective, and environmentally sound strategies aimed at reducing crop losses to pathogens will be needed. Current efforts to control pathogens and pests include widespread use of chemical sprays and pesticides. Large-scale use of such chemicals often results in environmental damage such as soil and groundwater contamination. In addition, chemical sprays and pesticides represent a large input cost for agricultural production. Estimated worldwide costs for pesticide applications totaled $26 billion annually (Food and Agriculture Organization, 1993). Traditional breeding efforts aimed at generating resistant crop varieties have resulted in varying levels of success. Agricultural practices of planting crops in large, genetically uniform plots place selective pressures on pathogens to overcome resistance conferred by single genes and a lack of basic understanding of the factors that contribute to durable, long-term resistance continues to impair efforts to generate crop varieties that display long-lasting pathogen resistance. Diseases caused by bacterial pathogens are difficult to control and can sometimes result in severe crop losses. Control of bacterial diseases often involves foliar sprays to control bacterial populations or use of insecticides to kill insects that act as vectors to spread bacteria from plant to plant. Gram-negative bacteria from the genera Pseudomonas, Ralstom'a, Agrobacterium, Xanthomonas, and Erwinia, and gram-positive bacteria from the genera Clavibacter and Streptomyces are the most common causal agents of bacterial disease in plants. Collectively, these bacteria infect a wide variety of plants and cause numerous types of diseases including leaf specks, wilts, scabs, blights, cankers, and rots. Examples of severe and agronomically important bacterial diseases of plants include fire blight of apples and pears caused by Erwinia amylovora, leaf blight of rice caused by Xanthomonas oryzae, and vvilts and rots of numerous crop plants caused by Ralstonia solanacearum. Life cycles and virulence strategies of phytopathogenic bacteria Outbreak of bacterial disease in crop plants is often highly variable from year to year depending on environmental conditions such as temperature, rainfall, insect populations, and other factors. Most phytopathogenic bacteria survive through winters in soil, crop residues and debris, seeds, perennial weeds, stem cankers, and sometimes even insect hosts. Infections and outbreaks are subsequently promoted by favorable environmental conditions and spread from plant to plant by rain drops, oozes, or insect vectors. Although different phytopathogenic bacteria employ a broad range of mechanisms for survival and spread, the disease cycles of most plant pathogenic bacteria are very similar and can be summarized by the following stages: entry into host tissue; suppression of host defenses and acquisition of nutrients to allow establishment of infection; aggressive multiplication within host intercellular spaces; release of bacteria from infected tissue and subsequent further infections of uninfected tissue from the same host plant or spread to other uninfected plants. One common characteristic shared by most phytopathogic bacteria is that, unlike many mammalian pathogens, they infect host tissue and complete the disease cycle without entering plant cells. Interestingly, the opportunistic bacterial pathogen Pseudomonas aeruginosa strain PA14 was observed to infect Arabidopsis thaliana and to actually perforate mesophyll cell walls and enter mesophyll cells (Plotnikova et al., 2000). More commonly, however, phytopathogenic bacteria remain in the intercellular spaces of plant tissues and must employ strategies to exploit host resources without actually entering host cells. Common strategies employed by bacteria to exploit plant resources include the secretion of toxins and injection of bacterial proteins into plant cells. These strategies likely promote the release of water and/or nutrients from the host cells into the apoplast, which is generally regarded as a harsh, nutrient poor environment that may not support aggressive bacterial multiplication. Additionally, some bacteria such as Ralstonia solanacearum and Erwim'a carotovora secrete hydrolytic enzymes that degrade the polysaccharides present in the plant cell walls (Huang and Allen, 2000; Reid and Collrner; 1986). Degradation products of the cell wall may provide nutrients for the bacteria and promote multiplication and such hydrolytic degradation of the cell wall is generally associated with rot and wilt diseases. In the following pages of this chapter, I will focus specifically on virulence mechanisms of the model bacterial plant pathogen Pseudomonas syringae. The Arabidopsis thaliana-Pseudomonas syringae model for studying plant disease During the past 15 years, a model system consisting of the bacterial plant pathogen Pseudomonas syringae pathovar tomato and the model host plant Arabidopsis thaliana has emerged as a powerful tool to facilitate inquiry into the molecular mechanisms underlying plant disease resistance and bacterial virulence. Arabidopsis thaliana is a small, dicotyledonous member of the Brassicaceae family and is the most commonly used model organism for the study of plant molecular biology. A large abundance of resources have been developed for the study of the biology of Arabidopsis including a completed genome sequence and a collection of T-DNA insertional mutant lines. With the emergence of Arabidopsis as a model system for the study of plant biology, Brian Staskawicz and colleagues sought to identify a bacterial pathogen that could infect Arabidopsis and potentially serve as a model for the study of bacterial diseases of plants. Pseudomonas syringae pathovar tomato (Pst), a bacterial pathogen of tomato, was found to infect Arabidopsis and one strain of Pst, strain DC3000, was found to cause severe disease symptoms (Whalen et al., 1991). The bacterial pathogen Pseudomonas syringae Pseudomonas syringae is a gram-negative bacterium belonging to the family Pseudomonadaceae. This family includes the genera Pseudomonas, Burkholderia, and Ralstonia and belongs to the gamma subgroup of proteobacteria, which includes other pathogenic bacteria including enteric mammalian pathogens such as Yersinia, Shigella, Escherichia, and Salmonella and other plant pathogens including Xanthomonas, Pantoea, and Erwinia. Pseudomonads are ubiquitous in the environment and inhabit a wide variety of niches. Pseudomonas syringae strains collectively infect a wide variety of plants and are classified as pathovars based on their host plant(s). Individual pathovars generally have a very limited host range with the ability to infect only one or a few plant species and often only specific ecotypes of a given plant species. At least 48 pathovars of Pseudomonas syringae have been identified based on host range and DNA analysis (Gardan et al., 1999). Pst DC3000 can infect both Arabidopsis and tomato and causes leaf speck disease on these hosts. Prior to infection, Pseudomonas syringae can generally live epiphytically on leaf surfaces. Pst enters leaf tissues through stomata, the surface apertures that allow plants to conduct gas exchange and regulate water relations, and through accidental openings in the plant cuticle such as wounds. Once inside the leaf intercellular space, Pst multiplies aggressively, attaining populations as high as 107-108 bacteria per square centimeter of leaf tissue in only 3 to 4 days, and causes disease symptoms characterized by necrotic lesions surrounded by chlorotic halos (Hirano and Upper, 2000). At least two well-characterized virulence factors are known to contribute to Pst pathogenicity and development of disease symptoms. These are the phytotoxin coronatine and the type III secretion system and battery of type III effector proteins. Bacterial type III secretion Type III secretion apparatuses are employed by a number of gram-negative bacteria to deliver proteins from the bacteria directly into the cytoplasm of the eukaryotic host cell. Type III secretion systems (TTSSs) are often key virulence determinants for pathogenic bacteria of both animals and plants and have been found in enteric mammalian pathogens such as Yersinia, Escherichia, Salmonella, and Shigella and in plant pathogens including Pseudomonas, Ralstonia, Xanthomonas, Erwinia, and others. In addition to pathogenic species, TTSSs are also found in plant and insect symbiotic bacteria which use the TTSS to interact with their hosts (Dale et al., 2001; Marie et al., 2001) In Pseudomonas, components of the TTSS are encoded by hrp (for hypersensitive response and pathogenicity) and hrc (for hrp gene conserved) genes. The hrp genes were originally identified as being required for bacteria to elicit the hypersensitive response (HR, a plant defense response) in resistant plant varieties and to cause disease on susceptible plants (Lindgren et al., 1986). It was subsequently discovered that the hrp genes play a role in secretion when they were found to be involved in secretion of bacterial proteins in culture (Arlat et al., 1994; He et al., 1993; Wei et al., 1993). The type III secretion apparatus is composed of a basal body that facilitates transport of proteins across the inner and outer bacterial envelopes, an extracellular needle or pilus structure that likely acts as a conduit for protein transport into the host cytosol, and accessory proteins that may be involved in translocation of proteins through the host membrane by forming pores and potentially docking the needle or pilus structure (He et al., 2004). Components of the basal body are encoded by 9 hrc genes that are conserved among all bacteria with TTSSs and share a high degree of sequence similarity (Hueck, 1998). Interestingly, a number of these conserved genes also share significant sequence similarity with genes encoding components of the flagellar secretion system (Blocker et al., 2003). Structural characterization of the TTSS basal body has been facilitated by the isolation of at least part of the structure from Salmonella enterica serovar typhimurium (Kubori et al., 1998). The basal body of the TTSS was found to be similar to the flagellar basal body and is made up of outer rings within the bacterial outer membrane and two inner rings interacting with the cytoplasmic membrane. In addition to the basal body, the TTSS apparatus also includes an extracellular appendage that, in mammalian pathogens, consists of a needle-like extension and in plant pathogens, is a longer pilus-like structure. The needle structure has been studied most extensively in mammalian pathogens such as Salmonella and Shigella. In Salmonella, the needle structure is composed of a major structural subunit encoded by the prgI gene and was found to be 80nm in length and to have a diameter of 8nm (Kimbrough and Miller, 2000; Kubori et al., 1998). A needle-like structure has not been found to be associated with the TTSS apparatus of plant pathogens. However, plant pathogenic bacteria have been found to assemble a longer, pilus-like extracellular structure associated with the TTSS (He and J in, 2003). These structures, referred to as Hrp pili, have the same diameter (8nm) as the TTSS needle of mammalian pathogens, but are significantly longer (several pm). This morphological difference between the extracellular components of mammalian and plant pathogen TTSSs is likely due to the unique cell type encountered by plant pathogens. Plant cells are surrounded by a thick cell wall, which must be penetrated by the phytopathogenic bacterium to reach the plasma membrane and cytosol of the host cell. The Hrp pilus of Pseudomonas syringae is composed of a major subunit encoded by the hrpA gene. Purified HrpA protein has been shown to assemble into a pilus-like structure in vitro and to be added to the tip of the nascent pilus during assembly (Li et al., 2002; Roine et al., 1997). Additionally, experimental evidence has demonstrated that type III effector proteins are secreted from the tip of the Hrp pilus, suggesting that the pilus acts as a conduit for the translocation of effectors into the host cell (J in and He, 2001; Li etal., 2002). Additional extracellular accessory proteins such as harpins are also necessary for function of the TTSS. These proteins may play a role in guiding the growing pilus through the plant cell wall and in forming pores in the host cell plasma membrane to allow docking of the pilus (He et al., 2004). Type III effector proteins Proteins secreted into host cells via the TTSS are generally thought to subvert host cell processes to create an environment favorable for bacterial multiplication (Grant et al., 2006). This can include the suppression of host defense responses to avoid detection or destruction by the host or the release of nutrients from the host cell to promote bacterial multiplication. Type III effector proteins from mammalian pathogens are generally less numerous than those from plant pathogens and many studies have shed light on the functions of these proteins. However, this section will focus on type III effectors from plant pathogens with an emphasis on the effectors encoded by Pseudomonas syringae. In 2003, the completed genome sequence of the model bacterial phytopathogen Pst DC3000 became available (Buell et al., 2003). The availability of a complete genome sequence for this bacterium prompted a number of studies aimed at cataloging the type III effector proteins in the Pst DC3000 genome. These studies used various bioinforrnatics and experimental approaches such as identification of a promoter element known as the hrp box which is recognized by the alternative sigma factor HrpL and used to coordinately regulate transcription of type III effectors and structural components of the TTSS (Guttrnan etal., 2002; Petnicki-chieja et al., 2002; Zwiesler-Vollick et al., 2002). One study suggested that Pst DC3000 may encode as many as 40 type III effectors (Schechter et al., 2004). However, more recent evidence gained using a combination or fluorescence-activated cell sorting, hrpL-dependent expression, and a reporter fusion system suggests that the number may be closer to 30 (Chang et al., 2005). Collectively, type III effector proteins are absolutely required for the pathogenicity of Pst DC3000 and other phytopathogenic bacteria such as Xanthomonas, Ralstonia, and Erwinia. Mutant bacteria such as hrp mutants that cannot translocate type III effectors into the host cell are rendered non-pathogenic and cannot cause disease on normally susceptible host plants (Lindgren et al., 1986). Because of their importance in pathogenicity, many studies have recently been aimed at discovering their function within the plant host. Type III effectors of phytopathogenic bacteria were originally characterized for their role in imparting avirulence on normally pathogenic bacteria and their ability to elicit the HR in a gene-for-gene manner on plants carrying a corresponding 10 resistance (R) gene. However, it has become clear that the primary function of type III effector proteins is to act as virulence factors to promote pathogenicity. Virulence functions of type III effectors Studying the function of type III effector proteins from plant pathogenic bacteria has proven to be difficult for a number of reasons. First, single pathovars of plant pathogenic bacteria encode large numbers of type III effectors, many of which seem to have overlapping functions in promoting virulence. Because of this, knockout mutants defective in the production of individual type III effectors generally display no effect or only a slight reduction in virulence, rendering the study of bacterial mutants relatively uninformative. Second, although sequence information is available for type III effectors from a variety of plant pathogenic bacteria, the nucleic acid and protein sequences share little or no homology to other genes of known function and, with only a few exceptions, provide little clue as to the virulence function. Recently, to circumvent these problems, a number of groups have employed a strategy of expressing bacterial type III effectors directly in plant cells to study their function. Such studies have begun to reveal clues as to the targets and functions of type III effectors from phytopathogenic bacteria. A number of type III effectors have been shown to suppress various aspects of plant defense responses. The effector protein AvrPto was originally characterized for its role in eliciting the HR in resistant tomato varieties carrying the Pto kinase (Ronald et al., 1992). Transgenic expression of AvrPto in Arabidopsis revealed that AvrPto could suppress the salicylic acid (SA)-independent deposition of callose in response to non- pathogenic type III secretion-deficient mutants of Pst DC3000 (Hauck et al., 2003). ll Callose, a [3-1-3 glucan, is deposited in cell wall reinforcements known as papillae in response to perception of bacteria by the plant (Brown et al., 1995). These cell wall reinforcements are thought to play a role in defense potentially by restricting bacterial access to host cells or by encapsulating bacteria and possibly facilitating their exposure to antimicrobial compounds. In addition to suppression of callose deposition, AvrPto transgenic Arabidopsis plants exhibited significant changes in gene expression. Interestingly, it was found that there was a clear bias in AvrPto plants toward the transcriptional repression of genes whose products were predicted to enter the plant cell secretory pathway. Collectively, these results suggest that the function of AvrPto in plant cells may be to block the assembly of defense-associated papillae. More recently, AvrPto was also found to block kinase activation and upregulation of the flagellin—induced receptor kinase 1 (frkI) gene in Arabidopsis cell cultures in response to treatment with the flg22 peptide, a synthetic peptide representing a highly-conserved region of the bacterial pathogen-associated molecular pattern (PAMP) flagellin (He et al., 2006). The precise molecular mechanisms by which AvrPto carries out these functions and the host target(s) on which it acts are still unknown. Recently, a molecular target of the Pst DC3000 effector Hole was identified. The hopMI gene lies within a conserved region of the Pst DC3000 genome known as the conserved effector locus (CEL) (Alfano et al., 2000). Partial deletion of the CEL in the ACEL mutant of Pst DC3000 results in a significant reduction in virulence and it was found that expression of either hopMI or the sequence-unrelated aer was sufficient to restore firll virulence (Alfano et al., 2000; DebRoy et al., 2004). These results suggest that the CEL~encoded effectors Hole and Aer have overlapping function and that this 12 function is critical for virulence on host plants. Transgenic expression of Hole in Arabidopsis revealed that this effector could suppress an SA-dependent pathway leading to deposition of papillae (DebRoy et al., 2004). Subsequently, it was discovered that Hole targets a plant ADP ribosylation factor (ARF) guanine nucleotide exchange factor (GEF) named AtMIN7 (Arabidopsis thaliana Hole interactor 7) for degredation via the plant proteasome (Nomura et al., 2006). ARF GEFs are regulatory components of vesicle trafficking in eukaryotic cells, suggesting that Hole may interfere with vesicle trafficking leading to deposition of papillae. Interestingly, Aer did not target AtMIN7 for degredation, suggesting that Aer functions in a manner mechanistically distinct from Hole. In addition to suppression of papillae-associated basal defenses, a number of Pst DC3000 type III effectors have been shown to suppress programmed cell death associated with the HR (Grant et al., 2006). A specific example of this type of effector is AvrPtoB. Like AvrPto, AvrPtoB is capable of triggering HR in resistant tomato varieties carrying Pto (Kim et al., 2002). However, AvrPtoB was found to suppress HR cell death initiated by Pto when both proteins, along with AvrPto, were transiently expressed in Nicotiana benthamiana (Abramovitch et al., 2003). Additionally, AvrPtoB was found to suppress the HR initiated by the Cf9 resistance protein in N. benthamiana and programmed cell death initiated by the pro-apoptotic mouse Bax protein in both N. benthamiana and yeast. It was subsequently discovered that AvrPtoB acts as an E3 ubiquitin ligase and that E3 ligase activity is required for suppression of HR-associated programmed cell death (Abramovitch et al., 2006; Janjusevic et al., 2006). However, other recent publications suggest a role for AvrPtoB in blocking basal defenses triggered 13 by perception of PAMPs such as flagellin. AvrPtoB was found to block formation of papillae and to promote virulence of the normally non-pathogenic Pseudomonas syringae pv phaseolicola strain RW60 on Arabidopsis and these functions were not dependent on E3 ligase activity (de Torres et al., 2006). AvrPtoB was also shown to suppress kinase activation and transcriptional upregulation of Arabidopsisfi'kl in response to perception of flg22 (He et al., 2006). In light of these recent results, the function(s) of AvrPtoB in promoting virulence are not clear. It is possible that AvrPtoB has multiple functions in the same plant host conferred by different regions of the protein. Alternatively, AvrPtoB may serve different functions in distinct host plants. Identification of specific host targets for AvrPtoB will likely help to sort out the exact virulence role of this interesting type III effector. Phytotoxins of plant pathogenic bacteria In addition to type III effectors, many plant pathogenic bacteria also produce one or more toxins that play a role in promoting virulence on host plants. The various Pseudomonas syringae pathovars produce a number of different toxins including coronatine produced by pv tomato and other pathovars, phaseolotoxin produced by pv phaseolicola, tabtoxin of pv tabaci, and syringomycin and syringopeptin of pv syringae (Bender et al., 1999). In many cases, phytotoxins have been demonstrated to contribute to the pathogenicity of the toxin-producing strains, however, relatively little is known about the specific mechanisms by which these toxins contribute to virulence. The phytotoxin coronatine (COR) is produced by at least 5 different pathovars of P. syringae including pvs tomato, glycinea, atropurpurea, maculicola, and morsprunorum 14 (Bender et al., 1999). COR is a polyketide toxin consisting of two distinct structural components, coronofacic acid and coronarnic acid (Ichihara et al., 1977). Structurally, COR is closely related to the plant hormone jasmonic acid (J A) and plants such as the COR-insensitive Arabidopsis coiI mutants are also resistant to the effects of IA, prompting researchers to hypothesize that COR may act as a molecular mimic of J A (Feys et al., 1994; Greulich et al., 1995; Koda et al., 1996; Weiler et al., 1994). It was observed that a COR-deficient mutant of Pst DC3000 could not cause disease when inoculated onto the surface of Arabidopsis leaves, but was capable of causing disease symptoms and multiplying to wild-type levels when infiltrated directly into the leaf intercellular spaces (Mittal and Davis, 1995). These results led the authors to speculate that COR may suppress an uncharacterized early defense response in Arabidopsis. In chapter 3 I will present evidence that this early defense response is bacterium-induced stomatal closure and that the primary virulence role for COR is the suppression of this stomatal defense response. The modes of action of several other P. syringae toxins have also been determined, however, in most cases the precise mechanisms by which these toxins promote virulence are not understood. For example, phaseolotoxin produced by pathovars phaseolicola and actinidiae has been shown to inhibit ornithine carbamoyl transferase (OCTase). OCTase is involved in the urea cycle in both prokaryotes and eukaryotes and catalyzes the conversion of ornithine and carbamoyl phosphate to citrulline. P. syringae pathovars that produce phaseolotoxin also produce resistant OCTase isoforms, rendering them insensitive to the effects of the toxin. The result of inhibition of OCTase is the accumulation of ornithine and a deficiency of intracellular arginine pools and leads to 15 chlorosis in plant tissues. However, it is not clear how phaseolotoxin contributes to virulence. It has been speculated that phaseolotoxin may actually be involved in direct competition with other epiphytic bacteria for niches in the phyllosphere (V olksch and Weingart, 1998). Syringomycins and syringopeptin are lipodepsipeptide toxins produced by pathovar syringae. These toxins contribute significantly to virulence and have been shown to cause necrosis in plant tissues (Hutchison and Gross, 1997; Quigley etal., 1993). The function of this class of toxins seems to be related to their ability to form pores in the host plasma membrane (Hutchison and Gross, 1997). Pore formation in plant cell membranes may be a mechanism for P. syringae pv syringae to acquire nutrients fiom the host. Plant defense responses Although some bacteria have evolved mechanisms to successfully colonize plants and cause disease, most bacteria are unable to infect plants. In addition, most pathovars of phytopathogenic bacteria are only able to infect a very limited number of plant hosts and are non-pathogenic on other plant species or cultivars. Therefore, plants are able to successfully defend themselves against the majority of bacteria to which they are exposed. Plants employ a number of mechanisms to successfully avoid infection by potentially pathogenic bacteria including preformed physical barriers such as the leaf cuticle and cell walls, a suite of basal defense responses including deposition of cell wall reinforcements and synthesis of antimicrobial compounds such as phytoalexins, a form of programmed cell death referred to as the hypersensitive response (HR), and a systemic immunity known as systemic acquired resistance (SAR). During the last 15 years, 16 significant progress has been made toward understanding the perception and signaling events associated with elicitation of the HR. In contrast, basal defenses have received less attention, and only within the last few years has the importance of basal defenses been recognized. It is likely that both basal defenses and the HR contribute to long-term, durable resistance of plants to a broad spectrum of potential pathogens. Therefore, a thorough understanding of both responses, including perception events, signal transduction, gene expression changes, and ultimately the biochemical processes leading to restriction of multiplication, will be required to develop strategies for engineering durable resistance to agronomically important plant pathogens. Plant basal defenses Basal defenses in plants are analogous to innate immunity in animals and insects (Staskawicz et al., 2001). Innate immunity is an evolutionarily ancient form of resistance to microbes and is characterized by the recognition of conserved microbial molecules referred to as pathogen-associated molecular patterns (PAMPs). PAMPs generally contribute to a function that is critical to the lifestyle of the organism and thus, are indispensable and are generally well conserved across a wide range of microbes (N umberger et al., 2004). A major breakthrough in the understanding of innate immunity in plants came when plants were found to perceive bacterial flagellin, the proteinaceous subunit that is the main component of the bacterial flagellum (Felix et al., 1999). Perception of flagellin or a 22 amino acid highly elicitor-active peptide, flg22, derived from a well-conserved domain of flagellin, was found to inhibit growth of Arabidopsis seedlings and elicit callose deposition in leaf tissue (Gomez-Gemez et al., 1999). 17 Additionally, treatment of suspension-cultured tobacco cells with flg22 led to alkalinization of the culture medium, suggesting that flg22 triggers production of reactive oxygen species (ROS). The receptor responsible for perception of flagellin by Arabidopsis was identified as an LRR receptor-like kinase and named FLS2 (G6mez- Gomez et al., 2000). It was subsequently discovered that signaling mediated by FLS2 proceeds through a mitogen—activated protein kinase (MAPK) cascade involving the Arabidopsis MAPKs AtMPK3 and AtMPK6 and leads to the rapid transcriptional induction of a number of genes including those encoding the WRKY transcription factors WRKY22 and WRKY29 (Asai et al., 2002). ArabidopsisflsZ mutants lacking the flagellin receptor were found to be slightly more susceptible to Pst DC3000 when the bacteria were inoculated on the leaf surface, but not when bacteria were infiltrated directly into the leaf apoplast, suggesting that flagellin perception may be important for triggering a defense active on the leaf surface (Zipfel et al., 2004). Expression profiling studies of Arabidopsis seedlings or cell cultures treated with flg22 identified a number of transcriptional changes associated with flagellin perception including the induction of genes whose products are associated with signal transduction, transcriptional regulation, and defense (Navarro et al., 2004; Zipfel et al., 2004). In addition to flagellin, Arabidopsis has subsequently been shown to perceive several other bacterial PAMPs including elongation factor Tu (EF-Tu) and lipopolysaccharides (LPS) and the receptor mediating perception of EF-Tu has been identified (Kunze et al., 2004; Zeidler et al., 2004; Zipfel et al., 2006). The above studies and others have shed significant light on PAMP perception and signaling in plants, however, the subsequent defense responses and their role in 18 restricting bacterial infection are still poorly understood. Two relatively well- characterized aspects of basal defenses are the deposition of cell wall appositions called papillae and the production of phytoalexins. However, the role of papillae and phytoalexins in restricting bacterial multiplication remains unclear. The formation of papillae in response to a pathogen was first observed over 100 years ago (reviewed in Aist, 1976). The majority of the literature concerning papilla formation in response to pathogens describes papilla deposition in response to attempted penetration by fungal pathogens. However, formation of papillae or papilla-like structures has also been observed in plants in response to bacteria, viruses, and nematodes (Bestwick et al., 1995; Allison and Shalla, 1974; Riggs et al., 1973). The most prevalent hypothesis for the function of papillae in response to fungal attack is that papillae serve as a structural barrier allowing the plant cell to resist mechanical or enzymatic penetration by the invading fitngus. Rapid formation of papillae has been correlated with enhanced resistance to fungal penetration, whereas, delayed papilla formation has been correlated with an increase in penetration frequency (Bayles et al., 1990; Collins et al., 2003). The structural components of papillae seem to vary between plant species, but compounds commonly found associated with papillae include: callose; lignin; phenolics; reactive oxygen species (ROS); peroxidases; and hydroxyproline-rich glycoproteins (Aist, 1976; McLusky et al., 1999). It is thought that these components are cross-linked, possibly through the action of ROS, into a dense, insoluble matrix that serves to strengthen the cell wall locally at the site of interaction with the invading pathogen. Although the papilla response has been studied for more than a century, very little is known about the elaboration of this response at the molecular level. Studies with 19 various pharmacological inhibitors suggest that transcription, translation, protein glycosylation, and microfilament polymerization are required for normal papilla deposition (Skalamera et al., 1997). The Arabidopsis callose synthase encoded by the GSL5 gene was shown to be required for deposition of callose in papillae and for wound- induced callose deposition (N ishimura et al., 2003; Jacobs et al., 2003). However, gsl5 mutant Arabidopsis plants were able to form papillae, which, although lacking callose, did not differ significantly in size and shape from papillae formed by wild type plants. Interestingly, gsl5 mutants showed enhanced resistance to several normally virulent species of powdery mildew. This enhanced resistance was found to be dependent on a functional salicylic acid (SA) defense signaling pathway, as double mutants impaired in SA defense signaling regained susceptibility. Additionally, the Arabidopsis SNARE (soluble NSF attachment protein receptor) protein PEN] (also known as SYP121) was found to be associated with deposition of papillae in response to perception of non-host fungal pathogens (Assaad et al., 2004; Collins et al., 2003; Lipka et al., 2005). Deposition of papillae has been observed in response to strains of several plant- associated bacteria including Pseudomonas syringae and Xanthomonas campestris in a variety of plant species such as lettuce, pepper, bean, and Arabidopsis (Brown et al., 1995; Bestwick et al., 1995; Brown et al., 1998). Whether components mediating the deposition of papillae in response to fungi, such as PEN], are also involved in formation of papillae in response to bacteria is currently not known. Of particular significance is the observation that saprophytic strains, avirulent strains and non-pathogenic mutants such as hrp mutants elicit deposition of papillae, whereas pathogenic strains do not elicit this response (Brown et al., 1995). This observation suggests that pathogenic strains have the 20 ability to suppress this defense response, and that this suppression is dependent on a functional type III secretion system, implicating the involvement of type III-secreted effector proteins in suppressing this response. As discussed previously, several type III effectors have already been implicated in suppressing this defense response. In chapter 4, I will present evidence that the Pst DC3000 effector HopD2 can suppress deposition of papillae. Gene-for-gene resistance In addition to recognition of PAMPs, plants have also evolved the ability to detect the presence of virulence factors such as type III effectors secreted directly into the host cell. Recognition of specific type III effectors is mediated by host resistance (R) proteins, most of which belong to a family of leucine-rich repeat (LR) and nucleotide-binding (NB) domain containing proteins with either a coiled-coil (CC) or Toll-interleukin-l receptor (TIR) domain at their N terminus. Activation of R proteins leads to a defense known as the hypersensitive response (HR), which involves the death of the affected cell and is thought to be a form of programmed cell death. The requirement of a specific single gene from the pathogen (generally referred to as an avirulence or avr gene) and a specific single gene from the host (the R gene) to confer resistance led H.H. Flor to describe this phenomenon as gene-for-gene resistance over 50 years ago (Flor, 1955). The most straightforward and simple model for recognition of bacterial virulence proteins and activation of plant R proteins is one in which the R protein acts as a receptor for the bacterial virulence protein and, through direct interaction, becomes activated and subsequently promotes initiation of the HR. In the case of several R-avr genetic 21 interactions, a direct physical interaction between R and Avr proteins has been demonstrated experimentally. Specific examples include recognition of the AVR—Pita protein from the rice blast fungus Magnaporthe grisea by the corresponding rice R protein Pi-Ta (J ia et al., 2000); recognition of Avr proteins of the flax rust (Melampsora Iini) Aer5 6 7 locus by the flax (Linum usitatissimum) R proteins L5, L6, and L7 (Dodds etal., 2006); and recognition of the PopP2 Avr protein of Ralstonia solanacearum by the Arabidopsis RRSl-R protein (Deslandes et al., 2003). However, for many other R-Avr pairs, a direct interaction has not been demonstrated and strong experimental evidence instead points to an alternative model referred to as the “guard hypothesis.” The guard hypothesis proposes that R proteins act as sentries within the plant cell to guard virulence targets of bacterial effector proteins by detecting the activity of the virulence proteins or the perturbation of the host target, thus indirectly detecting the virulence factors (N irnchuck et al., 2001). Recently, the guard hypothesis has gained significant experimental support. The best characterized example of indirect recognition of bacterial effector proteins is that of the Arabidopsis protein RIN4 (RPMl-interacting protein 4). Activation of at least two Arabidopsis R proteins, RPSZ and RPMl , has been shown to be associated with modification of RIN4. The R protein RPM] is activated in response to phosphorylation of RIN4 by either of two unrelated P. syringae effectors, AvrB or Aerme (Mackey et al., 2002). RPS2, on the other hand, is activated by the absence of RIN4 caused by its proteolytic degradation by the P. syringae effector Aerpt2 (Axtell and Staskawicz, 2003; Mackey et al., 2003). These results suggest that RIN4 may be an important virulence target of these effectors, however, the precise role of RIN4 in the plant response to bacteria is still not known. Recent work suggests that RIN 4 may play a 22 role in the plant basal defense system activated in response to PAMP perception (Kim et al., 2005). Another example of an Arabidopsis R protein guarding a potential host virulence target is that of RPSS. Similar to RPSZ, RPSS detects the proteolytic cleavage of a host protein, the PBS] kinase, by the P. syringae type III effector AverhB (Shao et al., 2003). The role of PBS] as a host virulence target is also currently unknown. Project summary and rationale One of the most important long-term goals underlying the study of plant-microbe interactions is to develop new strategies for reducing the impact of microbial pathogens on agricultural production. In order to successfully formulate and institute such strategies, it is necessary to understand the mechanisms employed successfully by plants to defend themselves against most potential pathogens and the strategies employed by successful pathogens to overcome these defenses. While gene-for-gene resistance and defenses associated with the hypersensitive response have received significant attention over the past two decades, resistance conferred by a single R gene is not likely to be durable and long-lasting due to significant selection pressure placed on the pathogen. Emerging evidence suggests that stable and durable resistance is likely conferred by a combination of basal defenses associated with PAMP perception and R gene-mediated defenses. In comparison to R gene-mediated defenses, relatively little was known about plant basal defenses at the onset of this study. One of the major goals of this work was to improve our understanding of plant basal defenses associated with perception of PAMPs. 