GENETIC AND TRANSCRIPTOMIC ANALYSES OF THE PSEUDOPERONOSPORA CUBENSIS-CUCUMIS SATIVUS INTERACTION By Elizabeth Ann Savory A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Plant Pathology 2012 ABSTRACT GENETIC AND TRANSCRIPTOMIC ANALYSES OF THE PSEUDOPERONOSPORA CUBENSIS-CUCUMIS SATIVUS INTERACTION By Elizabeth Ann Savory Pseudoperonospora cubensis [(Berkeley & M. A. Curtis) Rostovzev] is the causal agent of cucurbit downy mildew, a foliar disease responsible for devastating losses worldwide of cucumber, cantaloupe, pumpkin, watermelon, and squash. In the United States, cucurbit downy mildew is the major threat to cucumber (Cucumis sativus) production and has been a significant limiting factor since the 2004 and 2005 growing seasons. Prior to 2004, host resistance had been an effective means of controlling the disease and, as such, limited research had been conducted to study the biology, genetics, or virulence of Ps. cubensis. The first step to expand the knowledge base of Ps. cubensis was the generation of a 64.4 Mbp genomic assembly of the MSU-1 isolate and prediction of 23,519 loci and 23,522 gene models. Similar to other oomycete plant pathogens, Ps. cubensis utilizes RXLR and RXLR-like effector proteins, which can function as either virulence or avirulence determinants during the course of host infection. Using in silico analyses, 271 candidate effector proteins were identified with variable RXLR motifs, including 20 different amino acids at position R1. Of these, only 109 (41%) had a putative ortholog in Phytophthora infestans and evolutionary rate analysis of these orthologs shows that they are evolving at a significantly faster rate than most other genes. One Ps. cubensis effector protein, RXLR protein 1 (PscRXLR1) was characterized in detail. PscRXLR1 was shown to be up-regulated during the early stages of host infection, and elicits a cell death response in Nicotiana benthamiana. PscRXLR1 was also demonstrated to be a product of alternative splicing, marking this as the first example of an alternative splicing event in plant pathogenic oomycetes transforming a non-effector gene into a functional effector protein. We present the first large-scale global gene expression analysis of Ps. cubensis infection of a susceptible C. sativus cultivar, ‘Vlaspik’, and identification of both pathogen and host genes involved in infection and the defense response, respectively. Using mRNA-Seq, we captured differential expression of 2383 Ps. cubensis genes in sporangia and at 1, 2, 3, 4, 6, and 8 days post-inoculation (dpi). Co-expression analyses identified distinct modules of Ps. cubensis genes that were representative of early, intermediate, and late infection stages. Additionally, the expression of 15,286 cucumber genes was detected, of which 14,476 were expressed throughout the infection process from 1 dpi to 8 dpi. The rapid induction of key defense related genes, including catalases, chitinases, lipoxygenases, peroxidases, and protease inhibitors was detected within 1 dpi, suggesting recognition by C. sativus of the initial stages of Ps. cubensis infection. Co-expression network analyses revealed transcriptional networks with distinct patterns of expression including down-regulation at 2 dpi of known defense responsive genes suggesting coordinated suppression of host responses by the pathogen. In total, the work described herein presents an in-depth analysis of the interplay between host susceptibility and pathogen virulence in an agriculturally important pathosystem. I would like to dedicate this dissertation to my family. iv ACKNOWLEDGEMENTS I would like to thank my advisor, Dr. Brad Day for his unwavering support, advice, and guidance over the past 5 years. I would also like to thank my committee members Drs. C. Robin Buell, Raymond Hammerschmidt, and Rebecca Grumet for their support, collaboration, and advice over the course of my dissertation research. I would like to thank the Michigan State University Project GREEEN, Pickle Packers International, Inc, National Science Foundation East Asia and Pacific Summer Institutes and the United States Department of Agriculture National Institute of Food and Agriculture, Agriculture and Food Research Initiative for funding this reasearch. Finally, I would like to thank my lab mates, Caleb Knepper, Katie Porter, Alyssa Burkhardt, Patricia Santos, and Masaki Shimono who always made coming to lab worthwhile, who were always there whether the data looked good or bad, and without whom I would not have made it through. Thank you. v TABLE OF CONTENTS LIST OF TABLES...........................................................................................................viii LIST OF FIGURES...........................................................................................................ix CHAPTER 1 The cucurbit downy mildew pathogen Pseudoperonospora cubensis Summary................................................................................................................2 Introduction............................................................................................................4 Taxonomy and Morphology…………………………………………………………….5 Symptoms and Signs.............................................................................................7 Dispersal and Survival.........................................................................................10 Infection Mechanisms..........................................................................................12 Pathogenicity and Virulence................................................................................13 Disease Management..........................................................................................16 Future Prospects..................................................................................................19 Acknowledgements..............................................................................................21 . References...........................................................................................................22 CHAPTER 2 Alternative splicing of a multi-drug transporter from Pseudoperonospora cubensis generates an RXLR effector protein that elicits a rapid cell death Abstract................................................................................................................32 Introduction..........................................................................................................33 Results.................................................................................................................38 Discussion............................................................................................................57 Materials and Methods.........................................................................................63 Acknowledgements..............................................................................................72 Appendix………………………………………………………………………………..73 References...........................................................................................................87 CHAPTER 3 mRNA-Seq Analysis of the Pseudoperonospora cubensis transcriptome during cucumber (Cucumis sativus L.) infection Abstract…………………………………………………………………………………95 Introduction……………………………………………………………………………..96 Results and Discussion……………………………………………………………...100 Conclusions………………………...…………………………………………………120 Materials and Methods………………………………………………………………121 Appendix………………………………………………………………………………127 References……………………………………………………………………………132 vi CHAPTER 4 Expression profiling of Cucumis sativus in response to infection by Pseudoperonospora cubensis Abstract………………………………………………………………………………..141 Introduction……………………………………………………………………………142 Results and Discussion……………………………………………………………...145 Conclusions………………..………………………………………………………….162 Materials and Methods………………………………………………………………164 Acknowledgements…………………………………………………………………..169 Appendix………………………………………………………………………………170 References……………………………………………………………………………177 CHAPTER 5 Conclusions and Future Directions Conclusions.......................................................................................................185 Future Directions................................................................................................191 References.........................................................................................................195 vii LIST OF TABLES Table 3.1 Number of differentially expressed genes between each combination of time points and sporangia…………………………………………………...111 Table 4.1 Number of genes differentially expressed between different time points……………………………………………………………………….…159 viii LIST OF FIGURES Figure 1.1 Cucurbit downy mildew symptoms caused by Pseudoperonospora cubensis………………………………………………………………………….8 Figure 1.2 Morphology of Pseudoperonospora cubensis.............................................9 Figure 1.3 Life cycle of Pseudoperonospora cubensis………………………………..11 Figure 2.1 Strength of purifying selection on Pseudoperonospora cubensis effectors…………………………………….…………………………………..41 Figure 2.2 PscRXLR1 encodes a RXLR-containing effector protein with homology to a non-effector protein in Phytophthora infestans…………………………..47 Figure 2.3. Functional characterization of PscRXLR1 and PITG_17484……………..50 Figure 2.4. PscRXLR1 mRNA expression is up-regulated during Pseudoperonospora cubensis infection of cucumber………………………………………………54 Figure 2.5. PscRXLR1 is a splice variant of Psc_781.4………………………………..56 Figure S2.1 Signal peptide distribution among ortholog pairs…………………………..74 Figure S2.2 Relationship between PscRXLR1 and oomycete orthologs……………...75 Figure S2.3 Heterologous expression of PscRXLR1 specifically results in cell death in Nicotiana benthamiana………………………………………………………..81 Figure S2.4 Multiple sequence alignments of splice variant isoforms…………………82 Figure S2.5 Heterologous expression of Psc_781.4 in Nicotiana benthamiana……...86 Figure 3.1 Experimental design and sample collection…………………………… 102 Figure 3.2 Symptoms and microscopy images of Ps. cubensis infected Cucumis sativus cultivar 'Vlaspik' of time points used for transcriptome analysis………………………………………………………………………..104 Figure 3.3 Number of total RNA-seq reads, reads mapped, and number of genes expressed……………………………………………………………………..106 Figure 3.4 Correlation matrix of Pseudoperonospora cubensis expression profiles throughout a time course of Cucumis sativus infection…………………..109 ix Figure 3.5 Candidate effectors expressed at different timepoints…………………..113 Figure 3.6 CAZymes in Pseudoperonospora cubensis expressed during infection on Cucumis sativus………………………………………………………………114 Figure 3.7 Comparison of ribonucleic acid sequencing (RNA-seq) and microarray expression patterns…………………………………………………………116 Figure 3.8 Heat map of the eigengenes representing each gene module..………..119 Figure S3.1 Number of total RNA-seq reads, reads mapped, and number of genes expressed at different timepoints…………………………………………..128 Figure S3.2 Concordance of FPKM values of the genes expressed in two biological replicates of the Pseudoperonospora cubensis transcriptome………….129 Figure S3.3 Trend plots of the normalized gene expression values for each gene from six identified gene co-expression modules………………………………..131 Figure 4.1 Symptoms of Pseudoperonospora cubensis infection on susceptible Cucumis sativus cv. ‘Vlaspik’………………………………………………147 Figure 4.2 Experimental design and sample collection………………………………149 Figure 4.3 Comparison of total mRNA-Seq reads, reads mapped and number of genes expressed……………………………………………………………..151 Figure 4.4 Correlation matrix of Cucumis sativus expression profiles during infection by Pseudoperonospora cubensis…………………………………………..154 Figure 4.5 Comparison of orthologous gene expression in Cucumis sativus and Arabidopsis thaliana in a compatible interaction………………………..157 Figure 4.6 Heat map of eigengenes representing each gene module………………161 Figure 4.7 Trend plots of the normalized gene expression values from six identified gene co-expression modules………………………………………………163 Figure S4.1 Concordance of expression values in two biological replicates of Cucumis sativus during infection by Pseudoperonospora cubensis………………171 Figure S4.2 Trend plots for all 11 modules………………………………………………173 x CHAPTER 1 The cucurbit downy mildew pathogen Pseudoperonospora cubensis This review was previously published in Molecular Plant Pathology. Savory EA, Granke LL, Quesada-Ocampo LM, Varbanova M, Hausbeck M, and Day B. 2011. The cucurbit downy mildew pathogen Pseudoperonospora cubensis. Mol Plant Pathol 12:217-226. © 2011 BSPP and Blackwell Publishing Ltd. 1 SUMMARY Pseudoperonospora cubensis [(Berkeley & M. A. Curtis) Rostovzev], the causal agent of cucurbit downy mildew, is responsible for devastating losses worldwide of cucumber, cantaloupe, pumpkin, watermelon, and squash. Although downy mildew has been a major issue in Europe since the mid-1980s, in the United States, downy mildew on cucumber was successfully controlled for many years through host resistance. However, since the 2004 growing season, host resistance is no longer effective, and as a result, the control of downy mildew on cucurbits now requires an intensive fungicide program. Chemical control is not always feasible due to the high costs associated with fungicides and their application. Moreover, the presence of pathogen populations resistant to commonly used fungicides limits the long-term viability of chemical control. This review summarizes the current knowledge of taxonomy, disease development, virulence, pathogenicity and control of Ps. cubensis. In addition, topics for future research that aim to develop both short- and long-term control measures of cucurbit downy mildew are discussed. Taxonomy: Kingdom Straminipila; Phylum Oomycota; Class Oomycetes; Order Peronosporales; Family Peronosporaceae; Genus Pseudoperonospora; Species Pseudoperonospora cubensis. Disease Symptoms: Angular chlorotic lesions bound by leaf veins on the foliage of cucumber. Symptoms vary on different cucurbit species and varieties, specifically in 2 terms of lesion development, shape, and size. Infection of cucurbits by Ps. cubensis impacts fruit yield and overall plant health. Infection process: Sporulation on the underside of leaves results in the production of sporangia that are wind-dispersed. Upon arrival to a susceptible host, sporangia germinate in free water on the leaf surface, producing biflagellate zoospores that swim to and encyst on stomata where they form germ tubes. An appressorium is produced and forms a penetration hypha, which enters the leaf tissue through the stomata. Hyphae grow through the mesophyll and establish haustoria, specialized structures for the transfer of nutrients and signals between host and pathogen. Control: Management of downy mildew in Europe requires the use of tolerant cucurbit cultivars in conjunction with fungicide applications. In the U.S., an aggressive fungicide program with sprays every five to seven days for cucumber, and every seven to ten days for other cucurbits, has been necessary to control outbreaks and prevent crop loss. Useful websites: http://www.daylab.plp.msu.edu/pseudoperonospora-cubensis/ (Day Lab website with research advances in downy mildew), http://veggies.msu.edu/ (Hausbeck Lab website with downy mildew news for growers), http://cdm.ipmpipe.org/ (Cucurbit downy mildew forecasting homepage), http://ipm.msu.edu/downymildew.htm (Downy mildew information for Michigan’s vegetable growers). 