23 Successful pathogens have evolved mechanisms to overcome plant defenses. To achieve effective pathogen resistance in crop plants, it will also be important to understand the ways in which pathogens are able to block, break down, or circumvent normally effective host defenses. A second goal of this work was to investigate mechanisms by which the bacterial pathogen Pst DC3000 overcomes the basal defenses of the host plant Arabidopsis thaliana. It was known that two major virulence factors, COR and TTSS effectors, contribute to the ability of Pst DC3000 to overcome host defenses. However, the mechanism of action of COR or any specific effector protein was not known at the onset of this study. The work described here investigates aspects of both virulence systems. Chapter 2 describes microarray analyses of Arabidopsis gene expression changes associated with activation of basal defenses and with the activities of each of the two virulence factors of Pst DC3000, COR and the TTSS. Prior to the initiation of this study expression profiling of the interaction of Pst DC3000 with Arabidopsis had been reported (Tao et al., 2003). This study addressed the compatible interaction between Pst DC3000 and Arabidopsis and the incompatible interaction associated with Pst DC3 000 carrying an avirulence gene. However, gene expression changes associated with the activation of basal defense were not addressed in this study. Additionally, the specific roles of COR and the TTSS in modulating gene expression also were not addressed. To identify gene expression changes associated with the activation of basal defenses, I analyzed the expression profiles of Arabidopsis plants inoculated with non-pathogenic hrp mutants of Pst DC3000 or with the human pathogen E. coli 01 57:H7. Additionally, I sought to determine the role of flagellin perception in mediating the gene expression changes 24 associated with basal defenses. At the onset of this study, flagellin was the only bacterial PAMP known to be perceived by plants and it was conceivable that perception of flagellin may contribute specifically and uniquely to gene expression changes associated with the activation of basal defenses. To address this question, I compared the expression profiles of Arabidopsis plants inoculated with fliC mutant bacteria lacking flagella to those of plants inoculated with the wild-type, flagella producing strains. To address the specific roles of TTSS effectors and COR in modulating host gene expression, I collaborated with Dr. Roger Thilmony to analyze the expression profiles of plants inoculated with defined Pst DC3000 mutants defective specifically in COR biosynthesis or secretion of TTSS effectors. Chapter 3 reports the discovery of a novel component of plant basal defenses, closure of stomata. At the time this study was initiated, it was known that mutants of Pst DC3000 defective in COR biosynthesis were significantly compromised in virulence when inoculated onto the surface of host leaves, but were normally virulent when infiltrated directly into the leaf apoplast, suggesting the possible existence of a defense mechanism active on the leaf surface and a role for COR in overcoming this hypothetical defense (Mittal and Davis, 1995). Additionally, flsZ mutant Arabidopsis plants lacking the flagellin receptor were more susceptible to Pst DC3000 when inoculated on the leaf surface, but not when the bacteria were infiltrated directly into the leaf apoplast (Zipfel et al., 2004) This result suggested that flagellin perception may contribute to a defense mechanism active on the leaf surface. These results led us to hypothesize that stomata may play a role in basal defenses of plants against bacteria. To address this hypothesis, I collaborated with Dr. Maeli Melotto to investigate the behavior of plant stomata in 25 response to perception of bacteria and PAMPs and to investigate a potential role for COR in modulating stomatal behavior. Chapter 4 describes my research on the virulence function of a specific Pst DC3000 type III effector, HopD2. HopD2 shares some sequence similarity with protein tyrosine phosphatases, including a conserved active domain (Bretz et al., 2003; Espinoza et al., 2003). This potential enzyme activity provided a starting point for the study of this type III effector. To study the function of HopD2 in Arabidopsis, I created transgenic plant lines expressing wild-type HopD2 or a phosphatase-inactive mutant derivative under the control of an inducible promotor. Transgenic Arabidopsis plants expressing HopD2 allow enhanced multiplication of normally non-pathogenic hrp mutants of Pst DC3000. Additionally, these transgenic plants display reduced papillae-associated callose deposition in response to non-pathogens, suggesting that HopD2 may play a role in blocking basal defenses. This defense-suppressing activity of HopD2 is dependent on its tyrosine phosphatase activity as plants expressing the catalytically inactive mutant protein did not show these phenotypes. To investigate the possibility that HopD2 might exert its virulence function by inactivating Arabidopsis MAP kinases known to be involved in PAMP-triggered signaling leading to activation of basal defense, I collaborated with Dr. Shuqun Zhang to assess the activation of AtMPK3 and AtMPK6 in the HopD2 transgenic plants. I performed microarray and lipid profiling analyses to address the phosphatase- specific changes associated with HopD2 in host plants. These experiments provide evidence of a role for HopD2 in suppressing basal defenses in the host plant Arabidopsis thaliana. 26 References Abramovitch, R.B., Janjusevic, R., Stebbins, CE, and Martin, GB. (2006) Type III effector AvrPtoB requires intrinsic E3 ubiquitin ligase activity to suppress plant cell death and immunity. Proc. Natl. Acad. Sci. USA, 103, 2851-2856. Abramovitch, R.B., Kim, Y.J., Chen, S., Dickman, M.B., and Martin, GB. (2003) Pseudomonas type III effector AvrPtoB induces plant disease susceptibility by inhibition of host programmed cell death. EMBO J. 22, 60-69. Agrios, ON. (2005) Plant Pathology 5th edn. Academic Press, San Diego. Aist, J .R. (1976) Papillae and related wound plugs of plant cells. Annu. Rev. Phtyo. Pathol. 14, 145-163. Alfano, J .R., Charkowski, A.O., Deng, W.L., Bade], J .L., Petnicki-chieja, T., van Dijk, K., and Collmer, A. (2000) The Pseudomonas syringae Hrp pathogenicity island has a tripartite mosaic structure composed of a cluster of type III secretion genes bounded by exchangeable effector and conserved effector loci that contribute to parasitic fitness and pathogenicity in plants. Proc. Natl. Acad. Sci. USA, 97, 4856-4861. Allison, A.V., and Shalla, TA. (1974) The ultrastructure of local lesions induced by potato virus X: a sequence of cytological events in the course of infection. Phytopathology, 64, 784-93. Arlat, M., Gijsegem, F.V., Huet, J.C., Pemollet, J .C., and Boucher, CA. (1994) PopAl, a protein which induces a hypersensitivity-like response on specific Petunia genotypes, is secreted via the Hrp pathway of Pseudomonas solanacearum. EMBO J. 13, 543-553. Asai, T., Tena, G., Plotnikova, J ., Willmann, M.R., Chiu, W.L., Gomez-Gomez, L., Boller, T., Ausubel, F.M., and Sheen, J. (2002) MAP kinase signaling cascade in Arabidopsis innate immunity. Nature, 415, 977-983. Assaad, F.F., Qiu, J .L., Youngs, R, Ehrhardt, D., Zimmerli, L., Kalde, M., Wanner, G., Peck, S.C., Edwards, H., Ramonell, K., Somerville, C.R., and Thordal-Christensen, H. (2004) The PENl syntaxin defines a novel cellular compartment upon fungal attack and is required for the timely assembly of papillae. Mol. Biol. Cell, 15, 5118-5129. Axtell, M.J., and Staskawicz, B.J. (2003) Initiation of RPSZ-specified disease resistance in Arabidopsis is coupled to the Aerpt2-directed elimination of RIN4. Cell, 112, 369- 377. 27 Bender, C.L., Alarcon-Chaidez, F., and Gross, DC. (1999) Pseudomonas syringae phytotoxins: mode of action, regulation, and biosynthesis by peptide and polyketide synthetases. Microbiol. Mol. Biol. Rev. 63, 266-292. Bestwick, C.S., Bennet, M.H., and Mansfield, J .W. (1995) Hrp mutant of Pseudomonas syringae pv. phaseolicola induces cell wall alterations but not membrane damage leading to the hypersensitive reaction in lettuce. Plant Physiol. 108, 503-516. Blocker, A., Komoriya, K., and Aizawa, S. (2003) Type III secretion systems and bacterial flagella: insights into their fiinction from structural similarities. Proc. Natl. Acad. Sci. USA, 100, 3027-3030. Bretz, J .R., Mock, N.M., Charity, J .C., Zeyad, S., Baker, C.J., and Hutcheson, SW. (2003) A translocated protein tyrosine phosphatase of Pseudomonas syringae pv. tomato DC3000 modulates plant defence response to infection. Mol. Microbiol. 49, 389-400. Brown, 1., Mansfield, J ., and Bonas, U. (1995) hrp genes in Xanthomonas campestris pv. vesicatoria determine ability to suppress papilla deposition in pepper mesophyll cells. Mol. Plant Microbe Interact. 8, 825-836. Brown, I., Trethowan, J ., Kerry, M., Mansfield, J ., and Bolwell, G.P. (1998) Localization of components of the oxidative cross-linking of glycoproteins and of callose synthesis in papillae formed during the interation between non-pathogenic strains of Xanthomonas campestris and French bean mesophyll cells. Plant J. 15, 333-343. Buell, C.R., Joardar, V., Lindeberg, M., Selengut, J ., Paulsen, I.T., Gwinn, M.L., Dodson, R.J., Deboy, R.T., Durkin, A.S., Kolonay, J .F., Madupu, R., Daugherty, S., Brinkac, L., Beanan, M.J., Hafi, D.H., Nelson, W.C., Davidsen, T., Zafar, N., Zhou, L., Liu, J ., Yuan, Q., Khouri, H., Fedorova, N., Tran, B., Russell, D., Berry, K., Utterback, T., Van Aken, S.E., Feldblyum, T.V., D'Ascenzo, M., Deng, W.L., Ramos, A.R., Alfano, J .R., Cartinhour, S., Chatterjee, A.K., Delaney, T.P., Lazarowitz, S.G., Martin, G.B., Schneider, D.J., Tang, X., Bender, C.L., White, 0., Fraser, C.M., and Collmer, A. (2003). The complete genome sequence of the Arabidopsis and tomato pathogen Pseudomonas syringae pv. tomato DC3000. Proc. Natl. Acad. Sci. USA, 18, 10181-10186. Chang, J .H., Urbach, J .M., Law, T.F., Arnold, L.W., Hu, A., Gombar, S., Grant, S.R., Ausubel, F.M., and Dangl, J .L. (2005) A high-throughput, near-saturating screen for type III effector genes from Pseudomonas syringae. Proc. Natl. Acad. Sci. USA, 102, 2549- 2554. Collins, N.C., Thordal-Christensen, H., Lipka, V., Bau, S., Kombrink, E., Qiu, J .L., Huckelhoven, R., Stein, M., F reialdenhoven, A., Somerville, SC, and Schulze-Lefert, P. (2003) SNARE-protein-mediated disease resistance at the plant cell wall. Nature, 425, 973-977. 28 Dale, G., Young, S.A., Haydon, D.T., and Welburn, SC. (2001) The insect endosymbiont Sodalis glossinidius utilizes a type III secretion system for cell invasion. Proc. Natl. Acad. Sci. USA, 98, 1883-1888. de Torres, M., Mansfield, J .W., Grabov, N., Brown, I.R., Ammouneh, H., Tsiamis, G., Forsyth, A., Robatzek, S., Grant, M., and Boch J. (2006) Pseudomonas syringae effector AvrPtoB suppresses basal defence in Arabidopsis. Plant J. 47, 368-3 82. DebRoy, S., Thilmony, R., Kwak, Y.B., Nomura, K., and He, S.Y. (2004) A family of conserved bacterial effectors inhibits salicylic acid-mediated basal immunity and promotes disease necrosis on plants. Proc. Natl. Acad. Sci. USA, 101, 9927-9932. Deslandes, L., Olivier, J., Peeters, N., Feng, D.X., Khounlotham, M., Boucher, C., Somssich, I., Genin, S., and Marco, Y. (2003) Physical interaction between RRSl-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. Dodds, P.N., Lawrence, G.J., Catanzariti, A.M., Teh, T., Wang, C.I., Ayliffe, M.A., Kobe, B., and Ellis, J .G. (2006) Direct protein interaction underlies gene-for-gene specificity and coevolution of the flax resistance genes and flax rust avirulence genes. Proc. Natl. Acad. Sci. USA, 103, 8888-8893. Espinosa, A., Guo, M., Tam, V.C., Fu, Z.Q., and Alfano, J .R. (2003) The Pseudomonas syringae type III-secreted protein HothoD2 possesses protein tyrosine phosphatase activity and suppresses programmed cell death in plants. Mol. Microbiol. 49, 377-3 87. Felix, G., Duran, J .D., Volko, S., and Boller, T. (1999) Plants have a sensitive perception system for the most conserved domain of bacterial flagellin. Plant J. 18, 265-276. Feys, B.J.F., Benedetti, C.E., Penfold, C.N., and Turner, J .G. (1994) Arabidopsis mutants selected for resistance to the phytotoxin coronatine are male sterile, insensitive to methyl jasmonate, and resistant to a bacterial pathogen. Plant Cell, 6, 751-759. Flor, H.H. (1955) Host-parasite interactions in flax rust-its genetics and other implications. Phytopathology, 45, 680-685. Food and Agriculture Organization (1993) Production year book, F AO, Rome. Gardan, L., Shafik, H., Belouin, S., Broch, R., Grimont, F., and Grimont, P.A.D. (1999) DNA relatedness among the pathovars of Pseudomonas syringae and description of Pseudomonas tremae sp. nov. and Pseudomonas cannabina sp. nov. (ex Sutic and Dowson 1959). Int. J. Syst. Bacteriol. 49, 469-478. deez-Gémez, L., and Boller, T. (2000) FLS2: An LRR receptor-like kinase involved in the perception of the bacterial elicitor flagellin in Arabidopsis. Mol. Cell, 5, 1003-1011. 29 Gomez-Gomez, L., Felix, G., and Boller, T. (1999) A single locus determines sensitivity to bacterial flagellin in Arabidopsis thaliana. Plant J. 18, 277-284. Grant, S.R., Fisher, B.J., Chang, J .H., Mole, B.M., and Dangl, J .L. (2006) Subterfuge and manipulation: type III effector proteins of phytopathogenic bacteria. Annu. Rev. Microbial. 60, 425-449. Greulich, F ., Yoshihara T., and Ichihara, A. (1995) Coronatine, a bacterial phytotoxin, acts as a stereospecific analog of j asmonate type signals in tomato cells and potato tissues. J. Plant Physiol. 147, 359-366. Guttman, D.S., Vinatzer, B.A., Sarkar, S.F., Ranall, M.V., Kettler, G., and Greenberg, J .T. (2002) A functional screen for the type III (Hrp) secretome of the plant pathogen Pseudomonas syringae. Science, 295, 1722-1726. Hauck, P., Thilmony R., and He S.Y. (2003) A Pseudomonas syringae type III effector suppresses cell wall-based extracellular defense in susceptible Arabidopsis plants. Proc. Natl. Acad. Sci. USA, 100, 8577-8582. He, P., Shan, L., Lin, N.C., Martin, G.B., Kemmerling, B., Numberger, T., Sheen, J. (2006) Specific bacterial suppressor of MAMP signaling upstream of MAPKKK in Arabidopsis innate immunity. Cell, 125, 563-575. He, 8., Huang, A., and Collmer, A. (1993) Pseudomonas syringae pv. syringae harpinpssz a protein that is secreted via the Hrp pathway and elicits the hypersensitive response in plants. Cell, 73, 1255-1266. He, S.Y., and J in, Q. (2003) The Hrp pilus: learning from flagella. Curr. Opin. Microbial. 6, 15-19. He, S.Y., Nomura, K., and Whittam, TS. (2004) Type III protein secretion mechanism in mammalian and plant pathogens. Biochim. Biophys. Acta. 1694, 181-206. Hirano, SS, and Upper, CD. (2000) Bacteria in the leaf ecosystem with emphasis on Pseudomonas syringae-a pathogen, ice nucleus, and epiphyte. Microbial. Ma]. Biol. Rev. 64, 624-653. Huang, Q., and Allen, C. (2000) Polygalacturonases are required for rapid colonization and full virulence of Ralstonia solanacearum on tomato plants. Physio]. and Mol. Plant Pathol. 57, 77-83. Hueck, CJ. (1998) Type III protein secretion systems in bacterial pathogens of animals and plants. Microbial. Mol. Biol. Rev. 62, 379-433. 30 Hutchison, M.L., and Gross, DC. (1997) Lipopeptide phytotoxins produced by Pseudomonas syringae pv. syringae: comparison of the biosurfactant and ion channel- forming activities of syringopeptin and syringomycin. Ma]. Plant-Microbe Interact. 10, 347-3 54. Ichihara, A., Shiraishi, K., Sato, H., Sakamura, S., Nishiyama, K., Sakai, R., Furusaki, A., and Matsumoto, T. (1977) The structure of coronatine. J. Am. Chem. Soc. 99, 636-637. Jacobs, A.K., Lipka, V., Burton, R.A., Panstruga, R., Strizhov, N., Schulze-Lefert, P., and Fincher, GB. (2003) An Arabidopsis callose synthase, GSL5, is required for wound and papillary callose formation. Plant Cell, 15, 2503-2513. Janjusevic, R., Abramovitch, R.B., Martin, GB, and Stebbins, CE. (2006) A bacterial inhibitor of host programmed cell death defenses is an E3 ubiquitin ligase. Science, 311, 222-226. Jia, Y., McAdams, S.A., Bryan, G.T., Hershey, HR, and Valent, B. (2000) Direct interaction of resistance gene and avirulence gene products confers rice blast resistance. EMBO J. 19, 4004-4014. Jin, Q., and He, S.Y. (2001) Role of the Hrp pilus in type III secretion in Pseudomonas syringae. Science, 294, 2556-2558. Kim, M.G., de Cunha, L., Mcfall, A.J., Belkhadir, Y., DebRoy, S., Dangl, J.L., and Mackey, D. (2005) Two Pseudomonas syringae type III effectors inhibit RIN4-regulated basal defense in Arabidopsis. Cell, 121, 749-759. Kim, Y.J., Lin N.C., and Martin, GB. (2002) Two distinct Pseudomonas effector proteins interact with the Pto kinase and activate plant immunity. Cell, 109, 589-598. Kimbrough, T.G., and Miller, SJ. (2000) Contribution of the Salmonella typhimurium type III secretion components to needle complex formation. Proc. Natl. Acad. Sci. USA, 97,11008-11013. Koda, Y., Takahashi, K., Kikuta, Y., Greulich, F., Toshima, H., and Ichihara, A. (1996) Similarities of the biological activities of coronatine and coronafacic acid to those of jasmonic acid. Phytochemistry, 41, 93-96. Kubori, T., Matsushima, Y., Nakamura, D., Uralil, J ., Lara-Tejero, M., Sukhan, A., Galan, J .E., and Aizawa, SI. (1998) Suprarnolecular structure of the Salmonella typhimurium type 111 protein secretion system. Science, 280, 602-605. Kunze, G., Zipfel, C., Robatzek, S., Niehaus, K., Boller, T., and Felix, G. (2004) The N terminus of bacterial elongation factor Tu elicits innate immunity in Arabidopsis plants. Plant Cell, 16, 3496-3507. 31 Li, C.M., Brown, 1., Mansfield, J ., Stevens, C., Boureau, T., Taira M., and Taira S. (2002) The Hrp pilus of Pseudomonas syringae elongates from its tip and acts as a conduit for translocation of the effector protein HrpZ. EMBO J. 21, 1909-1915. Lindgren, P.B., Peet, R.C., and Panopoulos, NJ. (1986) Gene cluster of Pseudomonas syringae pv. phaseolicola controls pathogenicity of bean plants and hypersensitivity of nonhost plants. J. Bacterial. 168, 512-522. Lipka, V., Dittgen, J ., Bednarek, P., Bhat, R., Wiermer, M., Stein, M., Landtag, J ., Brandt, W., Rosahl, S., Scheel, D., Llorente, F., Molina, A., Parker, J ., Somerville, S., and Schulze-Lefert, P. (2005) Pre- and postinvasion defenses both contribute to nonhost resistance in Arabidopsis. Science, 310, 1180-1183. Mackey, D., Belkhadir, Y., Alonso, J .M., Ecker, J .R., and Dangl, J .L. (2003) Arabidopsis RIN4 is a target of the type III virulence effector Aerpt2 and modulates RPS2-mediated resistance. Cell, 112, 379-389. Mackey, D., Holt, B.F. 3rd, Wiig, A., and Dangl, J .L. (2002) RIN4 interacts with Pseudomonas syringae type III effector molecules and is required for RPMl-mediated resistance in Arabidopsis. Cell, 108, 743-754. Marie, C., Broughton, W.J., and Deakin, WJ. (2001) Rhizabium type III secretion systems: legume charmers or alarmers? Curr. Opin. Plant Biol. 4, 336-342. McLusky, S.R., Bennett, M.H., Beale, M.H., Lewis, M.J., Gaskin, P., and Mansfield, J .W. (1999) Cell wall alterations and localized accumulation of feruloyl-3’- methoxytyramine in onion epidermis at sites of attempted penetration by Batryis allii are associated with actin polarization, peroxidase activity and suppression of flavenoid biosynthesis. Plant J. 17, 523-534. Mittal, S.M., and Davis, KR. (1995) Role of the phytotoxin coronatine in the infection of Arabidopsis thaliana by Pseudomonas syringae pv. tomato. Mal. Plant-Microbe Interact. 8, 165-171. Navarro, L., Zipfel, C., Rowland, 0., Keller, 1., Robatzek, S., Boller, T., and Jones, J.D.G. (2004) The transcriptional innate immune response to flg22. Interplay and overlap with Avr gene-dependent defense responses and bacterial pathogenesis. Plant Physiol. 135,1113-1128. Nimchuk, Z., Rohmer, L., Chang, J .H., and Dangl, J .L. (2001) Knowing the dancer from the dance: R-gene products and their interactions with other proteins from host and pathogen. Curr. Opin. Plant Biol. 4, 288-294. Nishimura, M.T., Stein, M., Hou, B., Vogel, J.P., Edwards, H., and Somerville, SC. (2003) Loss of a callose synthase results in salicylic acid-dependent disease resistance. Science, 301, 969-972. 32 Nomura, K., Debroy, S., Lee, Y.H., Pumplin, N., Jones, J ., and He, S.Y. (2006) A bacterial virulence protein suppresses host innate immunity to cause plant disease. Science, 313, 220-223. Numberger, T., Brunner, F ., Kemmerling, B., and Piater, L. (2004) Innate immunity in plants and animals: striking similarities and obvious differences. Immuno. Rev. 198, 249- 266. Oerke, EC. (1994) Estimated crop losses due to pathogens, animal pests and weeds. In Crop Production and Crop Protection (Oerke, E.C., Dehne, H.W., Schonbeck, F ., and Weber, A., eds). pp. 72-741. Amsterdam: Elsevier Science, pp. 72-741. Petnicki-chieja, T., Schneider, D.J., Tam, V.C., Chancey, S.T., Shan, L., Jamir, Y., Schechter, L.M., Janes, M.D., Buell, C.R., Tang, X., Collmer, A., and Alfano, J .R. (2002) Genomewide identification of proteins secreted by the Hrp type III protein secretion system of Pseudomonas syringae pv. tomato DC3000. Proc. Natl. Acad. Sci. USA, 99, 7652-7657. Plotnikova, J .M., Rahme, LG, and Ausubel, F.M. (2000) Pathogenesis of the human opportunistic pathogen Pseudomonas aeruginosa PA14 in Arabidopsis. Plant Physio]. 124, 1766-1774. Quigley, N.B., Mo, Y.Y., and Gross, DC. (1993) SyrD is required for syringomycin production by Pseudomonas syringae pathovar syringae and is related to a family of ATP-binding secretion proteins. Mol. Microbiol. 9, 787-801. Reid, J .L., and Collmer, A. (1986) Comparison of pectic enzymes produced by Erwinia chrysanthemi, Erwinia carotavara subsp. carotovora, and Erwinia carotavara subsp. atroseptica. Appl. Environ. Microbial. 52, 305-310. Riggs, R.D., Kim, KS, and Gipson, I. (1973) Ultrastructural changes in Peking soybeans infected with Heterodera glycines. Phytopathalagy, 63, 76-84. Roine, E., Saarinen, J ., Kalkkinen, N., and Romantschuk, M. (1997) Purified HrpA of Pseudomonas syringae pv. tomato DC3000 reassembles into pili. FEBS Lett. 417, 168- 172. Ronald, P.C., Salmeron, J .M., Carland, F.M., and Staskawicz, B.J. (1992) The cloned avirulence gene avrPta induces disease resistance in tomato cultivars containing the Pto resistance gene. J. Bacteria]. 174, 1604-1611. Schechter, L.M., Roberts, K.A., Jamir, Y., Alfano, J .R., and Collmer, A. (2004) Pseudomonas syringae type III secretion system targeting signals and novel effectors studied with a Cya translocation reporter. J. Bacterial. 186, 543-545. 33 Shao, F., Golstein, C., Ade, J ., Stoutemyer, M., Dixon, J .E., and Innes, R.W. (2003) Cleavage of Arabidopsis PBSl by a bacterial type III effector. Science, 301, 1230-1233. Skalamera, D., Jibodh, S., and Heath, MC. (1997) Callose deposition during the interaction between cowpea (Vigna unguiculata) and the monokaryotic stage of the cowpea rust fungus (Uramyces vignae). New Phytol. 136, 511-524. Staskawicz, B.J., Mudgett, M.B., Dangl, J .L., Galan, J .E. (2001) Common and contrasting themes of plant and animal diseases. Science, 292, 2285-2289. Strange, RN, and Scott, RR. (2005) Plant disease: A threat to global food security. Ann. Rev. Phytopatha]. 43, 83-116. Tao, Y., Xie, Z., Chen, W., Glazebrook, J ., Chang, H.S., Han, B., Zhu, T., Zou, G., and Katagiri, F. (2003) Quantitative nature of Arabidopsis responses during compatible and incompatible interactions with the bacterial pathogen Pseudomonas syringae. Plant Cell, 15, 317-330. Volksch, B., Weingart, H. (1998) Toxin production by pathovars of Pseudomonas syringae and their antagonistic activities against epiphytic microorganisms. J. Basic Microbial. 38, 135-145. Wei, Z.M., and Beer, S.V. (1993) HrpI of Erwinia amylavara fitnctions in secretion of harpin and is a member of a new protein family. J. Bacterial. 175, 7958-7967. Weiler, E., Kutchan, W.T.M., Gorba, T., Brodschelm, W., Neisel, U., and Bublitz, F. (1994) The Pseudomonas phytotoxin coronatine mimics octadecanoid signaling molecules of higher plants. FEBS Lett. 345, 9-13. Whalen, M.C., Innes, R.W., Bent, A.F., and Staskawicz, B.J. (1991) Identification of Pseudomonas syringae pathogens of Arabidopsis and a bacterial locus determining avirulence on both Arabidopsis and soybean. Plant Cell, 3, 49-59. Zeidler, D., Zéihringer, U., Gerber, I., Dubery, 1., Hartung, T., Bors, W., Hutzler, P., and Durner, J. (2004) Innate immunity in Arabidopsis thaliana: Lipopolysaccharides activate nitric oxide synthase (NOS) and induce defense genes. Proc. Natl. Acad. Sci. USA, 101, 15811-15816. Zipfel, C., Kunze, G., Chinchilla, D., Caniard, A., Jones, J .D., Boller, T., and Felix, G. (2006) Perception of the bacterial PAMP EF-Tu by the receptor EFR restricts Agrabacterium-mediated transformation. Cell, 125, 749-760. Zipfel, C., Robatzek, S., Navarro, L., Oakeley, B.J., Jones, J .D.G., Felix, G., and Boller, T. (2004) Bacterial disease resistance in Arabidopsis through flagellin perception. Nature, 428, 764-767. 34 Zwiesler-Vollick, J ., Plovanich-Jones, A.E., Nomura, K., Bandyopadhyay, S., Joardar, V., Kunkel, B.N., and He, S.Y. (2002) Identification of novel hrp-regulated genes through fimctional genomic analysis of the Pseudomonas syringae pv. tomato DC3000 genome. Mal. Microbial. 45, 1207-1218. 35 Chapter 2 Genome-wide transcriptional analysis of the Arabidopsis thaliana interaction with the plant pathogen Pseudomonas syringae pv. tomato DC3000 and the human pathogen Escherichia coli 0157:117. This chapter has been previously published in The Plant Journal. Thilmony, R.,‘ Underwood, W.,. and He, S.Y. (2006) Genome-wide transcriptional analysis of the Arabidopsis thaliana interaction with the plant pathogen Pseudomonas syringae pv. tomato DC3000 and the human pathogen Escherichia coli 0157:H7. Plant J. 46, 34-53. I These authors contributed equally to this work. I would like to acknowledge Dr. Roger Thilmony for contribution of figures 2-3 and 2-4. 36 Abstract Pseudomonas syringae pv. tomato DC3000 (Pst) is a virulent pathogen, which causes disease on tomato and Arabidopsis. The type III secretion system (TTSS) plays a key role in pathogenesis by translocating virulence effectors from the bacteria into the plant host cell, while the phytotoxin coronatine (COR) contributes to virulence and disease symptom development. Recent studies suggest that both the TTSS and COR are involved in the suppression of host basal defenses. However, little is known about the interplay between the host gene expression changes associated with basal defenses and the virulence activities of the TTSS and COR during infection. In this study, we used the Affymetrix full genome chip to determine the Arabidopsis transcriptome associated with basal defense to Pst DC3000 hrp mutants and the human pathogenic bacterium Escherichia coli 01 57:H7. We then used Pst DC3 000 virulence mutants to characterize Arabidopsis transcriptional responses to the action of hrp-regulated virulence factors (e. g., TTSS and COR) during bacterial infection. Additionally, we used bacterial fliC mutants to assess the role of the PAMP flagellin in induction of basal defense-associated transcriptional responses. In total, our global gene expression analysis identified 2,800 Arabidopsis genes that are reproducibly regulated in response to bacterial pathogen inoculation. Regulation of these genes provides a molecular signature for Arabidopsis basal defense to plant and human pathogenic bacteria, and illustrates both common and distinct global virulence effects of the TTSS, COR, and possibly other hrp-regulated virulence factors during Pst DC3000 infection. 37 Introduction Pseudomonas syringae strains collectively infect hundreds of taxonomically diverse plant species and cause disease symptoms ranging from leaf spots to stem cankers (Hirano and Upper, 2000). P. syringae pv. tomato (Pst) strain DC3000 used in this study cause necrotic lesions that are ofien surrounded by chlorotic halos in susceptible tomato and Arabidopsis plants (Katagiri et al., 2002; Ma et al., 1991; Whalen et al., 1991). To successfully colonize plants, P. syringae strains have evolved a variety of virulence factors to subvert host defenses or to obtain nutrients. One common virulence mechanism is the hrp-gene-encoded type 111 protein secretion system (TTSS; He et al., 2004; J in et al., 2003). The TTSS is used by P. syringae to inject > 40 virulence effector proteins into the host cell (Chang et al., 2005; Collmer et al., 2002; Greenberg and Vinatzer, 2003; Nomura and He, 2005). Different P. syringae strains also produce a variety of phytotoxins (Bender et al., 1999). Athough phytotoxins are generally not required for bacterial pathogenicity, they do enhance pathogen virulence in host plants (Bender et al., 1999). Pst DC3000, for example, produces a polyketide toxin, coronatine (COR), which is required for full virulence in Arabidopsis and tomato plants (Brooks et al., 2004; Ma et al., 1991; Mittal and Davis, 1995; Zhao et al., 2003). Emerging evidence suggests that the production of the TTSS and COR is coordinately regulated in Pst DC3000 (Fouts et al., 2002; Penaloza-Vazquez et al., 2000) so that mutations in the regulatory hrp genes (e. g., hrpS, hrpL, and hrpA) could affect the expression of both the TTSS and COR biosynthetic enzymes. How TTSS effectors promote bacterial pathogenesis is poorly understood and remains an outstanding question in biology. In a previous study, we conducted 38 microarray analysis examining the effects of wild-type Pst DC3000 and its hrpS mutant (defective in hrp regulon expression producing a nonfunctional TTSS), Pst DC3118 COR' mutant (defective in COR toxin production), or the COR'hrpS double mutant (defective in the TTS S, COR toxin and potentially other hrp regulon controlled virulence factors) on the expression of about 7,200 Arabidopsis genes in the compatible Arabidopsis-Pst DC3000 interaction (Hauck et al., 2003). That study revealed that the TTSS is involved in biased suppression of Arabidopsis genes that encode putatively secreted proteins (Hauck etal., 2003). The TTSS-mediated suppression of these Arabidopsis genes is correlated with the ability of the TTSS and its effector AvrPto to suppress callose deposition (Hauck et al., 2003) and presumably other host responses that are associated with plant basal defense to hrp mutant bacteria (Bestwick et al., 1995) or to “pathogen-associated molecular patterns” (PAMPs) such as bacterial flagellin or EF- Tu (Gomez-Gomez et al., 1999; Kunze et al., 2004). However, it remains to be determined what role Type III secreted effectors play in the transcriptional regulation of other host physiological processes important for pathogenesis. The molecular mechanism by which COR facilitates Pst DC3000 virulence is also not well understood. COR shows a remarkable structural similarity to the plant hormone methyl jasmonic acid (MeJ A), which is involved in defense and wound response signaling (Bender et al., 1999; Benedetti et al., 1995; Feys et al., 1994; Uppalapati et al., 2005). Identification of the Arabidopsis JA/COR perception mutant coil, which is insensitive to both MeJA and COR, further suggests a similar mode of action of COR and MeJA (Xie et al., 1998; Kloek et al., 2001). The C011 gene encodes a subunit of an SCFC011 complex (E3 type ubiquitin ligase) presumably involved in proteasome-mediated 39 plOlC al.,l coli an} that prec host cult firs 2011 flag det int protein degradation (Devoto et al., 2002; Turner et al., 2002; Xu et al., 2002). Kloek et al., (2001) and Zhao et al., (2003) observed hyperexpression of a defense gene, PR-I, in P. syringae-inoculated caiI Arabidopsis and tomato plants, respectively. Resistance in caiI Arabidopsis plants is apparently caused by elevated SA-mediated defense, because a cat] Arabidopsis mutant expressing the nahG gene, which encodes an SA-degrading enzyme, fails to restrict bacterial growth (Kloek et al., 2001). These observations suggest that COR is involved in the suppression of an SA-dependent host defense. However, the precise nature of the COR-targeted host defense and the contribution of COR to global host gene expression during bacterial infection have not yet been determined. Recently, gene expression studies were performed using Arabidopsis suspension- cultured cells or seedlings treated with purified flagellin peptide (flg22), providing the first glimpse into the global gene expression changes during basal defense (Navarro et al., 2004; Zipfel et al., 2004). Whether the same set of Arabidopsis genes are regulated by flagellin and other PAMPs during bacterial infection of intact plant leaves remains to be determined. Recent studies suggest that the TTSS and COR have overlapping functions in the suppression of some PR genes and induction of some wounding response genes (He et al., 2004; Zhao et al., 2003). The extent of such overlap, however, has not yet been determined at the whole genome scale. Nor is it known how many of the PAMP-induced genes are subjected to transcriptional suppression by the TTSS and/or COR, respectively. In this study, we conducted genome-wide gene expression analysis of Arabidopsis plants treated with defined Pst DC3000 and E. coli 0157:H7 mutants to gain molecular insights into 1) the transcriptional changes associated with basal defense to live bacteria, which carry multiple PAMPs, 2) the contribution of flagellin to the regulation of the basal 4o defense transcriptome during infection, and 3) the global effects of hrp-regulated virulence factors, primarily the TTSS and COR, on the basal defense transcriptome and other host physiological processes. Previous global gene expression analyses have been performed in resistant Arabidopsis-Pst interactions (Tao et al., 2003) and other defense- related treatments (reviewed in Wan et al., 2002). We have conducted genome-wide analyses aimed at understanding the transcriptional responses of Arabidopsis plants to PAMPs, TTSS effectors, and COR from the same bacterium during infection. The results should complement other studies and provide a useful resource for future study of Pst DC3000 pathogenesis in plants. Experimental Procedures Plant Growth, Bacterial Strains and Bacteria Enumeration. Arabidopsis thaliana accession Col-O gl] plants were grown in soil in grth chambers with a day/night cycle of 12 h/12 h, a light intensity of 100 uE, and a constant temperature of 20°C. F our-week-old plants were used for experiments. Each biological replicate contains plants that were grown, inoculated and harvested side-by-side in the grth chamber entirely independent of the other replicates (typically separated in time by several weeks or months). Bacteria for inoculation were grown in low-salt Luria— Bertani broth (Katagiri et al., 2002) to the mid-logarithmic phase at 30°C. Bacterial cultures were centrifuged to recover bacteria, which were resuspended in sterile water to a final OD6oo of 0.002 (equivalent to 1x106 bacteria/mL), 0.1 (equivalent to 5x107 bacteria/mL) or 0.2 (equivalent to 1x108 bacteria/mL). Dilution plating was used to 41 confirm the number of bacteria present in the inoculum. The mock inoculum was water containing 0.004% Silwet L-77 surfactant (Osi Specialties, Friendship, WV). Fully expanded leaves were vacuum-infiltrated with bacterial suspensions, and in planta bacteria were enumerated as described by Katagiri et al., 2002. The mean values of the bacterial populations are plotted with the standard deviation displayed as error. The leaves of 10-15 plants grown in several different pots were harvested for each RNA sample used for microarray and RNA blot analysis. The regulation/secretion-defective Pst hrpS mutant, the Pst DC3118 COR'hrpS double mutant, the secretion-defective Pst hrpA mutant, E. coli 01 57:H7 and the E. coli TUV86-2 fliC mutant have been previously described (Berin et al., 2002; Hauck et al., 2003; Hayashi et al., 2001; Yuan and He, 1996). Inoculation with E. coli strains was staggered with the 108 dataset such that replicate number 1 of inoculations with E. coli OlS7:H7 and TUV86-2fliC was performed alongside replicate number 2 of the other 108 inoculations. To complete 3 replicates for each E. coli strain, we performed a final set of inoculations consisting of E. coli OlS7:H7, E. coli TUV86-2fliC, and mock inoculum. Therefore, in the publicly available dataset in the NASCArrays database, the E. coli inoculations are annotated as replicates 2, 3, and 4 to correspond with the other 108 inoculation data to which they are directly compared. RNA Isolation and Labeling Total RNA was isolated from Arabidopsis leaves using the Promega (Madison, WI) RNAgents total RNA isolation system. The RNA concentration was determined by absorbance at 260 nm and its quality was evaluated by separation on 2% formaldehyde 42 denaturing agarose gels. The total RNA was further purified using Qiagen (Valencia, CA) RNeasy minicolumns. Biotinylated target complementary RNA (cRNA) was prepared from 16ug of total RNA using the procedure described by the Affymetrix (Santa Clara, CA). Briefly, a primer encoding a T7 RNA polymerase promoter sequence fused to (dT)24 (Genset Oligos, La Jolla, CA) was used to prime double-stranded cDNA synthesis using the Invitrogen (Carlsbad, CA) SuperScript II Reverse Transcriptase. The resulting cDNA was transcribed in vitro using the BioArray High-Yield RNA Transcript Labeling Kit (Enzo Biochem, New York, NY) in the presence of biotinylated UTP and CTP to produce biotinylated target cRNA. Affymetrix GeneChip Hybridization and Data Collection The labeled target cRNA was purified, fragmented, and hybridized to Arabidopsis ATHl GeneChip arrays according to protocols provided by the manufacturer (Afiymetrix) in a Hybridization Oven model 640 (Affymetrix). The GeneChips were washed and stained with streptavidin-phycoerythrin using a GeneChip Fluidics Station model 400 and then scanned with a Gene Array Scanner (Hewlett-Packard, Palo Alto, CA). Affymetrix GeneChip data analysis The gene expression data was normalized and analyzed using the Affymetrix Microarray Program Suite (MAS 5.0 statistical algorithms and the Data Mining Tool (version 2.0)). Output from all GeneChip hybridizations was scaled globally such that its average intensity (TGT value) was equal to an arbitrary target intensity of 500 allowing 43 comparisons between GeneChips. Data was then compared between sample chips from the same biological replicate producing a Signal Log; Ratio (SLR) calculated from the GeneChip fluorescence signal intensity data. The software was used to determine whether there was a genuine change in mRNA accumulation (Change Call, D for Decrease, I for Increase) and Change Call p-value. SLRs, Change Calls and p-values were determined for each bacterially inoculated sample compared with its corresponding mock control or bacterial mutant treated sample. Six pair wise comparisons were made within the 1x106 bacteria/mL dataset, 3 pair wise comparisons for the 5x107 bacteria/mL dataset and 8 pairwise comparisons for the 1x108 bacteria/mL dataset (see Table 1). We used Significance Analysis of Microarrays (SAM; Tusher et al. 2001) to identify significant genes based on their differential expression between our set of samples. The SAM analysis feature contained in the TIGR Microarray Experiment Viewer v3.1 (Saeed et al., 2003; http://www.tm4.org/mev.htmL) was used to conduct the analysis. The normalized signal log; ratios for each replicate of each comparison were used for the analysis. Each inoculation level dataset (7 HPI 1x108, 10 HPI 5x107 and 24 HPI 1x106) was independently analyzed using the multiclass design, with the default settings. The delta value was set at level such that the median False Discovery Rate was less than 5%. This analysis identified a nonredundant list of 3,864 probesets. Microsofi (Redmond, WA) Access database management software was used to further filter and query the data. Reproducibly differentially expressed probe sets were selected from this list of 3,864, based on signal log; ratio of at least -1 .0, a gene expression change call of D (decrease) and p-value >0.995 or signal log; ratio of at least 1.0, change call of I (increase) and p-value of <0.005 for all 3 biological replicates for at 44 least one type of comparison. Probe sets that meet these rigorous selection criteria were further analyzed in detail. To our surprise, not a single probeset was reproducibly differentially regulated in the 1x106 bacteria/mL Pst COR'hrpS verses mock comparison (Table 1). This result illustrates the stringency of our selection criteria, because one would expect that a few probesets might be reproducibly differentially expressed just by chance. Our failure to identify Pst COR'hrpS (PAMP) differentially regulated genes following inoculation with 1x106 bacteria, suggest that the level of bacteria in this treatment is insufficient to elicit robust, detectable changes in host gene expression 24 HPI and supports our rationale to use higher doses of bacteria to identify PAMP- regulated gene expression in these comparisons. Hierarchical clustering and viewing was performed using either the Cluster v3.0 and TreeView programs (Eisen et al., 1998; http://rana.lbl.gov/EisenSofiware.htm) or the TIGR Microarray Experiment Viewer v3.1. The hierarchical clusters displayed in Figures 1 and 2 were analyzed using the Microarray Experiment Viewer with the default similarity metric settings (Euclidean Distance) with the Complete Linkage Clustering method selected. MapMan v1.6.0 (Thimm et al., 2004; Usadel et al., 2005) was used for analysis of the functional classes and metabolic pathways impacted following bacterial inoculation. The 944 COR toxin regulated genes were identified from the 1x106 bacteria/mL Pst DC3000 vs Pst COR' comparison, while the 791 hrp-regulated genes were identified from the 5x107 bacteria/mL Pst COR' vs Pst COR'hrpS comparison using the criteria listed above. The SLRs for the 3 biological replicates in the selected comparison were averaged and displayed using the MapMan software onto the Metabolism Overview, Secondary Metabolism and Photosynthesis displays. The 45 Nor the: its; W I also auto idea 56;! At; Nottingham Arabidopsis Stock Centre (NASC) web site (htygz//nasc.nott.ac.uk/) contains the raw microarray data in a MIAME compliant format for all of the experiments described, reference NASCArrays-340. RNA isolation and blot analysis Total RNA was isolated from Arabidopsis leaves using the Promega (Madison, WI) RNAgents Total RNA isolation system. The RNA concentration was determined by absorbance at 260 nm and separated on 2% formaldehyde denaturing agarose gels. The RNA was blotted onto Amersham (Piscataway, NJ) Hybond N+ nylon membranes using 10xSSC and UV crosslinked using a Stratagene (La Jolla, CA) Statalinker using the autocrosslink setting. The cDNA sequences of 15 differentially expressed genes identified by microarray analysis were used for RNA blot hybridization. Fourteen of the sequences were derived from ESTs acquired from the Arabidopsis Biological Resource Center stock center. Upon receipt of the ESTs, they were sequenced to confirm their identity. The cDNA probe sequences were liberated from the cloning vector using restriction digestion or PCR amplification with vector primers. The following cDNA sequences were used for labeling and RNA blot hybridization: At3 g163 70 from EST 175018T7, At1g12090 from EST 135D13T7, At2g38540 from EST 135H16T7, At1g29670 from EST 240F18T7, At2g19860 from EST E1C7T7, At1g03 870 from EST 172M22T7, At3g16240 from EST 206H6T7, At1g33240 from EST 88L16XP, At2g41940 from EST 163F20T7, At5g61590 from EST 215C17T7, Atlg72610 from EST 85C7T7, At2g10940 from EST 40D7T7, At5g24780 from EST 114D3XP, At1g52890 from EST 165P18T7, Atlg43160 from EST 132C14T7, At5g26340 from 46 ES“ 31: 99‘; fmr 10L and R63 R21 EST 181G24T7, At2g24850 from EST 245M18T7, At3g45140 from EST 304E6T7, At2g17500 from EST 178E2T7, Atl g62660' from EST 145A20T7, At4g27410 from EST 99D19T7, At4g02380 from EST E5E10T7, At4g02940 from EST 251H23T7, At3g16390 from EST 215J18T7. A 1.15 kb cDNA sequence for CLHI/CORII (At1g19670) was amplified with RT-PCR using the following primers: 5’- CAGAATTCAACACAACTCTTTAATTATC-3’ and 5’- ATCTCGAGTAACAAATGTTTTGATCGAGJ’. The amplified product was cloned into the EcoRI and XhoI sites of pBluescript SK+ and sequenced . cDNA sequences from the above listed clones were labeled with 32P-dCTP using the Stratagene Prime-It 11 Random Primer Labeling kit. RNA blot hybridizations were performed using Sigma (St. Louis, MO) PerfectHyb Plus hybridization buffer following the manufacturers protocol and then washed to 0.5x SCC and exposed to film. Results and Discussion Rationale and Experimental design All the experiments described were conducted using the Affymetrix ATHl GeneChip and four-week-old Arabidopsis leaves. For discovery of PAMP- TTSS- and COR-regulated genes, we used Pst DC3000, Pst DC3118 COR' mutant, and the Pst DC3118 COR‘hrpS double mutant bacteria, as described previously (Hauck et al., 2003). Because some of the strains have different growth potentials in plants, we used three different levels of bacterial inocula for gene expression profiling: low (1x106 bacteria/mL), moderately high (5x107 bacteria/mL) and high (1x108 bacteria/mL) (Table 47 2-1). At 1x106 bacteria/mL, Pst DC3000 reliably causes disease within 3 to 4 days and is used in this and other laboratories as a standard disease assay (Katagiri et al., 2002). This inoculum level has also previously been used to identify hundreds of Pst-regulated genes (Hauck et al., 2003). Gene expression changes at this inoculum level were examined after a 24 hour infection period (Figure A-l). During this time period, Pst DC3000 and the Pst COR’ mutant multiply similarly to levels 40-to-50-fold higher than that of the Pst hrpA, hrpS and COR'hrpS mutants (data not shown). Therefore, to minimize the effects the different bacterial population numbers have on host gene expression, and to examine samples with early, synchronous responses, we also used inoculum concentrations of 1x108 and 5x107 bacteria/mL. In these cases, the tissues could be harvested earlier (at 7 hours and 10 hours after inoculation respectively) when development of basal defense- associated papillae is occurring (Hauck et al., 2003). At the inocula and time points chosen for various strains, high and comparable expression levels of a known PAMP- induced gene, F lagellin-Induced Receptor Kinase 1 (F RKI ), were observed (Figure A-2). In total, we conducted eight pair wise comparisons for the 1x108 dataset, three for the 5x107 dataset, and six for the 1x106 dataset. A summary of the number of genes that passed our selection criteria (see Experimental Procedures) and were reproducibly differentially regulated in all 3 biological replicates for each comparison is shown in Table 2-1. These comparisons identified genes that are responsive to bacterial PAMPS of Pst hrp mutants and E. coli 01 57:H7 as well as those that are regulated by Pst DC3000 TTSS effectors and/or COR toxin (Table 2-1). Our global transcriptional profiling identified 2800 genes that were reproducibly differentially regulated. 48 Table 2-1: Arabidopsis genes reproducibly regulated by bacterial inoculation. Repressed Induced Comparlson Treatmenta Genes Genesb Total Expression Pattern Identified 1. Pst hrpA vs. mock W, 7 hrs 224 253 477 PAMP regulation 2. Pst hrpA fliC vs. mock 108. 7 hrs 5 95 100 PAMP regulation 3. Pst hIpA vs. Pst hrpA fliC 108. 7 hrs 11 9 20 Flagellin-specific regulation 4. E. coli O157:H7 vs. mock 108. 7 hrs 96 186 282 PAMP regulation 5. E. coli roves-2 fliC vs. mock 108. 7 ms 240 218 458 PAMP regulation 6. E. coli O157:H7 vs. E. coli TUV-86-2 MO 108, 7 hrs 0 0 0 Flagellin-specific regulation 7. Pst 003000 vs. mock 108, 7 hrs 901 495 1396 Pseudomonas regulation 8. P51003000 vs. Pst thA 108. 7 hrs 904 415 1319 Pst virulence factor regulation 1270 797 1901 Non-redundant Totals 9. Pst coa'mps vs. mock 5x107, 10 hrs 197 154 351 PAMP regulation 10. Pst COR' vs. mock 5x107, 10 hrs 829 399 1228 Pseudomonas regulation 11. Pst COR' vs. Pst COR-hrpS 5x107. 10 hrs 516 275 791 Type III effector Egulation 1018 549 1537 Non-redundant Totals 12. Pst COR’hrpS vs. mock 106. 24 hrs 0 o o PAMP regulation 13, PstCOR' vs. mock 106, 24 hrs 212 297 509 Pseudomonas regulation 14. Pst COR' vs. Pst COR'hrpS 106. 24 hrs 175 253 428 Type III effector regulation 15. Pst DC3000 vs. Pst COR' 106. 24 hrs 347 597 944 COR toxin regulation 16. Pst DC3000 vs. mock 106. 24 hrs 685 680 1365 Pseudomonas regulation 17. Pst DC3000 vs. Pst COR-hrpS 106, 24 hrs 777 700 1477 Pst virulence factoleation 955 1018 1929 Non-redundant Totals 1724 1430 2800 Overall Non-redundant Totals aSummary data from plants 7 hours after inoculation with 1x108 bacteria/mL (rows 1-8), 10 hours after inoculation with 5x107 bacteria/mL (rows 9-11), and 24 hours after inoculation with 1x10‘5 bacteria/mL Pst bacteria (rows 12-17) are shown. bCriteria for gene selection is described in the text. 49 \V ’et PAMP-induced transcriptional changes In order to identify the transcriptional changes resulting from PAMP perception, we inoculated Arabidopsis with three different bacteria: Pst DC3000 hrpA mutant, Pst DC3000 COR'hrpS mutants, and human pathogenic E. coli 01 57:H7. The Pst hrpA and COR'hrpS mutants lack the ability to secrete TTSS effectors into the host and are unable to multiply significantly in plant host leaves or produce disease symptoms. Furthermore, the COR-hrpS mutant is unable to produce COR toxin and similarly, the hrpA mutant is also likely defective in COR toxin production via a feedback mechanism. The hrpA mutant strain has repressed expression of the hrpS and hrpL genes, which globally control the expression of genes involved in the production of the TTSS, COR toxin, and possibly other virulence factors (F outs et al., 2002; Wei et al., 2000). Therefore, in the hrpA and COR'hrpS mutants, many, if not all, plant-specific virulence factors are not produced. E. coli 01 57:H7, a human pathogen, is also unable to multiply or cause any symptoms on Arabidopsis leaf tissue (data not shown) and is not expected to carry plant- specific virulence factors. We hypothesized that inclusion of these three bacterial strains in the analysis should give us a global view of common and distinct plant transcriptional responses through the perception of PAMPS displayed by human and plant pathogenic bacteria. To assess the contribution of flagellin perception in host gene regulation during bacterial infection, we examined Arabidopsis gene expression regulated by the hrpA fliC double mutant of Pst DC3000 (Hu et al., 2001) and a fliC mutant derivative of E. coli TUV86-2 (Gunzer et al., 1998). We identified a set of 736 PAMP-responsive genes that were differentially regulated at least 2-fold in 3 biological replicates for at least 1 out of the following 3 50 treatments with non-pathogenic bacteria: Pst hrpA vs mock 108 bacteria/mL, E. coli OlS7:H7 vs mock 108 bacteria/mL, and Pst COR'hrpS vs mock 5x107 bacteria/mL. A tally of the number of Arabidopsis genes regulated in response to each of the three bacterial treatments is presented in Table 2-1. Of the 736 reproducibly regulated genes, 347 were induced and 389 were repressed. Because we are particularly interested in genes whose products may be involved in the induction or elaboration of basal defenses, we chose to focus our analysis primarily on the set of 347 PAMP-induced genes. Hierarchical cluster analysis of PAMP-induced genes To provide an overview of the regulation of the 347 PAMP-induced genes across our set of treatments and biological replicates and to visualize the regulation of these genes in response to inoculation with virulent Pst DC3 000, we performed hierarchical clustering. The resulting cluster is displayed in Figure 2-1. The induction of Arabidopsis genes after inoculation with E. coli was strikingly similar to the induction elicited by the Pst hrp mutants. The correlation coefficients based on the PAMP induced genes for each of the 1x108 and 5x107 comparisons are shown in Table A-3. Although Pst hrpA inoculation resulted in the reproducible regulation of 195 more genes (that matched or exceeded our selection criteria) than inoculation with E. coli 01 57:H7 (Table 2-1), the expression profiles were remarkably similar for the two treatments as shown in Figure 2- 1 and their correlation coefficient was 0.79. We are particularly interested in identifying genes whose products may play a role in the plant basal defense response. Virulent Pst DC3000 is able to overcome or evade this host basal defense response and may do so by repressing or blocking the induction of host genes whose products are involved in the 51 elaboration of these defenses. Therefore, we expected that some PAMP-induced genes would be either repressed or simply not reproducibly induced upon inoculation with Pst DC3000. The blue bar in Figure 2-1 illustrates a cluster of genes that exhibit an expression pattern consistent with a role in basal defense. A total of 201 of the 347 PAMP-induced genes were either repressed or not significantly regulated (average fold change of <1.5) in response to Pst DC3000 inoculation. We propose that these genes may have a role in the Arabidopsis basal defense response to bacteria. 52 1 x 10‘ cfu/mL 5 x 107cfu/mL A B C D E F G H I '1‘ up: i 1 III ll I IQI! int. i l an ill Figure 2-1: Expression profile of 347 PAMP-induced Arabidopsis genes. Treatments: A- E. coli v mock; B- E. calfliC v mock; C- hrpA v mock; D- hrpAfliC v mock; E- DC3000 v mock; F- DC3000 v hrpA; G- corhrpS v mock; H- cor v mock; I- cor v carhrpS. Genes highlighted by the blue bar on the right are of particular interest as these genes are induced by PAMP perception, but are not reproducibly induced or are repressed by Pst DC3000 and/or Pst COR'. 53 Transcription factors and signaling components are induced in response to PAMPs. Of the 347 PAMP-induced Arabidopsis genes, 86 encode known or predicted transcription factors and kinases. Thirty transcription factors are reproducibly induced in response to PAMP perception, 17 of which are repressed or not significantly regulated in response to DC3000 (Table 2-2A). Of these 30 PAMP-induced transcription factors, 12 belong to the WRKY family and 8 of these are induced by PAMPS but not by Pst DC3000. WRKY22 (At4g01250) was previously shown to participate in the F LS2- mediated signaling cascade triggered by perception of the synthetic flagellin peptide flg22 (Asai et al., 2002). Additionally, 6 of these 12 WRKY transcription factor genes (shaded gray in Table 2-2A) were shown to be induced in Arabidopsis seedlings in response to treatment with flg22 (Zipfel et al., 2004). These data suggest that WRKY family transcription factors play an important role in regulating the Arabidopsis basal response to bacteria. Determining the target genes regulated by these PAMP-induced WRKY transcription factors should further our understanding of the processes involved in the plant basal defense response. In addition to WRKY transcription factors, 9 known or putative zinc finger transcription factors and 4 ERF/AP2 family transcription factors were induced by PAMPS. Two of the four ERF/AP2 family transcription factors and 6 of the 9 zinc finger proteins were previously shown to be induced by flg22 (Zipfel et al., 2004). Interestingly, the ethylene responsive AtERFI transcription factor was found to be induced by PAMPS. The plant hormone ethylene has been shown to play a role in plant defense against pathogens, and AtERFI was also found to be induced in response to flg22 treatment. However, our study shows that AtERF 1, while reproducibly induced approximately 2- 54 fbld b: pa'lrog induct ngnal flock: fllustr Pill) by P:\ inducc genes induce found hnase Innsll (illil irahh ' fold by PAMP perception, is even more strongly induced (~10-fold) by the virulent pathogen Pst DC3000. These data suggest that ethylene signaling may play a role in the induction of basal defenses, but that the basal defense responses triggered by ethylene signaling are either insufficient to prevent Pst DC3000 pathogenesis or that Pst DC3 000 blocks ethylene signaling or responses downstream of AtERF I induction. This example illustrates the value of determining the expression profile of these genes in response to Pst DC3 000 in addition to PAMPS rather than simply identifying genes that are induced by PAMPS. In addition to transcription factors, a number of other signaling components were induced by PAMP perception. Particularly abundant among PAMP-induced genes were genes known or predicted to encode kinases. F ifiy-six known or putative kinases were induced by PAMP perception (Table A-l). Of these 56 PAMP-induced kinases, 31 were found to be induced by flg22 treatrnent (shaded gray in Table A-l). Receptor-like kinases, including lectin family kinases, and wall-associated kinase-like proteins are the most highly represented classes of kinases induced by PAMPS. Wall-associated kinases (WAKs) and WAK-like kinases (WAKLs) have recently been implicated in the Arabidopsis response to pathogens. AtWakI was shown to be induced by P. syringae and SA (He et al., 1998) and WAKl was recently shown to interact with the glycine rich protein AtGRP3 and to phosphorylate oxygen evolving enhancer protein 2 (OEE2; Yang et al., 2003). Treatment of Arabidopsis protoplasts with purified AtGRP3 induced PR-I gene expression and inoculation of plants with avirulent Pst DC3000 enhanced WAKl - dependent phosphorylation of OEE2, suggesting that WAKl and GRP3 may play a role in the response to pathogens. Additionally, the WAK-like kinase WAKL22 was recently 55 found of i501 Al'thOt the pn recent role it recept Recog pereei PR 1 Zeidit is an I PRPs Path“ lCSpo rttep PAM paths DPICa “lib ll found to be identical to the resistance gene RfoI which confers to resistance to a number of isolates of the fungal pathogen F usarium axysparum (Diener and Ausubel, 2005). Although neither WAKI nor WAKL22 were specifically induced by PAMPS in our study, the prevalence of WAK-like kinases among PAMP-induced genes taken together with the recent results highlighted above suggest that this class of proteins may play an important role in the plant response to pathogens. Also of particular interest are a number of leucine-rich repeat (LRR)-containing receptor-like kinases. These receptor-like kinases may be candidates for Pattern Recognition Receptors (PRRs) in Arabidopsis. Although Arabidopsis is known to perceive a number of bacterial PAMPS including flagellin, LPS, and EF-Tu, only a single PR for bacterial PAMPS has been identified (Felix et al., 1999; Kunze et al., 2004; Zeidler et al., 2004). FLSZ, the PR which mediates flagellin perception in Arabidopsis, is an LRR receptor kinase (Gomez-Gomez and Boller, 2000). Identifying additional PRRs in Arabidopsis is a high priority and should lead to new insights into the signaling pathways initiated by perception of bacterial PAMPS and the downstream basal defense responses that protect plants against a wide array of bacteria. These PAMP-induced LRR receptor-like kinases provide a starting point in the search for other Arabidopsis PRRs. PAMPs induce genes that may be involved in secretion and cell wall modification. One of the most well characterized aspects of plant basal defense against pathogens is modification of the cell wall. Plants challenged with bacteria or fungi typically develop cell wall appositions referred to as papillae at the point of interaction with the invading pathogen. Formation of papillae is characterized by the deposition of 56 elecuolt i' 15h) 1101“. the test of proper lot the plant . 'tfioogen polarized potential: lonnatio: Ellit‘ptc be met Suretut during Peron, linkin Patio “373g 3484/ I 0o...» electron dense materials such as phenolic compounds, callose, and cell wall proteins such as hydroxyproline-rich glycoproteins. Callose is synthesized at the plasma membrane, but the rest of the components are synthesized within the cell and must be delivered to the proper location at the wall for deposition into papillae. Because of this, components of the plant secretory system are likely to play an essential role in basal defense against pathogens. Consistent with the predicted importance of cell wall modification and polarized secretion to basal defense, we found that a number of PAMP-induced genes are potentially involved in these processes. Table 2-2B contains 28 PAMP-induced genes that may have a role in cell wall modification or polarized secretion. Among these 28 genes are a large number of genes encoding putative hydrolase enzymes that may potentially play a role in reorganizing the structure of the cell wall during papilla formation. Additionally, 4 genes encoding proline-rich and hydroxyproline-rich glycoproteins are induced by PAMPS. Seven putative peroxidases are induced which may be involved in the oxidative cross-linking of proteins and polymers into the papilla structure. The peroxidase AT P24a (At5g39580) was previously shown to be upregulated during the RPP4 and RPP8-mediated resistance responses to the oomycete pathogen Peranaspora parasitica (Eulgem et al., 2004). Consistent with a role in oxidative cross- linking in the cell wall, all of these peroxidases are predicted to enter the secretory pathway by TargetP (www.cbs.dtu.dk/services/TargetP). Strikingly, 3 genes encoding potential components of the secretory pathway are upregulated by PAMP perception. Two members of the RabA family of small GTPases, RABA I a and RABA I e, were induced by PAMPS. Rab proteins are a class of small GTPases that also function in vesicle trafficking and regulate vesicle docking. Based on 57 work in post-go from p membl (At3gf su‘ount also St EXOT treatm induce these ; defen: may ; work in mammalian and yeast systems, this class of Rabs is predicted to participate in post-golgi vesicle trafficking (V emoud et al., 2003). Additionally, two RabA isoforrns from pea were suggested to have a role in delivery of cell wall components to the plasma membrane (Nagano et al., 1995). Also, a member of the EXO70 family of proteins (At3g55150) was upregulated in response to PAMPS. EXO70 family proteins are subunits of the exocyst protein complex which is required for exocytosis in yeast and is also suggested to play a role in docking of exocytotic vesicles (Cole et al., 2005). The EXO70 family member was previously found to be upregulated in response to flg22 treatment (Zipfel et al., 2004), however, RABAIa and RABA 1e represent novel PAMP- induced genes with functions related to secretion. Taken together, the upregulation of these genes supports a role for targeted vesicle trafficking in the Arabidopsis basal defense response to bacteria and identifies candidate secretory pathway components that may play a role in secretion of papilla components or other defense-related compounds. 58 Table 2-2: Selected PAMP- induced Arabidopsis genes. The average fold changes of each treatment compared to the mock control are shown. The hrpA, E. coli and DC3000 treatments were 7 hours post inoculation (HPI) with 1x108 bacteria/mL, while the COR hrpS treatment was at 10 HPI with 5x107 bacteria/mL. (A) Transcription factors induced by PAMP perception. (B) Genes encoding proteins known or predicted to be involved 1n cell wall modification or secretion. Lines within tables denote cutoff for PAMP-induced genes that are repressed or not significantly induced (less than 1.5-fold) by Pst DC3000. Genes below the line are also induced by DC3000. Grey shading indicates genes previously identified as induced by the bacterial PAMP flagellin. A Average Fold Change Array E. Element AGl Desert tlon h A COR'h S coll DC3000 254231 _af A14923810;'A1WRK¥534 ‘- ”'1' ' ' .5"? 1" Z 5, 2.09 "’ 2764 .. 2.00 30.56“ 267385_ at 1112944380 CHP—rl'ch zincfingerprotein putative: v . ~” 0 ‘_ ~ , 4.39 2.70 :. 4.00 '-6.06_,-_, 246993_ at A15g67450l AZF1CysZ/His2-typezincfingerprotein‘1 " n - .- ‘ ,1 2.96 ' 4.49 g 2.58 5 370,1 354159181. M9242LAMRWL 5 .5 4-6.x-” -f -55 . a. 42.70-125.58 421.48-232.51 255596_at At4gO1720 AtWRKY47 2.30 3.03 2.09 -2.58 25554731 At 01920 DC1 domain protein. similar to CHP-rich ' 1.55 2.41 1.38 -2.46 264616_ at AtZg17740 CHP- rich zinc fingeL protein, puta_t_ive 9.40 7461 J 579 -1. 38_1 255568:_at:A14901250 Ammo/2‘2 - : 1" “1'3—f :— 5.92 252 5.40 -1. 07 1 263797_ at Atzgzaszo AtWRKY17. .. f - ' " i . 276 .. 3,..03 3.25 1.00 3 2483.056 5apAt5052830-MNRKX27. 5 -5 5.- ' .5 5.14.459. 25.46;. 459511.11. 260783_ at At1gO6160 ERF subfamily ERF/APZtranscription factor family. 3.91 1.32 1.38 1.41 249690 at A15922570 AtWRKY36 5 5 1.41 4.19 1.62 1.48 25575’3_ at’ 1811916570" AMY§51 myb family trgpscription faotor T _ T ‘ 2.46 ‘— 274 :209 '1255 257919_ at 2113923250; AtMYB15rnyb fa'milytranscription faotor ~ . " 5.66 A. 2.64 2.96 1.74 ”t 25348.5. of 614931800 A'tWRKYte ' 5 . _‘ 1.78 2 3.17 . 1.62 200 ‘ 12.59153131_615515¢_3_95AMRKY§;2 .5 :-;:_..5 55-- .- . - _- .5 .-.-_..$2.9- 54922-5. 5.7.9.. 2.52.” 267588_ at At2942060 CHP-rich zinc finger protein, putative 6.81 7.46 4.81 2. 58 260432 __at At1_g68150 AtWRKYQ fl - 713 3 65 4 39 3. 25 1251948581 7111527730. AIZAItOsalttoleranceztncfinaflrpmtem_ _.-.5 2.00- 55 -230 755.246 5.882.. 246214_ at At4936990 AtHSF4 putative heat sh66k transcription_ factor 6 _-...... 317 .5. 6.17 2. 70 6. 81_ " 245252_at—At4912500-A15RF'1ewbnemsponsefaotoria f“ 2 '~ 2.565 5- 1.29 200 10.06; 245976_at A15g13060 AtWRKY75 7.29 3.17 3.73 11.56 DREB subfamily A5 of ERF/APZ transcription factor 252214_at At3g§0260 family. 1.87 2.56 1.78 13.30 267140_at A12938250 putative GT-1-Iike transcription factor 3.17 2.05 2.52 13.93 247264_at At5964530 no apical meristem (NAM) family protein 2.70 6.20 3.56 30.55 59 Table 2-2 (cont’d). B Avergge Fold Change Array E. Element 1 AG! n 4r“ -_ Description . hrpA COR ths coll DC3000 355524.9QM902339 hypothetical protein. jimilartopedinesterase . . 5. - .325. .8 3. 73......r 2.90.. 6.40 .i 245052,” at At2926449mput91ive pectinesteragg .__--5 “___ .5 _________4 _1 k2.__14___2_.96__1.74 __-_2_. 41 WW2!) .mmuymmma5 5--- .11.-.. 5.... 25.70.... “891:8”..4228-8 247954_ at At5_956870 hbeta-galactosidase. putative/lactase putative W . . 2. _‘24 _ 1. 23 4 2 70 M52129”. 3555M A14901700 dunnasa.puiauye,slmilartopeanuuypeii61111111858 _ 3. 03,-5.5 8.06 2.84 8.59.... 264669_ at At1909630 AtRABA1a Rab small GTPase 1.70 2.24 1.52 -1.52 259173 61 _At6gO6640 AtGLUC beta g_urcosidase_ .5.-- .--. .._...-..-...- ".1595...- 8389.1... "1. 74 -1.48 £59_4fi__t_At_110__2360 _ehitinaeefltativL555555 555.555 _55 _5 3.25 ”5,454.43 3517 «1.45;.- 253099_ at At4g37530 peroxidase putative .- ‘1.“ .91 2.19 . 2.14 M31. .05 551832.61. 858885158 mmubitiamzo mm... . 5. .1428... “.583.- 9:19 .8802; 354.943.23.88481-991959890.51! M52999 far-119185.991” tame-9.9199989. 53-69 - 97?... .859- 51.92...- 1260556_at At2943620 glycosyl hydrolase family 19 (chitinase) ' 1.29 6.35 1.55 1.07 i“ glycosyl hydrolase family 81; similar to beta-glucan-elicitor -' :246532_at A15915870 receptor 2.35 2.46 2.30 1.10 26549981 mmwoJMM9lueoawamteLase5-555 - 552.245 5314625 .1. 15 -. 254468_ at A14920460 ugff-glucose 4-epimerase-like grotein _ 3.73 2 30 3. 25 1.18 247327.318105954126 poiaxiaasq. putative. .. .8. 5. .5... .5 . .. 3.03 - 3.40 2858. 1.67:: 254673_at At4918430 AtRABA1e RAB GTPase 4.81 4.49 2.76 1.70 254262_at A14923470 hydroxyproline-rich giycoprotein family protein 1.32 2.52 1.35 2.14 264960_ at Attg76930 AtEXT1/4 Extensin hydroxyproline-rich glycoprotein 1.59 2.76 1.38 2.14 255111__ _a_t At4908780wperox1dase putative_ .__-___..." _ _.- 5..- “J _3262____ _6. 82 ___1. ”91 2. 60 AWL Wehmpmlinefimmnfl; - . .- .. - 4.81..- _._.3. 98;:4888: 2._89. proline-rich extensin-Iike family protein contains extensin- 259553__ at A11921310 like region; 55- 5 _W V HEM -__ ___ __ 666.656; _ 3.76 “6:26 _4 636 -- 261474_ at At1g14540 anionic peroxidase Similar to Iignin forming peroxidase 6.35 5.66 6.35 3.32 1 249489.61A15838588M248mmase -8..- _ -.._...- 8.1.85 .....919 ...8.7Z..A3.9...8 245148_ at 65132945229” _pectinestergse Lamilyw pr_otei11_ ______ 5--.--- 6.29. 5 3.40 4.70 4.92 _ 247467-81 1615682159 66888681768116de dommmimnwotem 53 25 2 76 a 03 81.988 246228_at At4936430 peroxidase. putative 11.06 3.40 11.5 8.98 255110_at At4908770 peroxidase, putative 54.44 9.40 14.5 24.82 60 Flagellin does not contribute uniquely to PAMP-induced transcriptional changes. Given the substantial overlap of the PAMP-induced genes identified by our microarray analysis with genes previously identified as induced by the synthetic flagellin peptide flg22 (131 of 347 PAMP-induced genes), we sought to determine the specific contribution of flagellin to the induction of these genes. To determine if perception of flagellin makes a specific contribution to the transcriptional changes observed in Arabidopsis afier bacterial inoculation, we compared the expression profiles of Arabidopsis plants inoculated with fliC mutant derivatives of Pst hrpA and E. coli OlS7:H7 to those of plants inoculated with the parent strains. fliC mutants are impaired in the production of the flagellin protein. Therefore, this comparison allows us to determine the specific contribution of flagellin perception to PAMP-induced transcriptional changes. A fliC mutant of E. coli OlS7:H7 was not available; therefore, a fliC mutant of the 01 57:H7 derivative strain TUV86-2 was used. To validate the use of this strain, we performed microarray analysis to compare the expression profiles of plants inoculated with E. coli 01 57:H7 or E. coli TUV86-2 and found that the expression profiles of these plants were highly similar (data not shown). The expression profiles of the 347 PAMP- induced genes from plants inoculated with fliC mutant bacteria were similar to those inoculated with the parental, flagellin-producing strains (Figure 2-1). The correlation values for the Pst hrpA vs mock and Pst hrpA fliC vs mock comparisons is 0.93, and for the E. coli vs mock and E. colifliC vs mock comparisons. it is 0.97. Consistent with the similarity apparent in the clustering analysis and the calculated correlation, we found that only 20 genes were reproducibly regulated more strongly by the flagellin-producing 61 strains than by the fliC mutant derivatives (Table 2-1). No genes were regulated more strongly by E. coli 01 57:H7 than by the fliC mutant derivative. However, when we compared the expression profile of plants inoculated with Pst hrpA fliC to mock inoculation, we found that using our selection criteria 377 fewer genes were reproducibly regulated by hrpA fliC than by hrpA (Table 2-1). This is a result of our stringent selection criteria as genes were regulated similarly, but not as robustly, by Pst hrpA fliC. This suggests that, at least in the case of hrpA, flagellin perception may make a quantitative contribution to the PAMP-induced regulation of transcription. However, this quantitative contribution is relatively minor as only 9 genes were reproducibly induced at least 2-fold more strongly by the flagellin-producing hrpA strain than by hrpA fliC (Table 2-1). These results suggest that although perception of flagellin induces transcriptional changes in Arabidopsis, flagellin does not make a substantial unique contribution to the PAMP- induced transcriptional changes observed after bacterial inoculation. Arabidopsis can perceive other PAMPS such as LPS and EF-Tu and our results suggest that perception of flagellin induces transcriptional changes that overlap almost completely with those induced by other PAMPS. Consistent with these results, multiplication of fliC mutants of Pst DC3000 and hrpA in Arabidopsis after vacuum-infiltration is similar to that of the flagellin-producing parent strains (Figure A-3). These results suggest that although flagellin may make a minor quantitative contribution to PAMP-induced transcriptional changes, it is only one of multiple PAMPS recognized by plants and that different PAMPS induce overlapping plant responses to bacteria within leaf tissue. The recent finding from Zipfel et al. that multiplication of Pst DC3000 is enhanced afier inoculation by dipping, but not after inoculation by vacuum-infiltration suggests that perception of bacterial 62 .Anal hrpS kxel indut 5x10 th TOge gene rEguf flagellin may make a unique contribution to defenses effective on the surface of the leaf, but does not make a unique contribution to elicitation of defenses when bacteria are infiltrated directly into the leaf intercellular spaces. Our data support this notion. Analysis of the TTSS- and COR-regulated genes To put the results of the PAMP gene regulation in perspective relative to a virulent pathogen capable of causing disease, and to identify host genes regulated by Pst hrp-dependent TTSS effectors and COR toxin, we examined host gene expression using wild-type Pst DC3000 and the virulence compromised mutants Pst COR' and Pst COR' hrpS at both low (1x106 bacteria/mL) and moderately high (5x107 bacteria/mL) inoculum levels (Table 2-1). Overall, we observed that relatively equal numbers of repressed and _ induced genes were identified in each comparison examined, with the exception of the 5x107 bacteria/mL Pst COR— vs mock and Pst COR' vs Pst COR' hrpS comparisons which identified approximately twice as many repressed genes as induced genes. Together the 5x107 and 1x106 datasets identified 2,439 host genes with Pst-responsive gene expression. Hierarchical cluster analysis of these 2,439 reproducibly differentially regulated genes is shown in Figure 2-2. Of these 2,439 genes, 1,027 (42%) were reproducibly differentially expressed in both datasets (Table 2—1; Figure 2-2). Inspection of Figure 2-2 illustrates that even though 1,412 of the genes were not contained in the lists of differentially regulated genes for both datasets (because they failed to surpass the selection criteria in both cases), many were regulated in a similar way. The overall reproducibility between the independent biological replicates is evident as indicated by the uniformity of the gene expression changes within the 3 biological replicates of each 63 comparison (Figure 2-2). COR toxin was a potent regulator, inducing or repressing the expression of 944 genes 24 HPI (Table 2-1 Row 4). The COR-induced genes comprise approximately 60% (597/101 8) of the reproducibly induced genes (Figure 2-2 clusters A1 and A2, Table 2-1), while only 347 genes are COR toxin repressed (Table 2-1). By comparing the samples inoculated with 5x107 Pst COR' and COR'hrpS bacteria/mL (to minimize bacterial population differences), a total of 791 genes exhibiting hrp/FT SS effector-dependent regulation were identified (Table 2-1, Row 11). These two strains lack COR and allow the identification of hrp/FT SS effector-dependent regulated genes, in the absence of the effect of COR toxin. Two hundred and seventy five of these genes exhibited hrp/Tr SS effector-dependent induction (Table 2-1; Figure 2-2, cluster B). Upon visual inspection of Figure 2-2, it is clear that many of the differentially expressed genes are coordinately regulated in response to Pst DC3 000 and the Pst COR' mutant (repressed by both strains or induced by both strains compared to the mock control), while other sets of genes are regulated in opposing directions. This result illustrates the importance of using both hrp and COR' mutant bacterial strains in discerning the specific effects of TTSS effectors and COR on host gene expression. Independent validation of the microarray expression data was pursued using RNA blot analysis. A total of 12 repressed and 13 induced genes were selected and cDNA probes for each gene were used for RNA blot hybridization. The results are shown in Figure A-l. The RNA samples used for blot analysis were from independent biological samples (different from those used for Affymetrix chip hybridization) and included a time-course following inoculation with 1x106 bacteria/mL. The results confirmed the expression patterns observed with the microarray and illustrate well that genes 64 differem early as [163111161 differentially expressed 24 HPI in our microarray experiments may exhibit changes as early as 6 hours and/or continue to be differentially expressed until at least 36 hours after treatment. 65 5x107 bacteria/mL 1x10‘5 bacteria/mL A B C D E F G H I -5“.- , ‘7 1. $577 «’17.... ‘ 'Ir 1’ W.,! ‘7 whiff" a J :1}:de 1‘. ll 1 I _I 82 q I 7' 1;? I J‘s-IA rein. Figure 2-2: The expression profile of the 2439 A rabidopsis genes reproducibly regulated by 1 x106 and/or 5x107 Pst bacteria/mL is shown. Treatments: A- corhrpS V mock; B- cor v mock; C- cor v corhrpS; D- corhrpS v mock; E- cor v mock; F- cor v corhrpS; G- DC3000 v cor; H- DC3000 v mock; G— DC3000 v corhrpS. 66 Impact of Pst pathogenesis on host metabolism An advantage of global gene expression profiling is the comprehensive view it can provide concerning the transcriptional regulation of genes involved in all aspects of plant physiology. In order to take advantage of this large dataset, we utilized MapMan to visualize how Pst pathogenesis impacted the expression of genes involved in various host metabolic pathways and processes (Thimm et al., 2004; Usadel et al., 2005). Some of the most dramatic global differences are illustrated in MapMan’s metabolism overview map displaying the 944 COR toxin regulated genes and 791 hrp-regulated genes. The most obvious bias within the COR-regulated genes is the induction of many genes involved in secondary metabolism (Figure 2-3A). Genes that encode enzymes in the phenylpropanoid, terpanoid, anthocyanin and glucosinolate pathways are all COR- induced (Figure 2-3A, Figure A-4). This result is consistent with recent data demonstrating that J A and other jasmonates induce the expression of numerous genes which encode enzymes involved in secondary metabolism (Sasaki-Sekimoto et al., 2005; Taki et al., 2005). In contrast, hrp-dependent TTSS effectors are responsible for the repression of cell wall genes, particularly pectin esterases, as well as fatty acid biosynthesis (in the chloroplast) and a dramatic overall repression of photosynthesis (Figure 2-3B). Many genes encoding chloroplast localized proteins involved in photosynthesis and the Calvin cycle as well as genes involved in photorespiration and tetrapyrrole synthesis are repressed by TTSS effectors (Figure 2-3B). 67 Figure 2-3 trelabolisr COR toxir HI’I. lxlO' I gates iden Mlcfiam Ftpressed. Figure 2-3: COR toxin and type III effectors target different aspects of plant metabolism. (A) The MapMan ‘Metabolism Overview’ display created using the 944 COR toxin-regulated genes identified from the Pst DC3000 vs Pst COR" comparison (24 HPI, 1x106 bacteria/mL) is shown. (B) The display created using the 791 hrp-regulated genes identified from the Pst COR' vs Pst COR' hrpS comparison (10 HPI, 5x10 bacteria/mL) is shown. The average fold change of the 3 biological replicates is displayed as illustrated in the fold change color bar in the lower right of each panel (note red is repressed, blue is induced). 68 Fig Figure 2-3 noz 22.0.8..— .358m .3! . ulu 2% 2.22 0:0 3:2. 2° Metabollsm Llplds Cell wall 69 T... .t A oAvU.‘ 3:9130 .09:- .Dbo . I‘a f-l‘- 50:3: 010 Figure 2-3 (cont’d). 2° Metabollsm 2235:. 5565 6:! Llplds 20.5%.“: poo—3:095 Cell wall 010 3:2. 