3 INTRODUCTION Cucurbit downy mildew (caused by Pseudoperonospora cubensis) is one of the most important foliar diseases of cucurbits, causing significant yield losses in the U.S., Europe, China and Israel [1]. The pathogen has a wide geographic distribution and has been reported in over 70 countries, including environments ranging from semi-arid to tropical. In addition, Ps. cubensis has a wide host range, infecting approximately 20 different genera of cucurbits [2,3]. The cucurbit crops grown in the U.S. that are susceptible to this aerially dispersed oomycete pathogen are valued at more than $246.2 million [4]. Control of downy mildew relies on application of fungicides and the use of host resistance. Nonetheless, fungicide-resistant Ps. cubensis populations have been documented throughout the world [3,5,6,7,8], and host resistance is no longer sufficient to control the disease as it once was in the U.S. [9]. The additional cost of fungicides, coupled with potential yield losses of up to 100% caused by downy mildew, threaten the long-term viability of cucurbit crop production [3,8,10]. Detailed knowledge of Ps. cubensis epidemiology, infection processes, and population genetics is currently lacking, but is necessary to guide future efforts in developing new resistant varieties and fungicides, as well as preventing pathogen populations from overcoming host resistance and chemical control. Studies of pathogen epidemiology and global population genetics, evaluation of fungicides for disease control, and 4 development of a high-coverage draft genome sequence will assist in our understanding of the pathogen, as well as in developing effective diagnostics and control measures for Ps. cubensis. The aim of this review is to briefly summarize what is currently known about the cucurbit downy mildew pathogen, Ps. cubensis, including taxonomy, disease development, virulence, pathogenicity and management. TAXONOMY AND MORPHOLOGY Ps. cubensis is the type species of the genus Pseudoperonospora, which includes five accepted species: Ps. cubensis, Ps. humuli, Ps. cannabina, Ps. celtidis and Ps. urticae, [11]. In addition, there are reports of a sixth species, Ps. cassiae, which, while rare, may also be a true species of Pseudoperonospora [12]. Originally named Peronospora cubensis when discovered in Cuba by Berkeley and Curtis [13], Ps. cubensis was reclassified in 1903 after further observations of sporangia germination [14]. Pseudoperonospora species have true sporangia that germinate via cytoplasmic cleavage to produce zoospores (Figure 1.2c); whereas, species of Peronospora have sporangia that germinate directly via a germ tube [2,14]. Recent work has shown that there are no significant morphological differences between Ps. cubensis and the hop (Humulus spp.) pathogen Ps. humuli; nonetheless, there is no evidence that Ps. humuli can infect cucurbits, and limited support for Ps. cubensis pathogenicity on hops [15]. Molecular evidence also shows conflictive results. Internal transcribed spacer (ITS) region sequences of both pathogens are highly similar, which 5 suggests that Ps. humuli could be a taxonomic synonym of Ps. cubensis [11]. However, a recent study using single nucleotide polymorphisms (SNPs) indicates that two nuclear and one mitochondrial gene support the separation of Ps. cubensis from Ps. humuli [15]. In addition, host range studies have demonstrated pathogenic differences between Ps. cubensis and Ps. humuli that further supports separation of these species [16]. Overall, these genetic, phenotypic and physiological characterizations of Ps. cubensis and Ps. humuli provide support for the distinction between these species. Further studies including evidence from hundreds of loci would be helpful to fully resolve the phylogeny of these closely related species. Morphological characters may not provide enough information for characterization of Ps. cubensis isolates or even for differentiating species of Pseudoperonospora [17]. Ps. cubensis sporangiophore morphology can vary with temperature, while sporangia dimensions are influenced by cucurbit host [12,18]. Recent work with a single isolate of Ps. cubensis inoculated onto six different cucurbit species has shown that the host cell matrix plays a role in influencing five morphological criteria, including sporangiophore length, length of ultimate branchlets, sporangial length and width, and the ratio between sporangial length and width [17]. The differences among these morphological characteristics were more obvious in phylogenetically unrelated hosts. These results indicate that it is desirable to include information from genetic markers when resolving phylogenetic relationships in species of Pseudoperonospora. 6 SYMPTOMS AND SIGNS Downy mildew of cucurbits is a foliar disease and is easily recognizable by the development of chlorotic lesions on the adaxial leaf surface, sometimes with necrotic centers. These lesions can be restricted by the leaf vein, as in cucumber, giving them an angular appearance (Figure 1.1). In other cucurbits, the symptoms may vary slightly in terms of shape and color. For example, in both melon and watermelon, foliar lesions are less defined than those on cucumber, and are not always bound by leaf veins [1]. As infection progresses, the chlorotic lesions expand and may become necrotic [19], with necrosis occurring more quickly in hot, dry weather [20]. Leaves colonized by Ps. cubensis undergo changes in temperature and transpiration rates, which vary during the course of infection and over the leaf surface [19,21]. Low temperatures can delay symptom development while still promoting colonization of the leaf tissue, while higher temperatures result in faster lesion chlorosis that may inhibit pathogen growth [22]. As the downy mildew disease progresses, entire leaves may die within days following the initial infection, as lesions expand and coalesce [1]. A reduced canopy leads to cessation of fruit development and increased sun exposure of the fruit, allowing for sun scald and secondary rots [23]. Ultimately, crop yield and fruit quality are affected (Figure 1.1). When temperatures are below those that allow lesion formation and relative humidity is ≥90%, sporulation, the eponymous “downy” appearance on the lower leaf surface, may be the first sign of disease (Figure 1.1) [24]. Hyaline sporangiophores (180-400 µm) 7 Figure 1.1 Figure 1.1 Symptoms caused by Pseudoperonospora cubensis. (A) Yellow angular lesions on a cucumber leaf. (B) Typical “downy” appearance on abaxial leaf surface caused by sporulation. Yields from untreated (C) and treated (D) cucumbers infected with Ps. cubensis. E. Field showing typical disease symptoms. Severe symptoms on untreated (F) versus treated (G) cucumbers. For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this dissertation. bearing papillate, lemon-shaped, gray-purple sporangia (20-40 x 14-25 µm) on sterigmata emerge in groups of 1-6 from stomata on the abaxial surface of infected leaves (Figure 1.2) [11,25]. While leaf wetness is prohibitory for sporangium production, a period of near-saturated relative humidity must occur for ≥6 hours to induce sporulation [26]. Sporulation, as in other downy mildews, is dependent on the diurnal cycle, and is enhanced by longer photoperiods [27]. Differentiation of sporangia requires a minimum 6 hour dark period [28]. The optimum temperature for sporangia production is 15-20°C, but sporangia may form on cucumber at temperatures from 5 to 8 Figure 1.2 Figure 1.2 Morphology of Pseudoperonospora cubensis. (A) Sporangiophore (bar = 50 µm). (B) Sporangia attached to distal end of sporangiophores. (C) Sporangia germinating via cytoplasmic cleavage. (D) Left panel: zoospores. Right panel: encysted zoospore with germ tube. (E) Intercellular growth: S, stomata; H, haustorium; IH, intercellular hyphae. (F) Scanning electron micrograph (SEM) of sporangiophore (bar = 20 µm). (G) SEM of multiple sporangiophores emerging through stomata (bar = 20 µm). Bars = 25 µm, except where noted. 30°C [1]. Other factors such as the host species, cultivar, host nutritional status, and host age may also affect sporulation [26]. 9 DISPERSAL AND SURVIVAL Ps. cubensis cannot overwinter in geographic locations with killing frosts. Instead, the pathogen overwinters in areas with mild winter temperatures that permit cucurbit hosts to be grown year round [29] or in greenhouses [1]. It was recently demonstrated that Ps. cubensis could infect a perennial member of the Cucurbitaceae, Bryonia dioica, in the laboratory, and the pathogen could potentially overwinter on this host in Central and Northern Europe [30]. However, this has not been supported with observations in the field [31], and it is unknown if B. dioica plays an important role in the life cycle of Ps. cubensis [30]. While oospores have been observed in both temperate and tropical regions including India, Japan, Austria, Russia, China, Italy, and Israel [2,32,33,34,35,36], the production of oospores is very rare [1,2]. The rare occurrence of thick-walled resting structures, ie., oospores, limits Ps. cubensis survival in the absence of a living host. It is currently unknown if oospores play an important role in the disease cycle (Figure 1.3). As an obligate biotroph Ps. cubensis requires live host tissue for reproduction and dispersal. Copious asexual sporangia are produced on infected foliage, which may be liberated to the air following a reduction in relative humidity when hygroscopic twisting movements of sporangiophores actively release sporangia into air currents [37]. Hence, airborne Ps. cubensis sporangia concentrations are greater in the morning and early afternoon when changes in relative humidity and leaf wetness tend to occur [27]. The distance sporangia travel depends upon where in the canopy the sporangia are 10 Figure 1.3 Figure 1.3 Life cycle of Pseudoperonospora cubensis. (A) Aerially dispersed lemon-shaped gray-purple sporangia land on the leaf surface and germinate in free moisture to form biflagellate zoospores. (B) Zoospores swim to and encyst in stomata, then penetrate the leaf surface via a germ tube. (C) Hyphae colonizes the mesophyll layer, establishing clavate-branched haustoria within plant cells. (D) The diurnal cycle triggers sporulation and up to 6 sporangiophores emerge through each stomate, bearing sporangia at their tips. Sporangia are dislodged from sporangiophores by changes in hydrostatic pressure and are picked up by wind currents that carry them to their next host. (E) Chlorotic, angular lesions bound by leaf veins are a symptom of Ps. cubensis infection visible on the adaxial leaf surface. On the lower leaf surface, sporulation is visible (inset). (F) The role of the sexual stage of Ps. cubensis is unknown. 11 produced, as well as wind conditions as they become airborne [38]. Like other downy mildews [39,40], Ps. cubensis sporangia may be wind-dispersed over long distances [1]. As such, it has been proposed that in the U.S., Ps. cubensis overwinters in frostfree areas of the southern states (e.g., Florida and Texas) and spreads northward each growing season via wind currents [41,42,43]. Likewise, sporangia that infect cucurbits in Central Europe originate from year-round production areas in Southeast Europe, and are transported via wind currents [44]. Subsequent local transport of secondary inoculum occurs primarily via wind, but sporangia also may be dispersed by rain splash or physical transfer on equipment within a field [1]. Sporangial survival during transport is limited to 1 to 16 days after dispersal [26,45] depending on temperature, relative humidity [1] and solar radiation [46]. Once a sporangium lands on a host plant, that sporangium must survive until environmental conditions are favorable for infection. Sporangial survival is favored by conditions of low relative humidity, lower temperature and cloudy days [45,46]. INFECTION MECHANISMS While liberation and dispersal of sporangia occurs under conditions of low leaf moisture, leaf wetness is necessary for the pathogen to successfully germinate and infect the host plant. At 15°C, the optimum temperature for infection, at least 2 hours of leaf wetness are required for infection when high levels of inoculum are present. Sporangia may also germinate and cause infection at temperatures ranging from 5-28°C [26], but longer 12 periods of leaf wetness are required [22,24] under these conditions or when less inoculum is present [26]. The incubation period depends on temperature, photoperiod, inoculum concentration, and leaf wetness duration, and can range from 4 to 12 days [1,22]. Sporangia germinate via cytoplasmic cleavage, resulting in the release of 2-15 motile, biflagellate zoospores [2], which preferentially swim to open stomata where they encyst [47] (Figure 1.2; Figure 1.3). Germ tubes form from encysted zoospores and produce appressoria. A penetration hypha develops from the appressorium and enters through the stomatal aperture into the leaf tissue. Hyaline coenocytic hyphae subsequently form and grow intercellularly through the mesophyll and palisade tissues. Clavate-branched haustoria are established within mesophyll cells where they invaginate the plant cell membrane [48,49] (Figure 1.2; Figure 1.3). These specialized structures are the site of nutrient uptake by the pathogen and delivery of effector proteins that could potentially function to redirect host metabolism and suppress defense responses [50,51]. PATHOGENICITY AND VIRULENCE Ps. cubensis has a broad host range, infecting over 49 species in 20 genera within the Cucurbitaceae, including 19 species in the genus Cucumis [2,52]. Cucumber (Cucumis sativus L.), melon (Cucumis melo L.), watermelon (Citrullus lanatus Matsum. and Nakai), and squash (Cucurbita spp.) are the four major food crops that are hosts to Ps. cubensis. Other cucurbits infected by Ps. cubensis include loofa (Luffa acutangula (L.) 13 Roxb.), bottle gourd (Lagenaria siceraria (Molina) Stand.), wax gourd (Benincasa hispida (Thunb.) Cogn.) and bitter melon (Momordica charantia L.) [2]. Ps. cubensis isolates show differences in virulence and pathogenicity depending on the cucurbit variety. To date, six physiological races have been identified in the U.S., Israel and Japan, and additional evidence suggests that many more exist in Europe [36,53,54]. Using Citrullus, Cucumis, and Cucurbita spp., Thomas et al. [55] identified five distinct physiological races of Ps. cubensis: 1 and 2 from Japan, 3 from Israel, and 4 and 5 from the U.S.. In 2003, Cohen et al. identified a sixth physiological race in Israel based on its pathogenicity to a wider range of susceptible cucurbits compared to race 3. All six physiological races that have been described are pathogenic on cucumber and muskmelon (C. melo var. reticulatus) but show differences in pathogenicity on watermelon, squash or pumpkin. Subsequently, Lebeda and Widrlechner (2003) developed a set of differential taxa that included 12 representatives from 6 genera, Benincasa, Citrullus, Cucumis, Cucurbita, Lagenaria and Luffa, which represent natural hosts of Ps. cubensis. Using this set of hosts, the authors evaluated the differences in pathogenicity of 22 additional isolates from the Czech Republic, Spain, France and the Netherlands, which were classified as 13 physiological races based on their virulence [52,53]. Collectively, these studies were the first to describe differences in virulence and pathogenicity of Ps. cubensis in Europe in detail; however, the genetic basis for differences among physiological races has not been established. 