70 Coronatine activates JA responses and secondary metabolism Analysis of the data has also uncovered that Pst challenge significantly targets several plant hormone signaling and stress-response pathways. The expression patterns observed for some genes involved in plant hormone signaling and stress responses are displayed in Figure 2-4. Consistent with the results from an analysis of the effects of purified COR in tomato (U ppalapati et al., 2005), known J A/COR toxin responsive genes including chlorophyllase (CLHI/CORII At1g19670; Benedetti et al., 1998; Tsuchiya et al., 1999) and COR13 (At4g23600; Lopukhina et al., 2001) were induced 24 HPI in a coronatine-dependent manner. Interestingly, 4 genes encoding enzymes involved in jasmonate biosynthesis (LOXZ At3g45140, AOS At5g42650, AOCI At3g25760 and 0PR3 At2g06050; Bell et al., 1995; von Malek et al., 2002; Schaller et al., 2000) are also COR toxin induced, suggesting activation of not only jasmonate responses but the host biosynthesis of jasmonates as well. As suggested by the analysis using MapMan, host secondary metabolism is activated by COR toxin including the induction of the PAP] transcription factor (At1g56650), a regulator of secondary metabolism including the production of anthocyanin pigments (Borevitz et al., 2000). As expected, CHS (At5g13930) and a putative anthocyanidin synthase (At2g3 8240) and UDP- anthocyanidin transferase (At4g27570) are strongly induced (Figure 2-4). The expression of PALI (At2g37040) involved in phenylpropanoid biosynthesis and GSTFI 1 (At3g03190) also displayed COR-dependent induction, consistent with their observed overexpression in PAP] activation-tagged plants (Borevitz et al., 2000). Several other genes involved in Phenylpropanoid metabolism also exhibited reproducible induction (F5H2 At5g04330, 71 4(19 . others A1595 confli. linalo shill: 4). 1}: metal 4CL9 At1g20510 and Eli3-2/CAD-BZ At4g37990), but somewhat surprisingly, numerous others were repressed (PAL3 At5g04230, F 5H1 At4g36220, 4CL1 Atl g51680, COMTI At5g54160 and CAD] At4g39330; Gachon et al., 2005). The physiological impact of this conflicting differential regulation on lignin biosynthesis is unclear. Two genes encoding enzymes putatively involved in terpenoid biosynthesis (S- linalool synthase At1g61 120 and T P503 At4g16740) and DHSI, the first enzyme of the shikimate pathway (At4g39980; Keith et al., 1991) are induced by COR toxin (Figure 2- 4). The shikimate pathway produces precursors for the synthesis of numerous secondary metabolites as well as chorismate for the production of aromatic amino acids. One metabolic pathway coordinately regulated is that of tryptophan (Trp) biosynthesis. Pst coordinately induces three genes that encode enzymes that convert chorismate to Trp (Figure A-5). Phosphoribosyl anthranilate isomerase (PAII /PA12 At1g07780/At5g05590, both represented by probe set 259770_s_at; He and Li 2001) is induced in a COR toxin- dependent way, while ASAI (At5g05730; Niyogi and Fink 1992) and ASBI/ASBZ (Atl g25220/At5g57890; Niyogi et al., 1993) are induced by both COR toxin and TTSS effectors (Figure A-S). The coordinated induction of these genes is consistent with the induced expression of A TR] (AtMyb34 At5g60890) and AT R2 (At5g46760) that are known to activate Trp and glucosinolate production in Arabidopsis(Ce1enza et al., 2005; Smolen et al., 2002). Since the tryptophan pathway activators ATRI and A TR2 are COR toxin induced, we examined whether other genes which encode enzymes in this pathway were differentially regulated, but did not meet our stringent selection criteria. Indeed, 5 genes, PAT (At5g17990) T SA] (At3g54640) TSBI/TSBZ (At5g54810/At4g27070; Radwanski etal., 1995) IGPS (At2g04400; Li et al.,1995) and a tryptophan sysnthase 72 a1. 30| {011ml pain“ cama Figur 1m by C tori: “til beta-like gene (At5 g3 8530) were also induced following bacterial inoculation (Figure A- 5). Tryptophan is likely being used as a substrate for the production of glucosinolates, and 1AA, since CYP79BZ (At4g39950; Zhao et al., 2002), CYP79B3 (At2g22330; Zhao et al., 2002), and NIT 3 (At3g44320; Vorwerk et al., 2001) are reproducibly induced following inoculation with 1x106 bacteria/mL Pst DC3 000 bacteria. The tryptophan pathway may also potentially contribute to the accumulation of the phytoalexin carnalexin since PAD3 is a known Pst induced gene (At3g26830; Zhou et al., 1999; Figure A-S). The synthesis of methionine-derived glucosinolates is also implied since CYP79F1/CYP79F2 (At1g16400/At1g16410; Chen et al., 2003; Reintanz etal., 2001; Tantikanjana et al., 2004), and AOP2 (At4g03060; Kliebenstein et al., 2001) are induced by COR toxin. The expression pattern displayed in Figure A-S also illustrates how COR toxin and TTSS effectors can both synergistically induce the expression of some genes while antagonistically regulating the expression of other genes. 73 Figure 2-4: Pst coordinately regulates hormone and stress response genes. The expression profile for 57 genes following Pst inoculation is shown. See Figure l legend for a description of the gene expression display. Gene annotation is listed on the right with background shading delineating a relationship to known pathways or responses. Orange- JA or COR toxin inducible, yellow-JA biosynthesis, red-secondary metabolism, green-ABA and abiotic stress response, violet-SA and defense responses, and blue-basal defense. 74 Figure 2-4 1x108 5x107 1x106 ABCDE‘FGH IJKLM 40 30 20151.015 2.0 6:0 At1g19640 JMT1 JA Methyl Transferase At2924850 TAT3 Tyrosine Amino Transferase At3911480 BSMT1 Benzoic acid and SA Methyl Transferase At1919670 CLHi/CORH Chlorophyilase At1952400 BLGL1 Beta Glucosidase At5924780 VSP1 Vegetative storage protein At3g16470 JR1 JA wounding responsive 1 At4923600 CORI3 COR-induced tyrosine aminotransferase At4g15440 HPL Hyperoxide Lyase CYP74BZ At5944420 PDF1.2 Defensin At3945140 LOX2 13-Lipoxygenase At5942650 AOS Allene oxide synthase At3925760 AOC1 Allene oxide cyclase At2906050 OPR3 12-OxoPhytodienoate-10,1 1-Roductase At1956650 PAP1 MYB75 R2R3 transcription factor At5913930 CHS Chalcone Synthase At2938240 putative anthocyanidin synthase At1961129 S—linalool synthase At4g16740 TPSO3 terpenoid synthase At4927570 anthocyanidin-a-gluooside mamnosyitransferase 1At4939980 onsr DHAP synthase At2937040 PAL1 phenylalanine ammonia lyase 75 Pst induces ABA/abiotic stress response genes COR toxin and, to some degree, TTSS effectors induce numerous ABA responsive genes, including a significant number of genes implicated in ABA/abiotic stress-responsive signaling (Figure 2-4; Bray 2002; F inkelstein and Rock 2002)./1811 (At4g26080; Leung et al., 1994) and A312 (At5g57050; Leung et al., 1997) are both induced in the absence of COR toxin at the 5x107 bacteria/mL inoculum level, suggesting that their expression is at least partly regulated by TTSS effectors (Figure 2-4). NCED3 (At3g14440; Iuchi et al., 2001), a 9-cis-epoxycarotenoid dioxygenase crucial for ABA synthesis, is also induced following Pst COR' inoculation (Figure 2-4). NCED4 (At4g19170), a related family member, is repressed by virulent Pst DC3000 and Pst COR' bacteria at the higher inoculum levels. A total of 6 transcription factors associated with ABA and abiotic stress responses (HE-12 At3g61890, HB-7 At2g46680, NAC3/ANAC055 At3g15500, RD26 ANAC072 At4g27410, ANAC019 At1g52890 and MY C2/RDZZBPI At1g32640) are also induced by Pst infection (Figure 2-4) (Abe et al., 2003; Fujita et al., 2004; Greve et al., 2003; Olsson et al., 2004; Soderman et al., 1996; Tran et al., 2004). These results imply that ABA responses are being activated by Pst and that both TTSS effectors and COR toxin are involved in this activation. Defense response-related gene expression Of particular interest is the role COR toxin and TTSS effectors have on SA- mediated defenses since those responses are necessary for resisting pathogen attack. Several SA-responsive genes involved in pathogen defense (EDS5 At4g39030, ICSI Atlg74710, PAD4 At3g33960, EDS] At3g48090, and NIMIN-I At1g02450) were 76 induced in a hrp-dependent manner (Figure 2-4) (Falk et al., 1999; J irage et al., 1999; Nawrath et al., 2002; Wildermuth et al., 2001; Weigel et al., 2005). The Pst-induced expression of these genes is consistent with previous published results, but it was somewhat surprising that their induction is enhanced between 1.5- and 20-fold by TTSS effectors. The expression of these genes is either not reliably induced or induced 2-to-5- fold by Pst COR'hrpS bacteria relative to the mock inoculated control in the 5x107 dataset. This suggests that the host plant is recognizing PAMPS present on the Pst COR’ hrpS mutant and potentially also recognizing specific TTSS effectors and activating defense gene expression (Figure 2-4). Interestingly, COR toxin actually dampens by approximately 4-fold the induction of at least one SA-related defense gene (NIMIN-I). In the samples examined, the defense response genes AIGI (At1g33960; Reuber and Ausubel 1996), and TONB (At2g32190; de Torres et al., 2003) are induced 6-to-20-fold. In contrast to the induction observed for the previously discussed SA/defense- related genes, virulent Pst bacteria repress three genes associated with basal defense responses. F lagellin Sensitive 2 (FLSZ At5g46330) which is required for the perception of the PAMP flagellin is repressed by both COR toxin and TTSS effectors at high inoculum concentrations (Figure 2-4) (Gomez-Gomez etal., 1999; Zipfel et al., 2004). The F lagellin induced Receptor Kinase 1 (FRKI Atl g19l90; Asai et al., 2002; de Torres etal., 2003) and WRKY22 (At4g01250; Asai et al., 2002) are induced by E. coli, Pst hrpA and Pst COR'hrpS mutant bacteria at high inoculum levels and that induction is blocked by TTSS effectors in the Pst COR" mutant (Figure 2-4). COR toxin produced by Pst DC3000 also contributes to the repression of FRKI , WRKY22 and FLSZ expression at the 77 1x106 inoculum level. These results are consistent with the idea that TTSS effectors and COR toxin contribute to pathogenesis by suppressing basal defense responses. Expression of auxin and cytokinin-related genes are altered during Pst infection It is clear that Pst virulence systems are modulating stress hormone response pathways (JA, SA and ABA), but plant host physiology also appears to be experiencing specific shifts in auxin and cytokinin responses as well. Thirty-one auxin-related genes are repressed while 13 other genes are induced following Pst inoculation (Figure A-6). In most cases the gene regulation appears to be partially dependent on both COR toxin and TTSS effectors. A few genes are also reproducibly differentially regulated in response to bacterial PAMPS present on E. coli and nonpathogenic strains of Pst as well. The 31 repressed genes include 4 A UX/LAA and 18 SA UR (Small Auxin Up-Regulated) genes, 6 auxin transporters (PIN3 Atl g70940, PIN4 At2g01420, PIN 7 Atl g23080, A UXI At2g38120, LAX] At5g01240, and LAX3 Atl g77690) and ARF18 (At3g61830) (Hagen and Guilfoyle, 2002; Okushima et al., 2005; Swarup et al., 2004). AUX/1AA genes are rapidly induced by auxin and are hypothesized to act in a feedback loop repressing the auxin response signaling pathway (Tiwari et al., 2004). Five SA UR genes, three 1AA- amino acid hydrolases (ILRI At3g02875, ILL5 At1g51780 and ILL6 Atl g44350), 3 [AA- arnido synthases (DFL2 At4g03400, GH3.3 At2g23170 and GH3.12 At5g13320), IAA18 (At3g16500) and NIT 3, are induced. NIT3 will synthesize 1AA (Figure A-SB) while IAA-arnino acid hydrolases and arnido synthases convert 1AA between an amino acid conjugate and free form (Staswick et al., 2005). Taken together, these results suggest that Pst is impacting host auxin signaling, potentially de-repressing the pathway, altering 78 auxin movement and activating biosynthesis of the hormone. These results are consistent with the observation that Pseudomonas syringae strains express biosynthetic enzymes that can produce 1AA from tryptophan (Buell et al., 2003; Glickmann et al., 1998). Pst DC3000 has been shown to elicit the accumulation of [AA in Arabidopsis leaves during infection and that activation of this pathway contributes to disease susceptibility (O’Donnell et al., 2003). Auxin and cytokinin are involved in many important aspects of plant growth and development. Pst pathogenesis appears to not only target auxin signaling but also genes involved in cytokinin accumulation and signaling. Seven type A response regulators (ARR4, 5, 6, 7, 9, I 5 and ARR16; D’Agostino et al., 2000; To et al., 2004) are repressed by COR toxin following inoculation with 1x106 Pst DC3000 bacteria/mL (Figure A-7). The other type A family members had low expression in the samples tested and in most cases were scored as absent (data not shown). One type B response regulator (ARR14 At2g01760) was repressed by Pst, but in this case repression appears to be dependent on TTSS effectors and not COR toxin. None of the other type B family members were reproducibly differentially regulated. Cytokinin synthesis is mediated by the isopentenyl transferase gene family, one member of that family; IPT3 (At3g63110) is COR toxin repressed (Kakimoto 2003a, Miyawaki et al., 2004). Two cytokinin oxidase genes (CKX4 At4g29740, CKX5 Atl g7 5450; Werner et al., 2003) are induced by TTSS effectors and COR toxin respectively. The repression of IPT3 together with CKX4 and CKX5 induction suggests a reduction in the levels of cytokinin within the leaves following inoculation. The type A response regulators are known to be rapidly induced by cytokinin and their repression would be consistent with a fall in the cytokinin levels. Members of the type A 79 response regulator family appear to be negative regulators of cytokinin signaling and may be part of feedback loop repressing induced responses (Kakimoto 2003b; To et al., 2004). The coordinated repression of this gene family suggests that cytokinin signaling may be de-repressed by Pst during infection. Conclusion Although the Arabidopsis-Pst DC3 000 interaction was initially developed as a model for studying disease resistance mechanisms (Whalen et al., 1991), this pathosystem is now also used widely for studying susceptible plant-pathogen interactions (N omura et al., 2005). A comprehensive transcriptional analysis of this interaction using defined Pst DC3000 virulence mutants, we thought, would provide the community a valuable resource for future understanding of this interaction at whole plant, tissue, cellular, and molecular levels. Indeed, this analysis revealed several interesting results. First, it appears that Arabidopsis plants respond similarly to PAMPS presented on live human and plant pathogenic bacteria, and that flagellin perception is not necessary for the regulation of PAMP-induced genes or bacterial multiplication in the context of Pst DC3 000 infection of Arabidopsis under our experimental conditions in which bacteria are infiltrated directly into the leaf intercellular spaces. We found that an impressive number of PAMP-induced genes belong to transcription factors, signaling components, and cell wall- and/or secretion-associated proteins, suggesting coordinate regulation of genes involved in signaling and secretion with those that encode secreted products during basal defense. Second, use of Pst DC3000 mutant strains deficient in COR toxin (PstDC3118) or COR and the TTSS (the hrpS COR' mutant) enabled us to obtain genome-wide 80 knowledge of both distinct and overlapping effects of these two important virulence factors in modulating host gene expression during actual bacterial infection. There is clear evidence for TTSS effector-mediated suppression of basal defense-associated genes. In contrast, many of the SA-regulated defense genes, including some of those associated with basal defense, are induced by TTSS effectors. The most dramatic effect of COR is the induction of J A-regulated genes, as expected, but COR also appears to exert a major impact on the expression of genes involved in secondary metabolism. Finally, both TTSS effectors and COR contribute to the dramatic regulation of genes that are responsive to plant hormones, such as auxin, ABA, and cytokinin. We hope that the results described in this paper and the online supplemental materials will facilitate future study of Pst DC3000 pathogenesis in Arabidopsis using a variety of genomic, cell biology, and biochemical approaches. Acknowledgements We gratefully acknowledge Annette Thelen and the Genomic Technology Support Facility at Michigan State University for valuable assistance with the Affymetrix microarray analysis, Dr. Thomas Whittam (Department of Microbiology, Michigan State University) and Dr. Arlette Darfeuille-Michaud (Pathogenic Bacterienne Intestinale, Clermont-Ferrand, France) for providing E. coli strains, and Dr. Thomas Whittam for providing lab space for work with human-pathogenic bacteria. Arabidopsis EST clones used for RNA blot hybridization were obtained from the Arabidopsis Biological Resource Center. This work was supported by research grants from the US Department of Energy (DE-FG02-91ER20021) and National Institutes of Health (1 R21A1060761 -01) 81 to S.Y.H., a postdoctoral fellowship to RT. from the US Department of Agriculture, a US Department of Education GAANN Graduate Fellowship and a Michigan State University Plant Science Graduate Fellowship to W.U., and funds from the Center of Microbial Pathogenesis at Michigan State University to S. Y. H. and Thomas Whittam. 82 References Abe, H., Urao, T., Ito, T., Seki, M., Shinozaki, K. and Yamaguchi-Shinozaki, K. (2003) Arabidopsis AtMYC2 (bHLH) and AtMYB2 (MYB) function as transcriptional activators in abscisic acid signaling. Plant Cell, 15, 63-78. Asai, T., Terra, G., Plotnikova, J ., Willmann, M.R., Chiu, W., Gomez-Gomez, L., Boller, T., Ausubel, F.M. and Sheen, J. (2002) MAP kinase signaling cascade in Arabidopsis innate immunity. Nature, 415, 977 - 983. Assaad, F.F., Huet, Y., Mayer, U., and Jiirgens, G. (2001). The cytokinesis gene KEULE encodes a See] protein that binds the syntaxin KNOLLE. J. Cell Biol. 152, 531—543 Assaad, F.F., Qiu, J ., Youngs, H., Ehrhardt, D., Zimmerli, L., Kalde, M., Wanner, G., Peck, S.C., Edwards, H., Ramonell, K., Somerville, CR. and Thordal-Christensen, H. (2004) The PEN] syntaxin defines a novel cellular compartment upon fungal attack and is required for the timely assembly of papillae. Mol. Biol. Cell, 15, 5118-5129. Bell, E., Creelman, RA. and Mullet, J.E. (1995) A Chloroplast lipoxygenase is required for wound-induced jasmonic acid accumulation in Arabidopsis. Proc. Natl. Acad. Sci. USA, 92, 8675-8679. Bender, C.L., Alarcon-Chaidez, F. and Gross, DC. (1999) Pseudomonas syringae phytotoxins: Mode of action, regulation, and biosynthesis by peptide and polyketide synthetases. Microbiol. Mol. Biol. Rev. 63, 266—292. Benedetti, C.E., Costa, C.L., Turcinelli, SR. and Arruda, P. (1998) Differential expression of a novel gene in response to coronatine, methyl jasmonate, and wounding in the C011 mutant of Arabidopsis. Plant Physiol. 116, 1037-1042. Benedetti, C.E., Xie, D. and Turner, J .G. (1995) C011 -dependent expression of an Arabidopsis vegetative storage protein in flowers and siliques and in response to coronatine or methyl jasmonate. Plant Physiol. 109, 567 -572. Berin, M.C., Darfeuille-Michaud, A., Egan, L.J., Miyamoto, Y. and Kagnoff, M.F. (2002) Role of EHEC Ol 57:H7 virulence factors in the activation of intestinal epithelial cell NF-KB and MAP kinase pathways and the upregulated expression of interleukin 8. Cell. Microbiol. 4, 635-647. Bestwick, C. S., Bennett, M. H. & Mansfield, J. W. (1995) Hrp mutant of Pseudomonas syringae pv. phaseolicola induces cell wall alterations but not membrane damage leading to the hypersensitive reaction in lettuce. Plant Physiol. 108, 503-516. 83 Borevitz, J .O., Xia, Y., Blount, J ., Dixon, RA. and Lamb, C. (2000) Activation tagging identifies a conserved MYB regulator of phenylpropanoid biosynthesis. Plant Cell, 12, 23 83-2393. Bray, EA. (2002) Abscisic acid regulation of gene expression during water deficit stress in the era of the Arabidopsis genome. Plant, Cell and Environ. 25, 153-161. Brooks D.M., Hernandez-Guzman G., Kloek A.P., Alarcon-Chaidez F., Sreedharan A., Rangaswamy V., Penaloza-Vazquez A., Bender CL. and Kunkel B.N. (2004) Identification of a well-defmed series of coronatine biosynthetic mutants of Pseudomonas syringae pv. tomato DC3000. Mol. Plant Microbe Interact. 17, 162-174. Buell, C.R., Joardar, V., Lindeberg, M. et al. (2003) The complete genome sequence of the Arabidopsis and tomato pathogen Pseudomonas syringae pv. tomato DC3000. Proc. Natl. Acad. Sci. USA, 100, 10181—10186. Celenza, J .L., Quiel, J .A., Smolen, G.A., Merrikh, H., Silvestro, A.R., Norrnanly, J. and Bender, J. (2005) The Arabidopsis ATRl Myb transcription factor controls indolic glucosinolate homeostasis. Plant Phys. 137, 253-262. Chang, J .H., Urbach, J .M., Law, T.F., Arnold, L.W., Hu, A., Gombar, S., Grant, S.R., Ausubel, F.M. and Dangl, J .L. (2005) A high-throughput, near-saturating screen for type III effector genes from Pseudomonas syringae. Proc. Natl Acad. Sci. USA, 102, 2540- 2554. Chen, S., Glawischnig, E., Jorgensen, K., Naur, P., Jorgensen, B., Olsen, C.E., Hansen, C.H., Rasmussen, H., Pickett, J .A. and Halkier, BA. (2003) CYP79F1 and CYP79F2 have distinct functions in the biosynthesis of aliphatic glucosinolates in Arabidopsis. Plant J. 33, 923-937. Cole, R.A., Synek, L., Zarsky, V. and Fowler, J .E. (2005) SEC8, a subunit of the putative Arabidopsis exocyst complex facilitates pollen germination and competitive pollen tube growth. Plant Physiol. 138, 2005-2018. Collins, N.C., Thordal-Christensen, H., Lipka, V., Bau, S., Kombrink, E., Qiu, J .L., Huckelhoven, R., Stein, M., F reialdenhoven, A., Somerville, SC. and Schulze-Lefert, P. (2003) SNARE-protein-mediated disease resistance at the plant cell wall. Nature, 425, 973-977. Colhner, A., Lindeberg, M., Petnicki-chieja, T., Schneider, DJ. and Alfano, J .R. (2002) Genomic mining type III secretion system effectors in Pseudomonas syringae yields new picks for all TTSS prospectors. Trends Microbiol. 10, 462-469. D'Agostino, I.B., Deruere, J. and Kieber, J .J . (2000) Characterization of the response of the Arabidopsis response regulator gene family to cytokinin. Plant Physiol. 124, 1706- 1717. ' 84 de Torres, M., Sanchez, P., Femandez-Delmond, I., and Grant, M. (2003). Expression profiling of the host response to bacterial infection: the transition from basal to induced defense responses in RPMl-mediated resistance. Plant J. 33, 665-676. Devoto, A., Nieto-Rostro, M., Xie, D., Ellis, C., Harrnston, R., Patrick, E., Davis, J., Sherratt, L., Coleman, M. and Turner, J .G. (2002) C011 links jasmonate signaling and fertility to the SCF ubiquitin-ligase complex in Arabidopsis. Plant J. 32, 457—466. Diener, AC. and Ausubel, F .M. (2005) RESISTANCE T0 F USARIUM OXYSPORUM 1, a dominant Arabidopsis resistance gene, is not race specific. Genetics, 171, 305-321. Eisen, M.B., Spellman, P.T., Brown, PC. and Botstein, D. (1998) Cluster analysis and display of genome-wide expression patterns. Proc. Natl. Acad. Sci. USA, 95, 14863- 14868. Eulgem, T., Weigman V.J., Chang, H., McDowell, J .M., Holub, E.B., Glazebrook, J ., Zhu, T. and Dangl, J .L. (2004) Gene expression signatures from three genetically separable resistance gene signaling pathways for downy mildew resistance. Plant Physiol, 135, 1 129-1 144. Falk, A., Feys, B.J., Frost, L.N., Jones, J .D., Daniels, MJ. and Parker, J.E. (1999)EDS1, an essential component of R gene-mediated disease resistance in Arabidopsis has homology to eukaryotic lipases. Proc. Natl. Acad. Sci. U S A, 96, 3292-3297. Felix, G., Duran, J .D., Volko, S. and Boller, T. (1999) Plants have a sensitive perception system for the most conserved domain of bacterial flagellin. Plant J. 18, 265-276. Feys, B.J.F., Benedetti, C.E., Penfold, ON. and Turner, J.G. (1994) Arabidopsis mutants selected for resistance to the phytotoxin coronatine are male sterile, insensitive to methyl jasmonate, and resistant to a bacterial pathogen. Plant Cell, 6, 7 51-759. Finkelstein, RR. and Rock, CD. (2002) Abscisic acid biosynthesis and response. In The Arabidopsis Book (Meyerowitz, E. M. and Somerville, C. R., eds). American Society of Plant Biologists, Rockville, MD, USA, DOI 10.1199/tab.0058. Fonts, D.E., Abramovitch, R.B., Alfano, J.R., Baldo, A.M., Buell, C.R., Cartinhour, S., Chatterjee, A.K., D'Ascenzo, M., Gwinn, M.L., Lazarowitz, S.G., Lin, N.C., Martin, G.B., Rehm, A.H., Schneider, D.J., van Dijk, K., Tang, X. and Coller, A. (2002) Genomewide identification of Pseudomonas syringae pv. tomato DC3 000 promoters controlled by the HrpL alternative sigma factor. Proc. Natl. Acad. Sci. USA, 99, 2275- 2280. 85 Fujita, M., Fujita, Y., Maruyama, K., Seki, M., Hiratsu, K., Ohme-Takagi, M., Tran, L.S., Yamaguchi-Shinozaki, K. and Shinozaki, K. (2004) A dehydration-induced NAC protein, RD26, is involved in a novel ABA-dependent stress-signaling pathway. Plant J. 39, 863- 876. Gachon, C.M., Langlois-Meurinne, M., Henry, Y. and Saindrenan, P. (2005) Transcriptional co-regulation of secondary metabolism enzymes in Arabidopsis: functional and evolutionary implications. Plant Mol. Biol. 58, 229-245. Glickmann, E., Gardan, L., Jacquet, S., Hussain, S., Elasri, M., Petit, A. and Dessaux, Y. (1998) Auxin production is a common feature of most pathovars of Pseudomonas syringae. Mol. Plant Microbe Interact. 11, 156-162. Gomez-Gomez, L. and Boller, T. (2000) FLS2: An LRR receptor-like kinase involved in the perception of the bacterial elicitor flagellin in Arabidopsis. Mol. Cell, 5, 1003-1011. Greenberg, J .T. and Vinatzer, BA. (2003) Identifying type III effectors of plant pathogens and analyzing their interaction with plant cells. Curr Opin Microbiol. 6, 20-28. Greve, K., La Cour, T., Jensen, M.K., Poulsen, F.M., and Skriver, K. (2003). Interactions between plant RING-H2 and plant-specific NAC (N AM/ATAF 1/2/CUC2) proteins: RING-H2 molecular specificity and cellular localization. Biochem. J. 371, 97—108. Gunzer, F., Bohn, U., Fuchs, S., Muhldorfer, 1., Hacker, J ., Tzipori S. and Donohue- Rolfe, A. (1998) Construction and characterization of an isogenic slt-ii deletion mutant of enterohemorrhagic Escherichia coli. Infect. Immun. 66, 2337-2341. Guo, F.Q., Okamoto, M. and Crawford, NM. (2003) Identification of a plant nitric oxide synthase gene involved in hormonal signaling. Science, 302, 100-103. Hagen, G. and Guilfoyle, T. (2002) Auxin-responsive gene expression: genes, promoters and regulatory factors. Plant Mol. Biol. 49, 373-385. Halachmi, N., and Lev, Z. (1996) The Secl family: a novel family of proteins involved in synaptic transmission and general secretion. J. Neurochem. 66, 889-897. Hauck, P., Thilmony, R. and He, S.Y. (2003) A Pseudomonas syringae type III effector suppresses cell-wall based extracellular defense in susceptible Arabidopsis plants. Proc. Natl. Acad. Sci. USA, 100, 8577-8582. 86 Hayashi, T., Makino, K., Ohnishi, M., Kurokawa, K., Ishii, K., Yokoyama, K., Han, C.G., Ohtsubo, E., Nakayama, K., Murata, T., Tanaka, M., Tobe, T., Iida, T., Takami, H., Honda, T., Sasakawa, C., Ogasawara, N., Yasunaga, T., Kuhara, S., Shiba, T., Hattori, M. and Shinagawa, H. (2001) Complete genome sequence of enterohemorrhagic Escherichia coli OlS7:H7 and genomic comparison with a laboratory strain K-12. DNA Res. 8, 11-22. He, S.Y., Nomura, K. and Whittam, S. (2004) Type 111 protein secretion mechanisms in mammalian and plant pathogens. Biochim. Biophys. Acta. 1694, 181-206. He, Y. and Li, J. (2001) Differential expression of triplicate phosphoribosylanthranilate isomerase isogenes in the tryptophan biosynthetic pathway of Arabidopsis thaliana (L.) Heynh. Planta, 212, 641-647. He, Z.H., Cheeseman, 1., He, D., Kohorn, ED. (1999) A cluster of five cell wall- associated receptor kinase genes, Wakl-S, are expressed in specific organs of Arabidopsis. Plant Mol. Biol. 39, 1189-1196. Hirano, SS. and Upper, CD. (2000) Bacteria in the leaf ecosystem with emphasis on Pseudomonas syringae-a pathogen, ice nucleus, and epiphyte. Microbiol. Mol. Biol. Rev. 64, 624-653. Iuchi, S., Kobayashi, M., Taji, T., Naramoto, M., Seki, M., Kato, T., Tabata, S., Kakubari, Y., Yamaguchi-Shinozaki. K. and Shinozaki, K. (2001) Regulation of drought tolerance by gene manipulation of 9-cis-epoxycarotenoid dioxygenase, a key enzyme in abscisic acid biosynthesis in Arabidopsis. Plant J. 27, 325-333. Jin, Q., Thilmony R., Zwiesler-Vollick, J. and He, S.Y. (2003) Type 111 protein secretion in Pseudomonas syringae. Microbes & Infect. 5, 301-310. Jirage, D., Tootle, T.L., Reuber, T.L., Frost, L.N., Feys, B.J., Parker, J.E., Ausubel, F.M. and Glazebrook, J. (1999) Arabidopsis thaliana PAD4 encodes a lipase-like gene that is important for salicylic acid signaling. Proc. Natl. Acad. Sci. U S A, 96, 13583-13588. Kakimoto, T. (2003a) Biosynthesis of cytokinins. J. Plant Res. 116, 233-239. Kakimoto, T. (2003b) Perception and signal transduction of cytokinins. Ann. Rev. Plant Biol. 54, 605-627. Katagiri, F., Thilmony, R. and He, S.Y. (2002) The Arabidopsis thaliana-Pseudomonas syringae interaction. In The Arabidopsis Book (Meyerowitz, E. M. and Somerville, C. R., eds). American Society of Plant Biologists, Rockville, MD, USA, DOI 10.1199-tab.0039. Keith, B., Dong X.N., Ausubel, F.M. and Fink, GR. (1991) Differential induction of 3- deoxy-D-arabino-heptulosonate 7-phosphate synthase genes in Arabidopsis thaliana by wounding and pathogenic attack. Proc. Natl. Acad. Sci. USA, 88, 8821-8825. 87 Kliebenstein, D.J., Lambrix, V.M., Reichelt, M., Gershenzon, J. and Mitchell-Olds, T. (2001) Gene duplication in the diversification of secondary metabolism: tandem 2- oxoglutarate-dependent dioxygenases control glucosinolate biosynthesis in Arabidopsis. Plant Cell, 13, 681-693. Kloek, A.P., Verbsky, M.L., Sharma, S.B., Schoelz, J .E., Vogel, J ., Klessig, DR and Kunkel, B.N. (2001) Resistance to Pseudomonas syringae conferred by an Arabidopsis thaliana coronatine-insensitive (coil) mutation occurs through two distinct mechanisms. PlantJ. 26, 509-522. Kunze, G., Zipfel, C., Robatzek, S., Niehaus, K., Boller, T. and Felix, G. (2004) The N terminus of bacterial elongation factor Tu elicits innate immunity in Arabidopsis plants. Plant Cell, 16, 3496-3507. Leung, J., Bouvier-Durand, M., Morris, P.C., Guerrier, D., Chefdor, F. and Giraudat, J. (1994) Arabidopsis ABA response gene ABII : features of a calcium-modulated protein phosphatase. Science, 264, 1448-1452. Leung, J ., Merlot, S. and Giraudat, J. (1997) The Arabidopsis ABSCISIC ACID- INSENSITIVE 2 (ABIZ) and ABII genes encode homologous protein phosphatases 2C involved in abscisic acid signal transduction. Plant Cell, 9, 759-771. Li, J., Chen, S., Zhu, L. and Last, R.L. (1995) Isolation of cDNAs encoding the tryptophan pathway enzyme indole-3-glycerol phosphate synthase from Arabidopsis thaliana. Plant Physiol. 108, 877-878. Li, X., Lin, H., Zhang, W., Zou, Y., Zhang, J ., Tang, X. and Zhou J. (2005) Flagellin induces innate immunity in nonhost interactions that is suppressed by Pseudomonas syringae effectors. Proc. Natl. Acad. Sci. USA, 102, 12990-12995. Lopukhina, A., Dettenberg, M., Weiler, E.W. and Hollander-Czytko, H. (2001) Cloning and characterization of a coronatine-regulated tyrosine aminotransferase from Arabidopsis. Plant Physiol. 126, 1678-1687. Ma, S.W., Morris, V.L. and Cuppels, DA. (1991) Characterization of a DNA region required for the production of the phytotoxin coronatine by Pseudomonas syringae pv. tomato. Mol. Plant Microbe Interact. 4, 69-74. Mittal, S. and Davis, KR. (1995) Role of the phytotoxin coronatine in the infection of Arabidopsis thaliana by Pseudomonas syringae pv. tomato. Mol. Plant Microbe Interact. 8, 1 65-1 71 . Miyawaki, K., Matsumoto-Kitano, M. and Kakimoto, T. (2004) Expression of cytokinin biosynthetic isopentenyltransferase genes in Arabidopsis: tissue specificity and regulation by auxin, cytokinin, and nitrate. Plant J. 37, 128-138. 88 Nagano Y., Okada Y., Narita H., Asaka Y. and Sasaki Y. (1995) Location of light- repressible, small GTP-binding protein of the YPT/rab family in the growing zone of etiolated pea stems. Proc. Natl. Acad. Sci. USA, 92, 6314-6318. Navarro, L., Zipfel, C., Rowland, 0., Keller, 1., Robatzek, S., Boller, T. and Jones, J .D. (2004) The transcriptional innate immune response to flg22. Interplay and overlap with Avr gene-dependent defense responses and bacterial pathogenesis. Plant Physiol. 135, 1113-1128. Nawrath, C., Heck, S., Parinthawong, N. and Metraux J .P. (2002) EDSS, an essential component of salicylic acid-dependent signaling for disease resistance in Arabidopsis, is a member of the MATE transporter family. Plant Cell, 14, 275-286. Niyogi, K.K. and Fink, GR. (1992) Two anthranilate synthase genes in Arabidopsis: defense-related regulation of the tryptophan pathway. Plant Cell, 4, 721-733. Niyogi, K.K., Last, R.L., Fink, GR. and Keith, B. (1993) Suppressors of trpl fluorescence identify a new Arabidopsis gene, TRP4, encoding the anthranilate synthase beta subunit. Plant Cell, 5, 1011-1027. Nomura, K., and He, S.Y. (2005) Powerful screens for bacterial virulence proteins. Proc. Natl. Acad. Sci. USA, 102, 3527-3528. Nomura, K., Melotto, M. and He, S.Y. (2005) Suppression of host defense in compatible plant-Pseudomonas syringae interactions. Curr. Opin. Plant Biol. 8, 361-368. Nl'ihse, T.S., Boller, T. and Peck, SC. (2003). A plasma membrane syntaxin is phosphorylated in response to the bacterial elicitor flagellin. J. Biol. Chem. 278, 45248- 45254. O'Donnell, P.J., Schmelz, B.A., Moussatche, P., Lund, S.T., Jones, J .B. and Klee, H.J. (2003) Susceptible to intolerance--a range of hormonal actions in a susceptible Arabidopsis pathogen response. Plant J. 33, 245-257. Okushima, Y., Overvoorde, P.J., Arima, K., Alonso, J .M., Chan, A., Chang, C., Ecker, J .R., Hughes, B., Lui, A., Nguyen, D., Onodera, C., Quach, H., Smith, A., Yu, G. and Theologis, A. (2005) Functional genomic analysis of the AUXIN RESPONSE FACTOR gene family members in Arabidopsis thaliana: unique and overlapping fimctions of ARF 7 and ARF 19. Plant Cell, 17, 444-463. Olsson, A.S., Engstrom, P. and Soderman, E. (2004) The homeobox genes AtHBIZ and AtHB7 encode potential regulators of growth in response to water deficit in Arabidopsis. Plant Mol. Biol., 55, 663-677. 89 Penaloza-Vazquez, A., Preston, G.M., Collmer, A., and Bender, CL. (2000) Regulatory interactions between the Hrp type III protein secretion system and coronatine biosynthesis in Pseudomonas syringae pv. tomato DC3000. Microbiology, 146, 2447- 2456. Radwanski, E.R., Zhao, J. and Last, R.L. (1995) Arabidopsis thaliana tryptophan synthase alpha: gene cloning, expression, and subunit interaction. Mol. Gen. Genet. 248, 657-667. Reintanz, B., Lehnen, M., Reichelt, M., Gershenzon, J ., Kowalczyk, M., Sandberg, G., Godde, M., Uhl, R. and Palme, K. (2001) Bus, a bushy Arabidopsis CYP79F1 knockout mutant with abolished synthesis of short-chain aliphatic glucosinolates. Plant Cell, 13, 351-367. Reuber, TL. and Ausubel, F.M. (1996) Isolation of Arabidopsis genes that differentiate between resistance responses mediated by the RPSZ and RPMl disease resistance genes. Plant Cell, 8, 241-249. Saeed, A.I., Sharov, V., White, J., Li, J ., Liang, W., Bhagabati, N., Braisted, J ., Klapa, M., Currier, T., Thiagarajan, M., Stum, A., Snuffin, M., Rezantsev, A., Popov, D., Ryltsov, A., Kostukovich, E., Borisovsky, 1., Liu, Z., Vinsavich, A., Trush, V. and Quackenbush, J. (2003) TM4: a free, open-source system for microarray data management and analysis. Biotechniques, 34, 374-378. Sasaki-Sekimoto, Y., Taki, N., Obayashi, T., Aono, M., Matsurnoto, F., Sakurai, N., Suzuki, H., Hirai, M.Y., Noji, M., Saito, K., Masuda, T., Takarniya, K-I., Shibata, D. and Ohta, H. (2005) Coordinated activation of metabolic pathways for antioxidants and defence compounds by jasmonates and their roles in stress tolerance in Arabidopsis. Plant J. 44, 653-668. Schaller, F., Biesgen, C., Mussig, C., Altmann, T. and Weiler, E.W. (2000) 12- oxophytodienoate reductase 3 (OPR3) is the isoenzyme involved in jasmonate biosynthesis. Planta, 210, 979-84. Smolen, G.A., Pawlowski, L., Wilensky, SE. and Bender, J. (2002) Dominant alleles of the basic helix-loop-helix transcription factor ATR2 activate stress-responsive genes in Arabidopsis. Genetics, 161, 1235-1246. Soderman, E., Mattsson, J. and Engstrom, P. (1996) The Arabidopsis homeobox gene AtHB-7 is induced by water deficit and by abscisic acid. Plant J., 10, 375-381. Staswick, P.E., Serban, B., Rowe, M., Tiryaki, 1., Maldonado, M.T., Maldonado, MC. and Suza, W. (2005) Characterization of an Arabidopsis enzyme family that conjugates amino acids to indole-3-acetic acid. Plant Cell, 17, 616-627. 90 Swarup, R., Kargul, J ., Marchant, A., Zadik, D., Rahman, A., Mills, R., Yemm, A., May, S., Williams, L., Millner, P., Tsurumi, S., Moore, 1., Napier, R., Kerr, ID. and Bennett, M.J. (2004) Structure-function analysis of the presumptive Arabidopsis auxin permease AUXl. Plant Cell, 16, 3069-3083. Taki, N., Sasaki-Sekimoto, Y., Obayashi, T., Aono, M., Matsumoto, F., Sakurai, N., Suzuki, H., Hirai, M.Y., Noji, M., Saito, K., Masuda, T., Takamiya, K-I., Shibata, D. and Ohta, H. (2005) lZ-oxo-phytodienoic acid triggers expression of a distinct set of genes and plays a role in wound-induced gene expression in Arabidopsis. Plant Physiol. 139, 1268-1283. Tantikanjana, T., Mikkelsen, M.D., Hussain, M., Halkier, BA. and Sundaresan, V. (2004) Functional analysis of the tandem-duplicated P450 genes SPS/BUS/CYP79F1 and CYP79F2 in glucosinolate biosynthesis and plant development by Ds transposition- generated double mutants. Plant Physiol. 135, 840-848. Tao, Y., Xie, Z., Chen, W., Glazebrook, J ., Chang, H.S., Han, B., Zhu, T., Zou, G. and Katagiri, F. (2003) Quantitative nature of Arabidopsis responses during compatible and incompatible interactions with the bacterial pathogen Pseudomonas syringae. Plant Cell, 15, 317-330. Thimm, O., Blasing, O., Gibon, Y., Nagel, A., Meyer, S., Kruger, P., Selbig, J., Muller, L.A., Rhee, S.Y. and Stitt, M. (2004) MapMan: a user-driven tool to display genomics data sets onto diagrams of metabolic pathways and other biological processes. Plant J. 6, 914-939. Tiwari, S.B., Hagen, G. and Guilfoyle. TJ. (2004) Aux/1AA proteins contain a potent transcriptional repression domain. Plant Cell, 16, 533-543. To, J .P., Haberer, G., Ferreira, F.J., Deruere, J ., Mason, M.G., Schaller, G.E., Alonso, J.M., Ecker, J .R. and Kieber, J .J . (2004) Type-A Arabidopsis response regulators are partially redundant negative regulators of cytokinin signaling. Plant Cell, 16, 658-671. Tran, L.S., Nakashima, K., Sakuma, Y., Simpson, S.D., Fujita, Y., Maruyama, K., Fujita, M., Seki, M., Shinozaki, K. and Yamaguchi-Shinozaki K. (2004) Isolation and functional analysis of Arabidopsis stress-inducible NAC transcription factors that bind to a drought- responsive cis-element in the early responsive to dehydration stress 1 promoter. Plant Cell, 16, 2481-2498. Tsuchiya, T., Ohta, H., Okawa, K., Iwamatsu, A., Shimada, H., Masuda, T. and Takamiya, K. (1999) Cloning of chlorophyllase, the key enzyme in chlorophyll degradation: finding of a lipase motif and the induction by methyl jasmonate. Proc. Natl. Acad. Sci. USA, 96, 15362-15367. Turner, J.G., Ellis, C. and Devoto, A. (2002) The jasmonate signal pathway. Plant Cell, 14, $153-$164. 91 Tusher, V.G., Tibshirani, R. And Chu, G. (2001) Significance analysis of microarrays applied to the ionizing radiation response. Proc. Natl. Acad. Sci. USA, 98, 5116-5121. Ueda, T. and Nakano, A. (2002) Vesicular traffic: an integral part of plant life. Curr. Opin. Plant Biol. 5, 513-517. Uknes, S., Mauch-Mani, B., Moyer, M., Potter, 8., Williams, S., Dincher, 8., Chandler, D., Slusarenko, A., Ward, E., and Ryals, J. (1992) Acquired resistance in Arabidopsis. Plant Cell, 4, 645-656. Uppalapati, S.R., Ayoubi P., Weng, H., Palmer, D.A., Mitchell, R.E., Jones, W. and Bender CL. (2005) The phytotoxin coronatine and methyl jasmonate impact multiple phytohorrnone pathways in tomato. Plant J. 42, 201-217. Usadel, B., Nagel, A., Thimm, O., Redestig, H., Blaesing, O.E., Palacios—Rojas, N., Selbig, J ., Hannemann, J., Piques, M.C., Steinhauser, D., Scheible, W.R., Gibon, Y., Morcuende, R., Weicht, D., Meyer, S. and Stitt, M. (2005) Extension of the visualization tool MapMan to allow statistical analysis of arrays, display of corresponding genes, and comparison with known responses. Plant Physiol. 