14 Differences in effector content could be a potential explanation for differences among physiological races. Effector proteins have been shown to play roles as both virulence and avirulence determinants in other oomycetes [56,57,58,59]. Oomycete effector proteins contain the RXLR motif that is located downstream of the signal peptide and are under diversifying selection at the C-terminal domain [60]. Preliminary sequence data of the Ps. cubensis genome has yielded 61 putative effector proteins [61,62]. Of these, 32 were secreted proteins containing the RXLR motif typical of previously identified oomycete effector proteins, while the remaining 29 had an R to Q substitution at the first amino acid (ie., QXLR). A family of QXLR containing effectors, PcQNE (Ps. cubensis QXLR nuclear localized effectors), was shown to localize to the nucleus and the C-terminal domain was under diversifying selection, as has been observed for RXLR effectors [61]. Understanding the diversity and role of effector proteins is key to understanding pathogenicity and the genetic basis for virulence differences between isolates. The pathogenic and genetic diversity of Ps. cubensis has been shown to vary temporally and geographically [2,55,63]. In a study by Lebeda and Urban [3], Ps. cubensis isolates were collected over a three-year period in the Czech Republic, and a general population shift from isolates with low pathogenicity to those with high pathogenicity was observed. This work also demonstrated that the variability of the population decreased from 33 different physiological races in 2001 to only 13 physiological races in 2003 [3]. An increased representation of a highly pathogenic physiological race that was capable of causing infection on all cucurbits studied was 15 also observed over the course of the experiments, indicating a population shift to high pathogenicity isolates [3]. Recently, using a combination of amplified fragment length polymorphism (AFLP) and ITS sequence analyses, molecular polymorphisms were identified among populations of Ps. cubensis from Greece, the Czech Republic, the Netherlands and France [54]. While there was no clear grouping of isolates based on their pathogenicity, AFLP analysis indicated genetic differentiation between the Greek isolates and those from central and western Europe [54]. The work by Sarris and colleagues (2009) suggested that the clustering of Ps. cubensis isolates from Greece corresponded to their geographical distribution rather than their pathogenicity or virulence on cucurbit hosts. Nevertheless, this study was supported on SNP evidence from one single loci and AFLP evidence with error rates of 1-4% due to band scoring. A similar study with information from several loci, a reliable polymorphism scoring system, and a more extensive sampling of regions of interest is needed to support or disprove what Sarris and colleagues (2009) proposed in their paper. DISEASE MANAGEMENT Ps. cubensis outbreaks over the past several decades have been responsible for annual yield losses of up to 80%, and as a result, cucurbit downy mildew is currently the most destructive disease of cucumbers for both field and greenhouse production in Europe [3,44]. Before 1984, downy mildew was not a major issue for Central [64,65] and Northern Europe [66,67]. However, from ~1985 onward, epidemics of downy mildew have been a challenge for cucurbit production in Europe [65]. In Europe, it has 16 been suggested that tolerant cultivars should be used in conjunction with fungicide applications when conditions are favorable for downy mildew [68,69]. In the U.S., host resistance introduced in the 1950s was effective in limiting losses caused by downy mildew without the use of fungicides. A resurgence of the disease was observed in many states along the Eastern Seaboard and the upper Midwest in 2004 and 2005, respectively [9]. This loss in resistance has led to the management of downy mildew through a spray program, with recommended fungicides applied every five to seven days on cucumbers [8,70] and every seven to ten days on other cucurbits [71]. An aggressive spray program is essential, as plants must have a protective barrier of fungicide prior to sporangium deposition to avoid yield losses. However, additional fungicide applications to control downy mildew greatly increase the cost of production. In Michigan, for example, major outbreaks of the disease have been observed since 2005, and the cost of additional fungicide sprays is over 6 million dollars annually (Hausbeck, unpublished data). Since fungicide applications are expensive, downy mildew sporangia trapping and forecasts can be useful tools to alert growers to when airborne sporangia are present or likely to be present so they can make an informed decision about when to initiate fungicide applications [43]. Delaying the initiation of a fungicide spray program may reduce management costs for growers and the amount of fungicides in the environment. The efficacy of chemical control measures may be diminished if Ps. cubensis populations develop resistance to key fungicides; Ps. cubensis was the first oomycete 17 with documented resistance to metalaxyl and reduced sensitivity to mancozeb [63,72]. In addition, populations of Ps. cubensis resistant to strobulurin fungicides have been described [73]. In the Eastern U.S. field and greenhouse fungicide trials, products containing mefenoxam and strobilurins as the active ingredients failed to provide adequate downy mildew disease control, indicating that resistance is widespread in this region [5,8,70,74]. Since Ps. cubensis has a high potential of developing resistance to fungicides [75], it is important that populations are carefully monitored for resistance to currently registered products and that new active ingredients are tested. While fungicide applications are currently necessary for adequate disease control [76], resistant varieties and cultural techniques are important components of a management strategy. The full genetic parameters controlling resistance to downy mildew in cucumber are unknown. The original source of host resistance (i.e., the recessive dm1 gene) was identified in cucumber accession PI 197087 and first described in India in 1954 [77]. The resistance response governed by dm1 is characterized by sparse pathogen sporulation, small necrotic lesions, tissue browning, and rapid cell death, indicative of the classical hypersensitive response (HR) type resistance. Since the 1950s, resistance conferred by dm1 has been widely used in commercial cultivars for cucumber production in the U.S. and was sufficient to prevent losses due to downy mildew until 2004 [9]. Cultivars containing the dm1 gene still show some level of resistance; unfortunately, the high level of resistance once observed is now lost. Additionally, susceptible cultivars without the dm1 gene become infected earlier in the season, and exhibit more severe damage than was previously observed [43]. 18 While the dm1 remains useful in a disease management program, a robust source of resistance is highly desirable. Current breeding research for resistance to downy mildew in cucumber is focused on the identification of resistant germplasm(s) and cultivars via large scale screening trials [78,79]. Tolerant and high yielding germplasm has been identified in these studies, but a source of complete resistance to downy mildew in cucumber has been elusive, likely due to limited genetic diversity for Ps. cubensis resistance in cucumber [52,79]. Other Cucumis spp., such as melon, may be more relevant for identifying effective sources of resistance [52]. Wild melon line PI 124111F [PI], for example, has been shown to be resistant to the 6 physiological races of Ps. cubensis via two constitutively expressed glyoxylate aminotransferase-encoding genes, At1 and At2 [80]. These two genes are known as enzymatic resistance (eR) genes which when expressed at high levels in either wild type or transgenic plants, confer complete resistance to infection by Ps. cubensis. FUTURE PROSPECTS The re-emergence of cucurbit downy mildew in the U.S. and its persistence across much of Europe and Asia, represents a significant threat to cucurbit production worldwide. While the disease was successfully managed for decades within the U.S. using host resistance (i.e., the dm-1 locus), severe epidemics have occurred since 2004. Whether this is due to a change in pathogen populations or a change in the environment is currently unknown. To this end, we need to investigate the changes in 19 pathogen populations, environmental factors, and how the pathogen-environment interaction affects the host-pathogen interaction and disease development. Studies have investigated some aspects of the basic biology of Ps. cubensis and its interactions with various cucurbit hosts. However, additional research is needed to further clarify the taxonomy, variations in virulence and pathogenicity among physiological races, and the pathogenicity determinants of the pathogen. A better understanding of each of these components will ultimately facilitate the development of durable host resistance. To this end, the forthcoming genome sequence will provide molecular tools for gene discovery and the development of molecular markers, which may then be used to investigate population and evolutionary biology of the pathogen. Such studies will yield information about possible migration events or evolutionary changes within pre-existing U.S. pathogen populations that resulted in strains with increased virulence. An integrated research approach that includes all factors affecting disease development (pathogen, host, and environment) is essential to control and predict future cucurbit downy mildew epidemics. First, it would benefit growers if new fungicides that are more economical and provide effective control were identified. Second, screening of breeding lines and wild germplasm will help identify durable sources of genetic resistance that would be preferable to an aggressive spray program. Finally, studies to determine the effects of environment on inocula and disease development will serve as a first step in developing rapid, and highly specific forecasting systems. In summary, research efforts contributing to the development of sustainable management strategies such as durable 20 host resistance are a priority to ensure the long-term viability of the cucurbit production industry. ACKNOWLEDGEMENTS The authors thank members of the Day lab for critical reading of the manuscript. Cucurbit downy mildew research in the Day lab is funded by the Michigan Agricultural Experiment Station (MAES), Project GREEEN (Award numbers GR06-0099D and GF07-077), the Michigan State University Office of the Vice President for Research and Graduate Studies, the National Science Foundation (Award number IOS-0641319), and a joint grant awarded to MKH and BD from the Agricultural Research Fund of Pickle Packers International Inc.. Work in the Hausbeck lab is supported by the Pickle and Pepper Research Committee of MSU, Fresh Vegetable Growers of Ontario, North Central IPM Center (Sub award 2003-51120-02111 S4256), and Project GREEEN (Award Numbers GR07-077 and GR06-0099D). 21 REFERENCES 22 REFERENCES 1. Thomas CE (1996) Downy mildew. In: Zitter TA, editor. Compendium of cucurbit diseases. Ithaca, NY: Cornell University Press. pp. 25-27. 2. Palti J, Cohen Y (1980) Downy mildew of cucurbits (Pseudoperonospora cubensis): the fungus and its hosts, distribution, epidemiology, and control. Phytoparasitica 8: 109-147. 3. Lebeda A, Urban J (2007) Temporal changes in pathogenicity and fungicide resistance in Pseudoperonospora cubensis populations. Acta Hort 731: 327-336. 4. Anonymous (2009) National Online Statistics. US Dep Agric Natl Agric Stat Serv. 5. Colucci SJ, Holmes GJ (2007) Fungicide insensitivity and pathotype determination of Pseudoperonospora cubensis, causal agent of cucurbit downy mildew. Phytopathology 97: S24. 6. Zhu S, Liu P, Liu X, Li J, Yuan S, et al. (2008) Assessing the risk of resistance in Pseudoperonospora cubensis to the fungicide flumorph in vitro. Pest Management Science 64: 255-261. 7. Mitani S, Araki S, Yamaguchi T, Takii Y, Ohshima T, et al. (2001) Biological properties of the novel fungicide cyazofamid against Phytophthora infestans on tomato and Pseudoperonospora cubensis on cucumber. Pest Manag Sci 58: 139-145. 8. Hausbeck MK, Cortright BD (2009) Evaluation of fungicides for control of downy mildew of pickling cucumber, 2007. PDMR 3: V112. 9. Holmes GJ, Thomas CE (2006) The history and re-emergence of cucurbit downy mildew. Phytopathology 99: S171. 10. Holmes G, Wehner T, Thornton A (2006) An old enemy re-emerges. American Vegetable Grower Feb.: 14-15. 11. Choi YJ, Hong SB, Shin HD (2005) A re-consideration of Pseudoperonospora cubensis and P. humuli based on molecular and morphological data. Mycol Res 109: 841-848. 23 12. Waterhouse GM, Brothers MP (1981) The taxonomy of Pseudoperonospora. Mycological Papers 148: 1-18. 13. Berkeley MS, Curtis A (1868) Peronospora cubensis. J Linn Soc Bot 10: 363. 14. Rostovzev SI (1903) Beitrage zur Kenntnis der Peronosporeen. Flora 92: 405-430. 15. Mitchell MN, Ocamb C, Gent D (2009) Addressing the relationship between Pseudoperonospora cubensis and Pseudoperonospora humuli by multigenic characterization and host specificity. Phytopathology 99: S87. 16. Gent DH, Mitchell MN, Holmes GJ (2009) Genetic and pathogenic relatedness of Pseudoperonospora cubensis and P. humuli: Implications fo detection and management. Phytopathology 99: S171. 17. Runge F, Thines M (In Press) Host matrix has major impact on the morphology of Pseudoperonospora cubensis. Eur J Plant Pathol: 1-10. 18. Iwata Y (1942) Specialization in Pseudoperonospora cubensis (Berk. et Curt.) Rostov. II. Comparative studies of the morphologies of the fungi from Cucumis sativus L. and Cucurbita moschata Duchesne. Ann Phytopathol Soc Japan 11: 172-185. 19. Oerke EC, Steiner U, Dehne HW, Lindenthal M (2006) Thermal imaging of cucumber leaves affected by downy mildew and environmental conditions. J Exp Bot 57: 2121-2132. 20. Cohen Y, Rotem J (1971) Rate of lesion development in relation to sporulating potential of Pseudoperonospora cubensis in cucumbers. Phytopathology 61: 265-268. 21. Lindenthal M, Steiner U, Dehne HW, Oerke EC (2005) Effect of downy mildew development on transpiration of cucumber leaves visualized by digital infrared thermography. Phytopathology 95: 233-240. 22. Cohen Y (1977) The combined effects of temperature, leaf wetness, and inoculum concentration on infection of cucumbers with Pseudoperonospora cubensis. Can J Bot 55: 1478-1487. 24 23. Keinath AP, Holmes GJ, Everts KL, Egel DS, Langston DB (2007) Evaluation of combinations of chlorothalonil with azoxystrobin, harpin, and disease forecasting for control of downy mildew and gummy stem blight on melon. Crop Protection 26: 83-88. 24. Rotem J, Cohen Y, Bashi E (1978) Host and environmental influences on sporulation in vivo. Ann Rev Phytopathol 16: 83-101. 25. Palti J (1975) Pseudoperonospora cubensis (Berk & M. A. Curtis) Rost. C M I Descr Pathog Fungi Bact 457: 1-2. 26. Cohen Y (1981) Downy mildew of cucurbits. In: Spencer DM, editor. The Downy Mildews. London: Academic Press. pp. 341-354. 27. Cohen Y, Rotem J (1971) Field and growth chamber approach to epidemiology of Pseudoperonospora cubensis on cucumbers. Phytopathology 61: 736-737. 28. Cohen Y (1977) Growth and differentiation of sporangia and sporangiophores of Pseudoperonospora cubensis on cucumber cotyledons under various combinations of light and temperature. Physiol Plant Pathol 10: 93-103. 29. Bains SS, Jhooty JS (1976) Over wintering of Pseudoperonospora cubensis causing downy mildew of muskmelon. Indian Phytopathol 29: 213-214. 30. Runge F, Thines M (2009) A potential perennial host for Pseudoperonospora cubensis in temperate regions. Eur J Plant Pathol 123: 483-486. 31. Lebeda A, Cohen Y (In Press) Cucurbit downy mildew (Pseudoperonospora cubensis)—biology, ecology, epidemiology, host-pathogen interaction and control. Eur J Plant Pathol: 1-36. 32. Bains SS, Sokhi SS, Jhooty JS (1977) Melothria maderaspatana - a new host of Pseudoperonospora cubensis. Indian J Mycol Plant Pathol 7: 86. 33. Hiura M, Kawada S (1933) On the overwintering of Peronoplasmopara cubensis Jap J Bot 6: 507-513. 25 34. D'Ercole N (1975) La peronoospora del cetriolo in coltura protetta. Inftor Fitopath 25: 11-13. 35. Bedlan G (1989) First detection of oospores of Pseudoperonospora cubensis (Berk et Curt.) Rost. on glasshousse cucumbers in Austria. Pflantzenschutzberichte 50: 119-120. 36. Cohen Y, Meron I, Mor N, Zuriel S (2003) A new pathotype of Pseudoperonospora cubensis causing downy mildew in cucurbits in Israel. Phytoparasitica 31: 458466. 37. Lange L, Eden U, Olson LW (1989) Zoosporogenesis in Pseudoperonospora cubensis, the causal agent of cucurbit downy mildew. Nord J Bot 8: 497-504. 38. Aylor DE (1990) The role of intermittent wind in the dispersal of fungal pathogens. Ann Rev Phytopathol 28: 73-92. 39. Aylor DE, Taylor GS (1982) Aerial dispersal and drying of Peronospora tabacina conidia in tobacco shade tents. Proc Natl Acad Sci USA 79: 697-700. 40. Wu BM, van Bruggen AHC, Subbarao KV, Pennings GGH (2001) Spatial analysis of lettuce downy mildew using geostatistics and geographic information systems. Phytopathology 91: 134-142. 41. Doran WL (1932) Downy mildew of cucumbers. Massachusetts Agr Exp Sta Bull 283. 42. Van Haltern F (1933) Spraying cantaloupes for the control of downy mildew and other diseases. Georgia Exp Sta Bull. 43. Holmes GJ, Main CE, Keever ZT (2006) Cucurbit downy mildew: a unique pathosystem for disease forecasting. In: Spencer-Phillips PTN, Jeger M, editors. Advances in downy mildew research. Dordrecht, The Netherlands: Kluwer academic publishers. 44. Lebeda A, Schwinn FJ (1994) The downy mildews - an overview of recent research progress. J Plant Dis Protection 101: 225-254. 26 45. Cohen Y, Rotem J (1971) Dispersal and viability of Pseudoperonospora cubensis. Trans Br Mycol Soc 57: 67-74. sporangia of 46. Kanetis L, Holmes GJ, Ojiambo PS (2009) Survival of Pseudoperonospora cubensis sporangia exposed to solar radiation. Plant Pathology 59: 313-323. 47. Iwata Y (1949) Studies on the invasion of cucumber plants by downy mildew. (In Japanese, with English summary). Ann Phytopathol Soc Japan 13: 60-61. 48. Voglmayr H, Riethmuller A, Goker M, Weiss M, Oberwinkler F (2004) Phylogenetic relationships of Plasmopara, Bremia and other genera of downy mildew pathogens with pyriform haustoria based on bayesian analysis of partial LSU rDNA sequence data. Mycol Res 108: 1011-1024. 49. Fraymouth J (1956) Haustoria of the Peronosporales. Trans Br Mycol Soc 39: 79107. 50. Hahn M, Mendgen K (2001) Signal and nutrient exchange at biotrophic plant-fungus interfaces. Curr Opin Plant Biol 4: 322-327. 51. Whisson SC, Boevink PC, Moleleki L, Avrova AO, Morales JG, et al. (2007) A translocation signal for delivery of oomycete effector proteins into host plant cells. Nature 450: 115-118 52. Lebeda A, Widrlechner MP (2003) A set of Cucurbitaceae taxa for differentiation of Pseudoperonospora cubensis pathotypes. J Plant Dis Protection 110: 337-349. 53. Lebeda A, Gadasova V (2002) Pathogenic variation of Pseudoperonospora cubensis in the Czech Republic and some other European countries. Acta Hort 588: 137-141. 54. Sarris P, Abdelhalim M, Kitner M, Skandalis N, Panopoulos N, et al. (2009) Molecular polymorphisms between populations of Pseudoperonospora cubensis from Greece and the Czech Republic and the phytopathological and phylogenetic implications. Plant Pathology Doi: 10.111/j.1365-3059.2009.02093.x. 55. Thomas CE, Inaba T, Cohen Y (1987) Physiological specialization in Pseudoperonospora cubensis. Phytopathology 77: 1621-1624. 27 56. Schornack S, Huitema E, Cano LM, Bozkurt TO, Oliva R, et al. (2009) Ten things to know about oomycete effectors. Molecular Plant Pathology 10: 795-803. 57. Hogenhout SA, Van der Hoorn RAL, Terauchi R, Kamoun S (2009) Emerging concepts in effector biology of plant-associated organisms. Mol Plant-Microbe Interact 22: 115-122. 58. Oliva R, Win J, Raffaele S, Boutemy L, Bozkurt TO, et al. Recent developments in effector biology of filamentous plant pathogens. Cellular Microbiology 12: 705715. 59. Thines M, Kamoun S (2010) Oomycete-plant coevolution: recent advances and future prospects. Curr Opin Plant Biol 13: 427-433. 60. Win J, Morgan W, Bos J, Krasileva K, Cano L, et al. (2007) Adaptive evolution has targeted the C-terminal domain of the RXLR effectors of plant pathogenic oomycetes. The Plant Cell 19: 2349-2369. 61. Tian M, Win J, Savory EA, Kamoun S, Day B (2011) 454 genome sequencing of Pseudoperonospora cubensis reveals effector proteins with a putative QXLR translocation motif. Mol Plant-Microbe Interact 24. 62. Savory EA, Tian M, Win J, Kamoun S, Day B (2009) Genome characterizarion and discovery of novel QXLR effector motif in the cucurbit downy mildew pathogen Pseudoperonospora cubensis 14th International ISM-MPMI Congress. Quebec, Canada. 63. Thomas CE, Jourdain EL (1992) Host effect on selection of virulence factors affecting sporulation by Pseudoperonospora cubensis. Plant Dis 76: 905-907. 64. Lebeda A (1986) Epidemic occurrence of Pseudoperonospora cubensis in Czechoslovakia. Temperate Downy Mildews Newsletter 4: 15-17. 65. Lebeda A, Schwinn FJ (1994) The downy mildews--an overview of recent research progress. J Plant Dis Protection 101: 225-254. 66. Forsberg AS (1986) Downy mildew-Pseudoperonospora cubensis in Swedish cucumber fields. Vaxtskyddsnotiser 50: 17-19. 28 67. Tahvonen R (1985) Downy mildew of cucurbits found for the first time in Finland. Vaxtskyddsnotiser 49: 42-44. 68. Chaban VS, Okhrimchuk VN, Sergienko VG (2000) Optimization of chemical control of Pseudoperonospora cubensis on cucumber in Ukraine. EPPO Bulletin 30: 213-215. 69. Urban J, Lebeda A (2006) Fungicide resistance in cucurbit downy mildewmethodological, biological, and population aspects. Annals of applied biology 149: 63-75. 70. Gevens AJ, Hausbeck MK (2006) Control of downy mildew of cucumbers with fungicides, 2005. F & N Tests: V062. 71. Hausbeck MK. Downy mildew and Phytophthora control in vine crops; 2009; Syracuse, NY. pp. 193-195. 72. Reuveni M, Eyal H, Cohen Y (1980) Development of resistance to Metalaxyl in Pseudoperonospora cubensis Plant Disease 64: 1108-1108. 73. Heaney SP, Hall AA, Davis SA, Olaya G. Resistance to fungicides in the QoI-STAR cross-resistance group: current perspectives; 2000. pp. 755-762. 74. Keinath AP, DuBose VB, Lassiter AW (2008) Evaluation of fungicides to manage downy mildew on pickling cucumber in Charleston, South Carolina. Plant Disease Management Reports 2: V024. 75. Russell PE (2004) Sensitivity baselines in fungicide resistance research and management. FRAC Monograph 3. 76. Gisi U (2002) Chemical control of downy mildews. In: Spencer-Phillips PTN, Gisi U, Lebeda A, editors. Advances in downy mildew research. Dordrecht, The Netherlands: Kluwer academic publishers. pp. 119-159. 77. Barnes WC, Epps WM (1954) An unreported type of resistance to cucumber downy mildew. Plant Disease Rep 38: 620. 29 78. Wehner TC, Shetty NV (1997) Downy mildew resistance of the cucumber germplasm collection in North Carolina field tests. Crop Science 37: 1331-1340. 79. Shetty NV, Wehner TC, Thomas CE, Doruchowski RW, Vasanth Shetty KP (2002) Evidence for downy mildew races in cucumber tested in Asia, Europe, and North America. Scientia Horticulturae 94: 231-239. 80. Taler D, Galperin M, Benjamin I, Cohen Y, Kenigsbuch D (2004) Plant eR genes that encode photorespiratory enzymes confer resistance against disease. Plant Cell 16: 172-184. 30 CHAPTER 2 Alternative splicing of a multi-drug transporter from Pseudoperonospora cubensis generates an RXLR effector protein that elicits a rapid cell death This chapter was originally published in PLoS ONE. Savory EA, Zou C, Adhikari BN, Hamilton JP, Buell CR, Shiu S-H, and Day B (2012) Alternative splicing of a multi-drug transporter from Pseudoperonospora cubensis. PLoS ONE 7(4): e34701. doi:10.1371/journal.pone.0034701. © 2012 Savory et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited. Author Contributions: Conceived and designed the experiments: EAS, S-HS, CRB, and BD. Performed the experiments: EAS, CZ, and BNA. Analyzed the data: EAS, CZ, BNA, JPH, S-HS, CRB, and BD. Contributed reagents/materials/analysis tools: EAS, CZ, BNA, and JPH. Wrote the paper: EAS and BD. 31 ABSTRACT Pseudoperonospora cubensis, an obligate oomycete pathogen, is the causal agent of cucurbit downy mildew, a foliar disease of global economic importance. Similar to other oomycete plant pathogens, Ps. cubensis has a suite of RXLR and RXLR-like effector proteins, which likely function as virulence or avirulence determinants during the course of host infection. Using in silico analyses, we identified 271 candidate effector proteins within the Ps. cubensis genome with variable RXLR motifs. In extending this analysis, we present the functional characterization of one Ps. cubensis effector protein, RXLR protein 1 (PscRXLR1), and its closest Phytophthora infestans ortholog, PITG_17484, a member of the Drug/Metabolite Transporter (DMT) superfamily. To assess if such effector-non-effector pairs are common among oomycete plant pathogens, we examined the relationship(s) among putative ortholog pairs in Ps. cubensis and P. infestans. Of 271 predicted Ps. cubensis effector proteins, only 109 (41%) had a putative ortholog in P. infestans and evolutionary rate analysis of these orthologs shows that they are evolving significantly faster than most other genes. We found that PscRXLR1 was up-regulated during the early stages of infection of plants, and moreover, that heterologous expression of PscRXLR1 in Nicotiana benthamiana elicits a rapid necrosis. More interestingly, we also demonstrate that PscRXLR1 arises as a product of alternative splicing, making this the first example of an alternative splicing event in plant pathogenic oomycetes transforming a non-effector gene to a functional effector protein. Taken together, these data suggest a role for PscRXLR1 in pathogenicity, and in total, our data provide a basis for comparative analysis of 32 candidate effector proteins and their non-effector orthologs as a means of understanding function and evolutionary history of pathogen effectors. INTRODUCTION The identification and characterization of secreted effector proteins from plant pathogens has anchored the recent evolution of molecular plant pathology [1,2,3]. As components of many pathogenic microorganisms’ secretomes, effector proteins represent a key component of phytopathogenicity, contributing to both the virulence and avirulence capacity of the invading pathogen. Numerous studies have identified and characterized the activities of secreted effector proteins from a broad range of phytopathogens [reviewed in 1,4,5]. Collectively, these works have revealed two primary functions for pathogen effector molecules. First, as virulence molecules, effector proteins can enhance a pathogen’s ability to cause disease, likely through abrogating host processes that would otherwise block pathogen infection, growth, and proliferation within the host [5,6]. Secondly, as avirulence determinants, effector proteins function to modulate the activation of host defense responses by perturbing the activity of host resistance (R) proteins [5,6]. For infection, colonization, and subsequent propagation within their hosts to occur, pathogens must dampen multiple layers of plant defense responses. Often described as basal resistance, the initial perception and elicitation of defenses requires the recognition of pathogen associated molecular patterns (PAMPs; e.g., chitin, flagellin, 33 LPS) [5,6], highly conserved molecules essential for the lifestyle and survival of the microorganism. The recognition of PAMPs, which are highly specific elicitors, occurs through receptors on the host membrane surface, and following initiation of this receptor-ligand interaction, a rapid first response known as PAMP-triggered immunity (PTI) is elicited [5,6]. Overall, the PTI response provides a basal level of resistance against a wide range of microorganisms, often utilizing conserved signaling pathways such as the up-regulation of the mitogen-activated protein kinase (MAPK) pathway, the generation of reactive oxygen species, and the induction of defense-related genes [5,6]. To overcome PTI, phytopathogens, including bacteria and oomycetes, rely on the delivery and activity of secreted effector proteins to abrogate this initial basal level of defense, as well as to further promote virulence [3,5,6]. In response, pathogen effectors can be recognized by R (resistance) proteins, leading to the activation of effectortriggered immunity (ETI) [5,6] best illustrated as an amplified and sustained layer of defense. ETI is a robust response that is often associated with the activation of a specific type of programmed cell death referred to as the hypersensitive response (HR) [5,6]. Over time, as this cycle of subversion and recognition evolves, host specificity and subsequent interactions between pathogen and host are modulated by the interplay between the activity and recognition of secreted pathogen effector molecules and their host counterparts. Oomycetes are a phylogenetically distinct eukaryotic lineage within the Stramenopiles, which as a group, are among the best-studied and most economically important plant pathogens. In recent years, the genomes of several agriculturally important oomycete 34 pathogens have been sequenced, including Phytophthora infestans, Phytophthora ramorum, and Phytophthora sojae, the causal agents of late blight of potato and tomato, sudden oak death, and soybean root rot, respectively [7,8]. The genomes of two other oomycete pathogens, Pythium ultimum, which causes damping off and root rot on a wide range of hosts, and Hyaloperonospora arabidopsidis, a pathogen of Arabidopsis thaliana, have also been sequenced [9,10]. These investigations, through the analysis of genome content and structure, have provided a wealth of information, both towards understanding the nature of the host-pathogen interaction (e.g., host specificity, virulence strategies), as well as insight into the evolution of the interaction itself. Central to the analysis of phytopathogen genomes, the identification and characterization of oomycete effector proteins has moved swiftly into the forefront in the field of plantpathogen interactions, due in large part to the aforementioned available genomic resources. At a primary level, the identification of a highly conserved N-terminal translocation motif (i.e., RXLR; Arg-X-Leu-Arg, where “X” is any amino acid) demonstrated to be necessary for effector delivery into host cells, has been a seminal discovery in the field of plant-oomycete interactions [11,12]. Similar in function to phytopathogenic bacterial effector proteins, oomycete RXLR-containing effectors have been demonstrated to suppress PTI [13], as well as to activate ETI [11,14,15,16,17,18]. Structurally, oomycete effector proteins display a modular organization, consisting of a N-terminal signal peptide, a conserved RXLR translocation motif, followed by a variable C-terminal effector domain [3]. It is the function and activity of the variable C-terminal effector domain that drives the activity of these molecules [3,4]. 