138, 1195-204. Vemoud, V., Horton, A.C., Yang, Z. and Nielsen, E. (2003) Analysis of the small GTPase gene superfamily of Arabidopsis. Plant Physiol. 131, 1191-1208. von Malek, B., van der Graaff, E., Schneitz, K. and Keller, B. (2002) The Arabidopsis male-sterile mutant dde2-2 is defective in the ALLENE OXIDE SYNTHASE gene encoding one of the key enzymes of the jasmonic acid biosynthesis pathway. Planta, 216, 187-92. Vorwerk, S., Biemacki, S., Hillebrand, H., Janzik, I., Muller, A., Weiler, E.W. and Piotrowski, M. (2001) Enzymatic characterization of the recombinant Arabidopsis thaliana nitrilase subfamily encoded by the N1T2/NIT1/NIT3-gene cluster. Planta, 212, 508-516. Wan, J ., Dunning, F .M. and Bent AF. (2002) Probing plant-pathogen interactions and downstream defense signaling using DNA microarrays. F unct. & Integr. Genomics, 2, 259-273. Weigel, R.R., Pfitzner, U.M. and Gatz, C. (2005) Interaction of NIMINl with NPRl modulates PR gene expression in Arabidopsis. Plant Cell, 17, 1279-1291. Werner, T., Motyka, V., Laucou, V., Smets, R., Van Onckelen, H. and Schmulling, T. (2003) Cytokinin-deficient transgenic Arabidopsis plants show multiple developmental alterations indicating opposite functions of cytokinins in the regulation of shoot and root meristem activity. Plant Cell, 15, 2532-2550. 92 Whalen, M.C., Innes, R.W., Bent, AF. and Staskawicz, B.J. (1991) Identification of Pseudomonas syringae pathogens of Arabidopsis and a bacterial locus determining avirulence on both Arabidopsis and soybean. Plant Cell, 3, 49-59. Wildermuth, M.C., Dewdney, J ., Wu, G. and Ausubel, F.M. (2001) Isochorismate synthase is required to synthesize salicylic acid for plant defense. Nature, 414, 562-565. Xie, D.X., F eys, B.F., James, S., Nieto-Rostro, M. and Turner, J .G. (1998) C011 : an Arabidopsis gene required for jasmonate-regulated defense and fertility. Science, 280, 1091 —1094. Xu, L., Liu, F., Lechner, E., Genschik, P., Crosby, W.L., Ma, H., Peng, W., Huang, D. and Xie, D. (2002) The SCF (COIl)ubiquitin-1igase complexes are required for jasmonate response in Arabidopsis. Plant Cell, 14, 1919-1935. Yang, B.J., Oh, Y.A., Lee, E.S., Park, A.R., Cho, S.K., Yoo, Y.J. and Park, OK. (2003) Oxygen-evolving enhancer protein 2 is phosphorylated by glycine-rich protein 3/wall- associated kinase 1 inArabidopsis. Biochem. Biophys. Res. Commun. 305, 862—868. Yuan, J. and He, S.Y. (1996) The Pseudomonas syringae Hrp regulation and secretion system controls the production and secretion of multiple extracellular proteins. J. Bacteriol. 178, 6399-6402. Zeidler, D., Zéihringer, U., Gerber, I., Dubery 1., Hartung, T., Bors, W., Hutzler, P. and Dumer, J. (2004) Innate immunity in Arabidopsis thaliana: Lipopolysaccharides activate nitric oxide synthase (NOS) and induced defense genes. Proc. Natl. Acad. Sci. USA, 101, 15811-15816. Zhao, Y., Hull, A.K., Gupta, N.R., Goss, K.A., Alonso, J ., Ecker, J.R., Normanly, J., Chory, J. and Celenza, J .L. (2002) Trp-dependent auxin biosynthesis in Arabidopsis: involvement of cytochrome P4503 CYP79B2 and CYP79B3. Genes Dev. 16, 3100-3112. Zhao, Y. Thilmony, R., Bender, C.L., Schaller, A., He, S.Y. and Howe, GA. (2003) Virulence systems of Pseudomonas syringae pv. tomato promote bacterial speck disease in tomato by targeting the jasmonate signaling pathway. Plant J. 36, 485-499. Zhou, N., Tootle, TL. and Glazebrook, J. (1999) Arabidopsis PAD3, a gene required for carnalexin biosynthesis, encodes a putative cytochrome P450 monooxygenase. Plant Cell, 11, 2419-2428. Zipfel, C., Robatzek, S., Navarro, L., Oakeley, B.J., Jones, J .D.G., Felix, G. and Boller, T. (2004) Bacterial disease resistance in Arabidopsis through flagellin perception. Nature, 428, 764-767. 93 Chapter 3 A novel function for plant stomata in innate immunity against bacteria Part of this chapter has been previously published in Cell. Melotto, M., Underwood, W., Koczan, J ., Nomura, K., and He, S.Y. (2006) Plant stomata function in innate immunity against bacteria] invasion. Cell, 126, 969-980. I would like to acknowledge Dr. Maeli Melotto for contribution of figures 3-1, 3-6A-C, 3-7A, 3-8, 3-9, and 3-10 and for contributing data to figure 3-6D. 94 Abstract Stomata are microscopic pores on the surface of plant leaves that allow the plant to conduct gas exchange and regulate water relations. Plants regulate the opening and closing of these pores through changes in turgor pressure within the guard cells, the pair of cells that creates a single stomata. Stomata have been instrumental in the evolution of land plants; however, such openings on the surface of the leaf can potentially render the plant vulnerable to invasion by pathogenic microbes. This is particularly relevant for bacterial pathogens that must enter the plant tissue to initiate pathogenesis and, unlike many fungal plant pathogens, are generally unable to directly penetrate the cuticle and cell wall. It has been assumed that stomata serve as passive ports of entry for bacteria, allowing potential pathogens to freely enter the leaf intercellular spaces. However, in this chapter I report that stomata play a role in plant innate immrmity by closing in response to perception of bacteria on the leaf surface. Specifically, stomata of Arabidopsis epidermal peels were found to close in response to both plant and human pathogenic bacteria, and in response to purified PAMPS. Bacteria- and PAMP-induced stomatal closure required ABA biosynthesis and components of the ABA signaling pathway. Unlike human pathogenic bacteria, the plant pathogen Pseudomonas syringae pv. tomato strain DC3 000 was able to re-open Arabidopsis stomata afier approximately three hours of incubation with epidermal peels. Re-opening of stomata by Pst DC3000 was dependent on production of the phytotoxin coronatine. 95 Introduction Entry into host tissue is often a critical first step in the establishment of infection by microbial pathogens. Foliar pathogens of plants face unique challenges in gaining entry into the host tissue. The epidermis and cuticle of the plant leaf must be traversed by microbial pathogens in order to gain access to the intercellular spaces and internal leaf tissues. Many fungal plant pathogens have the ability to directly penetrate the epidermis using cell wall-degrading enzymes, mechanical force, or both. Bacterial pathogens, on the other hand, generally cannot directly penetrate the leaf epidermis and are thought to enter leaf tissues through natural openings or accidental wounds on the leaf surface. Stomata, microscopic pores on the surface of the plant leaf that allow the plant to conduct gas exchange and regulate water relations, provide such natural openings into the leaf intercellular spaces through which bacteria may potentially enter. It is generally believed that openings on the leaf surface such as stomata serve as passive ports of entry for bacteria. However, emerging evidence suggests that plants may have a defense mechanism that acts at an early stage of pathogenesis such as entry into the host tissue. Recent studies have demonstrated the ability of plants to perceive pathogen-associated molecular patterns (PAMPS) such as bacterial flagellin and lipopolysaccharide (LPS) and that perception of PAMPS contributes to resistance to the bacterial phytopathogen Pseudomonas syringae pv. tomato strain DC3000 (Pst DC3000; Felix et al., 1999; Kunze et al., 2004; Zeidler et al., 2004; Zipfel et al., 2004). Interestingly, resistance to Pst DC3000 mediated by perception of flagellin by the FLS2 receptor was found to be effective against bacteria that had been inoculated onto the leaf surface, mimicking a natural infection, but not against bacteria that were artificially delivered into the leaf 96 tissue by vacuum-infiltration, bypassing the leaf surface (Zipfel et al., 2004). However, the mechanism by which PAMP perception contributes to limiting bacterial infection on the leaf surface is not yet known. Pst DC3000 infects both tomato and Arabidopsis and is widely used a model to study bacterial diseases of plants. Plant pathogenic bacteria such as Pst DC3000 have evolved a variety of virulence factors that contribute to their ability to colonize their hosts through subversion of host defenses or release of nutrients (Abramovitch and Martin, 2004; Nomura et al., 2005). The type III secretion system (TTSS) encoded by the hrp genes of Pst DC3000 is one such virulence factor (Alfano and Collmer, 2004; Buttner and Bonas, 2002; He et al., 2004; Mudgett, 2005; Staskawicz et al., 2001). Pst DC3000 uses the TTSS to inject a number of virulence effector proteins into host cells (Chang et al., 2005; Collmer et al., 2003; Greenberg and Vinatzer, 2003; Nomura and He, 2005). In addition to TTSS effector proteins, another virulence factor produced by Pst DC3 000, the phytotoxin coronatine (COR), seems to be required for full virulence. COR is a polyketide toxin that is structurally similar to the plant hormone jasmonic acid (JA). Interestingly, bacterial mutants impaired in COR biosynthesis were found to be significantly reduced in virulence when inoculated onto the surface of Arabidopsis leaves, but fully virulent when infiltrated directly into the leaf intercellular spaces (Mittal and Davis, 1995). This observation prompted the authors to speculate that COR may suppress an “early” defense in Arabidopsis. The nature of this early defense, however, has remained elusive. In this study, plant stomata were found to play an active role in innate immunity against bacteria. Stomata of intact Arabidopsis leaves and epidermal peals were found to 97 close in response to incubation with both plant- and human-pathogenic bacteria. Interestingly, the closure was transient after incubation with the plant pathogen Pst DC3000, but persistent after incubation with the human pathogen E. coli OlS7:H7. This result suggests that plant pathogens, but not human pathogens, have evolved mechanisms to overcome stomatal closure. We found that closure of stomata in response to bacteria or purified PAMPS required ABA synthesis and components of the ABA signaling pathway and that re-opening of stomata by Pst DC3000 was dependent on COR biosynthesis. These results suggest that stomata function as innate immunity gates to restrict entry of bacteria into plant tissues and that overcoming stomatal closure may be a key step required for bacterial pathogenicity on plants. Experimental Procedures Plant Materials Arabidopsis plants (ecotypes Col-0, Col-5, WS, Landsberg erecta [Ler] and mutant lines derived from these ecotypes) were grown in controlled growth chambers at 22°C with a 12hr photoperiod under light intensities of 100 uE/mZ/s. For all experiments, 5- to 6-week-old plants were used. Two independent T-DNA lines of the fls2 mutant (SAIL_691C4 and SALK_93905) were used for the experiments and they showed similar responses to treatments. 98 Chemicals Purified chemicals were used at the following concentrations: 10 uM abscisic acid (ABA, Sigma), 0.2 mM L-NNA (Nto-nitro-L-arginine, Sigma), 0.5 ng/pl coronatine (COR, purchased from C. Bender, Oklahoma State University), 100 ng/ul lipopolysaccharide (LPS from P. aeruginosa, Sigma), 5 uM flg22 peptide (Alpha Diagnostics, Inc. and EZBiolab) Chemicals were diluted in MES (2-(N-morpholino)- ethanesulfonic acid) buffer (25 mM MES-KOH [pH 6.15] and 10 mM KCl), except for LPS solution, which also contained 0.25 mM MgClz and 0.1 mM CaClz. Ultrapure LPS preparations from E. coli 055:B5 (Sigma) and Salmonella Minnesota R595 (Re, Calbiochem) were also tested with similar results. Concentrations of LPS, flg22, and COR were chosen based on previous studies (Zeidler et al., 2004; Zhao et al., 2003; Zipfel etal., 2004) and dose-response experiments. Assessment of Response of Stomata to Treatments To assure that most stomata were open before beginning experiments, we kept plants under light (100 umol.m'2.s") for at least three hours. Fully expanded, young leaves were immersed in water or bacterial suspension [108 colony forming units (CFU).ml'l in water]. At various time points, epidermis of three leaves was peeled off and immediately observed under a microscope (Zeiss Axiophot D-7082 photomicroscope with A3 fluorescence cube or laser scanning confocal microscope). Alternatively, epidermis was peeled from fully expanded leaves and placed on glass slides with the cuticle side in contact with water, MES buffer (25 mM MES-KOH, pH 6.15, and 10 mM KCl), chemical solutions in MES buffer or bacterial suspensions in water. At various 99 time points, pictures were taken of random regions. The width of the stomatal aperture was measured using the software Image-Pro, version 4.5 for windows (Cybernetics, Inc. Silver Spring, MD, USA). We found that stomata in intact leaves and epidermal peels responded similarly to various treatments. All stomatal aperture results reported here were from blind experiments, in which genotypes and treatments were unknown to the experimenters who measured stomatal responses until the completion of experiments. All experiments reported here were performed at least 3 times with similar results. Detection of Nitric Oxide Production in Guard Cells Epidermal peels of Arabidopsis plants were pre-incubated for 3 hr in MES buffer (25 mM MES-KOH, pH 6.15, and 10 mM KCl), soaked in 15 uM DAF-2 DA (4,5- diaminofluorescein diacetate, Sigma) diluted in MES buffer for 20 min, washed three times in MES buffer, and then incubated with chemicals or bacterial suspensions. To assess the COR effect on PAMP-induced NO production, peels were incubated with COR for 30 min prior to addition of purified PAMPS. MES buffer was used as control for purified chemicals and water was the control for bacterial suspensions. Photographs of guard cells were taken with a digital camera attached to a fluorescence microscope equipped with a 502-530 band pass filter. Bacterial Growth Assay Pst DC3000 and mutant derivatives were cultured at 30°C in Luria-Bertani (LB; Sambrook et al., 1989) medium supplemented with appropriate antibiotics until an OD6oo of 0.8 was reached. Bacteria were collected by centrifugation and resuspended in water 100 to the final concentration of 108 CFU.ml'l containing 0.05% Silwet L-77 (Osi Specialties, Friendship, WV). Arabidopsis plants were dipped in bacterial suspension and kept under high humidity until disease symptoms developed. Some plants were vacuum-infiltrated with bacterial suspension at a concentration of 106 CF U.ml°1 containing 0.004% Silwet L- 77. Infiltrated plants were allowed to dry until leaves were no longer water soaked and then covered with humidity domes until completion of the experiment. Bacterial population in the plant apoplast was determined as previously described (Katagiri et al., 2002). All bacterial multiplication assays reported here were performed at least 3 times with similar results. 101 Results Plant- and human-pathogenic bacteria induce closure of stomata Plant stomata have the ability to respond to a variety of environmental stimuli including light, humidity, and C02 concentration (Schroeder et al., 2001; Fan et al., 2004). However, the presence of pores on the surface of the plant leaf could potentially leave the plant vulnerable to invasion by bacteria. To determine if plant stomata can also respond to the presence of live bacteria, we first treated intact Arabidopsis leaves with the plant pathogen Pst DC3000. Prior to treatment, Arabidopsis leaves generally contained 70%-80% open stomata and 20%-30% closed stomata under our growth conditions. After treatment of detached leaves with 1 x 108 cfu/ml Pst DC3000, we observed a marked reduction in the number of open stomata, to ~30% after 2 hours of treatment (data not shown). This reduction in the percentage of open stomata was correlated with a significant decrease in the average width of the stomatal apertures (Figure 3-1A). Observations and measurements of stomatal behavior are commonly performed using epidermal peels rather than intact leaves. Epidermal peels allow stomata to be more easily viewed and recorded microscopically. We therefore tested the ability of stomata in Arabidopsis epidermal peels to respond to Pst DC3000 in a manner similar to those in intact leaves. Afier 1 hour of incubation with 1 x 108 cfu/ml Pst DC3000, stomata in epidermal peels of Arabidopsis leaves were found to behave similarly to those in intact leaves (Figure 3-1B), therefore, epidermal peels were used for all subsequent experiments. 102 In nature, plants are exposed to a variety of bacteria including highly adapted pathogens such as Pst DC3 000, but also epiphytes that survive on the leaf surface and even human or animal pathogens. To determine whether closure of Arabidopsis occurs as a general response to a wide variety of bacteria or is limited to highly adapted plant- pathogenic bacteria such as Pst DC3000, epidermal peels of Arabidopsis leaves were incubated with Escherichia coli 01 57:H7. E. coli 01 57:H7 is an enteric human pathogen that is often a causal agent of food poisoning associated with fresh fruits and vegetables. Despite this association, E. coli 01 57:H7 is not a pathogen of plants and is unable to multiply in plant tissues under normal growth conditions (Figure 3-2A). Stomata in Arabidopsis epidermal peels closed after incubation with 1 x 108 cfu/ml E. coli 01 57:H7, similar to the response to Pst DC3000 (Figure 3-2B). This result suggests that closure of stomata is likely to be a generalized response to perception of a wide variety of bacteria. Interestingly, however, unlike stomata in epidermal peels incubated with Pst DC3000, stomata incubated with E. coli OlS7:H7 remained closed throughout the entire duration of the experiment (8hrs; Figure 3-2B). This result suggests that the adapted plant pathogen Pst DC3000 may have evolved a mechanism (or mechanisms) to overcome the closure of stomata and gain access to the leaf intercellular spaces, whereas the human pathogen lacks this ability. 103 2.5 — A ‘ I Leaf/Water g ‘ ILeaf/DCSOOO E 3 t: OJ 0. <2: .73 (U E .9 (I) B I Peel/Water I Peel/DC3000 Stomatal Aperture (um) Figure 3-1: Pst DC3000 triggers closure of Arabidopsis stomata. Stomatal aperture in (A) intact leaves or (B) epidermal peels of Col-O plants exposed to water (red bars) or 1 x 108 cfu/ml Pst DC3000 (blue bars). Results in panel A are shown as mean (n = 120 stomata) :t SEM. In panel B and all subsequent experiments, results are shown as mean (11 = 60 stomata) 1 SEM. 104 1.0E+06 1 l i ICoI/E.co|i O157:H7 1.0E+05 4 i 1.0E+04 1‘ CFU/om2 1.0E+03 i i 1.0E+02 ‘ 0 2 4 Days Post Inoculation 5 ‘ IWater IO1572H7 Stomatal Aperture (pm) 2h 4h 8h Figure 3-2: (A) Multiplication of E. coli OlS7:H7 in Arabidopsis Col glI plants after inoculation at 1 x 106 cfu/ml of bacteria. Error bars represent standard deviation. (B) E. coli OlS7:H7 triggers persistent closure of Arabidopsis stomata. Stomatal apertures in epidermal peels of Col-0 after incubation with water (green bars) or 1 x 108 cfu/ml E. coli OlS7:H7 (blue bars). 105 Involvement of PAMPs in triggering stomatal closure Pathogen-associated molecular patterns (PAMPs) such as flagellin, lipopolysaccharides and elongation factor Tu (EF-Tu) are perceived by plants and trigger the plant’s innate immune response (Felix et al., 1999; Kunze et al., 2004; Zeidler et al., 2004). Recognition of PAMPs has been shown to contribute to disease resistance (Kim et al., 2005; Zipfel et al., 2004). However, the precise mechanisms triggered by PAMP perception that lead to resistance to infection have not yet been discovered. The ability of stomatal guard cells to respond to both plant- and human-pathogenic bacteria suggests that these cells can sense conserved bacterial molecules such as PAMPs. To test the involvement of PAMPs in bacteria-triggered stomatal closure, we incubated epidermal peels from Arabidopsis Col-0 plants with the purified PAMPs flg22 (a biologically active 22 amino acid peptide derived from a highly conserved domain of eubacterial flagellin) or LPS. We consistently observed a reduction in stomatal aperture after 1 hour of incubation with purified PAMPs (data not shown) and the closure was persistent for at least 4 hours of treatment (Figure 3-3A). Closure of stomata in response to the flg22 peptide was dependent on the FLSZ flagellin receptor, as stomata of fls2 mutant plants did not close in response to treatment with flg22. However, LPS (Figure 3-3A) and live E. coli OlS7:H7 bacteria (Figure 3-3B) were still able to induce stomatal closure on fls2 leaves. These results suggest that perception of flg22 by guard cells requires the flagellin receptor FLSZ, but that F LSZ is likely only one of a number of pattern recognition receptors that allow guard cells to perceive PAMPs and initiate closure. The involvement of PAMPs and FLSZ in stomatal closure suggests that closure of stomata may be an integral component of the Arabidopsis innate immune response to bacteria. 106 IMES Stomatal Aperture (um) Col-0 flsZ Iwater O157:H7 A 444 Stomatal Aperture (um) N Col-0 flsZ Figure 3-3: (A) Purified PAMPs trigger closure of Arabidopsis stomata. Stomatal apertures in epidermal peels of Arabidopsis Col-0 or flsZ mutant (SAIL_691C4) leaves after 4 hours of incubation with MES buffer (red bars), 5 uM flg22 (blue bars), or 100 ng/ul LPS (green bars). (B) Stomatal responses of wild type (Col-0) and flsZ mutant (SALK_93905) plants to 1 x 108 chl E. coli OlS7:H7. Stomatal aperture measurements were determined after 4 hours of exposure to bacteria. 107 ABA synthesis and signaling components are required for bacteria- and PAMP- induced stomatal closure. Closure of stomata in response to environmental cues and abiotic stresses has been studied extensively and is dependent on the plant hormone abscisic acid (ABA). The ABA signal in guard cells is transduced through a network of signaling components including the protein phosphatase 2Cs ABIl and ABIZ, a guard cell-specific kinase, OSTl, nitric oxide (N O), and H202 (Fan et al., 2004; Schroeder et al., 2001). The signaling events triggered by ABA ultimately lead to regulation of ion channels, such as the guard cell outwardly rectifying potassium channel GORKI , that modulate ion efflux from the guard cells (Hosy et al., 2003). Ion efflux from the guard cells drives the efflux of water and results in a change in guard cell turgor that causes the closure of the stomatal pore. To determine if ABA synthesis and/or signaling are required for PAMP- induced closure of stomata we tested a number of mutants defective in ABA synthesis or signaling for their ability to close stomata in response to PAMP perception. The Arabidopsis abaZ-I and aba3-I mutants are defective in ABA biosynthesis and impaired in stomatal closure in response to abiotic stresses (Leon-Kloosterziel et al., 1996; Merlot et al., 2002). Neither flg22 nor LPS could induce closure of stomata in epidermal peels from either aba2-I or aba3-I plants (Figure 3-4A), demonstrating that PAMP-induced closure of stomata requires ABA. To determine if PAMP-induced ABA signaling in guard cells leading to stomatal closure proceeds through known components of ABA signaling, we tested mutant plants defective in components of guard cell ABA signaling pathways for their response to flg22 and LPS. The protein phosphatase 2C proteins encoded by ABA Insensitive I (ABII) and ABIZ are required for stomatal closure in 108 response to ABA (Leung et al., 1997; Pei et al., 1997). Stomata of Arabidopsis abiI and abi2 mutant plants failed to respond to purified flg22 or LPS (Figure 3-4B). Additionally, the Arabidopsis mutants ostI (Mustilli et al., 2002), lacking a guard cell-specific kinase, and gorkI (Hosy et al., 2003), lacking a guard cell outwardly rectifying potassium channel, both failed to close stomata in response to LPS and 0st] failed to close in response to flg22 (Figure 3-5). The response of gorkl to flg22 was not tested as the gorkI mutation is in the WS genetic background that naturally lacks the FLSZ flagellin receptor. Production of the small signaling molecules nitric oxide (N O) and hydrogen peroxide (H202) has also been associated with ABA signal transduction leading to closure of stomata (Fan et al., 2004; Schroeder et al., 2001). The NADPH oxidase double mutant rbohD/F was previously shown to be partially impaired in ABA-induced stomatal closure (Kwak et al., 2003). Similarly, this mutant showed an intermediate response to both flg22 and LPS (Figure 3-5). To test the involvement of NO in PAMP-induced stomatal closure, we performed DAF-2 DA staining on epidermal peels treated with flg22, LPS, or MES buffer to detect production of NO in guard cells. Treatment of epidermal peels with flg22 and LPS rapidly (within 10 min) induced production of NO in guard cells of stomata that eventually closed (Figures 3-6B and 3-6C). Additionally, treatment of epidermal peels with Nco-nitro-L-arginine (L-NNA), an inhibitor of nitric oxide synthase (NOS), prevented closure of stomata in response to flg22, LPS, and E. coli 01 57:H7. Taken together, these results clearly demonstrate that ABA synthesis and downstream signaling components of the ABA signal transduction network in guard cells are required for bacteria- and PAMP-induced stomatal closure. 109 A IMES g lflg22 j; ILPS S t' (D Q. < .79 (U E .9. (I) Col-O aba2—1 aba3-1 B A 4'5 1 IMES E, 41 Ifl922 3; 3-5 1 ILPS ‘5 34, E 2.5 7‘ < 24, .79 1.54, (U e 11 (+5 0.5 4‘ o 1 Col-0 abi1-1 abi2-1 Figure 3-4: Stomatal responses in Arabidopsis ABA (A) biosynthetic mutants aba2 and aba3, and (B) signaling mutants abi1-I and abi2—I . Stomatal aperture measurements were determined after 4 hours of incubation with MES buffer (red bars), 5 uM flg22 (blue bars), or 100 ng/ul LPS (green bars). 110 I MES I LPS Ifl922 Stomatal Aperture (um) Figure 3-5: Stomatal responses of Arabidopsis ABA response mutants 0st], gorkI, and rbohD/F. Stomatal aperture measurements were determined after 4 hours of incubation with MES buffer (red bars), 5 uM flg22 (blue bars), or 100 ng/ul LPS (green bars). Col-0 was wild-type for rbohD/F, Landsberg eracta (Ler) was wild-type for 0st] , and WS was wild-type for gorkI . The response of WS and gorkI to flg22 was not tested as the WS genetic background naturally lacks the FLSZ flagellin receptor. 111 S” 01 1 oo o I—l—l l-—I-—| N or H-I H—l TH til-1 1+! 8 1+: H—i Stomatal Aperture (pm) 2; '5’ SD 01 I .0 o Figure 3-6: Involvement of nitric oxide (N O) in PAMP-induced stomatal closure. (A-C) NO production in guard cells of Col-0 treated with MES buffer (A), 5 uM flg22 (B), or 100 ng/ul LPS (C). Brightfield images are displayed on the left, with fluorescent images displayed on the right. NO production was not observed in the control treatment; therefore, only a black screen was seen under the fluorescent microscope. (D) Effect of the NOS inhibitor L-NNA (0.2mM) on stomatal closure induced by PAMPs (5 pM flg22 or 100 ng/ul LPS) or 1 x 108 cfu/ml E. coli OlS7:H7. L-NNA represents the control treatment of L-NNA dissolved in MES for treatment with purified PAMPs whereas L- NNA“ represents the control treatment of L-NNA dissolved in water for treatment with E. coli. Stomatal aperture measurements were determined after 2 hours of treatment. 112 Pst DC3000 overcomes stomatal closure by production of the phytotoxin coronatine. The observation that Pst DC3000 causes a transient stomatal closure (Figure 3-1), whereas E. coli 01 57:H7 causes persistant closure of stomata (Figure 3-2B) suggests that Pst DC3000 may have evolved a mechanism (or mechanisms) to overcome PAMP- induced stomatal closure and that the non-plant-pathogen E. coli 01 57:H7 lacks this ability. It is likely that Pst DC3000 encodes a virulence factor that can promote re- opening of stomata that initially closed in response to PAMP perception. Pst DC3000 is known to encode two well-characterized virulence factors: a TTSS that secretes a battery of effector proteins, and the phytotoxin COR. To determine if either the TTSS or COR are required for re-opening of stomata by Pst DC3 000 we examined stomatal responses to mutants defective either in the TTSS (nonpolar hrcC”; Penaloza-Vazquez et al., 2000) or in COR synthesis (cor'; Ma et al., 1991). The TTSS-deficient hrcC mutant was not impaired in the ability to re-open closed stomata, behaving similarly to the wild-type Pst DC3000 (Figure 3-7B). The cor' mutant, however, was unable to re-open stomata, similar to E. coli 01 57:H7 (Figure 3-7A). Furthermore, purified COR at a concentration of 0.5 ng/ul could interfere with stomatal closure elicited by both flg22 and LPS (Figure 3-8A). To determine if COR acts downstream of ABA or at an earlier stage in guard cell signaling such as PAMP perception or initial signaling from PAMP receptors, we tested the ability of COR to interfere with stomatal closure induced by exogenous ABA. Stomata of epidermal peels pretreated with 0.5 ng/ul COR and subsequently incubated with 0.5 ng/ul COR + 10 uM ABA failed to close (Figure 3-8B), suggesting that COR acts downstream of ABA to interfere with stomatal closure. COR interference of ABA- induced stomatal closure was not observed in the COR-insensitive Arabidopsis mutant 113 coil. COR did not effect production of NO in guard cells elicited either by ABA or PAMPs (Figure 3-9), suggesting that COR acts downstream or independent of NO synthesis. 114 a" 3 - DWater I DC3000 I 0031 18 N 01 1 EH N l A l Stomatal Aperture (pm) 8 5‘. 0 1h 3h 9" O1 1 El Water I hrcC (um) N no: IL H—l 1.5 _k ____l__ Stomatal Aperture O in j- 1h 3h 1”; Figure 3-7: Stomatal responses to (A) cor' (DC3118) and (B) hrcC mutants of Pst DC3000. Epidermal peels of Col-0 leaves were exposed to water (white bars in both figures), or 1 x 108 cfu/ml Pst DC3000 (red bars), cor' mutant Pst DC31 18 (blue bars), or TTSS-deficient hrcC. Stomatal aperture measurements were determined after 3 hours of treatment. 115 E25 9 2 E 01.5 o. < 1 530.5 E 2 0 (I) go B E 3— 92.5— % 2‘ 0.1.5- < Te 1‘ to - E05 II .9 OJ (0 (D a: < a: < 1: a 8 i a it < < 0‘.) m < < Col-0 coi1-20 Figure 3-8: Purified COR blocks both ABA- and PAMP-induced closure of stomata. (A) Epidermal peels of Col-0 leaves were incubated with MES buffer, 5 uM flg22, or 100 ng/ul LPS with our without 0.5 ng/ul COR. Apertures were measured afier 3 hours of treatment. (B) Epidermal peels of Col-0 or coil -20 mutant leaves were incubated with MES buffer, 0.5 ng/ul COR, 10 M ABA, or 0.5 ng/ul COR + 10 M ABA. For COR + ABA experiments, peels were pre-incubated with COR for 30 min. The COR solution was then replaced with COR + ABA. Apertures were measured after 1 hour of treatment. 116 Figure 3-9: COR does not effect NO production in guard cells in response to ABA or PAMPs. (A-B) Fluorescence micrographs of Col-0 epidermal peels incubated with 10 uM ABA (A) or 10 uM ABA and 0.5 ng/ul COR (B). (C-E) Brightfield (left) and fluorescence (right) images NO production in guard cells of Col-0 epidermal peels incubated with 0.5 ng/ul COR (C), 0.5 ng/ul COR + 100 ng/ul LPS (D), or 0.5 ng/ul COR + 5 uM flg22 (E). 117 Biological relevance of stomatal closure in bacterial infection. The observation that plant stomata close in response to live bacteria and purified PAMPs compellingly suggests that closure of stomata may be an important component of the plant innate immune response and provide a defense barrier against bacterial ingress. However, for this to be the case, stomata must effectively restrict entry of bacteria into the host tissue. Thus, it is important to demonstrate that closure of stomata plays a biologically relevant role in restricting bacterial entry into the leaf and protecting the plant from potential infection. To address this important question, we placed GFP-labeled Pst DC3000 or cor' Pst DC3118 directly underneath an epidermal peel (with the cuticle side in contact with the bacterial suspension) and examined the ability of the bacteria to move through stomata and reach the upper surface of the peel. After 3 hours of incubation, numerous GFP-labeled Pst DC3000 were observed in clusters on the upper surface of Col-0 epidermal peels (Figure 3-10A). In contrast, very few cor' mutant bacteria were observed on the upper surface of epidermal peels (Figure 3-10B). This result directly demonstrates that PAMP-induced stomatal closure can effectively restrict the movement of cor’ mutant bacteria through stomata and that COR-producing Pst DC3000 can overcome stomatal closure and move through stomata. To further test the relevance of PAMP-induced stomatal closure in the context of bacterial infection of intact plants, we examined the ability of the cor' mutant to infect Arabidopsis mutants defective in stomatal closure. When vacuum-infiltrated directly into the apoplast of Arabidopsis Col-0 leaves, cor' mutants achieved a population level similar to DC3000 after 3 days (Figure 3-11A). However, when inoculated on the surface of leaves by dipping, multiplication of cor' mutants in wild-type Col-0 or Landsberg erecta 118 (Ler) leaves was significantly reduced compared to wild-type Pst DC3000 (Figure 3- 11B). These results suggest that the cor' mutant is defective in entering the leaf tissue, but is fully virulent once inside the apoplast. Remarkably, cor' mutant bacteria were able to achieve a population level similar to that of wild-type Pst DC3000 after surface inoculation onto leaves of aba3 or 0st] Arabidopsis mutants, which are impaired in PAMP-induced stomatal closure (Figure 3-11B). Taken together , these results suggest that the primary virulence function of COR is to promote opening of stomata to allow bacterial entry into the leaf intercellular spaces and that this COR-mediated suppression of plant stomatal defense is critical for pathogenicity of Pst DC3000. 119 l—-t '3“ pm Figure 3-10: Movement of GFP-labeled (A) Pst DC3000 and (B) cor' Pst DC3118 through stomata of Arabidopsis Col-0 epidermal peels. Bacterial suspensions (1 x 107 cfu/ml) were placed in contact with the cuticle surface of epidermal peels. Microscopic images show the relative number of bacteria (green) that moved through stomata to reach the upper side of the peel after 3 hours of incubation. 120 CFU/cm2 CFU/crrf Figure 3-1 1: bars) in Col-0 afier vacuum-infiltration of 1 x 106 CFU/ml bacteria. (B) Multiplication of Pst DC3000 and cor' Pst DC3118 in Arabidopsis Col-0 (red bars), ABA-deficient aba3-I (blue bars), Landsberg erecta (grey bars), and ost1-2 (green bars), alter dipping in 1 x 108 CFU/ml bacterial suspensions. Col-0 is the wild-type control for aba3-1, Ler is the control for ost1-2. Multiplication was assessed 3 days after inoculation. Error bars represent standard deviation. 1.0E+08 — I ocsooo 1.oa+o7 . I DC3118 10E+06- 105+05— 1oe+o4- 1.0E+03 — Day 0 Day 3 1.0E+08 - I Col-0 l I aba3-1 1.0E+07 1 3 Le, . tLZ loeme— '” 1 10E+051 i 1.0E+04 f 1 1.0E+03 ‘ DC3118 DC3000 (A) Multiplication of Pst DC3000 (red bars) and cor' Pst DC3118 (blue 121 Discussion In nature, plants encounter and interact with a wide variety of microbes. Many of these microbes may be potential pathogens against which the plant must defend itself. In recent years, a new view of plant defenses against microbes as a multi-tiered system consisting of several layers of defense responses has begun to emerge. The defenses associated with the hypersensitive response (HR) have received the most attention over the past decade, however, the discovery that plants perceive PAMPs has led to an increased appreciation for so-called innate immrmity. In this chapter I have presented evidence for a new layer of innate immunity in plants. We found that stomatal guard cells present on the surface of Arabidopsis leaves could sense the presence of live bacteria and close to prevent bacterial entry into the leaf intercellular spaces. Prior to this discovery, stomata were widely regarded as passive ports of bacterial entry into the host, allowing bacteria to indiscriminately enter the plant. The emerging picture is that of a gauntlet of layered plant defenses that must be overcome in order for bacteria to infect a host plant and cause disease. Pathogenic bacteria must live epiphytically on the leaf surface prior to entering the leaf through wounds or stomata, and during this period, potential pathogens are likely to encounter competition from other epiphytic microbes normally present on the surface of the leaf. Relatively little is known about the initial stages of infection, when bacteria transition from living epiphytically on the leaf surface to a pathogenic lifestyle within the host plant. This is largely due to the use of artificial inoculation procedures in the laboratory, such as vacuum-infiltration, that bypass this transition and introduce bacteria directly into the leaf tissue. The discovery that stomata play a role in defense against 122 bacteria is likely to open up many new areas of inquiry. It is likely that potential bacterial pathogens face challenges before and directly after entry into the host through stomata. A recent discovery suggests that immediately upon entry through stomata, pathogenic bacteria may be challenged with a battery of antimicrobial compounds in the substomatal cavity. The Arabidopsis ATP-binding cassette (ABC) transporter encoded by PEN3/PDR8 has been shown to contribute to penetration resistance against a number of non-host fungi (Stein et al., 2006). Unexpectedly, Arabidopsis pen3 mutant plants were found to be resistant to the normally virulent powdery mildew Erisyphe cichoracearum. This resistance was correlated with chlorosis, cell death, and activation of the SA pathway and was partially suppressed by a mutation at another locus called PEN2. PEN2 encodes a peroxisome-localized glycosyl hydrolase that the authors speculate is involved in the synthesis of an antimicrobial compound that is subsequently exported by the ABC transporter encoded by PEN3. A recent study suggests that PEN3 may also participate in the response to bacterial pathogens and, strikingly, found that PEN3 was strongly and constitutively expressed in the cells of the substomatal cavity (Kobae et al., 2006). This result suggests that plants may have a second line of defense, in the form of secreted antimicrobial compounds, waiting for those bacteria that can overcome stomatal defenses. In addition to the novel discovery of stomata as a relevant battleground for plant- bacteria interactions, our results have also shed light on the mystery of the role of coronatine in bacterial virulence. Our results demonstrate that COR is required for Pst DC3000 to overcome stomatal defenses, as the cor‘ mutant was unable to re-open closed stomata. Additionally, purified COR was sufficient to interfere with stomatal closure 123 induced by PAMPs and ABA. The observation that Arabidopsis mutants defective in closure of stomata, such as aba3 and 0st], allow cor' mutant bacteria to achieve population levels similar to wild-type Pst DC3000 suggests that the primary firnction of COR in bacterial virulence is to overcome stomatal defenses. This finding provides an explanation for the observation of Mittal and Davis (1995) over a decade ago that the cor' mutant was defective in virulence when inoculated on the leaf surface, but not when infiltrated directly into the leaf tissue. Our results demonstrate that the “early defense” postulated by Mittal and Davis is stomatal closure. We did not observe a role for the TTSS and associated effectors in re-opening stomata. Whether the TTSS contributes to bacterial fitness and virulence prior to entering the leaf tissue or is only involved in manipulating host cells after the bacteria have entered through stomata is not yet clear. It will be interesting to determine if the genes encoding TTSS components are expressed on the leaf surface and if the TTSS plays a role in virulence prior to bacterial entry. Only five pathovars of P. syringae are known to produce COR. This raises the question of whether stomatal closure is a general defense response conserved in a broad range of plants, and if so, how other P. syringae pathovars overcome stomatal defense. To begin to address these questions, Dr. Maeli Melotto tested the stomatal response of tomato to Pst DC3000 and LPS and the response of tobacco to P. syringae pv. tabaci. The response of tomato to both LPS and Pst DC3000 was similar to that of Arabidopsis (Melotto et al., 2006). Additionally, P. syringae pv. tabaci, which does not produce COR, induced closure of stomata initially but was able to re-open stomata at later timepoints similar to Pst DC3000. These results suggest that stomatal closure may be a common plant defense response initiated by perception of bacterial PAMPs and that coronatine is 124 only one of a number of virulence factors used by bacteria to overcome stomatal defenses. Collectively, P. syringae pathovars produce a variety of phytotoxins and it will be interesting to determine if any of these other toxins are also involved in the suppression of stomatal defenses. During the course of our observations, we noticed that in experiments where leaves and epidermal peels were incubated with bacteria or purified PAMPs, a population of stomata (generally 25-3 5%) remained open regardless of the treatment. The concentrations of bacteria used were clearly high enough to coat the entire leaf surface with bacteria and we would expect that the leaf surface was evenly exposed to the purified PAMPs, therefore, it is somewhat unexpected that some stomata would remain open in the presence of a stimulus that should elicit closure. Stomata are required to respond to a number of stimuli including light, humidity, CO2 concentration, bacteria, and the circadian clock. How these inputs are prioritized by guard cells and the mechanisms of prioritization are unknown but are interesting and important questions for future study. Plant leaves are generally colonized by epiphytes and potential pathogens, therefore, stomata are likely to perceive bacteria on a regular basis. However, it is likely that only a proportion of stomata are interacting with bacteria at any given time, allowing other stomata to normally carry out their roles in regulating water relations and gas exchange. It is interesting to note that most severe outbreaks of bacterial disease in crop plants are associated with periods of heavy rain or high humidity. It is possible that, under these environmental conditions, stomatal defenses are compromised due to the plant’s need to prioritize stomatal responses to other stimuli, therefore allowing more bacteria to enter the leaf tissue and promoting infection. Further experiments will be 125 needed to determine how plants prioritize guard cell signaling events initiated by multiple external stimuli and whether stomatal defense against bacteria is compromised under conditions of rain or high humidity. Our results demonstrate that stomata close initially in response to PAMPs present on the surface of Pst DC3000 and that Pst DC3 000 can re-open stomata after approximately 3 hours. This observation clearly demonstrates a role for stomata in the early stages of bacterial infection. However, the movements of stomata throughout the entire course of infection have not yet been investigated. It is possible that stomata may also play a role later in the infection process, permitting exit of bacteria from infected leaves and spread to uninfected tissue or to other plants. It will be interesting to determine the role of stomata throughout the entire disease cycle and to determine if COR or TTSS effectors potentially play a role in modulating stomatal movements at later stages of infection. The results presented in this chapter reveal a novel aspect of plant innate immunity against bacteria. Closure of stomata is likely to be an important, and potentially first line of defense against foliar bacterial pathogens and may be a widespread defense mechanism in vascular plants. Therefore, to be a successful foliar pathogen of plants, bacteria must either possess mechanisms to counteract stomatal defenses, or rely on environmental conditions under which stomatal defenses may be compromised in order to enter plant leaves. The discovery of PAMP-induced stomatal closureas a defense mechanism effective against bacterial plant pathogens and of COR as a suppressor of stomatal defense represents a significant advance in our understanding of bacterial 126 pathogenesis of plants and opens up many new avenues for potential future research on bacterial pathogenesis, guard cell signaling, and microbial ecology. 127 References Abramovitch, RB, and Martin, GB. (2004). Strategies used by bacterial pathogens to suppress plant defenses. Curr. Opin. Plant Biol. 7, 356-364. Alfano, J .R., and Collmer, A. (2005). Type III secretion system effector proteins: double agents in bacterial disease and plant defense. Annu. Rev. Phytopathol. 42, 385-414. Buttner, D., and Bonas, U. (2002). Getting across--bacterial type III effector proteins on their way to the plant cell. EMBO J. 21, 5313-5322. Chang, J .H., Urbach, J .M., Law, T.F., Arnold, L.W., Hu, A., Gombar, S., Grant, S.R., Ausubel, F.M., and Dangl, J .L. (2005). A high-throughput, near-saturating screen for type III effector genes from Pseudomonas syringae. Proc. Natl. Acad. Sci. USA, 102, 2549- 25 54. Collmer, A., Lindeberg, M., Petnicki-chieja, T., Schneider, D.J., and Alfano, J.R. (2002). Genomic mining type III secretion system effectors in Pseudomonas syringae yields new picks for all TTSS prospectors. Trends Microbiol. 10, 462-469. Fan, L.M., Zhao, Z., and Assmann, SM. (2004). Guard cells: a dynamic signaling model. Curr. Opin. Plant Biol. 7, 537-546. Felix, G., Duran, J .D., Volko, S., and Boller, T. (1999) Plants have a sensitive perception system for the most conserved domain of bacterial flagellin. Plant J. 18, 265-276. Greenberg, J .T., and Vinatzer, BA. (2003). Identifying type III effectors of plant pathogens and analyzing their interaction with plant cells. Curr. Opin. Microbiol. 6, 20- 28. He, S.Y., Nomura, K., and Whittam, T. (2004). Type III secretion in mammalian and plant pathogens. Biochim. Biophys. Acta. 1694, 181-206. Hosy, E., Vavasseur, A., Mouline, K., Dreyer, 1., Gaymard, F., Poree, F., Boucherez, J., Lebaudy, A., Bouchez, D., Very, A.A., Simonneau, T., Thibaud, J .B., and Sentenac, H. (2003) The Arabidopsis outward K+ channel GORK is involved in regulation of stomatal movements and plant transpiration. Proc. Natl. Acad. Sci. USA, 100, 5549-5554. Katagiri, F., Thilmony, R. and He, S.Y. (2002) The Arabidopsis thaliana-Pseudomonas syringae interaction. In The Arabidopsis Book (Meyerowitz, E. M. and Somerville, C. R., eds). American Society of Plant Biologists, Rockville, MD, USA, DOI 10.1199-tab.0039. Kim, M.G., de Cunha, L., Mcfall, A.J., Belkhadir, Y., DebRoy, S., Dangl, J.L., and Mackey, D. (2005) Two Pseudomonas syringae type III effectors inhibit RIN4-regulated basal defense in Arabidopsis. Cell, 121, 749-759. 128 Kobae, Y., Sekino, T., Yoshioka, H., Nakagawa, T., Martinoia, E., and Maeshima, M. (2006) Loss of AtPDR8, a plasma membrane ABC transporter of Arabidopsis thaliana, causes hypersensitive cell death upon pathogen infection. Plant Cell Physiol. 47, 309- 318. Kunze, G., Zipfel, C., Robatzek, S., Niehaus, K., Boller, T., and Felix, G. (2004) The N terminus of bacterial elongation factor Tu elicits innate immunity in Arabidopsis plants. Plant Cell, 16, 3496-3507. Kwak, J .M., Mori, I.C., Pei, Z.M., Leonhardt, N., Torres, M.A., Dangl, J .L., Bloom, R.E., Bodde, S., Jones, J.D., and Schroeder, J .I. (2003) NADPH oxidase AtrbohD and Atrbth genes function in ROS-dependent ABA signaling in Arabidopsis. EMBO J. 22, 2623- 2633. Leon-Kloosterziel, K.M., Gil, M.A., Ruijs, G.J., Jacobsen, S.E., Olszewski, N.E., Schwartz, S.H., Zeevaart, J.A.D., and Koomneef, M. (1996). Isolation and characterization of abscisic acid-deficient Arabidopsis mutants at two loci. Plant J. 10, 655-661. Leung, J., Merlot, S., and Giraudat, J. (1997) The Arabidopsis ABSCISIC ACID- INSENSITIVE2 (AB12) and ABIl genes encode homologous protein phosphatases 2C involved in abscisic acid signal transduction. Plant Cell, 9, 759-771. Ma, S.W., Morris, V.L. and Cuppels, DA. (1991) Characterization of a DNA region required for the production of the phytotoxin coronatine by Pseudomonas syringae pv. tomato. Mol. Plant Microbe Interact, 4, 69-74. Melotto, M., Underwood, W., Koczan, J., Nomura, K., and He, S.Y. (2006) Plant stomata function in innate immunity against bacterial invasion. Cell, 126, 969-980. Merlot, S., Mustilli, A.C., Genty, B., North, H., Lefebvre, V., Sotta, B., Vavasseur, A., and Giraudat, J. (2002) Use of infrared thermal imaging to isolate Arabidopsis mutants defective in stomatal regulation. Plant J. 30, 601-609. Mittal, S.M., and Davis, KR. (1995) Role of the phytotoxin coronatine in the infection of Arabidopsis thaliana by Pseudomonas syringae pv. tomato. MoI. Plant-Microbe Interact. 8, 165-171. Mudgett, MB. (2005). New insights to the function of phytopathogenic bacterial type III effectors in plants. Annu. Rev. Plant. Biol. 56, 509-531. Mustilli, A.C., Merlot, S., Vavasseur, A., Fenzi, F., and Giraudat, J. (2002). Arabidopsis OSTl protein kinase mediates the regulation of stomatal aperture by abscisic acid and acts upstream of reactive oxygen species production. Plant Cell, 12, 3089-3099. 129 Nomura, K., and He, S.Y. (2005). Powerful screens for bacterial virulence proteins. Proc. Natl. Acad. Sci. USA, 102, 3527-3528. Nomura, K., Melotto, M., and He, S.Y. (2005). Suppression of host defense in compatible plant-Pseudomonas syringae interactions. Curr. Opin. Plant Biol. 8, 361-368. Pei, Z.M., Kuchitsu, K., Ward, J .M., Schwarz, M., and Schroeder, J .I. (1997) Differential abscisic acid regulation of guard cell slow anion channels in Arabidopsis wild-type and abi1 and abi2 mutants. Plant Cell, 9, 409-423. Penaloza-Vazquez, A., Preston, G.M., Collmer, A., and Bender, CL. (2000). Regulatory interactions between the Hrp type 111 protein secretion system and coronatine biosynthesis in Pseudomonas syringae pv. tomato DC3000. Microbiology, 146, 2447- 2456. Sambrook, J ., Fritsch, E., and Maniatis, T. (1989). Molecular Cloning: a Laboratory Manual. Cold Spring Harbor Laboratory, Cold Spring Harbor, NY, USA. Schroeder, M., Allen, G.J., Hugouvieux, V., Kwak, J .M., and Waner, D. (2001). Guard cell signal transduction. Annu. Rev. Plant Physiol. Plant Mol. Biol. 52, 627—658. Staskawicz, B.J., Mudgett, M.B., Dangl, J.L., and Galan, J.E. (2001). Common and contrasting themes of plant and animal diseases. Science, 292, 2285-2289. Stein, M., Dittgen, J., Sanchez-Rodriguez, C., Hou, B.H., Molina, A., Schulze-Lefert, P., Lipka, V., and Somerville, S. (2006) Arabidopsis PEN3/PDR8, an ATP binding cassette transporter, contributes to nonhost resistance to inappropriate pathogens that enter by direct penetration. Plant Cell, 18, 523-528. Zeidler, D., Zéihringer, U., Gerber, I., Dubery, I., Hartung, T., Bors, W., Hutzler, P., and Dumer, J. (2004) Innate immunity in Arabidopsis thaliana: Lipopolysaccharides activate nitric oxide synthase (NOS) and induce defense genes. Proc. Natl. Acad. Sci. USA, 101, 15811-15816. Zhao, Y. Thilmony, R., Bender, C.L., Schaller, A., He, S.Y. and Howe, GA. (2003) Virulence systems of Pseudomonas syringae pv. tomato promote bacterial speck disease in tomato by targeting the jasmonate signaling pathway. Plant J., 36, 485-499. Zipfel, C., Robatzek, S., Navarro, L., Oakeley, B.J., Jones, J.D.G., Felix, G., and Boller, T. (2004) Bacterial disease resistance in Arabidopsis through flagellin perception. Nature, 428, 764-767. 130 Chapter 4 Characterization of the virulence function of HopD2, a type III effector tyrosine phosphatase of Pst DC3000 that suppresses basal defenses in Arabidopsis thaliana 131 Abstract The bacterial pathogen Pseudomonas syringae infects plants by colonizing leaf tissue and multiplying aggressively in the leaf intercellular spaces. To promote multiplication, Pst DC3 000 secretes a battery of effector proteins into host cells via a type III secretion system. Type III effectors are critical for virulence of Pst DC3000 as mutants that cannot secrete effectors are no longer virulent on normally susceptible hosts. However, little is currently known about the molecular mechanisms by which individual type III effectors promote virulence. Most type III effectors identified in the Pst DC3000 genome have little or no similarity to known proteins. However, the effector HopD2 has similarity to protein tyrosine phosphatases, including a conserved active domain. To investigate the in planta function of HopD2 I generated transgenic Arabidopsis plants expressing HopD2 under the control of an inducible promotor. Expression of HopD2 suppresses basal defense- associated callose deposition elicited by the Pst DC3 000 hrpA mutant and flg22-induced resistance to Pst DC3000, suggesting that HopD2 blocks PAMP-induced innate immunity. Transgenic plants expressing HopD2 also allow the normally non-pathogenic hrpA mutant to multiply to high levels within the leaf tissue. The virulence function of HopD2 is dependent on an intact phosphatase catalytic site, as transgenic plants expressing a catalytically inactive derivative do not show these effects. Interestingly, expression of the catalytically inactive HopD2 has a dominant-negative effect on the function of the wild-type HopD2. Remarkably, genome-wide expression profiling analysis suggests that HopD2 acts on host targets without affecting host gene expression. 132 Introduction Type 111 protein secretion systems (W885) and associated secreted virulence effector proteins play a major role in the infection of eukaryotic hosts by bacterial pathogens belonging to many diverse genera including Yersinia, Salmonella, Escherichia, Pseudomonas, Xanthomonas, and Ralstonia (Comelis and Van Gij segem, 2000). The virulence functions of TTSS effectors from mammalian and human pathogens such as Yersinia enterocolitica and Salmonella enterica have been studied extensively and have been shown to participate in manipulation of host cytoskeleton and vesicle trafficking systems to allow bacterial entry into host cells and suppression of host defenses such as phagocytosis (Abrahams and Hansel, 2006; Navarro et al., 2005). In contrast to mammalian and human pathogens, relatively little is known about the virulence functions of TTSS effectors encoded by plant pathogenic bacteria. The recent availability of genome sequences for a number of plant pathogenic bacteria such as P. syringae pv. tomato (Buell et al., 2003), X. campestris pv. campestris (da Silva et al., 2002), X. campestris pv. vesicatoria (Thieme et al., 2005), and R. solanacearum (Salanoubat et al., 2002) has allowed researchers to mine these genomes for potential TTSS effector genes using a variety of bioinformatics and functional approaches (Chang et al., 2005; Fonts et al., 2002; Guttman et al., 2002; Noel et al., 2001; Petnicki-chieja et al., 2002; Roden et al., 2004; Schecter et al., 2004; Zwiesler-Vollick et al., 2002). The results of these studies showed that plant pathogenic bacteria, in general, encode and secrete many more TTSS effector proteins than their mammalian- and human-pathogenic counterparts. For example, the human pathogen Yersinia enterocolitica secretes 6 TTSS effectors (Navarro et al., 2005) compared to 30-40 effectors secreted by P. syringae pv. tomato strain 133 DC3000 (Pst DC3000; Chang et al., 2005; Schechter et al., 2004). The importance of TTSS effectors in promoting virulence is evidenced by the observation that bacterial mutants defective in assembly of the TTSS and secretion of effectors, such as hrp (hypersensitive response and pathogenicity) mutants of P. syringae, are unable to cause disease on normally susceptible hosts (Lindgren et al., 1986). Given the important role of TTSS effector proteins in the ability of plant-pathogenic bacteria to cause disease, understanding the function of these proteins within the host cell should significantly advance the study of plant-microbe interactions. Studying the virulence functions of TTSS effector proteins from plant pathogenic bacteria has proven to be difficult for a number of reasons. First, bacterial mutagenesis approaches to study the function of single effector knock-out mutants have been largely unsuccessful. Single effector mutants generally show little or no reduction in virulence, presumably due to functional redundancy among the many TTSS effectors encoded by a given plant pathogenic bacterium. Second, the gene and protein sequences of TTSS effectors share little or no homology to proteins of known function, rendering comparative approaches uninformative and hindering the ability to generate testable hypotheses for the function of individual effectors. However, limited sequence similarity, such as the presence of conserved enzyme active sites in some TTSS effectors, and new approaches such as transgenic expression of effectors directly in host cells have led to insights into the function of some effectors in the host. The Pst DC3 000 effectors Aerpt2 and AverhB have both been shown to be cysteine proteases that mediate the degradation of specific host proteins. Aerpt2 causes the degradation of the Arabidopsis RIN4 protein, whereas AverhB degrades the Arabidopsis PBSl kinase (Kim et al., 134 2005; Shao et al., 2003). It is not yet clear how the degradation of RIN4 and PBSl contributes to the virulence of Pst DC3000. Other studies suggest that a major virulence function of TTSS effectors from plant pathogens is the suppression of host defense responses. Several Pst DC3000 effectors have been demonstrated to suppress programmed cell death (PCD) associated with the hypersensitive response (HR; Abramovitch et al., 2003; Jarnir et al. 2004). Most notably, AvrPtoB was found to suppress PCD and to possess E3 ubiquitin ligase activity, and this activity was required for PCD supression (Abramovitch et al., 2006; Janjusevic et al., 2006). In addition to HR suppression, other Pst DC3000 effectors such as Hole, AvrPto, Aerme, and Aerpt2 have been shown to suppress defense readouts associated with PAMP-induced innate immunity (DebRoy et al., 2004; Hauck et al., 2003; Kim et al., 2005). The Pst DC3000 TTSS effector HopD2 (also known as HothoDZ and HopAOl) contains a conserved motif (HCxxGxxRS/T) characteristic of protein tyrosine phosphateses (PTPs) and exhibits such enzymatic activities in vitro (Bretz etal., 2003; Espinoza et al., 2003). This potential enzymatic activity provides a starting point for studying the virulence function of this effector protein. HopD2 has previously been shown to contribute to virulence of Pst DC3000 as deletion mutants lacking HopD2 were reduced in virulence on Arabidopsis (10-100 fold reduction in multiplication; Bretz et al., 2003). Additionally, when expressed ectopically in P. syringae pv. phaseolicola (Espinosa et al., 2003) or Pst strain Psy6l (Bretz et al., 2003), HopD2 caused a delay in the onset of HR in the non-host plants Nicotiana tabacum and N. benthamiana respectively. When transiently expressed in N. tabacum, HopD2 was also able to suppress HR-associated PCD initiated by expression of a constitutively active form of the 135 mitogen-activated protein kinase kinase (MAPKK) NtMEK2 (Espinosa et al., 2003). However, in a screen for effectors that could suppress HR-associated PCD elicited by the effector HopPsyA, HopD2 failed to suppress PCD in either N. tabacum or Arabidopsis (Jamir et al., 2004). The contradictory nature of these results leaves open the question of the true virulence function of HopD2 in susceptible host plants. To study the virulence filnction of HopD2 in the susceptible host Arabidopsis, I created transgenic plants expressing HopD2 or a catalytically inactive derivative under the control of an inducible promotor. In this chapter, I report that HopD2 does not suppress the HR when expressed transgenically in Arabidopsis. However, transgenic expression of HopD2 does enhance the multiplication of the normally non-pathogenic Pst DC3 000 hrpA mutant and suppresses defense responses associated with PAMP-induced innate immunity. The phosphatase activity of HopD2 is required for these functions, as transgenic plants expressing the catalytically inactive derivative do not show these phenotypes. To address the firnction of HopD2 in suppressing host defenses, I investigated the activation of defense-associated MAPKs, gene expression changes, and lipid profiles in the transgenic plants. 136 Experimental Procedures Bacterial strains and media Bacterial cultures of Escherichia coli (E. coli) were grown at 37°C in low-salt (Sg/L NaCl) Luria Bertani (LB) medium (Sambrook et al., 1989). P. syringae cultures were grown in low salt LB at 30°C. Antibiotics were used at the following concentrations unless otherwise specified - rifampicin 60 mg/L, kanamycin 50 mg/L, ampicillin 100 mg/L, spectinomycin, 50 mg/L. Plant growth and bacterial enumeration Arabidopsis thaliana plants were grown in controlled growth chambers at a constant temperature of 20°C with a 12 hr photoperiod under a light intensity of 100 uE/mz/s. Four- to six-week-old plants were used for all experiments. Bacteria for plant inoculations were grown to the mid-logarithmic phase, centrifuged at 2500 x g, and re- suspended in sterile H2O to the specified inoculum density. For vacuum-infiltration, Silwet L-77 (Osi Specialties, Friendship, WV) was added to bacterial suspensions at a concentration of 0.004%. After inoculation, plants were left uncovered until leaves were no longer water soaked, then covered with humidity domes until completion of experiments. Bacterial population in the plant apoplast was determined as previously described (Katagiri et al., 2002). The mean values of the bacterial populations are plotted 137 with the standard deviation displayed as error. To assess flg22-induced resistance to Pst DC3000, plants were sprayed with 30 uM dexamethasone 12 hrs prior to syringe- infiltration of 5 uM flg22 or water (control). After 24 hrs, plants were vacuum-infiltrated with Pst DC3000 (1 x 106 cfu/ml) and bacterial multiplication was assessed 2 days after inoculation. All bacterial multiplication assays reported here were performed at least 3 times with similar results. Generation of transgenic plants Genomic DNA was extracted from Pst DC3 000 as described previously (Chen and Kuo, 1993). PCR was used to amplify the hopD2 coding sequence using HIFI Taq polymerase (Invitrogen, Carlsbad, CA) and the following primers: sense primer 5’-TACTCGAGCGAGATAGTTCATACAGCTATG-3 ’ antisense primer 5 ’-CAACTAGTGCGAGAAACACTAAAGGGC-3 ’. The hopD2 gene was cloned into the pBD vector (a gift from Jeff Dangl, University of North Carolina, Chapel Hill), allowing for inducible expression of the transgene after application of the rat glucocorticoid hormone dexamethasone (DEX). The HopD2 insert was sequenced and pBD/hopD2 was transformed into E. coli DHSa and transferred into Agrobacterium tumefaciens strain C58C1 by tri-parental mating using the E. coli helper strain pRK600. Four pots each of A. thaliana Col gII plants (5 plants per pot) were transformed with A. tumefaciens carrying pBD/hopD2 using the floral dip method (Clough and Bent, 1998). Seeds collected from each pot were kept separate to ensure that independently transformed lines could be isolated. T1 pBD/hopD2 seeds were sowed in 138 soil and allowed to germinate. Two-week-old Tl seedlings were sprayed with a 0.2% solution of the herbicide BASTA (glufosinate-ammonium, trade name Finale, AgroEvo Environmental Health, Montvale, NJ) to select transformants. Site-directed mutagenesis of HopD2 A phosphatase catalytically-inactive derivative of hopD2 (hopD2C378S) was generated by site-directed mutagenesis using the QuickChange XL mutagenesis kit (Stratagene, La J olla, CA) according to the manufacturers protocol. The mutagenesis primers used to change the conserved catalytic Cysm to Ser were: . sense primer 5’- GAGTCGCTAGTTGTGCACAGTAACGGCGGTCGGGGCCG-3’ antisense primer 5 ’ -CGGCCCCGACCGCCGTTACTGTGCACAACTAGCGACTC-3 ’. Production of HopD2 antibody The hopD2 gene was amplified by PCR from Pst DC3000 genomic DNA using the following primers: sense primer 5 ’-ATCATATGCCTTTCGTCAGCCAA'ITACTC-3 ’ antisense primer 5’-AGCTCGAGAAACACTAAAGGGCTG-3’. The gene was cloned into the pET28(a) vector (Invitrogen, Carlsbad, CA) and transformed into E. coli BL21. Protein was induced by addition of 1 mM IPTG to a mid- log culture and incubated for 4 hrs at 37°C. HopD2 protein was extracted from E. coli cells using a standard protocol (Qiagen, Valencia, CA) and purified using Ni-NTA agarose. Purified HopD2 protein was analyzed by SDS-PAGE and the band corresponding to HopD2 was excised for further purification. The gel slice was destained 139 for 2 hrs in 30% MeOH and 10% acetic acid followed by rinsing in H2O for 1 hr. The gel slice was then cut into pieces and further macerated by centrifugation in a 1 ml pipette tip loaded with copper wire coil. HopD2 protein was eluted from the macerated gel overnight at room temperature by rocking in 50mM Tris + 1% SDS. The final purified HopD2 protein was analyzed by SDS-PAGE and used to raise antibodies in rabbit at Cocalico Biologicals, Inc., Reamstown, PA. Pre- and post-immune sera was obtained and checked for the ability to recognize the antigen using immunoblot analysis. DEX induction of transgene expression Dexamethasone (Sigma Aldrich, St. Louis, MO) was dissolved in 100% ethanol at a concentration of 30 mM and stored at -20°C. This stock solution was diluted in water to the concentrations specified for individual experiments. Microarray analysis Pots of Arabidopsis Col glI, HopD2, or HopD2C3788 were sprayed with 30 uM DEX to induce the HopDZ transgene. After 24 hrs, pots were vacuum-infiltrated with a bacterial suspension of 1 x 108 cfu/ml of the Pst DC3000 hrpA mutant or a mock inoculum of sterile H2O with 0.004% Silwet L77 (08] Specialties, Friendship, WV). Plants were left uncovered until leaves were no longer water soaked and then covered with a humidity dome. After 7 hrs, all fully expanded leaves from 10-12 plants for each treatment were harvested, frozen in liquid N2, and stored at -80°C for subsequent RNA isolation. For bacterial mutant analysis, Arabidopsis Col-0 plants were vacuum-infiltrated with bacterial suspensions of 1 x 108 cfu/ml Pst DC3000, Pst DC3000 AhopDZ mutant, 140 or a mock inoculum of sterile H2O with 0.004% Silwet L77. Infiltrated plants were left uncovered until leaves were no longer water soaked and then covered with a humidity dome. Tissue was harvested for RNA isolation 7 hrs after infiltration in the manner described above. Total RNA was isolated from Arabidopsis leaves using the Promega (Madison, WI) RNAgents total RNA isolation system. The RNA concentration was determined by absorbance at 260 nm and its quality was evaluated by separation on 2% denaturing agarose gels containing formaldehyde. The total RNA was further pmified using Qiagen (Valencia, CA) RNeasy minicolumns. Biotinylated target complementary RNA (cRNA) was prepared from 16pg of total RNA using the procedure described by Affymetrix (Santa Clara, CA). Labeled target cRNA was subsequently purified, fragmented, and hybridized to Arabidopsis ATHl GeneChip arrays according to protocols provided by the manufacturer (Affymetrix) in a Hybridization Oven model 640 (Affymetrix). The GeneChips were washed and stained with streptavidin-phycoerythrin using a GeneChip Fluidics Station model 400 and then scanned with a Gene Array Scanner (Hewlett-Packard, Palo Alto, CA). The gene expression data was normalized and analyzed using the Affymetrix Microarray Program Suite (MAS 5.0 statistical algorithms and the Data Mining Tool (version 2.0)). Output from all GeneChip hybridizations was scaled globally such that its average intensity (TGT value) was equal to an arbitrary target intensity of 500 allowing comparisons between GeneChips. Data was then compared between sample chips fi'om the same biological replicate producing a Signal Log2 Ratio (SLR) calculated from the GeneChip fluorescence signal intensity data. The software was used to determine whether there was a genuine change in mRNA accumulation (Change Call, D for Decrease, I for Increase) and Change Call p-value. 141 SLRs, Change Calls and p-values were determined for each bacterially inoculated sample compared with its corresponding mock control or bacterial mutant treated sample. Significance Analysis of Microarrays (SAM; Tusher et al. 2001) was used to identify significant genes based on their differential expression between sets of samples. The SAM analysis feature contained in the TIGR Microarray Experiment Viewer v4.0 (Saeed et al., 2003; http://www.tm4.org/mev.html) was used to conduct the analysis. The normalized signal log2 ratios for each replicate of each comparison were used for the analysis. The two datasets (transgenic plants and bacterial mutant) were independently analyzed using the multiclass design, with the default settings. The delta value was set at a level such that the median False Discovery Rate was less than 5%. Access database management software was used to further filter and query the data. The criteria used to identify reproducibly differentially expressed probe sets were a signal log2 ratio of at least -1.0, a gene expression change call of D (decrease) and p-value >0.995 or signal log2 ratio of at least 1.0, change call of I (increase) and p-value of <0.005 for all 3 biological replicates for at least one type of comparison. Lipid profiling analysis Arabidopsis Col g1], HopD2, or HopD2C3788 plants were sprayed with 30 uM DEX to induce the transgene. After 24 hrs, 5 fillly expanded leaves were harvested for extraction of total lipids. Harvested leaves were quickly immersed in isopropanol + 0.01% butylated hydroxytoluene (BHT; Sigma Aldrich, St. Louis, MO) preheated to 75°C in a water bath and incubated for 15 min. Samples were removed from water bath and 1.5 ml of chloroform and 0.6 ml water were added with vortexing. Samples were 142 incubated with shaking at room temperature for 1 hr. Lipid extracts for each sample were transferred to a separate tube and 4 ml chloroformzmethanol (2:1) + 0.01% BHT was added to the leaf samples. Samples were incubated for 30 min. at room temperature with shaking. Lipid extracts were again removed and combined with the first extract. Extraction with 2:1 chloroformzmethanol was repeated 4 times and extracts for each sample were combined. Extracts were rinsed with 1 ml of 1M KCl followed by 2 ml of H20. Solvents were evaporated by streaming nitrogen into the tubes and lipids were resuspended in 0.5 ml chloroformzmethanol (2:1) + 0.01% BHT and stored at -80°C. Extracted leaves were dried overnight in an oven and weighed to obtain dry weights. Dry weights for each sample were 8-10 mg. Lipid samples were analyzed using high pressure liquid chromatography (HPLC) coupled to a Waters MicroMass dual quadrupole mass spectrometer using electrospray ionization (ESI) in the negative ion mode. To detect phosphatidylinositol lipids, a parent ion method was used to scan the mass spectrum for lipids that fragmented to yield parent ions of 241 m/z (representing the phosphatidylinositol headgroup with no reversible phosphorylation) or 321 m/z (representing the phosphatidylinositol headgroup with a single reversible phosphorylation). Scans were conducted for 3 min. per sample. Data on P1 content reported here are from single MS/MS analyses of single lipid extracts for each sample. Callose assay Arabidopsis Col glI, HopD2, or HopD2C378$ plants were sprayed with 30 uM DEX and after 24 hrs were vacuum-infiltrated with a suspension of the type III secretion- defective Pst DC3000 hrpA mutant (Wei et al., 2000) at a concentration of 1 x 108 cfu/ml 143 (OD600 = 0.2). Plants were left uncovered after infiltration and leaves were harvested after 7 hrs. Harvested leaves were cleared of pigment by vacuum-infiltrating alooholic lactophenol (1 :1 :1 :1 :2 phenol:glycerol:lactic acidzwaterzethanol) followed by 30 min. of incubation at 65°C. The leaves were then transferred to fresh alcoholic lactophenol solution and incubated overnight at room temp. Cleared leaves were rinsed briefly in 50% ethanol, then water, and stained with aniline blue solution (0.01% aniline blue [Sigma] in 150 mM K2HPO4, pH 9.5). Leaves were examined with a Zeiss Axiophot D-7082 fluorescence microscope with an excitation filter of 365 :1: 25 nm, a 400 nm dichroic mirror, and a 450 nm longpass emission filter. The number of callose deposits (visualized as white spots against a blue background) was determined using Quantity One software (BioRad, Hercules, CA). More than 10 adjacent fields of view along the length of the leaf were analyzed and averaged. Average numbers of callose deposits are plotted with standard deviation displayed as error. Callose assays reported here were performed 3 times with similar results. In-gel kinase assay Arabidopsis Col glI, HopD2, or HopD2C3788 plants were sprayed with 30 11M DEX and after 24 hours were inoculated by vacuum-infiltration with a bacterial suspension of 1 x 108 cfu/ml of the Pst DC3000 hrpA mutant or a mock inoculum of H20. Leaf tissue samples were collected at 0 (immediately prior to vacuum-infiltration), 1, 2, 3, and 6 hrs after inoculation. Four leaf discs (1.5 cm2) per sample were collected using a #5 cork borer and immediately frozen in liquid N2. For analysis of kinase activation after inoculation of bacterial mutants, Arabidopsis Col gII plants were inoculated with 144 bacterial suspensions of l x 108 cfu/ml of Pst DC3000, a Pst DC3 000 hopD2 deletion mutant (AhopDZ; Bretz et al., 2003), or Pst DC3000 carrying the avirulence gene aerpt2. Leaf tissue samples were collected at 0 (immediately prior to vacuum- infiltration), 3, 6, and 9 hours after inoculation as described above. Frozen leaf tissue was ground using plastic micro pestles and 200 pl of extraction buffer (100 mM Hepes, pH 7.5, 5 mM EDTA, 5 mM EGTA, 10 mM DTI‘, 10 mM Na3V04, 10 mM NaF, 50 mM [3- glycerolphosphate, 1 mM phenylmethylsulfonyl fluoride, 5 ug/mL antipain, 5 ug/mL aprotinin, 5 pg/mL leupeptin, 10% glycerol) was added to each sample followed by further homogenization until no large leaf pieces remained. Homogenized samples were centrifuged at 12,000 x g at 4°C for 30 min and supematants were transferred to new tubes and stored at -80°C. Protein concentration for each sample was determined using a Pierce BCA Protein Assay Kit (Pierce, Rockford, IL) according to the manufacturers instructions. 10 pg of protein for each sample was separated on 10% SDS polyacrylamide gels containing 0.25 mg/ml of myelin basic protein (MBP) and in-gel kinase assays were performed as described previously (Zhang and Klessig, 1997). All in-gel kinase assays reported here were performed two times on independent samples with similar results. Immunoblotting and HopD2 protein fractionation Following DEX treatment, one cm2 of leaf tissue was collected using a #2 cork borer, homogenized in 100 pl of protein extraction buffer (0.125M Tris-HCl, pH 6.8, 4% SDS, 20% glycerol, 10% B-mercaptoethanol) and denatured at 100°C for 10 min. Equal volumes of each sample were separated on 12% SDS-polyacrylarnide gels and proteins were transferred onto Immobilon-P blotting membrane (Millipore, Billerica, MA) using a 145 Semi-Phor semi-dry transfer apparatus (Hoefer Scientific Instruments, San Francisco, CA). Immunoblotting was carried out using HopD2 antiserum and anti-rabbit alkaline phosphatase or horseradish peroxidase conjugated secondary antibodies. HopD2 protein bands were visualized by a standard color reaction using Sigma Fast BCIP/NBT reagent (Sigma Aldrich, St. Louis, MO) or by enhanced chemiluminescence using a Pierce SuperSignal chemiluminescent detection kit according to the manufacturers protocol (Pierce, Rockford, IL). For localization of HopD2 to membrane or soluble fractions of extracts of HopD2 transgenic Arabidopsis plants, plants were sprayed with 30 uM DEX to induce the transgene. After 24 hrs, 2.5 cm2 of leaf tissue was collected using a #5 cork borer and homogenized in 200 pl extraction buffer (50 mM Tris-HCI, pH 8.0, 150 mM NaCl, 10% glycerol, 10 uM B-mercaptoethanol, 1x Sigma plant protease inhibitor cocktail). Following homogenization, an additional 200 pl of extraction buffer was added to each sample. Samples were centrifuged at 10,000 x g for 30 min at 4°C and supernatants were transferred to new tubes. Samples were then centrifuged at 100,000 x g for 1 hour at 4°C. The supernatant (soluble fraction) was removed and the pellet (membrane fraction) was resuspended in 400 ill extraction buffer. Protein fractions were concentrated approximately 8-fold using Centricon microcentrifuge units (Millipore, Bellerica, MA) with a 10,000 MW cutoff. After concentration, 20 pl of each sample was mixed with 5x SDS-PAGE loading buffer (0.2M Tris-HCl pH 6.8, 40% glycerol, 10% SDS, 0.0005% bromophenol blue, 15% B-mercaptoethanol), loaded and separated on 12% SDS polyacrylamide gels, and blotted as described above. HopD2 was detected as described above. 146 Results Arabidopsis thaliana transgenic plants expressing HopD2 exhibit chlorosis and necrosis upon induction. To study the function of HopD2 in promoting virulence on susceptible host plants, I created transgenic Arabidopsis Col glI plants expressing the hopD2 gene or the catalytically inactive derivative hopD2C378S under the control of a DEX-inducible promotor. Transgenic lines were selected on the basis of their resistance to the herbicide BASTA. Seven independent HopD2 transgenic lines and 11 independent HopD2C3788 lines were analyzed by western blotting to detect the presence of the HopD2 protein after induction with DEX. Each line was scored for the development of phenotypes after DEX treatment. Two HopD2-expressing lines and two HopD2C37ss-expressing lines were subsequently assayed for multiplication of the Pst DC3 000 hrpA mutant. HopD2 line 141 and HopD2C378$ line 193 were chosen for further characterization and the experiments reported here were performed using these transgenic lines. Un-induced HopD2 and HopD2C3788 plants were phenotypically normal and indistinguishable from the parental Col glI line. However, by 3 days after induction of the HopD2 protein by spraying with 30 pM DEX, HopD2 141 plants began to develop leaf chlorosis (Figure 4-1A). By 5 to 6 days after DEX treatment, leaves of HopD2 141 plants began to exhibit a necrotic phenotype and eventually collapsed, resulting in death of the transgenic plants (Figure 4- 1B). In contrast, HopD2C3785 193 plants did not develop these severe phenotypes after DEX induction and were generally indistinguishable from Col glI plants (Figure 4-1 A 147 and B). Occasionally, HopD2C3785 developed mild chlorosis after several days of DEX treatment. 