35 Alternative splicing (AS) of pre-mRNA drives the generation of multiple protein isoforms through assembly of different combinations of splice sites within a single gene. In total, this process represents a conserved mechanism found in eukaryotes which drives proteome complexity within organisms with a finite number of genes [19]. In oomycetes, there are few reports of intron processing [20,21,22], and to date, these analyses has been strictly in silico [20,22], with little functional validation [21]. Costanzo et al. [21] characterized alternative processing in P. sojae family 5 endoglucanases revealing the generation of both coding and non-coding RNA isoforms. Additionally, based on their large-scale analysis of intronic structure and alternative splicing in P. sojae, Shen and colleagues [22] validated splice variants leading to premature translation termination. Ps. cubensis is an obligate biotrophic oomycete pathogen of cucurbits (i.e., cucumber, melon, squash, watermelon, etc.), and is the causal agent of cucurbit downy mildew, an economically important foliar disease [23]. Capable of rapid defoliation of fields in short periods of time (i.e., <10 days), Ps. cubensis is the primary factor limiting cucurbit production in the United States. Despite obvious economic importance, very little is known about the genetic determinants of virulence and pathogenicity of Ps. cubensis, as well as the molecular-genetic basis of resistance in the cucurbits. Similar to related oomycete pathogens of plants, Ps. cubensis possesses a suite of effector proteins that likely function to promote virulence and suppress host defense responses [3,24]. Recent work by Tian et al. [24] identified and characterized a preliminary set of effector proteins from a draft genome sequence of Ps. cubensis 36 obtained using 454 pyrosequencing. In brief, this set of 61 candidate effectors included a large class of variants with sequence similarity to the canonical RXLR motif found in other oomycete plant pathogens [24]. Specifically, this work characterized the function of a QXLR-containing effector, designated PcQNE, which was shown to be a member of a large family of Ps. cubensis QXLR nuclear-localized effectors, up-regulated during infection of cucumber. Additionally, internalization of PcQNE was shown to require the QXLR-EER motif, thereby establishing a basic homology to the well-characterized Phytophthora spp. effector proteins. In the current study, we describe the identification and evolutionary potential of the Ps. cubensis effector repertoire. First, through characterization of a RXLR effector protein, PscRXLR1, we investigated the localization and in planta activity, and similarly to some oomycete effector proteins described to date, PscRXLR1 induces a rapid cell death response when delivered into plant cells. Additionally, using whole transcriptome sequencing analyses, as well as RT-PCR, we show that PscRXLR1 is a product of alternative splicing of the Psc_781.4 gene which encodes a putative multi-drug transporter. Coupled with the induction and expression of PscRXLR1 mRNA during Ps. cubensis infection of cucumber, as well as a complement of bioinformatic, cell biology and in vivo analyses, we provide evidence suggestive of a virulence role for PscRXLR1. Finally, we used PscRXLR1 as template for assessing the conservation and evolutionary potential of oomycete effector proteins from Ps. cubensis, identifying and analyzing orthologous pairs of Ps. cubensis effector proteins and P. infestans noneffector proteins. Using more robust methods, we identified additional candidate 37 effectors from Ps. cubensis for these analyses and showed that, like other oomycete effectors, they tend to be influenced by positive selection. Assessment of evolutionary rate and conservation of secretion signals between orthologous pairs revealed that Ps. cubensis effectors are undergoing adaptive evolution and conservation of signal peptides are similar among effector and non-effector proteins in Ps. cubensis. Overall, our study provides support for the investigation of relationships among oomycete effectors and their non-effector orthologs, and in total, the analysis presented herein establishes a foundation for understanding the evolution of effector repertoires and host-pathogen specificity. RESULTS Genome sequencing of Ps. cubensis Next generation sequencing with the Illumina Genome Analyzer II platform was used to generate an assembly of the Ps. cubensis MSU-1 genome. A total of 4.5 Gb of cleaned paired end reads from two libraries were used to generate the assembly using Velvet, a de novo short read assembler [25]. The final assembly contains 35,546 contigs with an N50 contig size of 4.0 Kbp representing 64.4 Mbp. Protein coding genes in the draft assembly were annotated using MAKER [26] which incorporated ab initio gene predictions, protein evidence, and transcript evidence from other sequenced oomycete genomes. In total, 23,519 loci and 23,522 gene models were predicted. 38 Identification of the Ps. cubensis effector repertoire Our initial analysis of the effector complement of Ps. cubensis in an earlier draft assembly [24] identified 61 sequences containing the conserved RXLR, or novel QXLR, motif found in known oomycete effector proteins. This number is significantly less than the effector count predicted for other plant pathogenic oomycetes (i.e., 563 effectors in P. infestans, 396 in P. sojae, 374 in P. ramorum, and 134 in H. arabidopsidis; [7,10,27]), and is likely the result of limited coverage generated from an initial 454 pyrosequencing [24]. Generation of genomic sequences using the Illumina Genome Analyzer platform and their subsequent assembly generated a more comprehensive dataset. Using this resource, 269 additional sequences were identified as putative effector proteins. Interestingly, the putative Ps. cubensis effectors showed more variation at the R1 position of the RXLR motif than previously shown [24], with 18 amino acids predicted at the R1 position, in addition to R and Q (Table S2.1; Available at www.plosone.org, e34701). Moreover, we have evidence for expression for at least one predicted effector with any one of 19 amino acids (except Y, Tyr) at position R1, during the course of infection on a susceptible cucumber cultivar [28], supporting the hypothesis of an expanded translocation motif repertoire in Ps. cubensis. In total, including the previously characterized PcQNE, the current predicted effector complement of Ps. cubensis contains 271 members. 39 Nature of selection on Ps. cubensis paralogs Based on comparative genomic analyses of several oomycete plant pathogens, positive selection has been postulated to act disproportionately on effectors gene[s] compared to other genes in the genome [27,29]. To this end, we examined the strength of selection acting upon the predicted effector complement of Ps. cubensis by estimating ω, the ratio of the non-synonymous substitution rate (Ka) to the synonymous substitution rate (Ks). Among all Ps. cubensis effector paralogs, the median ω is 0.54, which is significantly higher than that of Ps. cubensis paralogous genes in general (ω = 0.24, Wilcox Rank Sum Test, p < 2.2e-16). For comparison, we also examined P. infestans effectors and arrived at the same conclusion (Wilcox Rank Sum Test, p < 4.0e-14). Because more recent duplicates tend to have elevated ω, we examined if the higher ω values among effector paralogs can be attributed to recent duplication. We found that, using Ks as a proxy of time, the ω values for effector pairs are in general significantly higher than other paralogs in Ps. cubensis in a Ks bin (Figure 2.1A). Thus, the elevated ω values among effectors are not exclusively due to relaxation of selection among recent duplicates. The results for P. infestans effectors are similar (Figure 2.1B), although the ω values of P. infestans are in general higher than those in Ps. cubensis. Taken together, Ps. cubensis effectors either have experienced a significantly lower degree of selective constraints, or tend to be positively selected. Consistent with the latter, 6.3% of Ps. cubensis effector paralog pairs have ω > 1, compared to 3.2% of all other paralogous gene pairs. In parallel to our observations in Ps. cubensis, 4.6% of P. 40 Figure 2.1 41 Figure 2.1 (cont'd) Strength of purifying selection on Pseudoperonospora cubensis effectors. Comparison of selective constraints on effector paralogs (blue) and all other genes (red) in (A) Ps. cubensis and (B) P. infestans. Frequency distributions (C) of ω, the ratio between the nonsynonymous substitution rate (Ka) and the synonymous substitution rate (Ks) of Ps. cubensis and P. infestans sequence pairs. Distributions of Ks values (D) of Ps. cubensis and P. infestans sequence pairs. Green symbols indicate Ps. cubensis effector – P. infestans non-effector ortholog pairs. Blue symbols represent other orthologous gene pairs between the two species. infestans effector paralog pairs have ω > 1, compared to 3.2% for all other paralogs. Although there is no clear evidence suggesting that most effectors are subjected to positive selection, it is interesting that even among relatively ancient effector duplicates, the rate of evolution among effectors is significantly higher than most genes. Given that older duplicates that survive for tens to hundreds of millions of years tend to be subjected to substantially stronger selective constraints than young duplicates [30], this would suggest that, perhaps, effectors function in a way that do not require as strong a constraint on their primary sequence. Alternatively, it is also possible that pathogen effectors, even those having undergone ancient duplication events, experience some degree of continuous positive selection. Relationship between Ps. cubensis effectors and their P. infestans orthologs Subsequent in silico analysis of candidate Ps. cubensis effectors and comparisons to annotated genes in P. infestans revealed that there were a number of orthologs between Ps. cubensis effector proteins and both effector and non-effector genes in P. infestans. We hypothesized that this scenario (i.e., effector with non-effector ortholog) 42 may provide a foundation for the analysis of the evolution of effectors from non-effector proteins. Therefore, we identified orthologous pairs of predicted Ps. cubensis effector proteins and their non-effector counterparts in P. infestans considering sequence similarity and synteny (see Materials and Methods). With this approach, 11,601 orthologous gene pairs were identified between Ps. cubensis and P. infestans for comparison. Of 271 Ps. cubensis effector sequences, 109 had a predicted ortholog in P. infestans (TABLE S2.2; Available at www.plosone.org, e34701). As shown in Figure 2.1C, the Ps. cubensis effector P. infestans (PscE-Pi) ortholog pairs have significantly higher ω values as compared to the baseline pairs (Kolmogorov-Smirnov test, p < 7.9e06), consistent with what was found with the effector paralogs (Figure 2.1A). Additionally, the distribution of ω for the PscE-Pi pairs appears multi-modal. Given that the first effector ortholog peak (at ω ~0.15) is mostly overlapping with that of the other orthologs, these effector paralogs are more highly conserved. The second peak at ω ~0.3 likely indicates the presence of a group of effectors that are more quickly evolving (Figure 2.1C). However, we cannot rule out the possibility that these peaks are present simply due to the small effector ortholog sample size. To determine if the overall higher ω value among effector orthologs is an artifact due to mis-identification of orthologous genes, we examined if putative effectors, as well as the other orthologs, have similar "age". As shown in Figure 2.1D, the distributions of Ks values for the effector and the other orthologs are highly similar and are statistically indistinguishable. Thus, mis-identified orthologous pairs likely do not significantly impact our findings. 43 Signal peptide conservation among ortholog pairs Signal peptides are essential components of oomycete effector proteins, as they are required for translocation of the protein from the pathogen haustorium to the extrahaustorial matrix prior to uptake by the host cell membrane [3]. As such, all 109 of the Ps. cubensis effector sequences in the PscE-Pi dataset are predicted to have signal peptides (Figure S2.1). However, only 71 (65%) of the corresponding P. infestans orthologs were predicted to be secreted proteins. For comparison, predictions of signal peptides were made for 10,383 of the 11,601 ortholog pairs. Of these, there were 688 (6.63%; Psc-sec/Pi-sec) ortholog pairs where both members were predicted to have signal peptides, 428 (4.12%; Psc-sec/Pi-non) pairs where the Ps. cubensis protein was predicted to be secreted and the P. infestans ortholog was not, and 622 (5.99%; Pscnon/Pi-sec) where the Ps. cubensis sequence did not have a predicted signal peptide and its corresponding P. infestans sequence was predicted to be secreted. Additionally, there were 8,645 (83.3%; Psc-non/Pi-non) ortholog pairs where neither member was predicted to be secreted. For statistical analysis, the Psc-sec/Pi-sec and Psc-sec/Pi-non datasets from the Ps. cubensis effector-P. infestans orthologs were compared to their respective genome-wide datasets. Using the Fisher’s exact test, no significant difference (p = 0.5354) was identified between the two datasets, indicating that presence of signal peptide prediction is not a good indicator of potential selection for effector peptide evolution. 44 Identification of Pseudoperonospora cubensis effector PscRXLR1 Using the RXLR effector identification pipeline [29], we previously identified 61 candidate effector protein sequences from a draft genome assembly of Ps. cubensis [24]. Initial analysis of these sequences using the Basic Alignment Analysis Search Tool (BLAST) against the proteome of P. infestans revealed that only 7 of these sequences aligned with annotated proteins within the P. infestans genome database; moreover, only 1 of these was predicted to be a secreted RXLR effector. Of these sequences, one (contig01871_F1) had 75% amino acid identity to P. infestans protein PITG_17484, a putative member of the drug/metabolite transporter (DMT) superfamily (CLO184; 2.2). Additional cloning via 3' RACE PCR and subsequent analysis revealed that the Ps. cubensis coding sequence, hereafter referred to as Ps. cubensis RXLR protein 1 (PscRXLR1), appeared significantly shorter (i.e., 132 amino acids), compared to its P. infestans ortholog PITG_1784 (i.e., 332 amino acids). This apparent truncation in PscRXLR1 results in a protein coding sequence lacking the EamA functional domain (PF00892; formerly called DUF6) found in members of the DMT family [31]. SignalP analysis of PscRxLR1 identified a putative 26 amino acid signal peptide at the N-terminus of the protein (Figure 2.2). Based on the conserved features and domain organization of oomycete effectors, the presence of a canonical RXLR motif was identified at amino acid position 70 (Figure 2.2). However, unlike several previously characterized oomycete effector proteins, PscRXLR1 does not contain an EER motif, which has also been implicated in oomycete effector translocation into the host cell 45 [12,24,27]. The P. infestans ortholog, PITG_17848, while not having a predicted signal peptide, does contain an RXLR-like motif (i.e., RFMR; Figure 2.2A). To eliminate the possibility that PITG_17484 was mis-annotated and did in fact contain a signal peptide upstream of the predicted coding sequence, the region 500 bp upstream of the annotated PITG_17484 sequence was examined and a canonical signal peptide sequence was not identified. We therefore concluded that PITG_17484 is not an RXLR effector protein. The absence of a predicted signal peptide in PITG_17848 suggests that PscRXLR1 may have evolved this function independently. To address this possibility, and to further explore the ancestral function of these proteins, orthologous PscRXLR1 sequences in additional plant oomycete pathogen species were identified, including P. sojae, P. ramorum, and Py. ultimum. Not surprisingly, the sequences from P. sojae and P. ramorum were more similar to PscRXLR1 than those from Py. ultimum (76% and 72%, respectively, compared to 59%; Figure S2.2). Additionally, while none of these orthologs had predicted signal peptides, they did contain EamA functional domains, indicating that they were also members of the DMT superfamily (Figure S2.2). Phylogenetic analysis to infer evolutionary relationships between PscRXLR1 and orthologs from P. sojae, P. ramorum, and Py. ultimum supported these observations (Figure 2.2C). 