148 Col glI HopD2 HopDzC378$ Col glI HopD2 HopD2C3788 Figure 4-1: Leaves of Col glI or HopD2 transgenic plants 3 days (A) or 7 days (B) after treatment with 30 uM DEX. Plants were sprayed with DEX every 48 hours. Leaves of HopD2 transgenic plants developed chlorosis and eventually collapsed. Note that leaves of HopD2C37 3 plants did not develop this phenotype. 149 HopD2 transgenic plants allow enhanced multiplication of the Pst DC3000 hrpA mutant. Collectively, TTSS effectors secreted by Pst DC3000 into host cells function to promote multiplication of the bacteria within the apoplastic space. Mutants that are defective in assembly of the TTSS such as the hrpA mutant cannot secrete effectors and are rendered completely non-pathogenic. Previous studies have demonstrated that transgenic expression of single Pst DC3000 TTSS effectors in Arabidopsis can be sufficient to promote significant multiplication of normally non-pathogenic hrp mutants (Hauck et al., 2003; Kim et al., 2005). To determine if transgenic expression of HopD2 in Arabidopsis could promote multiplication of normally non-pathogenic TTSS-deficient bacteria, HopD2 and HopD2C3788 plants were inoculated with the Pst DC3000 hrpA mutant. HopD2 transgenic plants allowed hrpA to multiply to significantly higher population levels than the parental Col glI plants (Figure 4-2A). Irnportantly, this enhanced multiplication was dependent on an intact HopD2 phosphatase catalytic site as multiplication of hrpA in HopD2 C3783 was indistinguishable from Col gII . Although hrpA was able to achieve population levels approaching that of Pst DC3 000 after 3 days, the TTSS-deficient mutant was not able to cause typical disease symptoms such as water soaking, chlorosis, or necrosis. DEX treated HopD2 plants inoculated with hrpA were visually similar to un-inoculated, DEX treated plants. Expression of HopD2 had no effect on the multiplication of Pst DC3000 (Figure 4-2B). Interestingly, while HopD2C3788 plants did not allow the hrpA mutant to multiply, transgenic expression of HopD2 C3783 had a dominant-negative effect on Pst DC3000 multiplication, causing a 5-10-fold reduction in population levels (Figure 4-2B). To 150 determine if this effect is a specific, dominant-negative effect on the virulence fimction of the native HopD2 secreted by Pst DC3 000 during the course of an infection, the multiplication of Pst DC3 000 AhopDZ, a deletion mutant lacking HopD2, was investigated in C01 gll, HopD2 and HopD2 C3785 plants. Expression of wild-type HopD2 in planta complemented the AhopDZ mutant, allowing the mutant bacteria to achieve a population level similar to wild-type Pst DC3000. Multiplication of AhopDZ was similar in C01 g1] and HopD2 C3785 (Figure 4-2B) suggesting that: 1) Expression of HopD2 C3785 cannot complement the virulence defect of the AhopDZ mutant and, 2) HopD2 C3788 likely interferes specifically with the function of wild-type HopD2 delivered by Pst DC3000 as the virulence of the AhopDZ mutant is not further reduced in HopD2 C3783 plants. 151 1.0E+07 . 1 lCol gl1 1.0E+06 l ' H°pDZ 1 I HopD2 C3788 N ‘1 E 1.0E+05 1 U 1 \ D 1 LL 1.0E+04 1 0 1 1 1.0E+03 4‘ 1 1.0E+02 T o 3 Days Post inoculation B 1 E+09 1 I Col gl1 1.E+08 1‘ I H°PDZ 1 I HopD2 C3788 1 E+07 7—1 " 1 5 1 5 1 E+06 1 LI- 1 O 1 1.E+05 1 1 1 E+04 1 1.E+03 DC3 000 AhopD2 Figure 4-2: Bacterial multiplication in HopD2 transgenic plants. (A) Multiplication of the Pst DC3000 hrpA mutant in C01 glI (red bars), HopD2 (blue bars), or HopD2 "‘5 (green bars) plants. Plants were vacuum-infiltrated with a bacterial suspension of 1 x 106 cfu/ml. Error bars represent standard deviation for both charts. (B) Multiplication of Pst DC3000 or the AhopDZ mutant in C01 glI (red bars), HopD2 (blue bars), or HopD2C 3785 (green bars) plants. Plants were vacuum-infiltrated with bacterial suspensions of 1 x 106 cfu/ml and multiplication was assessed 3 days after inoculation. Bacterial populations immediately after inoculation (Day 0) were averaged and are represented by the solid horizontal line with standard deviation represented by the dashed horizontal lines. 152 Transgenic expression of HopD2 suppresses defense responses associated with PAMP-induced innate immunity. TTSS effectors such as HopD2 can potentially promote multiplication and/or symptom development through several possible mechanisms. Effectors may contribute to virulence by suppressing host defense responses such as those associated with PAMP- induced innate immunity or the HR, or by manipulating host cells to promote the creation of an apoplastic environment that is more suitable for bacterial multiplication, potentially through the release of water and/or nutrients. Transgenic plants expressing HopD2 allow multiplication of the hrpA mutant, which is normally unable to multiply in Arabidopsis presumably due, at least in part, to it’s inability to suppress PAMP-induced innate immunity. Therefore, I investigated the ability of HopD2 transgenic plants to mount defense responses associated with PAMP-induced innate immunity. One such defense response is the deposition of callose-rich papillae at the cell well (Gomez-Gomez et al., 1999; Soylu et al., 2005). Perception of PAMPs present on non-pathogens such as the hrpA mutant leads to the deposition of papillae that can be visualized by aniline blue staining for callose. After inoculation with the hrpA mutant, Col glI leaves exhibited characteristic callose deposits (white spots on blue background) when stained with aniline blue (Figure 4-3A). In contrast, HopD2 plants exhibited a significant reduction in callose deposition (Figure 4-3A) to less than 50% of the number of callose deposits observed in C01 g1] leaves (Figure 4-3B). An intact phosphatase catalytic site was required for HopD2 suppression of callose deposition as the number of callose deposits in HopD2C378 leaves was not significantly different than in C01 glI (Figure 4-3B). 153 To further test the effect of HopD2 on PAMP-induced innate immunity, the ability of treatment with the flg22 peptide to confer partial resistance to Pst DC3000 was investigated in C01 glI , HopD2, and HOpDZC3788 plants. Infiltration of Arabidopsis leaves with the flg22 peptide prior to or along with Pst DC3000 has been shown to reduce bacterial multiplication (Zipfel et al., 2004). Consistent with published results, treatment of Col glI leaves with 5 llM flg22 24 hrs prior to infiltration with Pst DC3000 resulted in an ~10-fold reduction in bacterial multiplication compared to H2O-treated control leaves (Figure 4-4). Multiplication of Pst DC3000 in HopD2 transgenic plants did not differ significantly between leaves treated with flg22 and H2O-treated (control) leaves, suggesting that HopD2 can block defenses activated by flg22. Pst DC3000 population levels in H2O-treated HopD2C3788 leaves were ~5 fold less than in H2O-treated Col glI leaves, consistent with our prior observation of a dominant-negative effect of HopD2C3788 on Pst DC3000 multiplication. Multiplication of Pst DC3000 was further reduced in flg22-treated HopD2C37°S leaves, suggesting that HopD2 phosphatase activity is required for suppression of flg22-induced resistance to Pst DC3000. 154 HopD2 HopD2C378S H2O hrpA 160 ICol 911 140 4 , IHoppz 120 j IHopoz C3788 é 8 Average Callose Deposits o 8 8 8 Water hrpA Figure 4-3: Papilla-associated callose deposition in HopD2 transgenic plants. Col glI, HopD2, and HopD2C3788 plants were inoculated with Pst DC3000 hrpA mutant bacteria (1 x 108 cfu/ml) or sterile H2O. Leaf tissue was collected after 8 hrs, cleared, and stained with aniline blue to detect callose. (A) Representative microscope images, callose deposits appear as white spots on the blue background. (B) Quantitation of callose deposition. Callose spots were counted for at least 10 microscope fields of view per treatment. Average numbers of deposits per field are plotted with standard deviation displayed as error. 155 CFUIcm2 1.00E+08 7 .Water 1 Ifl922 1.00E+07 1 1 1.00E+06 j i 1.00E+05 1 1 1.00E+04 j l- _ 1.00E+03 Col gl1 HopD2 HopDZ 03788 Figure 4-4: Transgenic expression of HopD2 blocks flg22-induced resistance to Pst DC3000. Multiplication ofPst DC3000 in C01 glI, HopD2, or HopD2 C3733 plants after treatment with 5 11M flg22 (blue bars) or water (red bars) as a control. Leaves were syringe-infiltrated with flg22 or water 24 hrs prior to vacuum-infiltration with Pst DC3000 (1 x 108 cfu/ml). Multiplication was assessed 2 days after inoculation. Error bars represent standard deviation. Bacterial populations immediately after inoculation (Day 0) were averaged and are represented by the solid horizontal line with standard deviation represented by the dashed horizontal lines. 156 Development of the HR is not effected in HopD2 transgenic plants. In previous studies, transient expression of HopD2 was sufficient to block or delay the onset of HR-associated PCD elicited by avirulent bacteria in the non-host plants N. benthamiana and N. tabacum (Bretz et al., 2003; Espinosa et al., 2003). However, in a screen for Pst DC3000 TTSS effectors that could suppress PCD elicited by the effector HopPsyA in N. tabacum, HopD2 failed to suppress or delay HR-associated PCD (Jarnir et al., 2004). These experiments were all carried out by transiently expressing HopD2 in non-host Nicotiana plants. To determine if HopD2 could suppress or delay the onset of HR-associated PCD in stable transgenic Arabidopsis plants, we investigated the ability of the transgenic plants expressing HopD2 to mount the HR in response to Pst DC3000 carrying the avirulence gene AerptZ. Aerpt2 is secreted by Pst DC3000 into Arabidopsis cells where it is recognized by the R protein RPS2, leading to the HR (Kunkel et al., 1993; Yu et al., 1993). Under our laboratory conditions, recognition of Aerpt2 by RPS2 led to the onset of the HR in C01 glI plants approximately 14-16 hours after bacterial inoculation (Figure 4-5). Development of the HR in HopD2 and HopD2C3788 transgenic plants was indistinguishable from Col glI in both the severity and timing of PCD-associated tissue collapse. Similar results were obtained when plants were inoculated with Pst DC3000 carrying the avirulence genes AvrB or AverhB (data not shown). These results suggest that the virulence function of HopD2 in the susceptible host Arabidopsis is not related to suppression of HR-associated PCD. 157 Col glI HopD2 HopD2C3788 16 hrs 24 hrs Figure 4-5: Onset of HR in C01 g1] and HopD2 transgenic plants. Col glI , HopD2, and HopDZC378$ plants were sprayed with 30 11M DEX 24 hrs prior to vacuum-infiltration with bacterial suspensions of 1 x 108 cfu/ml Pst DC3000 carrying the avirulence gene aerptZ. Onset of HR was observed by monitoring tissue collapse associated with cell death. Representative leaves were detached for imaging at 16 hrs (upper panel) and 24 hrs (lower panel) after inoculation. Note that neither the timing, nor severity of the HR was affected by transgenic expression of HopD2. 158 Activation of the Arabidopsis MAPKs AtMPK3 and AtMPK6 is enhanced in HopD2 transgenic plants. Our observations that transgenic expression of HopD2 allows multiplication of the hrpA mutant, reduces papilla-associated callose deposition, and blocks flg22-induced resistance to Pst DC3000 suggest that the virulence function of HopD2 in host plants is to suppress PAMP-induced innate immunity by dephosphorylating a host target or targets. There are several potential mechanisms through which HopD2 could act to block defenses associated with PAMP-induced innate immunity including blocking early signaling events initiated by pattern recognition receptors (PRRs) such as FLS2, altering gene expression changes associated with innate immunity, or blocking later events such as vesicle trafficking leading to deposition of papillae. Activation of mitogen-activated protein kinases (MAPKs) is known to be an early step in the signal transduction cascade initiated by the FLS2 receptor (Asai et al., 2002). The MAPKs AtMPK3 and AtMPK6 are activated in response to perception of flg22 by FLS2. Because MAPKs require tyrosine phosphorylation to become activated, we hypothesized that HopD2 may block PAMP-induced signaling initiated by PRRs by dephosphorylating and thus, inactivating AtMPK3, AtMPK6, or both. To test this hypothesis, we investigated the activation of AtMPK3 and AtMPK6 in HopD2 transgenic plants and in wild-type Col plants after inoculation with Pst DC3000 or the AhopDZ mutant. We observed no significant difference in activation of AtMPK3 or AtMPK6 in C01 gl I plants inoculated with Pst DC3000 or the AhopDZ mutant. Both Pst DC3000 and the AhopDZ mutant began to elicit activation of AtMPK6 above background levels by 9 hrs after vacutun-infiltration with more significant activation by 12 hrs after infiltration (Figure 4-6A). Activation of 159 AtMPK3 was not observed in response to either Pst DC3 000 or the A hopD2 mutant at the time-points observed in our experiments. As a control, Col glI plants were inoculated with Pst DC3000 carrying the avirulence gene aerpt2. This strain elicits the HR and caused strong activation of both AtMPK3 and AtMPK6 as early as 3 hours after infiltration. To our surprise, HopD2 transgenic plants exhibited significant activation of AtMPK6 and slight activation of AtMPK3 compared to C01 glI or HopD2C37°° prior to bacterial inoculation (0 hrs; Figure 4-6B). In addition to their role in signaling initiated by the FLS2 receptor, AtMPK3 and AtMPK6 are activated by multiple biotic and abiotic stresses including wounding. Constitutive activation of these kinases in plants expressing HopD2 may result from general stress as these plants develop chlorosis and necrosis after DEX treatment. AtMPK3 and AtMPK6 were activated even more strongly in HopD2 plants 1 hr after vacuum-infiltration with the Pst DC3000 hrpA mutant. HopD2C37°S plants also exhibited slightly higher levels of kinase activation 1 hr after inoculation compared to C01 glI . Taken together, our results with transgenic plants and the AhopDZ bacterial mutant suggest that HopD2 does not exert its virulence activity by inactivating the MAPKs AtMPK3 or AtMPK6. These results derived from bacterial infection of stable transgenic plants are consistent with a recently published study in which HopD2 failed to block flg22-elicited activation of AtMPK3 or AtMPK6 in Arabidopsis protoplasts (He et al., 2006), suggesting that HopD2 acts downstream or independent of kinase activation to block PAMP-induced innate immunity. 160 Pst DC3000 Pst DC3000 1’“ DC3000 AhopDZ (aerptZ) HP10369120369120369 err-“j. r," ,“""‘*‘W’iifi‘* "‘_‘-”.“'"-““ ”’VPTTWT*WW 3‘ _ 1 AtMPK6 .1" " ~=*~-*i'-- -----r In- L.,... s. .__-.."s....-.... -4‘ ... ._ __ L' — ' ‘ “the *i ‘“ AtMPK3 B Col glI HopD2 HopD2C3788 l l HP1012360123601236 H20 * - ” rung L.,. ”a”, 1’- ! "l. . . . , Figure 4-6: (A) Activation of the Arabidopsis MAPKs AtMPK3 and AtMPK6 in C01 gl 1 leaves after vacuum-infiltration with 1 x 10° cfu/ml suspensions of Pst DC3000, Pst DC3000 AhopDZ, or Pst DC3000 (aerptZ). Leaf tissue was collected at the indicated time-points and protein extracts were used for in-gel kinase assays. (B) Activation of AtMPK3 and AtMPK6 in C01 gll , HopD2, or HopD2°°7°S leaves after inoculation with 1 x 10° cfir/ml hrpA mutant bacteria or a mock inoculum of sterile H2O. AtMPK3 and AtMPK6 were identified by their electrophoretic mobility. In all panels, upper bands represent AtMPK6 and lower bands represent AtMPK3. 161 HopD2 is a non—transmembrane protein tyrosine phosphatase and is present in both the soluble and membrane fractions of Arabidopsis thaliana protein extracts. Protein tyrosine phosphatases can be categorized into distinct subgroups based on several aspects of their sequences, subcellular localizations, and substrate specificity (Tonks, 2006). These proteins are classified as either transmembrane, receptor-like phosphatases, or intracellular phosphatases. HopD2 lacks transmembrane domains, and thus, is expected to be an intracellular phophatase. However, more specific information about the localization of HopD2 in Arabidopsis cells could potentially provide clues to help determine the substrate(s) of HopD2. Many TTSS effectors of Pst DC3000, including AvrPto and Hole , have been discovered to be peripherally membrane associated through modifications such as myrisoylation and other unknown mechanisms (N omura et al., 2006; Shan et al., 2000). The effector AvrPto is associated with membranes and membrane association is required for its ability to elicit the HR in resistant tomato plants carrying the Pto kinase (Shan et al., 2000) and seems to be required for its virulence function in Arabidopsis (Nomura and He, unpublished). To determine if HopD2 is localized to membranes in Arabidopsis cells, protein extracts from DEX-induced HopD2 and HopD2C378 plants were separated into soluble and membrane fractions by ultracentrifugation. Western blots with HopD2 antibodies revealed that HopD2 is detectable primarily in the soluble fraction of Arabidopsis protein extracts (Figure 4-7A). However, some HopD2 was detected in the membrane fraction and migrated at a slightly lower mobility, suggesting a potential post-translational modification. Non-transmembrane protein tyrosine phosphatases are further categorized into two classes based on substrate specificity. Classical protein tyrosine phosphatases 162 (PTPs) possess catalytic activity toward phosphotyrosine residues only, whereas dual specificity phosphatases (DSPs) can act on phosphoserine or phosphothreonine residues in addition to phosphotyrosine. Some DSPs have even been shown to act on non-protein substrates such as phosphorylated lipids (Robinson and Dixon, 2006). The amino acid sequence of HopD2 bears little similarity to known tyrosine phosphatases outside of the conserved catalytic motif. However, the conserved catalytic domain bears greater similarity to the canonical motif of PTPs than to that of DSPs (Figure 4-7B). 163 HopD2 HopD2C37°° Soluble Membrane Soluble Membrane 5.! mu- 4.“ --' ‘45s. - ‘._. mr‘nxr: W} _. 4 . ‘ ' .. . - Z 4‘ :l _ ‘p i I ' .12 ‘ f ' . ‘ ‘ ‘= . . ,‘ ‘ .I ~ .. . . . h 1 ‘ - "' ’-".'.'~ i . ' .31H’l'fl'i-EII 1t", B PTP-- PIVVHCSAGVGRTG DSP-- rleHCqAGISRSa HopD2-- SLVVHCNGGRGRTT Figure 4-7: (A) Western blot analysis of soluble and membrane fractions of protein extracts from HopD2 and HopD2C 7°° leaves. Plants were treated with 30 uM DEX 24 hrs prior to extraction of total proteins. Total proteins were fractionated by ultracentrifugation and fractions were analyzed by western blotting and detection with HopD2-specific polyclonal antibodies. (B) Amino acid sequence alignment of the consensus catalytic motifs of PTPs and DSPs (F auman and Saper, 1996) with the catalytic region of HopD2. Capital letters in the FTP and DSP motifs represent highly conserved residues present in 90% or more of family members. 164 Microarray analysis of HopD2 function using transgenic plants and bacterial mutants. As a protein tyrosine phosphatase, HopD2 has the potential to alter or disrupt signaling pathways in the host. Given that transgenic expression of HopD2 in Arabidopsis suppresses numerous defense readouts associated with PAMP-induced innate immunity, we hypothesized that HopD2 may alter host signaling pathways and changes in gene expression leading to the activation of basal defenses. To investigate the effect of HopD2 on gene expression changes in Arabidopsis, we conducted microarray analyses to profile gene expression changes associated with transgenic expression of HopD2 in Arabidopsis, and with the presence or absence of HopD2 in bacteria using the AhopDZ bacterial deletion mutant. To determine if transgenic expression of HopD2 altered gene expression changes associated with perception of the non-pathogenic hrpA mutant (PAMP perception), Col gll, HopD2, and HopD2C°7°° plants were inoculated with the hrpA mutant (1 x 10° cfu/ml) or a mock inoculum of sterile H20 24 hrs after spraying with 30 11M DEX. Leaf tissue was collected 7 hrs after inoculation for RNA isolation, labeling, and analysis of transcripts using Affymetrix ATHl genechips. Comparison of mock-inoculated samples also allowed the identification of HopD2-dependent changes in gene expression in the absence of bacterial challenge. A summary of the experiments and comparisons performed using HopD2 transgenic plants is presented in Table 4-1. To determine if transgenic expression of HopD2 is associated with gene expression changes in unchallenged plants, expression profiles of mock-inoculated HopD2 plants were compared to those of mock-inoculated Col gll plants. Remarkably, only 11 genes were reproducibly differentially regulated in plants expressing HopD2 compared to wild-type 165 plants (Table 4-1, row 1). Furthermore, these changes in gene expression were not dependent on phosphatase activity, as these genes were regulated similarly in HopD2°°7°S plants (Table 4-1, row 2). These results suggest that HopD2 phosphatase activity does not effect gene expression in unchallenged plants. To determine if HopD2 expression suppresses or alters gene expression changes induced by PAMP perception, the expression profiles of HopD2 plants inoculated with hrpA were compared to those of hrpA-inoculated Col gl 1. This comparison revealed only 2 genes that were reproducibly differentially regulated (Table 4-1, row 3). Similar to the previous comparison, the alteration of expression observed for these genes was not dependent on HopD2 phosphatase activity (Table 4-1, row 4). Genes differentially regulated in HopD2 and HopD2C3788 plants compared to C01 gll (Table 4-1, rows 1 and 3) are displayed in Table 4-2. To verify the quality and reproducibility of our dataset, the genes reproducibly regulated in C01 gll by inoculation of the hrpA mutant were compared to the independently identified set of genes regulated by the same treatment described in Chapter 2. These datasets were found to be highly similar. We also compared the expression profiles of Col gll plants inoculated with either Pst DC3000, the AhopDZ mutant, or a mock inoculum of H20. Similar to the results with our transgenic plants, we found no HopD2-specific gene expression changes in comparisons of plants inoculated with the AhopDZ bacterial mutant to those inoculated with wild-type Pst DC3000. Taken together, these results suggest that the virulence function of HopD2 in Arabidopsis is not associated with alteration of gene expression or blocking early signaling events mediated by PAMP perception and that HopD2 likely blocks basal defenses at a relatively late stage, downstream of gene expression changes. 166 Repressed Induced Comaflson Treatment Genesb Genesb Total Expression Pattern Identified HopD2 vs. Col gl1 mock, 7 hrs. 3 8 11 HopDZ regulation HopDZ vs. l-lopDan78$ mock, 7 his. 0 0 O Phophatase—dependent regulation HopDZ vs. Col 9" 1x10° cfu/ml hrpA, 7 hrs. 2 0 2 HopDZ alteration of PAMP regulation HopD2 vs. HopDZ03788 1x108 cfu/ml hrpA, 7 hrs. 0 0 0 Phosphatase-dependent regulation Table 4-1: Microarray analysis of HopD2 transgenic plants. Summary of treatments and comparisons used to identify gene expression changes resulting from transgenic expression of HopD2 and changes dependent on the phosphatase activity of HopD2. ° Criteria for gene selection is described in the “Experimental Procedures” section of the text. 167 Average Fold ("notifies HopDZ HopDz Array vs. vs. Element AGI Description Col 9" Col {I1 266385_at_ At2914810 ~PR-1_ pathogenesis-related protein] 3.9 4.8 265943_at At2919570 .AtCDA1 cytidine deaminase 1 f 2.1 2.1 284648_at .At_1909080 putativelunfiljnalpingding prgteigflh _ fl H.,“... 3.5 3.3 264153_at A11965390 Putative TIR class disease resistance protein -4.9 -2.9 254975_at ‘At4g10500 oxidoreductase, ZOG-Fe(ll) oxygenase family protein; 4.3 2.5 254042_at At4925810 AtXTR6 xylogluean endotransglycosylase-related protein 5.7 13.0 251625_at At3957260 PR-2 pathogenesis-related protein 2, beta 1,3-glucanase 4.0 3.6 249032_at 7At5944910 putative disease resistance protein 7 -2.3 -2.3 248921_at ,At5945950 ,GDSL-motigf lipase/hydrolasg-likeprotein -2.3 -5.2 247925_at At5957560 AtTCH4 xylogluean endotransglycosylase 3.3 5.0 246302_at Atgq51860 AtCAX3 vacuolar CaZ+lH+-exchanging protein 4.8 4.4 B Average Fold Chalclge’s HopDZ HopDZ Array vs. vs. Element AGI Description Col 9" Col git 262399_at At1g49500 "expressed protein -3.4 -2.2 260101_at M3260 ‘trypsin and protease inhibitor family protein -4.9 -2.1 Table 4-2: Genes differentially regulated by transgenic expression of HopD2 and HopD2C37°°. (A) Differentially regulated genes in HopD2 vs. Col gll and HopD2C°7°° vs. Col gll comparisons after mock inoculation with sterile H2O (no bacterial challenge). (B) Differentially regulated genes in HopD2 vs. Col gll and Ho D2C°7°° vs. Col gll comparisons after inoculation with the hrpA mutant (1 x 10 cfu/ml). 168 Lipid profiling analysis of HopD2 transgenic plants. Recent studies of protein tyrosine phosphatases have revealed that some DSPs, particularly those of the myotubularin family in mammalian cells, can act on phosphorylated lipids such as phosphoinositides (Robinson and Dixon, 2006). Phosphatidylinositol lipids can be reversibly phosphorylated and exist in a number of distinctly phosphorylated forms where the phosphotidylinositol headgroup can be reversibly singly, doubly, and triply phosphorylated. Phosphoinositides (PIs) have been shown to play an important role in regulating vesicle trafficking through interactions with protein components of the membrane trafficking machinery (Di Paolo et al., 2004; Wenk and Camilli, 2004). To determine if HopD2 dephosphorylates a PI lipid substrate in Arabidopsis, the content of phosphotidylinositol species in lipid extracts from Col gll , HopD2, and HopD2C37°° leaves was analyzed by mass spectrometry. Previous studies have analyzed phosphotidylinositol species in plants that fragment to give a parent ion of mass-to-charge ratio (m/z) 241 using tandem mass spectrometry (Welti and Wang, 2004). This parent ion represents the phosphotidylinositol head group with no additional phosphate groups (ie. not reversibly phosphorylated). Using parent ion mode ESI- MS/MS, we analyzed lipid extracts from Col gll, HopD2, and HopD2C°7°S for lipids that fragmented to yield a parent ion of 241 m/z. In our lipid samples, the primary phosphotidylinositol species detected were PI 34:3 (831 m/z) and PI 34:2 (833 m/z; Figure 4-8). We did not observe any significant differences in the abundance of PI species between HopD2 transgenic plants and Col gll (Figure 4-8). This result suggests that HopD2 does not dephosphorylate a mono-phosphorylated PI lipid species. We also examined the spectrum of lipids that fragmented to yield a parent ion of 321 m/z to 169 determine if HopD2 acts on a bi-phosphorylated PI lipid. Again, we found no significant differences in the samples derived from HopD2 plants compared to C01 gll plants (data not shown), suggesting that HopD2 does not dephosphorylate a bi-phosphorylated PI lipid species. Further analysis of our lipid extracts to detect PI species that fragment to yield a parent ion of 401 m/z will be required to determine if HopD2 acts on a tri- phosphorylated PI lipid. However, these initial results of analyzing the PI lipid content of HopD2 transgenic plants combined with the observation that the conserved active motif of HOpD2 is more similar to that of PTPs rather than DSPs suggests that HopD2 is more likely to act on a protein substrate, and, more specifically, a tyrosine-phosphorylated protein substrate, than a phosphorylated lipid species. 170 A 1 ' 83‘18 3338 827.2 1 .33 3097 313381575175 823.4 . 335° 3418 844.9 848.8 553.9 8557.9 8°?28627 ~. and ‘Vl‘m- ‘ 1. at“ ' ““"c-"e,1\,\,"‘\,.\J\ _J ‘o’jwt ,2 W V ,‘ --~.\- 011‘:U'—1""1'* 1 s. ' 1*1 1 '*’-**°'1*"ru I B 100 835.6 f333,3 . 815.0 5197 3232 i .6 855.5 °\ ”938100 j°1°f° 828.3 1'21“ 537.5 8434 M8475 848785578552 85397 351° ,1 a"! l V J vW‘ “'I'\,~ ‘*.-'°.‘»"' '4 'V‘r —- N“.'."‘»j ' ‘ w ‘m' 1..M “A ‘2 “MN ”\1‘ -"x "' o .,...Tf.v ....--, c. as - .. -..,... , ..,.-c., - a, 100 8°16 (3335 815.0 519.7 3232 1'. 5346 855.5 5 8095110081“ , . 828.3 ‘ 5375 8434 847.5 8487 855711562 8597 °°1‘° 1‘1. .57,"1'E~J\Vx\~\/“~I~»J°'~°l/ t' \“Il‘m 4v...“ ”ax—f -.__,' WN\AL ”‘55”wa .~‘.*..- ' 0...--.-,. ...- -. - .s-.,..a.. Figure 4-8: Phos sghoinositide (PI) lipid content in extracts of (A) Col gll, (B) HopD2, and (C) HopD2C3 °° leaves 24 hrs after treatment with 30 11M DEX. Lipid extracts were analyzed by tandem mass spectrometry using a parent ion method to detect lipids fragmenting to yield a parent ion of 241 m/z. Peaks at 831 m/z represent PI 34:3 and peaks at 833 m/z represent PI 34:2. These data are from single MS/MS analyses of single lipid extracts for each sample. 171 Discussion Effector proteins delivered into host cells by plant pathogenic bacteria through TTSSs are often essential virulence determinants. Emerging evidence suggests that a major role of TTSS effectors in promoting bacterial virulence is the suppression of host defense responses. However, the specific molecular mechanisms through which individual TTSS effectors function to suppress host defenses are poorly understood. The Pst DC3000 effector HopD2 is required for full virulence on the host plant Arabidopsis and is a protein tyrosine phosphatase (Bretz et al., 2003; Espinosa et al., 2003). Previous studies of HopD2 function carried out in non-host Nicotiana species suggested that HopD2 may promote bacterial virulence through suppression of the HR (Bretz et al., 2003; Espinosa et al., 2003). In this chapter I report that HopD2, when expressed transgenically in the host plant Arabidopsis, does not suppress the HR elicited by Pst DC3000 carrying the avirulence genes aerptZ, avrB, or averhB, suggesting that the virulence function in the host plant is not related to suppression of the HR. Instead, transgenic expression of HopD2 suppressed several defense responses associated with PAMP-triggered innate immunity. Plants expressing HopD2 exhibited a reduced ability to deposit callose associated with papillae in response to inoculation with the non- pathogenic hrpA mutant and were completely impaired in resistance to Pst DC3000 induced by treatment with the flg22 peptide. Additionally, transgenic expression of HopD2 was sufficient to promote significant multiplication of hrpA mutant bacteria. Importantly, these effects of HopD2 expression were entirely dependent on an intact phosphatase catalytic site as transgenic plants expressing HopD2‘33788 exhibited responses similar to the parental Col gII plants. 172 HopD2, as a protein tyrosine phosphatase, may disrupt host signaling pathways leading to the activation of basal defenses. MAPKs were obvious candidates for virulence targets of HopD2 given their requirement for tyrosine phosphorylation for activation and their roles in signal transduction. Specifically, the MAPKs AtMPK3 and AtMPK6 have been shown to participate in signaling initiated by perception of bacterial flagellin by the FLS2 receptor kinase (Asai et al., 2002). However, contrary to this expectation, we found that transgenic expression of HopD2 caused enhanced activation of AtMPK3 and AtMPK6. Even more strikingly, expression of HopD2 had no effect on gene expression levels in Arabidopsis. No phosphatase-dependent alterations of gene expression were observed in transgenic Arabidopsis plants in microarray experiments, nor did HopD2 effect changes in gene expression in response to perception of hrpA bacteria. These results are in strong contrast with previous studies of transgenic plants expressing the Pst DC3000 effector AvrPto. Transgenic expression of AvrPto was found to have significant effects on gene expression and to cause changes in expression that were strikingly similar to those induced by inoculation with Pst DC3000 (Hauck et al., 2003). The results of our microarray experiments with HopD2 transgenic plants suggest that alteration of gene expression is not a universal characteristic of transgenically expressed TTSS effector proteins. Additionally, these results suggest that HopD2 acts at a significantly later stage to suppress basal defenses, downstream of initial signaling events and gene expression changes. Protein tyrosine phosphatases are a broad group of enzymes including transmembrane, receptor-like phosphatases and intracellular phosphatases that act on diverse substrates including tyrosine-phosphorylated as well as serine- and threonine- 173 phosphorylated proteins and even phosphorylated lipids. HopD2 does not contain any predicted transmembrane domains and was localized primarily to the soluble fraction of protein extracts from HopD2 transgenic plants. However, some HopD2 protein was also detected in the membrane fraction, suggesting that HopD2 may be peripherally membrane associated. Several Pst DC3000 effectors, including Hole , AvrPto, and Aerme , have been shown to localize to host membranes (N imchuk et al., 2000; Nomura et al., 2006; Shan et al., 2000). Both AvrPto and Aerme carry potential myristoylation signals that promote their membrane localization. HopD2 does not have any predicted myristoylation signals. It will be interesting to determine how HopD2 is targeted to host membranes and whether the membrane-localized pool of HopD2 is, in fact, the active form. Alternatively, the presence of HopD2 in membranes may be an artifact of overexpression in the plant cells. Because our results suggest that HopD2 acts at a late stage to block PAMP- triggered innate immunity, we suspect that HopD2 may potentially interfere with host vesicle trafficking pathways leading to the deposition of papillae or secretion of antimicrobial compounds. Reversibly phosphorylated PI lipids are hypothesized to play a role in regulating vesicle trafficking events. We therefore examined the PI lipid content of our HopD2 transgenic plants using mass spectrometry. Our results suggest that HopD2 does not act on mono- or bi-phosphorylated PI lipids. However, we have not investigated the PI content of wild-type and HopD2 transgenic plants after induction of basal defenses by the hrpA mutant or flg22. It is possible that the substrate of HopD2 is only phosphorylated in response to detection of bacterial elicitors; therefore, further analysis of the PI lipids will be required to rule out these molecules as targets of HopD2 174 phosphatase activity. However, it is noteworthy that the conserved PTP catalytic domain of HopD2 bears greater similarity to that of tyrosine-specific phosphatases than to DSPs. Taken as a whole, these results point toward a tyrosine-phosphorylated protein(s) as the likely virulence target of HopD2. Clearly, identifying the host target(s) of HopD2 phosphatase activity will be the next step toward a mechanistic understanding of its virulence function. Additionally, identifying the virulence target of HopD2 may provide novel information about the mechanisms of PAMP-triggered innate immunity. Currently, relatively little is known about the cell biological events leading to resistance to bacteria or about the compounds that ultimately antagonize bacterial multiplication. During the past 5 years, significant progress has been made toward understanding the early events of PAMP recognition including the identification of the flagellin receptor, signal transduction components such as MAPKs and transcription factors, and gene expression changes associated with innate immunity (Asai et al., 2002; Gomez-Gomez et al., 2000; Thilmony et al., 2006; Zipfel et al., 2004). However, the later events leading to enhanced resistance remain enigmatic. Deposition of callose-rich papillae has long been correlated with restriction of bacterial multiplication (Brown et al., 1995). However, the mechanisms by which plant cells target defense-related compounds to specific locations at the site of interaction with invading bacteria are unknown as are the specific compounds that act to restrict bacterial proliferation. Because HopD2 acts at a late stage to block PAMP-induced defenses, identifying the target(s) of HopD2 may also improve our understanding of these enigmatic components of innate immunity in Arabidopsis. Interestingly, the human pathogen Yersinia pseudotuberculosis also secretes a TTSS effector, YopH, that is a PTP. 175 Investigators were able to identify protein substrates of YopH by co-immunoprecipitation using a catalytically inactive derivative that was found to trap substrates in unproductive complexes in the absence of phosphatase catalytic activity (Black and Bliska, 1997). Our observation that transgenic expression of the catalytically inactive HopD2C3788 has a dominant-negative effect on the function of wild-type HopD2 delivered by Pst DC3000 suggests that HopD2C3788 may also form unproductive complexes with target proteins. Therefore, it may be possible to co-immunoprecipitate substrates of HopD2 using the catalytically inactive form as a substrate trap. Identifying the host target(s) of HopD2 should provide novel mechanistic insights into both bacterial virulence strategies and host innate immune responses. Acknowledgements I would like to acknowledge Annette Thelen and the Research Technology Support Facility at Michigan State University for valuable assistance with the Affymetrix microarray analysis, Dr. Daniel Jones of the MSU mass spectrometry facility for training and assistance with mass spectrometry and lipid profiling analyses, Dr. Shuqun Zhang (Department of Biochemistry, University of Missouri) for performing in-gel kinase assays reported in figure 4-6, and Dr. Steven Hutcheson (Department of Cell Biology and Molecular Genetics, University of Maryland) for the kind gift of the Pst DC3000 AhopDZ mutant. 176 References Abrahams, G.L., and Hansel, M. (2006) Manipulating cellular transport and immune responses: dynamic interactions between intracellular Salmonella enterica and its host cells. Cell. Microbiol. 8, 728-737. Abramovitch, R.B., Janjusevik, R., Stebbins, CB, and Martin, GB. (2006) Type III effector AvrPtoB requires intrinsic E3 ubiquitin ligase activity to suppress plant cell death and immunity. Proc. Natl. Acad. Sci. USA, 103, 2851-2856. Abramovitch, R.B., Kim, Y.J., Chen, S., Dickman, M.B., and Martin, GB. (2003) Pseudomonas type III effector AvrPtoB induces plant disease susceptibility by inhibition of host programmed cell death. EMBO J. 22, 60-69. Asai, T., Tena, G., Plotnikova, J ., Willmann, M.R., Chiu, W.L., Gomez-Gomez, L., Boller, T., Ausubel, F .M., and Sheen, J. (2002) MAP kinase signaling cascade in Arabidopsis innate immunity. Nature, 415, 977-983. Black, D.S., and Bliska, J.B. (1997) Identification of p130“IS as a substrate of Yersinia YopH (Yop51), a bacterial protein tyrosine phosphatase that translocates into mammalian cells and targets focal adhesions. EMBO J. 16, 2730-2744. Bretz, J.R., Mock, N.M., Charity, J.C., Zeyad, S., Baker, C.J., and Hutcheson, SW. (2003) A translocated protein tyrosine phosphatase of Pseudomonas syringae pv. tomato DC3000 modulates plant defence response to infection. Mol. Microbiol. 49, 389-400. Brown, 1., Mansfield, J., and Bonas, U. (1995) hrp genes in Xanthomonas campestris pv. vesicatoria determine ability to suppress papilla deposition in pepper mesophyll cells. Mol. Plant Microbe. Interact. 8, 825-836. Buell, C.R., Joardar, V., Lindeberg, M., Selengut, J ., Paulsen, I.T., vainn, M.L., Dodson, R.J., Deboy, R.T., Durkin, A.S., Kolonay, J .F ., Madupu, R., Daugherty, S., Brinkac, L., Beanan, M.J., Hafi, D.H., Nelson, W.C., Davidsen, T., Zafar, N., Zhou, L., Liu, J ., Yuan, Q., Khouri, H., Fedorova, N., Tran, B., Russell, D., Berry, K., Utterback, T., Van Aken, S.E., Feldblyum, T.V., D'Ascenzo, M., Deng, W.L., Ramos, A.R., Alfano, J .R., Cartinhour, S., Chatterjee, A.K., Delaney, T.P., Lazarowitz, S.G., Martin, G.B., Schneider, D.J., Tang, X., Bender, C.L., White, 0., Fraser, C.M., and Collmer, A. (2003). The complete genome sequence of the Arabidopsis and tomato pathogen Pseudomonas syringae pv. tomato DC3000. Proc. Natl. Acad. Sci. USA, 18, 10181-10186. Chang, J .H., Urbach, J .M., Law, T.F., Arnold, L.W., Hu, A., Gombar, S., Grant, S.R., Ausubel, F.M., and Dangl, J.L. (2005) A high-throughput, near-saturating screen for type III effector genes from Pseudomonas syringae. Proc. Natl. Acad. Sci. USA, 102, 2549- 2554. 