46 Figure 2.2 47 Figure 2.2 (cont'd) Figure 2.2 PscRXLR1 encodes a RXLR-containing effector protein with homology to a non-effector protein in Phytophthora infestans. (A) Schematic representations of PscRXLR1, Psc_781.4 and PITG_17484 from P. infestans. (B) Multiple sequence alignment of PscRXLR1, Psc_781.4 and PITG_174984. Protein sequences were aligned using CLUSTALW and shaded for consensus using BOXSHADE. The asterisk indicates a stop codon. (C) The full length protein sequences of PscRXLR1, Psc781.4 and their orthologs from P. infestans (PITG_17484), Phytophthora ramorum (P. ramorum 96701 0 3435), Phytophthora sojae (P. sojae 156165) and Pythium ultimum (PYU_T005955) were aligned using Muscle and evolutionary history was inferred by using the Maximum Likelihood method based on the JTT matrix-based model [1] using MEGA5[2]. 500 bootstrap runs were performed. Functional validation of the PscRXLR1 signal peptide A primary characteristic of oomycete effector proteins is signal peptide-mediated secretion from the haustoria into the extrahaustorial matrix (EHM) prior to translocation into the host cell [3]. PscRXLR1 contains a 26 amino acid signal peptide as predicted by SignalP 3.0 (HMM Probability, 0.966), whereas its closest P. infestans ortholog, PITG_17484, does not have a predicted signal peptide. To determine if the predicted 48 signal peptide from PscRXLR1 is functional, we used a yeast signal trap assay based on the requirement of invertase secretion for yeast growth on media with raffinose as the sole carbon source [32]. This assay has been used previously to confirm predicted signal peptide sequences in candidate effector proteins from both P. infestans and Ps. cubensis [24,33]. As shown in Figure 2.3A, pSUC2-PscRXLR1 (column 4) is able to grow on medium containing only raffinose (YPRAA), indicating that the signal peptide of PscRXLR1 is sufficient for secretion of invertase. As a second confirmation of signal peptide function, we monitored the reduction of 2,3,5-triphenyltetrazolium chloride (TTC) to the red-colored compound triphenylformazan. Again, pSUC2-PscRXLR1 (column 4) was confirmed as having a functional signal peptide. In contrast, neither the control yeast strains (i.e., YPK12 [column 1] or pSUC2 [column 2]), nor the pSUC2PITG_17484 (column 3) yeast strain containing a PITG_17484-invertase fusion construct, grew on YPRAA, nor were they able to reduce TTC. Our positive control, PcQNE-SP1 (5) was, as shown previously [24], both able to grow on YPRAA medium and reduce TTC. These data support the annotation of PscRXLR1 as a secreted RXLR effector protein and confirm that its P. infestans ortholog is a non-secreted protein. PscRXLR1 and PITG_1784 localize to the plant plasma membrane In planta localization of effector proteins has been successfully used to guide functional analysis and to infer the function itself [34,35,36,37]. To identify a possible role for PscRXLR1 in the pathogenicity of Ps. cubensis, and to provide clues as to its putative role in planta, we investigated the localization of PscRXLR1. To this end, a C-terminal 49 Figure 2.3 50 Figure 2.3 (cont'd) Functional characterization of PscRXLR1 and PITG_17484. (A) PscRXLR1 has a functional signal peptide validated by the yeast signal trap assay. Yeast strains were grown on media with raffinose (YPRAA) as the sole carbon source or in the presence of 2,3,5-triphenyltetrazolium chloride (TTC). Yeast strains YTK12 and YTK12:pSUC2 EV both lack a functional invertase gene and cannot grow on YPRAA medium or reduce TTC to red formazan. The functional signal peptide of PscRXLR1, when fused in-frame to the mature yeast invertase (pSUC2-PscRXLR1), allows for secretion of invertase, resulting in growth on YPRAA medium, as well as reduction of TTC to red formazan. PITG_17484, as predicted, does not have a functional signal peptide (pSUC2-PITG_17484). (B) Both PscRXLR1_CFP (top row) and PITG_17484_CFP (bottom row) co-localize with a plasma membrane-specific AtPIP2A_YFP marker. Left panels: C-CFP fusion protein only. Center panels: AtPIP2A_YFP. Right panels: Merge of CFP and YFP images. Scale bar = 10 µm. (C, D) Heterologous expression of PscRXLR1 induces chlorosis and necrosis in Nicotiana benthamiana. Agrobacterium tumefaciens C58-C1 expressing PscRXLR1_CFP, PITG_17484_CFP, or an AtPIP2A-YFP construct were infiltrated into N. benthamiana. The chlorosis and necrosis phenotype of PscRXLR1_CFP infiltrated leaves (C) was photographed at 2, 3, and 7 days post-inoculation (dpi). Leaf areas infiltrated with either PITG_17484_CFP or AtPIP2A-YFP (D) as designated by the dash line circles did not shown any phenotype at 3 dpi. CFP-fusion protein (i.e., PscRXLR1-CFP) was transiently expressed in Nicotiana benthamiana, and protein localization visualized using laser scanning confocal microscopy. Based on the similarity of PscRXLR1 with members of the DMT superfamily in Phytophthora spp. and Py. ultimum, PscRXLR1 was predicted to localize to the plant plasma membrane, despite the absence of a predicted transmembrane domain when analyzed using InterProScan. To confirm this, PscRXLR1-CFP was coexpressed with a YFP-tagged construct encoding the aquaporin gene AtPIP2A, a marker for plasma membrane localization [38]. As predicted, PscRXLR1-CFP colocalized with AtPIP2A-YFP (Figure 2.3B), confirming that PscRXLR1 localizes to the plasma membrane in planta. Additionally, a C-terminal CFP fusion was made to the P. 51 infestans ortholog PITG_17484, which was also confirmed to be plasma membrane localized (Figure 2.3B). PscRXLR1, but not its P. infestans ortholog, elicits a rapid cell death response when expressed in Nicotiana benthamiana The obligate nature of a plant pathogen often presents challenges towards functional characterization of virulence and susceptibility within their respective host(s). To circumvent this limitation, transient heterologous systems have been developed and have proved valuable in their functional characterization [34,39,40,41,42]. To investigate the activity of PscRXLR1 in planta, we utilized heterologous expression in N. benthamiana as means to characterize and determine the function of PscRXLR1. As shown in Figure 2.3C, expression of PscRxLR1 resulted in leaf chlorosis throughout the entire infiltration zone by 2 dpi, followed by browning and initiation of necrosis at 4 dpi. By 7 dpi, the zone of infiltration was completely dehydrated. In comparison, neither infiltration with PITG_17484, nor pMDGFP, resulted in any detectable cell death-type phenotype in N. benthamiana leaves at 4 dpi (Figure 2.3D). PscRXLR1 expression is induced during Ps. cubensis infection of cucumber The function of pathogen effector molecules is to enhance the virulence of the pathogen during its lifecycle, as well as to dampen host defense responses activated during infection. In this regard, the temporal expression of effector molecules during infection 52 and pathogen development often signals critical stages in the host-pathogen interaction. Expression of PscRXLR1 mRNA was measured using quantitative reverse transcription (qRT)-PCR following infection of Ps. cubensis on the susceptible cucumber cultivar ‘Vlaspik’. As shown in Figure 2.4, expression of PscRXLR1 was significantly (p < 0.001) induced during infection, beginning at 1 dpi and continuing through 4 dpi, as compared to the basal expression level in sporangia. Induction of gene expression at 1 dpi corresponds with zoospore encystment in the stomata, the first stage of pathogen entry into the host (Figure 2.4B, left panel). Subsequent expression observed through 4 dpi corresponds with hyphal penetration through the stomata, growth throughout the mesophyll, and formation of haustoria (Figure 2.4B, center and right panels). This pattern of expression supports a potential role for PscRXLR1 in initial establishment of the infection possibly through dampening host defense responses. Additionally, this pattern is consistent with the expression patterns observed in other oomycete plant pathogen effectors, further supporting the prediction of PscRXLR1 as an effector protein with a role in infection and disease development. PscRXLR1 is a splice variant of Psc_781.4 Automated annotation of the Illumina-generated Ps. cubensis assembly described in this study resulted in Psc_781.4, a gene model at the PscRXLR1 locus that more closely mirrored PITG_17484 than our prediction for PscRXLR1 and what was obtained through molecular cloning (Figure 2.5), with the primary difference between the two predictions being that intron 1 is either spliced in Psc_781.4, or retained in PscRXLR1 53 Figure 2.4 Figure 2.4 PscRXLR1 mRNA expression is up-regulated during Pseudoperonospora cubensis infection of cucumber. (A) Expression levels of PscRXLR1 in sporangia (SP) and at 1, 2, and 4 days post-inoculation (dpi). Expression is displayed as fold sporangia expression and all time points are significantly different compared to SP control (*** indicates p < 0.001) using Two-way ANOVA. Error bars represent the standard error of the mean of 2 technical replicates from each of 2 biological replicates. (B) Differential interference (DIC) microscopy images of stages of Ps. cubensis infection on cucumber where e = encysted zoospore; a = appressorium; and h = haustorium. Scale bars = 25 µm. 54 (Figure 2.5A). Empirical whole transcriptome sequence data (RNA-seq) from Ps. cubensis sporangia (unpublished results) provides support for both isoforms at this locus. When the first intron is retained, a stop codon is also brought into frame, yielding a truncated transcript (i.e., putative effector PscRXLR1), and subsequently, a smaller protein, which, as described in Figure 2.2, lacks the EamA functional domain (Figure 2.5A). Based on our in silico predictions we confirmed which gene model, or both, was represented in vivo. Using an RT-PCR-based approach, we were able to amplify both splice variants from purified sporangia (SP), as well as from infected leaf material harvested at 4 and 8 days post-inoculation (dpi) (Figure 2.5B), suggesting that both isoforms are present throughout the infection process. As an added control, the resultant PCR products were cloned and sequenced to confirm that they corresponded to the appropriate splice variant (Figure S2.4). Additional functional analysis of Psc_781.4 confirmed transient expression in N. benthamiana does not elicit a cell death response in planta, indicating that it likely has no virulence function in Ps. cubensis (Figure S2.5). In total, these independent methods confirm our conclusion that Psc_781.4 is alternatively spliced leading to generation of a functional RXLR effector protein. 55 Figure 2.5 Figure 2.5 PscRXLR1 is a splice variant of Psc_781.4. (A) Schematic representation of intron and exon structures of PITG_17484, Psc_781.4, and PscRXLR1. (B) RT-PCR analysis of alternative splicing in Psc_781.4 in sporangia (SP) and at 4 and 8 days post-inoculation (dpi). RT-PCR products were subcloned into the pGEM vector and DNA was amplified by PCR using intron spanning primers as shown in (A). S, transcript with spliced intron 1. NS, non-spliced transcript. 56 DISCUSSION In this study, we describe a candidate RXLR-type effector from Ps. cubensis that results from a splice variant of a putative multi-drug transporter protein, and additionally expands the scope of our understanding of the function and evolutionary history of the Ps. cubensis effector repertoire. While Ps. cubensis is an oomycete pathogen of worldwide economic importance, insight into the mechanism(s) underlying its virulence and pathogenicity remain limited [43]. A recent study has provided a foundation for investigating the genetic basis for virulence and pathogenicity in Ps. cubensis through generation of a large scale genomic dataset [24]. We build upon this previous work using a combination of in silico analyses, gene expression studies, and cell biology to functionally characterize PscRXLR1 and establish a potential role in promoting Ps. cubensis infection and proliferation. Alternative splicing has been previously described in oomycetes pathogens of plants; specifically related to the family 5 endoglucanases (EGL5) from P. sojae [21], as well as in gene families such as Crinklers (CRNs), protein kinases, and transcription factors [22]. In P. sojae, EGL5 proteins have a role in infection of soybean and are highly upregulated during the early stages of infection. As part of these analyses, four different mechanisms of alternative splicing were described: intron skip, exon skip, alternative donor site, and alternative acceptor site, with intron skip, where the intron is retained, being the most commonly observed mechanism [22]. In agreement with this previous observation, we propose that the Psc_RXLR1 transcript is generated via a retained 57 intron from Psc_781.4, yielding an RXLR-type effector. From an evolutionary standpoint, alternative-splicing functions to expand the capacity of an organism's proteome, thus enabling the generation of multiple functional isoforms from a single coding sequence. Over time, new isoforms may be maintained if they have a beneficial function [44], or lost, if their function is not critical to the lifecycle of the organism. In the case of plant pathogens, this process could potentially serve an adaptive role to allow for generation of isoforms of "housekeeping" type genes that gain new function(s), potentially to promote virulence or infection. Alternatively, the PscRXLR1 splice variant could represent a step in evolutionary time, as it is generated from the same coding sequence as Psc_781.4 and maintained in the coding repertoire, but has not been duplicated or retained as a separate sequence. Like other oomycete effectors characterized to date, PscRXLR1 has a functional 26 amino acid signal peptide necessary for secretion from the haustorium into the extrahaustorial matrix prior to translocation into the host cytoplasm. Interestingly, PscRXLR1 is also localized to the host plasma membrane, despite the lack of a predicted transmembrane domain or localization signal (Figure 2.3). To further examine this, we surveyed the genomes of additional oomycete plant pathogens for orthologous sequences. We hypothesize that if functional characterization data for orthologs in any of these better characterized species (e.g., P. infestans, P. ramorum and P. sojae) existed, it might provide insight into both the function and conservation of PscRXLR1. Through BLAST analysis of the P. infestans, P. sojae, P. ramorum and Py. ultimum genomes, we identified orthologous sequences in each of the pathogens, all of which 58 were annotated as members of the Drug/Metabolite Transporter (DMT) superfamily [31]. The DMT superfamily encompasses 19 families; the orthologs described here are members of the EamA family, named for the O-acetylserine/cysteine export gene in Escherichia coli [45]. While PscRXLR1 is lacking the EamA transmembrane domains that are characteristic of these transporter proteins, our data clearly demonstrated plasma membrane localization. Monitoring the expression of both pathogen and host genes during infection can provide insight into the interplay between resistance and susceptibility [46]. Using qRT-PCR, we demonstrated that expression of PscRXLR1 mRNA is up-regulated during the early stages of infection of cucumber. While we were unable to distinguish between isoforms, expression was induced nearly 10-fold at 1, 2 and 4 days post-inoculation (dpi), corresponding with zoospore encystment, appressorium formation and penetration, and proliferation and haustoria establishment, respectively (Figure 2.5). Several effector proteins from P. infestans have also been demonstrated to have distinct temporal patterns of expression, and are often expressed during the pre- and early stages of infection, representative of the biotrophic phase of the P. infestans life cycle [12]. Based on the robust induction of PscRXLR1 mRNA during the early stages of infection, as well as the aggressive nature of the necrosis-inducing activity observed in N. benthamiana, we hypothesize that the expression pattern of PscRXLR1 could support a role in the initial infection process, possibly through dampening of host defense responses. Indeed, effectors from other oomycete plant pathogens, including PcQNE from Ps. cubensis and members of the CRN family from P. infestans, have also been shown to 59 elicit similar phenotypes when transiently expressed in N. benthamiana [3,24,47], supporting the classification of PscRXLR1 as an effector protein with a putative role in virulence. In the current study, ongoing analysis of the Ps. cubensis genome has expanded the candidate effector complement of Ps. cubensis to 271 sequences, revealing significant variation in the conserved translocation motif. While previous analyses revealed a near equal distribution of RXLR:QXLR motifs in Ps. cubensis, our current work, based on a higher coverage draft genome sequence and predicted protein sequences rather than open reading frames (ORFs), predicts sequences with 20 different amino acid possibilities at the R1 position of the XXLR motif. Of these 20 predicted R1 substitutions, 19 have expression support (Table S2.1). While all 20 R1 substitutions have yet to be functionally validated, it is not surprising that Ps. cubensis effectors may in fact utilize a more diverse set of translocation motifs compared to the Phytophthora spp., given its obligate lifestyle and relatively narrow host range. Among Phytophthora spp., the conservation of the RXLR motif is well-established, yet there are additional classes of oomycete effectors, such as the CRN family, that appear to utilize disparate translocation motifs [34,47]. Moreover, analysis of the Py. ultimum genome has identified an additional predicted translocation motif, YxSL[RK] [9]. Indeed, divergence of transport signal sequences is even more pronounced between oomycetes and the true fungi, which have no obvious conserved motifs that could function in transport and show high degrees of variation even within the same species [48]. For example, the effectors AvrM and AvrL567 from Melampsora lini, an obligate rust fungi with a similar 60 lifestyle to Ps. cubensis, rely on unique N-terminal sequences for uptake [48]. These sequences, while different in regards to sequence similarity from the RXLR motif observed in Phytophthora spp., are similar in that they feature positively charged residues, implying that secondary protein structure may be a factor contributing to uptake of these proteins. Both M. lini and Ps. cubensis are obligate biotrophs with specific host ranges, which may have influenced the evolution of their effector repertoires to select for unique translocation motifs compared to those found in Phytophthora spp. Preliminary analysis of the Ps. cubensis effector repertoire reveals minimal orthology with annotated effector proteins from P. infestans, similar to what has been observed when comparing the effector complement from P. infestans with P. ramorum, P. sojae or H. arabidopsidis [8,27,29]. Through extensive analysis using both evolutionary and comparative genomics, Phytophthora RXLR effector genes have been shown to be undergoing accelerated rates of birth and death evolution as well as both widespread gene duplication and loss events [7,8]. As such, among the predicted RXLR effector genes from P. infestans, P. sojae, and P. ramorum, there are few genes with high degrees of sequence similarity or 1:1:1 orthology [7]. Similarly, the same phenomenon was observed in comparing candidate effector sequences from Ps. cubensis to those of P. infestans. Of 271 predicted Ps. cubensis sequences, less than half (41%) of these had significant similarity (e-value < 1e-5) to predicted P. infestans proteins, and only 3 of these sequences had similarity to annotated P. infestans RXLR effector proteins. These results indicate that the effector repertoire Ps. cubensis utilizes to promote its virulence 61 and pathogenicity on its hosts is, as could be predicted, very different than that utilized by P. infestans, and likely the other Phytophthora spp. as well. This is likely due to differing selective pressures on Ps. cubensis resulting from host specificity as well as differences in lifestyle between the two pathogens (i.e., obligate vs. non-obligate). In this study, we identified minimal conservation between the predicted Ps. cubensis effector complement and effectors from P. infestans. We hypothesize that the identification and analysis of effector to non-effector relationships among oomycete plant pathogens is a valid measure to assess conservation and rates of evolution. Additionally, with the identification of PscRXLR1, a splice variant of a non-effector gene, we posit that these types of analyses as well as a more thorough analysis of alternative splicing may provide a preliminary baseline to not only investigate evolutionary differences among oomycete plant pathogens, but to also infer the relationship between effector repertoire and the host-pathogen specificity and lifestyle. We have used several criteria (i.e., prediction of selection pressure, secretion, etc.) to identify and analyze the relationship between predicted Ps. cubensis effectors and their orthologs in P. infestans. As observed for other effector proteins, some Ps. cubensis effectors may have experienced stronger positive selection than most other proteins within the genome. Interestingly, in addition to varying significantly from the genome average, the distribution of ω for the PscE-Pi pairs has two distinct peaks, representing groups of effectors under different levels of selection pressure. Thus, it appears that aside from acting as effectors during infection, some of these slower evolving genes may have additional, "housekeeping" roles that are yet to be uncovered. Despite computational 62 evidence indicating that these slower evolving genes are likely effectors, their role(s) in pathogenesis remain to be established. MATERIALS AND METHODS Ps. cubensis culture and growth conditions Ps. cubensis was maintained on Cucumis sativus cv. ‘Vlaspik’ as previously described [24]. Cucumber plants were grown at 22 °C with a 12 h light/dark photoperiod. For Ps. cubensis inoculation, sporangia were collected from heavily sporulating leaves by washing with cold sterile distilled water and collecting sporangia in a centrifuge tube. Sporangia were enumerated with a hemocytometer and suspended to a concentration of 1 x 105 sporangia/ml in sterile distilled water. The underside of fully expanded 2 nd or 3rd true leaves of 4-week-old cucumber plants were spray-inoculated, until run-off, with the suspension, and incubated for 24 h at 100% humidity in the dark. After 24 h, inoculated plants were moved to a growth chamber (22 °C with a 12 h light/dark photoperiod). DNA and RNA extraction Genomic DNA of Ps. cubensis was isolated from sporangia of isolate MSU-1 using the DNeasy Plant Mini kit (Qiagen, Germantown, MD) with slight modifications. Sporangia were washed from heavily sporulating leaves with sterile distilled water and filtered 63 through a 40 m nylon cell strainer to remove residual plant debris. The resultant sporangia suspension was centrifuged, and the supernatant decanted. Sporangia were suspended in buffer AP1 containing RNase and 5 l of Proteinase K and incubated at 37 °C for 20 min. 50 l of 425-600 µm acid washed beads were added to the sporangia suspension and sporangia disrupted for 3 min using a vortex. Subsequent DNA extraction steps were performed according to manufacturer’s instructions. Ps. cubensis total RNA was isolated as follows: sporangia were collected as described above for DNA isolation, yet re-suspended in buffer RLT from the RNeasy Plant Mini Kit (Qiagen, Germantown, MD) and disrupted as above. RNA isolation was performed according to the manufacturer’s instructions. RNA samples were treated with DNase (Promega, Madison, WI) prior to use. Sequence, assembly, and annotation of the Ps. cubensis genome Genomic DNA was isolated from Ps. cubensis MSU-1 as described above and libraries constructed using the Illumina Genomic DNA Sample kit (Illumina, San Diego, CA). Two separate paired-end libraries were end sequenced using an Illumina Genome Analyzer II (Illumina, San Diego, CA) at the UC-Davis Genome Center. The first library was sequenced with 84bp reads and an insert size of 180bp yielding 7.8 Gbp of sequence. The second library was sequenced with 100bp reads and an insert size of 480bp yielding 5.5 Gbp of sequence. Illumina reads were trimmed to 51 bp to remove low quality regions at the 3’ end of the reads. Reads with more than one N base or a base 64 with a quality score less than 20 were removed. The reads were then searched against the Cucumber genome assembly [49] with Bowtie v0.12.7 [50] and matching reads were removed; 4.5 Gbp of sequence was retained following trimming and cleaning the reads. The trimmed and cleaned reads were assembled using Velvet v1.0.14 [25]. Three Velvet runs were performed with hash lengths of 31, 41, and 51 and coverage cutoffs of 7, 3.6, and 2, respectively. A minimum contig size cutoff of 200bp was used for all the assembly runs. The contigs from each Velvet run were then merged into one assembly using the Minimus2 (http://sourceforge.net/apps/mediawiki/amos/index.php?title=Minimus2). pipeline Contaminant- containing and mitochondrial contigs were removed; the final assembly contains 35,546 contigs with an N50 contig size of 4.0 Kbp; the total assembly is 64.4 Mbp. Reads were deposited in the Sequence Read Archive at the National Center for Biotechnology Information under study number SRP011018. The assembly is available at NCBI under accession AHJF00000000 and via http://www.daylab.plp.msu.edu/wp- content/uploads/psc_merged_contigs.fasta.zip. The annotation can be downloaded at http://www.daylab.plp.msu.edu/wp-content/uploads/psc_merged_contigs.gff3.zip. The assembly was annotated using the MAKER [26] annotation pipeline. The FGENESH gene finder [51] was used with the Phytophthora matrix to produce the initial gene calls for the pipeline. All transcript and protein sequences from sequenced oomycete genomes were provided to MAKER to improve the quality of the annotation. In total, 23,519 loci and 23,522 gene models were predicted. Putative functional 65 annotation was assigned by searching the gene models against UniRef100 using BLASTX (cutoff: 1E-5) and transferring the first hit with informative annotation. Identification and cloning of PscRXLR1, Psc_781.4 and PITG_17484 Amplification of the coding sequence of PscRXLR1 was performed using DNA primers that correspond to the open reading frame of PscRxLR1 (Figure 2.5). Subsequent isolation and cloning of PscRxLR1 was performed by PCR using gene-specific primers and genomic DNA from Ps. cubensis sporangia. Resultant amplicons were cloned into the TA cloning vector pGEM-T-Easy (Promega), generating pGEM_PscRXLR1. To ensure identification of a complete coding sequence, as well as to verify the absence of introns in the sequence, 3 RACE (Rapid Amplification of cDNA Ends) was performed using the SMARTer RACE cDNA Amplification Kit (Clontech, Mountain View, CA). Amplification of Psc_781.4 was performed using 3' RACE as described above and the final coding sequence was amplified using gene specific primers (Figure 2.5). Fidelity of all sequences was confirmed by DNA sequencing using the ABI 3730 Genetic Analyzer (Applied Biosystems, Foster City, CA). P. infestans clone PITG_17484 was amplified from cDNA of P. infestans by PCR using gene-specific primers. Amplicons were subcloned into pGEM-T-Easy and sequences confirmed by sequencing. 66 DNA cloning and construct preparation To validate the predicted signal peptide (SP) of PscRXLR1, the yeast signal trap assay was used [24,32,33]. A DNA fragment corresponding to the predicted 26 amino acid signal peptide (including start codon) was amplified by PCR using gene specific primers (Table S3) modified to include 5 EcoR1 and 3 XhoI restriction sites. Resultant amplicons were cloned into TA cloning vector pGEM-T-Easy (Promega), to yield pGEMPscRXLR1-SP. The plasmid pGEM-PscRXLR1-SP was digested with EcoR1 and XhoI, the 84 bp SP fragment was gel purified, and ligated into the EcoR1/XhoI sites of the yeast signal trap vector pSUC2T7M13ROI [32], generating pSUC2-PscRXLR1. The SP sequence of PITG_17484 was similarly amplified and cloned, generating pSUC2-PITG_17484. Plasmids pSUC2T7M13ROI, pSUC2-PscRXLR1 and pSUC2PITG_17484 were transformed into the yeast SUC2 minus strain YTK12 using the Frozen-EZ Yeast Transformation IITM kit (Zymo Research, Orange, CA) following the manufacturer’s instructions. Transformants were selected on CMD-W plates (0.67% yeast nitrogen base without amino acids, 0.075% tryptophan dropout supplement, 2% sucrose, 0.1% glucose, and 2% agar) for 3 days at 28 C. To confirm signal peptide function, YTK12 containing the pSUC2 constructs were grown on raffinose-containing YPRAA plates (1% yeast extract, 2% peptone, 2% raffinose, 2 g/L Antimycin A, and 2% agar). Yeast strains were replicated onto YPDA plates (1% yeast, 2% peptone, 2% glucose, 0.003% adenine hemisulfate, and 2% agar) and CMD-W plates as equal viability controls. TYK12 without pSUC2 was used as a negative control. The detection 67 of the secreted invertase activity with 2,3,5-triphenyltetrazolium chloride (TTC) was performed as described by Tian et al. [24]. For construction of plasmids used for localization and phenotype studies, the open reading frames of PscRXLR1, Psc_781.4 and PITG_17484 (minus stop codons) were amplified and cloned into the Gateway entry vector pENTR/D-TOPO (Invitrogen, Carlsbad, CA), yielding pENTR-PscRXLR1-GW, pENTR-Psc_781.4-GW, and pENTRPITG_17484-GW, respectively. The destination vector pVKH18En6gw-cCFP [24] was used to create the C-terminal CFP fusions, using heterologous recombination via LR Clonase, as per the manufacturer’s instructions (Invitrogen). Transient expression and localization in N. benthamiana Infiltration and transient expression in N. benthamiana using A. tumefaciens was performed on 4-6 week old plants as described in Tian et al. [24]. A. tumefaciens strains were grown overnight at 28 °C on Luria-Bertani (LB) plates containing 50 µg/mL rifampicin and 25 g/mL kanamycin. A. tumefaciens clones were re-suspended in induction buffer (10 mM MES, pH 5.6, 10 mM MgCl2, 150 m acetosyringone) and incubated at room temperature, shaking in the dark, for 2 hours prior to infiltration. A. tumefaciens suspensions were infiltrated at a final concentration of OD600 = 0.8. A. tumefaciens-mediated transient expression in N. benthamiana for localization of PscRXLR1-CFP, Psc_781.4-CFP, and PITG_17484-CFP with AtPIP2A-YFP [38] was 68 performed as described above. Visualization of fluorescently tagged proteins was observed