177 Clough, 8.1., and Bent, A.F. ( 1998) Flora] dip: a simplified method for Agrobacterium- mediated transformation of Arabidopsis thaliana. Plant J. 16, 73 5-743. Comelis, G. R., and Van Gijsegem, F. (2000) Assembly and function of type III secretory systems. Annu. Rev. Microbiol. 54, 735-774. da Silva, A.C., Ferro, J .A., Reinach, F.C., Farah, C.S., Furlan, L.R., Quaggio, R.B., Monteiro-Vitorello, C.B., Van Sluys, M.A., Alrneida, N.F., Alves, L.M., do Amara], A.M., Bertolini, M.C., Camargo, L.E., Camarotte, G., Cannavan, F., Cardozo, J ., Chambergo, F., Ciapina, L.P., Cicarelli, R.M., Coutinho, L.L., Cursino-Santos, J .R., El- Dorry, H., Faria, J .B., Ferreira, A.J., Ferreira, R.C., Ferro, M.I., Formighieri, E.F., Franco, M.C., Greggio, C.C., Gruber, A., Katsuyama, A.M., Kishi, L.T., Leite, R.P., Lemos, E.G., Lemos, M.V., Locali, E.C., Machado, M.A., Madeira, A.M., Martinez- Rossi, N.M., Martins, E.C., Meidanis, J ., Menck, C.F., Miyaki, C.Y., Moon, D.H., Moreira, L.M., Novo, M.T., Okura, V.K., Oliveira, M.C., Oliveira, V.R., Pereira, H.A., Rossi, A., Sena, J .A., Silva, C., de Souza, R.F., Spinola, L.A., Takita, M.A., Tamura, R.E., Teixeira, E.G., Tezza, R.I., Trindade dos Santos, M., Trufii, D., Tsai, S.M., White, F.F., Setubal, J .C., and Kitajima, J.P. (2002) Comparison of the genomes of two Xanthomonas pathogens with differing host specificities. Nature, 417, 459-463. DebRoy, S., Thilmony, R., Kwak, Y.B., Nomura, K., and He, S.Y. (2004) A family of conserved bacterial effectors inhibits salicylic acid-mediated basal immunity and promotes disease necrosis in plants. Proc. Natl. Acad. Sci. USA, 101, 9927-9932. Di Paolo, G., Moskowitz, H.S., Gipson, K., Wenk, M.R., Voronov, S., Obayashi, M., F lavell, R., Fitzsimonds, R.M., Ryan, T.A., and De Camilli, P. (2004) Impaired PtdIns(4,5)P2 synthesis in nerve terminals produces defects in synaptic vesicle trafficking. Nature, 431, 415-422. Espinosa, A., Guo, M., Tarn, V.C., Fu, Z.Q., and Alfano, J .R. (2003) The Pseudomonas syringae type III-secreted protein HothoD2 possesses protein tyrosine phosphatase activity and suppresses programmed cell death in plants. Mol. Microbiol. 49, 377-3 87. Fauman, BB, and Saper, MA. (1996) Structure and function of the protein tyrosine phosphatases. Trends Biochem. Sci. 21, 413-417. Fouts, D.E., Abramovitch, R.B., Alfano, J .R., Baldo, A.M., Buell, C.R., Cartinhour, S., Chatterjee, A.K., D'Ascenzo, M., Gwinn, M.L., Lazarowitz, S.G., Lin, N.C., Martin, G.B., Rehm, A.H., Schneider, D.J., van Dijk, K., Tang, X., and Collmer, A. (2002) Genomewide identification of Pseudomonas syringae pv. tomato DC3000 promoters controlled by the HrpL alternative sigma factor. Proc. Natl. Acad. Sci. USA, 99, 2275- 2280. Gomez-Gomez, L., and Boller, T. (2000) FLSZ: an LRR receptor-like kinase involved in the perception of the bacterial elicitor flagellin in Arabidopsis. Mol. Cell, 5, 1003-1011. 178 Gomez-Gomez, L., Felix, G., and Boller, T., (1999) A single locus determines sensitivity to bacterial flagellin in Arabidopsis thaliana. Plant J. 18, 277—284. Guttman, D.S., Vinatzer, B.A., Sarkar, S.F., Ranall, M.V., Kettler, G., and Greenberg, J .T. (2002) A functional screen for the type III (Hrp) secretome of the plant pathogen Pseudomonas syringae. Science, 295, 1722-1726. Hauck, P., Thilmony R., and He S.Y. (2003) A Pseudomonas syringae type III effector suppresses cell wall-based extracellular defense in susceptible Arabidopsis plants. Proc. Natl. Acad. Sci. USA, 100, 8577-8582. Jamir, Y., Guo, M., Oh, H.S., Petnicki-chieja, T., Chen, S., Tang, X., Dickman, M.B., Collmer, A., and Alfano, J .R. (2004) Identification of Pseudomonas syringae type III effectors that can suppress programmed cell death in plants and yeast. Plant J. 37 , 554- 565. Janjusevic, R., Abramovitch, R.B., Martin, G.B., and Stebbins, CE. (2006) A bacterial inhibitor of host programmed cell death defenses is an E3 ubiquitin ligase. Science, 311, 222-226. Katagiri, F., Thilmony, R. and He, S.Y. (2002) The Arabidopsis thaliana-Pseudomonas syringae interaction. In The Arabidopsis Book (Meyerowitz, E. M. and Somerville, C. R., eds). American Society of Plant Biologists, Rockville, MD, USA, DOI 10.1 l99-tab.OO39. Kim, H.S., Desveaux, D., Singer, A.U., Patel, P., Sondek, J., and Dangl, J .L. (2005) The Pseudomonas syringae effector AerptZ cleaves its C-terminally acylated target, RIN4, from Arabidopsis membranes to block RPM] activation. Proc. Natl. Acad. Sci. USA, 102, 6496-6501. Kim, M.G., da Cunha, L., McFall, A.J., Belkhadir, Y., DebRoy, S., Dangl, J .L., and Mackey, D. (2005) Two Pseudomonas syringae type III effectors inhibit RIN4-regulated basal defense in Arabidopsis. Cell, 121, 749-759. Kunkel, B.N., Bent, A.F., Dahlbeck, D., Innes, R.W., and Staskawicz, B.J. (1993) RPSZ, an Arabidopsis disease resistance locus specifying recognition of Pseudomonas syringae strains expressing the avirulence gene aerptZ. Plant Cell, 5, 865-875. Lindgren, P.B., Peet, R.C., and Panopoulos, NJ. (1986) Gene cluster of Pseudomonas syringae pv. phaseolicola controls pathogenicity of bean plants and hypersensitivity of nonhost plants. J. Bacteriol. 168, 512-522. Navarro, L., Alto, M.N., and Dixon, J.E. (2005) Functions of the Yersinia effector proteins in inhibiting host immune responses. Curr. Opin. Microbiol. 8, 21-27. 179 Nimchuk, Z., Marois, E., Kjemtrup, S., Leister, R.T., Katagiri, F ., and Dangl, J .L. (2000) Eukaryotic fatty acylation drives plasma membrane targeting and enhances function of several type III effector proteins from Pseudomonas syringae. Cell, 101, 353-363. Noel, L., Thieme, F., Nennstiel, D., and Bonas, U. (2001) cDNA-AFLP analysis unravels a genome-wide hrpG-regulon in the plant pathogen Xanthomonas campestris pv. vesicatoria. Mol. Microbiol. 41, 1271-1281. Nomura, K., DebRoy, S., Lee, Y.H., Pumplin, N., Jones, J ., and He, S.Y. (2006) A bacterial virulence protein suppresses host innate immunity to cause plant disease. Science, 313, 220-223. Petnicki-chieja, T., Schneider, D.J., Tam, V.C., Chancey, S.T., Shan, L., Jarnir, Y., Schechter, L.M., Janes, M.D., Buell, C.R., Tang, X., Collmer, A., and Alfano, J .R. (2002) Genomewide identification of proteins secreted by the Hrp type III protein secretion system of Pseudomonas syringae pv. tomato DC3000. Proc. Natl. Acad. Sci. USA, 99, 7652-7657. Robinson, F .L., and Dixon, J.E. (2006) Myotubularin phosphatases: policing 3- phosphoinositides. Trends Cell Biol. 16, 403-412. Roden, J .A., Belt, B., Ross, J .B., Tachibana, T., Vargas, J ., and Mudgett, MB. (2004) A genetic screen to isolate type III effectors translocated into pepper cells during Xanthomonas infection. Proc. Natl. Acad. Sci. USA, 101, 16624-16629. Salanoubat, M., Genin, S., Artiguenave, F., Gouzy, J ., Mangenot, S., Arlat, M., Billault, A., Brottier, P., Camus, J .C., Cattolico, L., Chandler, M., Choisne, N., Claudel-Renard, C., Cunnac, S., Demange, N., Gaspin, C., Lavie, M., Moisan, A., Robert, C., Saurin, W., Schiex, T., Siguier, P., Thebault, P., Whalen, M., Wincker, P., Levy, M., Weissenbach, J ., and Boucher, CA (2002) Genome sequence of the plant pathogen Ralstonia solanacearum. Nature, 415, 497-502. Sambrook, J ., Fritsch, E., and Maniatis, T. (1989). Molecular Cloning: a Laboratory Manual. Cold Spring Harbor Laboratory, Cold Spring Harbor, NY, USA. Schechter, L.M., Roberts, K.A., Jamir, Y., Alfano, J.R., and Collmer, A. (2004) Pseudomonas syringae type III secretion system targeting signals and novel effectors studied with a Cya translocation reporter. J. Bacteriol. 186, 543-545. Shan, L., Thara, V.K., Martin, G.B., Zhou, J .M., and Tang, X. (2000) The Pseudomonas AvrPto protein is differentially recognized by tomato and tobacco and is localized to the plant plasma membrane. Plant Cell, 12, 2323-2338. Shao, F., Golstein, C., Ade, J ., Stoutemyer, M., Dixon, J.E., and Innes, R.W. (2003) Cleavage of Arabidopsis PBSl by a bacterial type III effector. Science, 301, 1230-1233. 180 Soylu, S., Brown, I., and Mansfield, J .W. (2005) Cellular reactions in Arabidopsis following challenge by strains of Pseudomonas syringae: From basal resistance to compatibility. Mol. Plant Pathol. 66, 232-243. Thilmony, R., Underwood, W., and He, S.Y. (2006) Genome-wide transcriptional analysis of the Arabidopsis thaliana interaction with the plant pathogen Pseudomonas syringae pv. tomato DC3000 and the human pathogen Escherichia coli 01 57:H7. Plant J. 46, 34-53. Thieme, F ., Koebnik, R., Bekel, T., Berger, C., Boch, J., Buttner, D., Caldana, C., Gaigalat, L., Goesmann, A., Kay, S., Kirchner, 0., Lanz, C., Linke, B., McHardy, A.C., Meyer, F., Mittenhuber, G., Nies, D.H., Niesbach-Klosgen, U., Patschkowski, T., Ruckert, C., Rupp, 0., Schneiker, S., Schuster, S.C., Vorholter, F.J., Weber, H., Puhler, A., Bonas, U., Bartels, D., and Kaiser, O. (2005) Insights into genome plasticity and pathogenicity of the plant pathogenic bacterium Xanthomonas campestris pv. vesicatoria revealed by the complete genome sequence. J. Bacteriol. 187, 7254-7266. Tonks, N.K. (2006) Protein tyrosine phosphatases: from genes, to function, to disease. Nature Rev. Mol. Cell Biol. 7, 833-846. Wei, W., Plovanich-Jones, A., Deng, W.L., Jin, Q.L., Collmer, A., Huang, HQ, and He, S.Y. (2000) The gene coding for the Hrp pilus structural protein is required for type III secretion of Hrp and Avr proteins in Pseudomonas syringae pv. tomato. Proc. Natl. Acad. Sci. USA, 97, 2247-2252. Welti, R., and Wang, X. (2004) Lipid species profiling: a high-throughput approach to identify lipid compositional changes and determine the function of genes involved in lipid metabolism and signaling. Curr. Opin. Plant Biol. 7, 337-344. Wenk, M.R., and De Camilli, P. (2004) Protein-lipid interactions and phosphoinositide metabolism in membrane traffic: Insights frOm vesicle recycling in nerve terminals. Proc. Natl. Acad. Sci. USA, 101, 8262-8269. Yu, G.L., Katagiri, F ., and Ausubel, F.M. (1993) Arabidopsis mutations at the RPSZ locus result in loss of resistance to Pseudomonas syringae strains expressing the avirulence gene aerptZ. Mol. Plant Microbe. Interact. 6, 434-443. Zhang, S., and Klessig, DP. (1998) The tobacco wounding-activated mitogen-activated protein kinase is encoded by SIPK. Proc. Natl. Acad. Sci. USA, 95, 7225-7230. Zipfel, C., Robatzek, S., Navarro, L., Oakeley, B.J., Jones, J .D.G., Felix, G., and Boller, T. (2004) Bacterial disease resistance in Arabidopsis through flagellin perception. Nature, 428, 764-767. 181 Zwiesler-Vollick, J ., Plovanich-Jones, A.E., Nomura, K., Bandyopadhyay, S., Joardar, V., Kunkel, B.N., and He, S.Y. (2002) Identification of novel hrp-regulated genes through functional genomic analysis of the Pseudomonas syringae pv. tomato DC3 000 genome. Mol. Microbiol. 45, 1207-1218. 182 Chapter 5 Conclusions and Future Perspectives 183 The long-term goals of the study of plant-pathogen interactions are to improve our understanding of plant defenses against pathogens and the mechanisms used by pathogens to overcome these defenses and infect plants and to use this knowledge to improve agricultural production. There is significant potential for improvement in agricultural production through enhanced resistance to pathogens, as crop losses due to disease have been estimated to reduce worldwide agricultural productivity by 12% (Food and Agriculture Organization, 1993). In addition to increasing overall productivity and agricultural output, development of more resistant crop varieties also has the potential to reduce input costs associated with applications of pesticides and foliar sprays to control pathogen spread and proliferation. Reducing the reliance on chemical sprays and pesticides to control pathogens will also lessen the negative environmental impacts of using such chemicals. Increasing agricultural productivity and efficiency is likely to become very important in the near future, as increased demands will be placed on agriculture for the food needs of a steadily increasing human population. In addition to food needs, agriculture may also be relied on to provide other resources such as biofuels. An increase in demand for agricultural output in the form of fuel crops will likely challenge our agricultural systems to become more efficient in order to supply the demands for both food and fuel. To develop and institute novel strategies for controlling crop losses due to pathogens will require a thorough understanding of the molecular nature of plant defenses, the mechanisms by which successful pathogens overcome these defenses, and the effects of varying environmental conditions on the effectiveness of defenses. It has become clear that plants defend themselves from pathogens by employing multiple layers 184 of defense responses. It is likely that durable pathogen resistance, ofien described as non- host resistance, is a combination of several successful layers of defense responses including surface defenses and preformed barriers, apoplastic innate immune responses, the hypersensitive response, and systemic acquired resistance. It will also be important to understand why these defense responses are effective against many potential pathogens, but fail to prevent infection by successful pathogens. The aims of the studies described here were to improve our understanding plant innate immunity triggered by perception of pathogen-associated molecular patterns (PAMPs) and the mechanisms used by a successful bacterial pathogen to overcome these defenses. When these studies were initiated, relatively little was known about PAMP- induced innate immunity in plants. The discovery that Arabidopsis thaliana could perceive bacterial flagellin was a significant breakthrough that sparked an increased interest in innate immunity in plants (Felix et al., 1999). This discovery led to the identification of the flagellin receptor and several downstream signaling components and to the identification of several other PAMPs perceived by plants (Asai et al., 2002; Gomez-Gomez et al., 2000; Kunze et al., 2004; Zeidler et al., 2004). An understanding of PAMP perception and subsequent signaling events was beginning to take shape, however, downstream events such as gene expression changes and activation of defense responses were still poorly understood. We sought to investigate gene expression changes in Arabidopsis thaliana associated with perception of PAMPs presented on live bacteria by whole-genome microarray analysis. Additionally, we sought to analyze the specific effects of the two well-characterized Pseudomonas syringae pv. tomato DC3000 (Pst DC3000) virulence factors coronatine (COR) and the type III secretion system (TTSS) on 185 gene expression changes, including the effect of these virulence factors on expression changes associated with PAMP-induced innate immunity. Our expression profiling analyses described in chapter 2 identified 736 PAMP-regulated genes including 347 genes induced by PAMP perception. These expression profiles provide a molecular signature of PAMP-induced innate immunity in Arabidopsis. Additionally, our analyses identified candidate genes whose products may be involved in the downstream events leading to defense responses such as deposition of papillae. Two particularly interesting examples are the EXO70H1 and EXO70H2 genes that were strongly induced by PAMP perception (5-15 fold). EXO70 proteins are components of a complex known as the exocyst, which is composed of 8 subunits (Wang and Hsu, 2006). In yeast and mammalian cells, the exocyst complex has been shown to be involved in targeted exocytosis (Wang and Hsu, 2006). Although plants encode all 8 subunits of the exocyst complex (Elias et al., 2003), the role of this protein complex in plants has not been extensively characterized. Interestingly, a significant expansion of the EXO70 family of proteins has occurred in plants. The genome of the yeast Saccharomyces cerevisiae contains only a single EX 070 gene, while the Arabidopsis genome contains 22 EXO70 isoforms (Synek et al., 2006). The PAMP-induced EXO70H1 and EXO70H2 genes are closely related and form a distinct clade (Synek etal., 2006). We are currently working to identify T-DNA insertional mutants for each gene and to cross the mutants to determine if these EXO70 proteins have a specific role in responses to pathogens such as facilitating targeted vesicle trafficking leading to deposition of papillae. Our identification of transcriptional changes associated with PAMP-induced innate immunity and with 186 specific effects of COR and the TTSS should facilitate further inquiry into the molecular mechanisms of plant defense and COR and TTSS effector fimction in bacterial virulence. Chapter 3 describes our discovery of bacterium-induced stomatal closure as an integral part of the plant innate immune response. Prior to this study, stomata had been generally regarded as passive ports of entry for bacteria. Although closure of stomata had previously been observed in response to fungal elicitors such as chitosan (Lee et al., 1999) and modulation of guard cell K+ currents was observed during the hypersensitive response (HR) elicited by the Cladosporiumfulvum avr protein Avr9 (Blatt et al., 1999), this study provided the first evidence that stomatal closure is a component of the plant innate immune response and effectively restricts entry of bacteria into leaf tissue. We found that stomata have the ability to perceive PAMPs present on the surface of bacteria such as flagellin and LPS. PAMP-induced closure of stomata was dependent on the presence of abscisic acid (ABA) and downstream components of the ABA signaling pathway. Additionally, we found that the virulent pathogen Pst DC3000 can overcome PAMP-induced stomatal closure. The phytotoxin COR was found to be the virulence factor that allows Pst DC3000 to overcome stomatal closure and enter leaf tissue to initiate infection. These findings open up many new areas for future research. Although ABA and components of the ABA signal transduction pathway are required for bacterium-induced stomatal closure, it is not yet clear if perception of PAMPs by pattern recognition receptors such as F LS2 leads to increased biosynthesis of ABA or relocalization of previously existing ABA pools. ABA and components of the ABA signaling pathway mediate plant responses to numerous environmental stresses. However, the signaling events upstream of ABA synthesis or mobilization are poorly 187 understood. Genetic screens are currently being pursued to identify signaling intermediates linking activation of pattern recognition receptors by PAMPs to the ABA signaling pathway (Zeng and He, unpublished). The mode of action of COR in blocking stomatal closure or promoting opening of closed stomata is not yet clear. COR did not block synthesis of nitric oxide (N 0) induced by PAMP perception, suggesting that COR acts downstream or independent of NO synthesis. However, the host target(s) of COR are yet to be discovered. Additionally, only five P. syringae pathovars are known to produce COR, yet bacterium-induced stomatal closure seems to be conserved among a variety of plant species, suggesting that other pathovars and species of phytopathogenic bacteria have evolved distinct virulence factors to overcome stomatal defense. Phytopathogenic bacteria are known to produce numerous toxins and it will be interesting to determine if any of these toxins are also involved in overcoming stomatal closure. Guard cells are required to respond to numerous environmental stimuli including light, C02 concentration, humidity, and the presence of bacteria. It is likely that, in some cases, the plant may perceive conflicting signals necessitating a prioritization of the stomatal response. How guard cells prioritize input signals is not known, but is of significant relevance to plant-bacteriurn interactions. Severe outbreaks of bacterial disease in plants are often associated with conditions of rain or high humidity. It is possible that these conditions favor opening of stomata and allow more bacteria to enter leaf tissues, promoting infection. A thorough understanding of the response of stomata to numerous environmental inputs may facilitate the development of new strategies to control outbreaks of bacterial disease in crop plants. 188 Chapter 4 summarizes my work to characterize the virulence function of the Pst DC3000 TTSS effector HopD2 in Arabidopsis. HopD2 contains a conserved protein tyrosine phosphatase (PTP) catalytic motif and previous studies have shown that HopD2 possesses tyrosine phosphatase activity in vitro (Bretz et al., 2003; Espinoza et al., 2003). Additionally, the results of these studies suggested that HopD2 may promote bacterial virulence by blocking or delaying activation of the hypersensitive response (HR). However, I found that transgenic expression of HopD2 in Arabidopsis did not affect the timing or severity of the HR elicited by several P. syringae avr proteins. This result suggests that the virulence function of HopD2 in Arabidopsis is not related to suppression of the HR. Instead, I found that transgenic expression of HopD2 resulted in the suppression of defense responses associated with PAMP-induced innate immunity. Transgenic expression of HopD2 suppressed flg22-induced resistance to Pst DC3000, reduced papilla-associated callose deposition in response to inoculation with the Pst DC3000 hrpA mutant, and promoted significant multiplication of the normally non- pathogenic hrpA mutant bacteria. I also created transgenic Arabidopsis lines expressing the catalytically inactive HopD2C3788 and found that the effects of HopD2 were dependent on an intact phosphatase catalytic site. Interestingly, transgenic expression of HopD2C3785 had a dominant-negative effect on wild-type HopD2 delivered from Pst DC3000, suggesting that the catalytically inactive HopD2C3788 may form unproductive complexes with host virulence targets of HopD2. PTPs in yeast and mammalian cells are often found to be involved in regulation of signal transduction pathways through the dephosphorylation of targets such as mitogen-activated protein kinases (MAPKs). We tested the ability of HopD2 to block 189 activation of the Arabidopsis MAPKs AtMPK3 and AtMPK6, which are known to be involved in signaling initiated by the FLS2 flagellin receptor (Asai et al., 2002). Contrary to our expection, activation of AtMPK3 and AtMPK6 was enhanced in HopD2 transgenic plants. Additionally, global gene expression profiling by microarray analyses revealed that HopD2 does not effect gene expression in Arabidopsis. These results suggest that HopD2 does not block signal transduction or gene expression changes leading to the activation of innate immunity, but likely acts downstream of these events at a later stage to block basal defenses. Some PTPs, specifically those of the myotubularin family, have been shown to act on phosphotidylinositol lipid species (Robinson and Dixon, 2006). Phosphoinositides (PIs) are involved in numerous cell signaling and regulatory processes including the regulation of targeted vesicle trafficking events (Di Paolo et al., 2004; Wenk and Camilli, 2004). To determine if HopD2 targets Arabidopsis PIs, I analyzed the PI lipid content of the HopD2 transgenic plants by tandem mass spectrometry. The results of these analyses show that HopD2 does not act on mono- or di-phosphorylated PIs, however, the effect of HopD2 on tri-phosphorylated PIs has not yet been determined. The next step toward understanding the virulence function of HopD2 will be the identification of the Arabidopsis target(s) that are dephosphorylated by this effector. Several approaches have already been pursued to identify physical interactors of HopD2 20785 as bait (Jin and He, including yeast-two-hybrid analyses using HopD2 or HopD unpublished) and co-immunoprecipitation experiments using the polyclonal HopD2 antibody (Underwood and He, unpublished). Thus far, our attempts to identify HopD2- interacting Arabidopsis proteins have not been successful. However, the observation that 190 HopDZC3788 has a dominant-negative effect on HopD2 fimction suggests that this catalytically-inactive protein may form unproductive complexes with host targets. The Yersinia enterocolitica TTSS effector YopH is also a PTP and a catalytically-inactive variant of YopH was found to trap substrates in unproductive complexes, allowing the identification of virulence targets by Co-IP (Black and Bliska, 1997). Further attempts at Co-IP, potentially using affinity-tagged versions of HopD2C3788, may facilitate the identification of the virulence target(s) of HopD2. Additionally, DEX induction of the HopD2 transgene in Arabidopsis seedlings results in lethality. This seedling-lethal phenotype could be used as the basis for a suppressor screen to identify genetic interactors of HopD2. However, this approach assumes that the seedling-lethal phenotype is directly related to the virulence function of HopD2. Our observation that HopD2 does not affect gene expression changes resulting from activation of basal defenses as a result of perception of the Pst DC3000 hrpA mutant suggests that HopD2 acts at a late stage to block PAMP-induced innate immunity. Relatively little is known about the molecular events leading to the activation of basal defense or the specific components of innate immunity that antagonize bacterial proliferation. Because HopD2 is likely to act at a late stage to block PAMP-induced innate immunity, identifying the virulence target(s) of HopD2 and their role in basal defenses may shed light on the cellular components mediating innate immunity in plants. 191 References Asai, T., Tena, G., Plotnikova, J ., Willmann, M.R., Chiu, W.L., Gomez-Gomez, L., Boller, T., Ausubel, F .M., and Sheen, J. (2002) MAP kinase signaling cascade in Arabidopsis innate immunity. Nature, 415, 977-983. Black, D.S., and Bliska, J .B. (1997) Identification of p130cas as a substrate of Yersinia YopH (Yop51), a bacterial protein tyrosine phosphatase that translocates into mammalian cells and targets focal adhesions. EMBOJ. 16, 2730-2744. Blatt, M.R., Grabov, A., Brearley, J., Hammond-Kosack, K., and Jones, J .D. (1999) K+ channels of Cf-9 transgenic tobacco guard cells as targets for Cladasporiumfulvum Avr9 elicitor-dependent signal transduction. Plant J. 19, 453-462. Bretz, J .R., Mock, N.M., Charity, J .C., Zeyad, S., Baker, C.J., and Hutcheson, SW. (2003) A translocated protein tyrosine phosphatase of Pseudomonas syringae pv. tomato DC3000 modulates plant defence response to infection. Mol. Microbiol. 49, 389-400. Di Paolo, G., Moskowitz, H.S., Gipson, K., Wenk, M.R., Voronov, S., Obayashi, M., Flavell, R., F itzsimonds, R.M., Ryan, T.A., and De Camilli, P. (2004) Impaired PtdIns(4,5)P2 synthesis in nerve terminals produces defects in synaptic vesicle trafficking. Nature, 431, 415-422. Elias, M., Drdova, E., Ziak, D., Bavlnka, B., Hala, M., Cvrckova, F., Soukupova, H., and Zarsky, V. (2003) The exocyst complex in plants. Cell Biol. Int. 27, 199-201. Espinosa, A., Guo, M., Tam, V.C., Fu, Z.Q., and Alfano, J .R. (2003) The Pseudomonas syringae type III-secreted protein HothoD2 possesses protein tyrosine phosphatase activity and suppresses programmed cell death in plants. Mal. Microbial. 49, 377-3 87. Felix, G., Duran, J .D., Volko, S., and Boller, T. (1999) Plants have a sensitive perception system for the most conserved domain of bacterial flagellin. Plant J. 18, 265-276. Food and Agriculture Organization (1993) Production year book, FAO, Rome. Gomez-Gomez, L., and Boller, T. (2000) FLSZ: An LRR receptor-like kinase involved in the perception of the bacterial elicitor flagellin in Arabidopsis. Mol. Cell, 5, 1003-1011. Kunze, G., Zipfel, C., Robatzek, S., Niehaus, K., Boller, T., and Felix, G. (2004) The N terminus of bacterial elongation factor Tu elicits innate immunity in Arabidopsis plants. Plant Cell, 16, 3496-3507. Lee, S., Choi, H., Suh, S., Doo, I.S., Oh, K.Y., Choi, E.J., Taylor, A.T.S., Low, PS, and Lee, Y. (1999) Oligogalacturonic acid and chitosan reduce stomatal aperture by inducing 192 the evolution of reactive oxygen species from guard cells of tomato and Commelina cammunis. Plant Physiol. 121, 147-152. Robinson, F.L., and Dixon, J .E. (2006) Myotubularin phosphatases: policing 3- phosphoinositides. Trends Cell Biol. 16, 403-412. Synek, L., Schlager, N., Elias, M., Quentin, M., Hauser, M.T., and Zarsky, V. (2006) AtEXO70A1, a member of a family of putative exocyst subunits specifically expanded in land plants, is important for polar growth and plant development. Plant J. 48, 54-72. Wang, S., and Hsu, SC. (2006) The molecular mechanisms of the mammalian exocyst complex in exocytosis. Biochem. Soc. Trans. 34, 687-690. Wenk, M.R., and De Camilli, P. (2004) Protein-lipid interactions and phosphoinositide metabolism in membrane traffic: Insights from vesicle recycling in nerve terminals. Proc. Natl. Acad. Sci. USA, 101, 8262-8269. Zeidler, D., Zé’thringer, U., Gerber, I., Dubery, I., Hartung, T., Bors, W., Hutzler, P., and Durner, J. (2004) Innate immunity in Arabidopsis thaliana: Lipopolysaccharides activate nitric oxide synthase (NOS) and induce defense genes. Proc. Natl. Acad. Sci. USA, 101, 15811-15816. 193 Appendix A This appendix contains supplementary information for Chapter 2 I would like to acknowledge Dr. Roger Thilmony for contribution of figures A-l, A-4, A- 5, A-6, and A-7. 194 Average Fold Change Amy Element h A COR‘ hrpS E. coll V267165_at At2937710 lectin protein kinase putative ,. p 1.70 2.89 1.70 -2.76 261836 at At116090 AtWAKL7 WAK-Iike kinase 267436_at At2919190 AtFRK1 Flagellin-induced receptor-like kinase 9.40 4.19 5.53 -1.10 266203_at At2902230 putative phloem-specific lectin _ g 3.32 2.89 3.10 -1.10 256181_at At1951820 leucine-rich repeat protein kinase. putative 9.40 3.25 5.53 -1.07 246373 at At1951860 receptor-like urotein kinase. cutative ‘ 4.29 3.48 4.70 -1.05 . . F . ‘ . . . , A . "‘7 9 ~ ‘ . 263232_at At1gO5700 putative light repressible receptor protein kinase 6.35 7.29 7.13 1.32 254271_at At4923150 serine/threonlne kinase - like protein 2.00 3.40 1.91 1.35 255340_ at Al4gO4490 putative receptor-like protein kinase 2.52 2.89 1.95 1.41 247145 at A15965600_ receptor protein kinase- like Lrotein 8.19 4 92 ___2. 6__4__ 1 45 . -—, *z—r - — 24595326 ell-5425.440!- mateiilkmiefamomein an» at. " i- .14:—..ér1xZ§.-: ' 2:41 ‘ w 1.59.: 195 254660 at At4-_81250 receator serine/threonine kinase-like erotein 1 _3. 56 _ 3. 91 _ 2.14 _ 1.48 ’11-"- -- Wmmmmm 1.: 1m MINI—m Iflhml .-. -k - ‘_ . ‘ leucine- rich repeat family protein/protein kinase family ‘ . .256981_at At3g133804 protein 4 7 7 ., _ _ 1.95 2.89 2.19 1.70 29159.12: 461.192.8399- .argtgin uqmmmogin" _. __ __ _ -_ _ 3.82 1.91 3.10 1.70 ‘251063_at At5gO1850 protein kinase. putative 1.45 2.35 -1.35 1.74 .-I.‘ ,7 .c' 1 6. .- . , ,,_., .- ,_.: , , . . _f . .. .__ _ .7 .. . . . W , ' ’ _ - :' , . .- c - - 3'1“‘J.'_‘3‘ '1": .22 .42.: :2 ’2' . : -32" , 4 ' » ’ ' I254410_at At4921410 serine thneonine kinase - like protein; 7 5.28 2.58 4.09 2.46 259640 _at A_t_5go715_0 leucine-rich repeat kinase far_n_ilyprotein_ _ _ _. ,_ 1 2.14 6.65 1.95 2.64 252511123. A13946280 protein kinase-related 8 19 4. 92 6.06 3.17 ' ‘1'. J T ' Us“ l-i' " ' .E‘Q‘W ’. mmmmm‘... 2:: 2.... .2... .... .2... 219m . ' 248381_at At5951830 pike-type carbohydrate kinase family protein 2.76 2.70 1.87 6.50 248090_at_ At5955990 putative protein; similar to NP_K1__ related protein kinase 7.82 3.03 5.53 14.59 Table A-1: Known and predicted kinases induced by PAMP perception. Line within table denotes cutoff for PAMP-induced genes that are repressed or not significantly induced (less than 1.5-fold) by Pst DC3000. Genes below the line are PAMP- and Pst DC3000-induced. Grey shading indicates genes previously identified as induced by the bacterial PAMP flagellin. 196 Figure A-l 7.1 7.5 489 40 27 0.43 0.93 o m _L m In d” a) N A u . a: _ a: _I N _h mt» N a. w a: :- o c 3 '2 At4927410 R026 ANAC072 Transcription FactOI At3945140 Lipoxygenase 2 A11 962660 Beta-Fructofuranosidase 3 M391 6390 jacalin-like lectin At1943160 . RAP2.6 transcription factor ANACO19 transcription factor At2917500 Auxin efflux carrier At1g1QS70 CLH1/CORI1 Chlorophyllase A1391 6370 Proline Rich Protein if“ At2919860 ,- Hexokinese2 At2941940 ZFP8 Zinc Finger Protein At5961 590 ERF Transcription Factor At1933240 _' GTL2 Trihelix Transcription Factor A139 1 3920 eif4A constitutive 197 4.5 4.1 39 0.14 0.24 0.19 2.2 25 39 0.25 0.36 Pst hrpS Pst DC3000 Pst COR 0 12 24 36 12 24 12 24 36 hoursPl At5926340 M881 Hexose Transporter ' A149023BO (321 w ' . At4902940 v." Oxidoreductase . 1‘ At2924850 Tyrosine Aminotransferase At5924780 vsp1 At1g72610 Gennin Like Protein 1 At2938540 ‘ Lipid Transfer Protein 1 An 9 1 2090 Extensin Like Protein Lipase/hydrolase At291 0940 Praline-rich protein ..a”h~.‘0 It... . ...'.‘ e I Figure A-l: RNA blot analysis of differentially expressed genes identified by microarray analysis. (A) Samples were harvested from uninoculated, Pst hrpS and Pst DC3000 inoculated tissues 6, 12, 18, 24 and 36 hours after infiltration with 1x106 bacteria/mL. The results for 8 induced genes, 5 repressed genes, and a constitutively expressed control is shown. The average microarray ratio for a similar comparison is shown on the lefi. (B) Samples were harvested from uninoculated, Pst hrpS, Pst DC3000 and Pst DC31 l8 COR" inoculated tissues 12, 24 and 36 hours afler infiltration with 1x106 bacteria/mL. The average microarray ratio for the Pst DC3000 vs Pst COR'hrpS and Pst DC3000 vs Pst COR' comparison is shown on the left. The AGI identifier and gene annotation is listed on the right. 198 I'm” a. . at. armrm ‘ 3 . 13g13920 eff-IA Figure A-2: RNA blot analysis of the PAMP-responsive gene F lageIlin-Induced Receptor Kinase 1 (FRKI). RNA samples were isolated from leaf tissues harvested at 0, 2, 7, 10, and 24 hours after inoculation with the following. A. Mock inoculum; B. Pst hrpA, 108 cfu/ml; C. Pst COR hrpS, 106 cfu/ml; D. Pst COR hrpS, 5x107 cfu/;ml E. Pst DC3000, 106 cfu/ml; F. Pst DC3000, 108 cfu/ml. No 24 hour sample was collected for treatment F as tissue collapse had begun to occur by this time point. Please note that at 2 HPI, FRKI is induced in all treated tissues, presumably due to the vacuum-infiltration procedure. 199 7 I PstDC3000 5 . I PstDCBOOOfIIC 5 I PstDC3000hrpA 4 I PstDC3000 hrpA/flIC 3.. Bacterial number [Iog(CFUIcm’)] 0 2 4 days post inoculation Figure A-3: In planta multiplication of bacterial flagellin mutants. Arabidopsis Col-0 leaves were vacuum infiltrated with 1x106 bacteria/mL suspensions of Pst DC3000 (black), the Pst DC3000fliC mutant (green), the Pst DC3000 hrpA mutant (red), or the Pst DC3000 hrpA fliC (blue) double mutant bacteria. Bacterial numbers per cm2 leaf tissue was determined 0, 2, and 4 days after inoculation. 200 Non-MVA pathway Shikimatie pathway . c '5' u 3 Te Ienoida TocoI erol Fl \ Sim ole henols l Sulfur containin» Figure A-4: COR toxin induces genes involved in secondary metabolism and TTSS effectors repress genes involved in photosynthesis. The average fold change (of the 3 biological replicates) is shown as illustrated in the fold change color bar in each panel. (A) The MapMan ‘Secondary Metabolism’ display created using the 944 COR toxin- regulated genes identified from the Pst DC3000 vs Pst COR' comparison (24 HPI, 1x106 bacteria/mL) is shown. 201 Figure A-5: Pst regulates the gene expression of enzymes in the tryptophan, glucosinolate and related biosynthetic pathways. A. The expression profile of the genes for 7 enzymes and the regulatory transcription factors ATRI and AT R2 following Pst inoculation is shown. B. The expression profiles of 7 other genes which encode enzymes within the pathways displayed in C are shown. Treatments: A- E. coli v mock; B- hrpA v mock; C- DC3000 v mock; D— DC3000 v hrpA; E- corhrpS v mock; F- car v mock; G- cor v corhrpS; H- corhrpS v mock; 1- car v mock; J - cor v corhrpS; K- DC3000 v cor; L- DC3000 v mock; M- DC3000 v hrpA. C. The enzymes encoded by the genes displayed in A and B are placed on their corresponding biosynthetic pathways. See Figure 2-1 legend for a description of the gene expression display. 202 Figure A-5 8E 052.28% 3.935 E: _ma< 3:58 cease 5:839 8233 ow: 05283.3 3:35 BE 28%? 025.5 2292.2 9833-285 E2 83.95% as 025:5 553a? 8255 23mm? 2% 025:5 553.5 25223 $3.33. ugcomgazzmontozamoi 8238528 #4.; catwflax m 3.232 E2 8353. 0359086 «5238 888=_woxo-~ EO< ouomowvsx 8E 2828.5 295559.50 83:» :< 83 8:98.95 43.222 2.5.. 8883. 820882 035.553.392.885 «Eng: 9:832 8E 852525 89.2.6 38%? can 05259.9 89.25 883.2 83 035:5 32:55:. Nmm<:mm< can» .2 fizaaémeeznscofimofi 22:52.5 $2.03. Sea 8:98:82 55 95. 834mg 72%.. oaaim 32225.2 1%.... 3.8839. .252 oasim 03:555. 23. 8.2.8»? 88 ON WEI OFWHIMVNIBIOVBI 2 A M a =an .2: H I O n— m D U m :55? Bum ___—3.89 ac:— < < 203 Figure A-5 (cont’d). chon'smate ASA1 ASB1/ASBZ camilexin anthran'late : | ?: PAD3 PAT PAH T381 CYP7982 : ° PAIZI IGPi indoie-3- TSA . TSBZ CYP7983 MGM NIT . ho honbos p arft‘firanilatey‘. cerolphospha'lfi-‘k"lame--’tryptOphan _'acetalgoxime _p 4 GU30" 0AA) Y CYP79F1 ' . . CYP79F2 3-methylthio CY'383A13-methylthio ADP? . homomethIonIne —"propanaldoxime —I> hydroximate —> —> —> _.p glucosInolates 204 J Figure A-6: Bacterial inoculation alters auxin-related gene expression. The expression profile of 44 auxin-related genes is shown. See Figure 2-1 legend for a description of the gene expression display. Treatments: A- E. coli v mock; B- hrpA v mock; C- DC3000 v mock; D- DC3000 v hrpA; E- corhrpS v mock; F- cor v mock; G- car v corhrpS; H- corhrpS v mock; 1- car v mock; J- cor v corhrpS; K- DC3000 v cor; L- DC3000 v mock; M- DC3000 v hrpA. 205 At4g38840 SAUR-14 small auxin up-regulated RNA At4g38860 SAUR-16 small auxin upregulated RNA At1929460 SAUR-65 small auxin up—regulated RNA At2921200 SAUR-7 smdl auxin up—regulated RNA A11929440 SAUR-63 small auxin upregulated RNA. GH3-lite ABg53250 SAUR-57 small auxin up-regulated RNA At4g12980 AlR12 Anion-Induced in Root cultures 12 At4g34770 SAUR-1 smdl auxin upregulated RNA Angoauo SAUR-27 small auxin up-regulated RNA A11929500 SAUR-66 snail auxin Lip-regulated RNA At1929450 SAUR-64 snnll aIxin up-regulated RNA At192951o SAUR-67 small auxin upregulated RNA A11929430 SAUR-62 small andn upregulated RNA At5918060 SAUR-23 small auxin up-regulated RNA N4936110 SAUR- 9 small auxin up-regulated RNA Atig75580 SAUR- 51 small auxin upregulated RNA A6953590 $AUR—30 small auxin upregulated RNA 206 Figure A-7: Pst alters host cytokinin-related gene expression. Seven type A response regulator genes, one type B response regulator (ARR14) and 1PT3, a cytokinin synthase are repressed following Pst inoculation. Two cytokinin oxidase genes (CKX4 and CKX5) are induced. See Figure 2-1 legend for a description of the gene expression display. Treatments: A- E. coli v mock; B- hrpA v mock; C- DC3000 v mock; D- DC3000 v hrpA; E- corhrpS v mock; F- cor v mock; G- car v corhrpS; H- corhrpS v mock; 1- car v mock; J- car v corhrpS; K- DC3000 v cor; L- DC3000 v mock; M- DC3000 v hrpA. 207 Figure A-7 no.5 Sahara 320 859.2 $E< §§fl< E 93383 ”E. are; E 2.83% E 89% E A M _. ___—Eco ea:— : o mi .559 .2 m D xm O m < :550 .3: 208 MICHIGAN STATE UNIVERSITY LIBRARIES ill/WI!ill/llllill/lill/Il/llllllll/l 3 1293 02845 57