using an Olympus Fluoview 1000 laser scanning confocal microscope. Images were adjusted for contrast in Canvas X (ACD Systems). Quantitative real time PCR First strand cDNA was synthesized from 1 g total RNA using the first-strand cDNA synthesis kit (USB, Cleveland, OH). Quantitative RT-PCR was performed using a Mastercycler ep Realplex real-time PCR (Eppendorf, Hamburg, Germany) using HotStart SYBR Green qPCR Master Mix (2x) (USB), as previously described [52]. For amplification of PscRXLR1 transcripts, gene specific primers were designed to amplify a 50 bp fragment (Forward: 5'-TGCGTAGCATCGCCAACCGA-3 and Reverse: 5'TCTTGCCAGCTGCATCGCGA-3'). Primers specific for the Ps. cubensis internal transcribed spacer (ITS) region were used as an endogenous control (Table S3). Cycle parameters were as follows: 95 °C for 2 min, followed by 40 cycles of: 95 °C (15 sec), 60 °C (15 sec) and 72 °C (45 sec). Fold expression was calculated based on expression in sporangia. Data were analyzed by two-way ANOVA using Prism 4 (GraphPad Software). Splice variant analysis Primers spanning the region of intron 1 were used (Figure S2.4) to amplify RT-PCR products from SP, 4, and 8 dpi cDNA samples and resultant products were cloned into 69 the TA cloning vector pGEM-T-Easy. Fidelity of all sequences was confirmed by DNA sequencing as described above. Ortholog identification and sequence analysis Candidate effector proteins were identified from the predicted proteome of Ps. cubensis generated from the draft genome assembly using a modified RXLR effector identification pipeline [29]. Orthologs of PscRXLR1 were identified in P. infestans (http://www.broadinstitute.org/), P. sojae (http://genome.jgi-psf.org), P. ramorum (http://genome.jgi-psf.org) and Py. ultimum (http://pythium.plantbiology.msu.edu) with BLAST [53]. Signal peptides were http://www.cbs.dtu.dk/services/SignalP/) predicted and protein using SignalP motifs 3.0 identified ([54], using InterProScan ([55], http://www.ebi.ac.uk/Tools/pfa/iprscan/). Amino acid alignments were generated using ClustalW2 ([56], http://www.clustal.org), and resultant figures generated using BoxShade v3.21 (http://www.ch.embnet.org/software/BOX_form.html). P. infestans effector sequences from Haas et al. [7] were used for analyses and P. infestans protein models were obtained from the Phytophthora infestans Sequencing Project (http://www.broadinstitute.org/). Defining paralogs and orthologs and evolutionary rate estimates Synonymous and non-synonymous substitution rates (Ks and Ka, respectively) were determined using the yn00 program in PAML [57]. Protein sequences were aligned first 70 and “back-translated” to coding sequence alignments. Very few pairs had run errors (e.g., NAN in PAML output), and those with run errors were excluded. Sequence pairs that were too similar (Ks ≤ 0.005) or too divergent (Ks > 3) were also excluded from further analysis. For each Ps. cubensis or P. infestans effector protein, the closest paralogous genes were identified using within-species BLAST searches and used for rate calculation. Rates between putative orthologs were calculated as well. Putative orthologs were identified globally between Ps. cubensis, P. infestans or P. ultimum by first determining pairwise sequence similarities between all genes in these species. For each Ps. cubensis protein, X, a protein in a second species, Y, is considered an ortholog if the following two conditions are met: 1) X is the reciprocal best match of Y and 2) X is located in a syntenic block where Y is found. Syntenic regions were established using Multiple Collinearity Scan [58], with 1e-5 as an alignment significance threshold, match size ≥5, and average intergenic distance. Molecular phylogenetic analysis The full-length protein sequences of PscRXLR1 and its orthologs were aligned using default parameters with MUSCLE [59]. The multiple sequence alignment was used to infer phylogenetic relationships between PscRXLR1 and its orthologs using the Maximum Likelihood method, based on the JTT matrix-based model [60] with MEGA5 [61]. Bootstrap values (based on 500 replicates) for each node are given if >25%. 71 ACKNOWLEDGEMENTS We thank members of the Day lab for critical reading of the manuscript. Dr. Joe Win (The Sainsbury Laboratory) is gratefully acknowledged for his contributions in the preliminary analysis of genomic sequences from Ps. cubensis. 72 APPENDIX 73 Figure S2.1 Figure S2.1 Signal peptide distribution among ortholog pairs. (A) Distribution of secreted or non-secreted proteins in the Pseudoperonospora cubensis – Phytophthora infestans ortholog baseline dataset. (B) Distribution of P. infestans orthologs of Ps. cubensis effectors that are predicted to be secreted. Psc-sec = Ps. cubensis secreted protein. Psc-non = Ps. cubensis non-secreted protein. Pi-sec = P. infestans secreted protein. Pi-non = P. infestans non-secreted protein. 74 Figure S2.2 PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 1 1 1 1 1 1 MVWLQLKKSGLGFTMSLSAVYGAVYAAAN-SVPAGKIDSGKKAMRHLENLPLLLASDSLE --------------MSLSAVYGAVYAAAN-SVPAGKIDSGKKAMRHLENLPLLLASDSLE --------------MPLPAVYGAVYATATSSAPAGKIDSGKKAMRHLENLPLLVASDSLE --------------MPLPAVYGAVYASATSSAPAG-----KKALRHLENLPLLVASDSLE --------------MPLPAVYGAVYATATRSAPAGKIDSGKKAMHHLENLPLLIASDSLD --------------MPAPAAVGIAVAAAA------NTDSGKKAMKQLEELPLLAVRHASD PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 60 46 47 42 47 41 SVSTEGKWLPRFLRQAIMRSIANRIAGIILVSTSAFLASCIATLVKEDAVKLAPVEILFW SVSTEGKWLPRFLRQAIMRSIANRIAGIILVSTSAFLASCIATLVKEDAVKLAPVEILFW SMSTDGKWLPRFMRRAVVRSVANRIAGLILVGTSAFLASCIATLVKDDAFKLSTVETLFW SMTTEGKWLPRFMRRAVVRSVANRIAGLILVGTSAFLASCITTLVKDDTIKLSAIEALFW SMSTEGKWLPRFMRRAVVRSVANRIAGLILVGTSAVLASCIATLTKDADFKLSPVETLFW AAIRTGSRLPKLVQKWYSKLSATQLEGVVLVAASAFTFSLLSTLIKYASQSMPSMETVFW PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 120 106 107 102 107 101 RSLVSWLLTLVSS*---------------------------------------------RSLVSWLLTLVAITTTGVKTRLKKEYYRPIVLRSFTGCIATTLTIIMLQELAVSNAIAIT RSLVSWLLTLAAIAATGVKMRVKKEFQRPLLLRCFTGCIALTLTVLVIQKLEVSNATAIT RSLVSWFLTVAALATTSTKMRVKKEFNRPLTLRCVFGCISTTLTIGVLEKLAVSNATAVT RSLVSWLLTLAAIAATGIRMRVKKEFHRPLVLRCVTGCVAMTLTLLVLQTLAVSNATAIT RSFVAWLLNLVAVWQ-----------------------------------MVLADASVLI PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 166 167 162 167 126 -----------------------------------------------------------YFSPLLAFAMAAKFLKEKPKLFAVACSVMCVIGAVLVVRPVFLFGKSGSTDASWYRRSMT YLSPLLAFAMAAFFLKEKPGAFTLACSALCVVGAVQVVRPAFVFGKNGSTDAKWYRRSMT YVSPLLAFAMATFFLKEKPGVFTAVCSALCVAGAVLVVRPAFLFGKSGSTDAKWYHRSMA YFSPLLAFAMAALFIKEKPDIFTVACSVVCVAGAILVVRPAFLFGKDGSTDAKWYRRSMA FTSPVMTFLLGAMVLKEKIDPVNMGYALFSFVGVICVVRPSFIFGNDHTTAG-------75 Figure S2.2 (cont'd) PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 226 227 222 227 178 -----------------------------------------------------------SFVTSYLFGESLAIGCAIIVVFMQAGAYVSLRSLQKVPHLVVMHYYLVTTTLVSLP*--SFVTSYLFGESLAIGCAVVTVVMQAGAYVSLRSLQKVPHLVVMNYFLLTMTLVSLIAILV SFVTSNLFGESLAVGCAVVVAFMQAGAYVSLRSLHKVEYLVVMQYYLFTMTLAALAAMLG SFVTSSLFGESLAIGCAAVVVFMQAGAYVSLRSLQKVPHLVVMHYFLLSMTLVSLVAVLV ------TDGSVFAIMCALLGAAAQAVAYVSMRRLQQVNYLVVINYFLLTSSVMSALSLLL 287 282 287 232 ----------------------------------------------------------------------------------------------------------------------VQHGKFKAGLSVETWGAILGTGALAFAEQLFLTRGFQFDGAGVLAATRLLHVGYEFVWGV IQHGKFKAGTSLETWGAIVGTGALAFVEQLFLTRGFQFDGAGVLAATRLLHVSCEFAWGV VQHGKFKTDLSLGTWSAILGTGALAFAEQLFLTRGFQFDGAGVLAATRLLHVGCEFAWGV VQR-KFVIKMSLDVWLAVLGTGFLGFIGQLFLTRGFQLESAGTASVMRYLDVVFVFVWDI PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 347 342 347 291 ----------------------------------------------------------------------------------------------------------------------ALLGTALNPWSASGAAATAAGVLFLALRRTARSREALAPHKNAYLH--------FSHARR ILLGTALNPWSAGGAGVTAAGVLFLALRRVHTHWAARRSLRRILQ*-------------ILLGTALNPWSASGAAATAAGVLFLALRRVHTHWAARRSLRRMAAKPHKNAYLHFSSARR TLLHERINAWSAVGALIICGSAIAIAIRKMQS*--------------------------- PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 ----------------------------------------------------------------------------------------------------------------------399 DELAAENPSWSVQQVSAELGRQWKALSAAERKPWVELAQFDKARFHTEAHHHV-NQQQSD -----------------------------------------------------------407 EQLAEANPAWSVQQVSAELGRQWKSLAAVERKPWVELAQFDKARYHTEAHQHMRQQTDEQ ------------------------------------------------------------ PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum96701_0_3435 PYU1_T005955 76 Figure S2.2 (cont'd) PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 ----------------------------------------------------------------------------------------------------------------------458 EQPEQAPPKRKKQSNEPRQPDTAYICFWKSQRPEVVAANPFLAAPLVSKEVGRQWRALSD -----------------------------------------------------------467 PERPHLPTKRKKRPNEPRQPDTAYICFWKSRRPEVIAENPLLAAPSVSREEIARFQP-------------------------------------------------------------- PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 ----------------------------------------------------------------------------------------------------------------------518 DERQPTLAATTPDMLSALKTPLKDPFAPKPAKTAFQLFMSHNRESFMLLNMTINEFRAEM -----------------------------------------------------------524 ------ALVATSEIPPALKAPLKDPFAPKPAKTGFQLFMSHNRESFTLLNMTINEFRTEM ------------------------------------------------------------ PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 ----------------------------------------------------------------------------------------------------------------------578 SQLWKRLSDADKAEWHELAKEDQRRYDTEMNAYKPPAYMDLVVQRSHKRMEELRRLARED -----------------------------------------------------------578 SQLWKRLSDADKNEWYELAKLDERRYETEMNAYKPPAYMESAVQRAHKRLDELRRLARRD ------------------------------------------------------------ PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 ----------------------------------------------------------------------------------------------------------------------638 SAAPRLPMNAYNCYLSAKRQELVDRRPGRKNPEIMREIGVTWKALSDDERAVYQRKADED -----------------------------------------------------------638 AAAPRLPMNAYNCYLSKERQELAVQRPDLKNPEIMREIGVTWKALSEDERASFQRKAEDD -----------------------------------------------------------77 Figure S2.2 (cont'd) PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 ----------------------------------------------------------------------------------------------------------------------698 VERFRAEMEAHIAKKNEEEAANPLTKRRPRKRKEFDDEEELVKTPTVPRKKRKSGPPRRP -----------------------------------------------------------698 VERFRADMEAYLTQQEEQRAVQEEQDVEPEVVREVVHEVVKEP--VVGRKKRKSVSPRRP ------------------------------------------------------------ PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 ----------------------------------------------------------------------------------------------------------------------758 KTAYNLMYMSKRTELLSTYQMSHNECSALCGKLWRQMSEAEREPYKRMAAEDKHRYEAEL -----------------------------------------------------------756 KTAYNLMYMSKRAELLSTYQMSHNECSALCGRLWRQMSEEEREPYKRMAAEDKRRYETEM ------------------------------------------------------------ PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 ----------------------------------------------------------------------------------------------------------------------818 QVYNAQQEEANNKTLRDSAGFRHFLEAKRRENEAISSDEAAAIWQEMTEPHQLLWTELAR -----------------------------------------------------------816 EIYNAEVDAANKKTLRESAGFSYFLEAKRRENEQISEGEAAAIWRDMLAPHQMLWTELAS ------------------------------------------------------------ PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 ----------------------------------------------------------------------------------------------------------------------878 DNKHKTSVERTAVDVLDTLL*-------------------------------------------------------------------------------------------------876 DTPAPASSNATHQKGQTGMNGGRWTEQEHQSFLAGLRLYGREWKKVAAKIKTRTSAQIRS -----------------------------------------------------------78 Figure S2.2 (cont'd) PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------936 HAQKYFAKLARDDEMRKHSGLSMIMAGSIGYFSDGGSSVAQNSGDDDAEASDASRQMARA ------------------------------------------------------------ PscRXLR1 Psc_781.4 P. sojae 156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------996 RSAGQSKGTAAILIAPMGSAVSGLYKQTTGATKKRARAAVTGFDGQLEMGAATSSFPYKL ------------------------------------------------------------ PscRXLR1 Psc_781.4 Psojae_156165 PITG_17484 P. ramorum 96701_0_3435 PYU1_T005955 --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------QKRQNDATRVEYLPSQEELLAKASPNLRHRLSSLIEAELCALQVLSCYAMLQQQEQISAP ------------------------------------------------------------ PscRXLR1 --------------------------------Psc_781.4 --------------------------------Psojae_156165 --------------------------------PITG_17484 --------------------------------P. ramorum 96701_0_3435 1116 RQKTKRQGSAKASTLGLPMLSTEQMPPTSSIY* PYU1_T005955 --------------------------------79 Figure S2.2 (cont'd) Relationship between PscRXLR1 and oomycete orthologs. Alignment of PscRXLR1, Psc_781.4, PITG_17484 (P. infestans), PYU_T005955 (Py. ultimum), P. ramorum 96701_0_3435, and P. sojae 156165 amino acid sequences were generated using ClustalW and represented with BoxShade. PscRXLR1 signal peptide is boxed in blue. The RXLR or RXLR-like domains are boxed in red. The green boxes represent the EamA domains found in each protein sequence. Stop codons are represented by asterisks. 80 Figure S2.3 Figure S2.3 Heterologous expression of PscRXLR1 specifically results in cell death in Nicotiana benthamiana. Infiltration of PscRXLR1_CFP with or without the plasma membrane marker construct AtPIP2A-YFP results in chlorosis and necrosis 4 days post-inoculation (dpi). Circles mark the infiltration zones, visible at 0 dpi. Infiltration with AtPIP2A-YFP alone does not result in any observable phenotype in N. benthamiana leaves. 81 Figure S2.4 A 781_4 exon 1(last 100 bp) 781_4 exon1 F primer 4 dpi #6-8 8 dpi #4-12 SP #5-9 INTRON 1 781_4 intron 1 span R primer 781_4 exon 2 (first 100 bp) TGCCAGCTGC ---------------------------------------------------------------- ATCGCGACGT ---------------------------------------------------------------- TGGTGAAAGA ---------------------------------------------------------------- GGACGCCGTT ---------------------------------------------------------------- 781_4 exon 1(last 100 bp) 781_4 exon1 F primer 4 dpi #6-8 8 dpi #4-12 SP #5-9 INTRON 1 781_4 intron 1 span R primer 781_4 exon 2 (first 100 bp) CCGTGGAGAT ---------------------------------------------------------------- TTTGTTTTGG ------TTGG ------TTGG ------TTGG ------TTGG ---------------------------- CGCTCACTCG CGCTCACTCG CGCTCACTCG CGCTCACTCG CGCTCACTCG ---------------------------- TGTCTTGGCT GCTAACGCTT TGTCTTG-------------TGTCTTGGCT GCTAACGCTT TGTCTTGGCT GCTAACGCTT TGTCTTGGCT GCTAACGCTT ---------- ------------------- ------------------- ---------- 781_4 exon 1(last 100 bp) 781_4 exon1 F primer 4 dpi #6-8 8 dpi #4-12 SP #5-9 INTRON 1 781_4 intron 1 span R primer 781_4 exon 2 (first 100 bp) ------------------GTAAGCTCTT GTAAGCTCTT GTAAGCTCTT GTAAGCTCTT ------------------- ------------------GACCCACTGA GACCCACTGA GACCCACTGA GACCCACTGA ------------------- ------------------TATTGTACGA TATTGTACGA TATTGTACGA TATTGTACGA ------------------- ------------------TGATTGCCTA TGATTGCCTA TGATTGCCTA TGATTGCCTA ------------------- 82 AAATTAGCAC ---------------------------------------------------------------- ------------------ACAAATTCTT ACAAATTCTT ACAAATTCTT ACAAATTCTT ------------------- Figure S2.4 (cont'd) 781_4 exon 1(last 100 bp) 781_4 exon1 F primer 4 dpi #6-8 8 dpi #4-12 SP #5-9 INTRON 1 781_4 intron 1 span R primer 781_4 exon 2 (first 100 bp) ------------------GGTGATTGAT GGTGATTGAT GGTGATTGAT GGTGATTGAT ------------------- ------------------TACAAGGTTG TACAAGGTTG TACAAGGTTG TACAAG------------G ------GTTG ------------------CAATCACGAC CAATCACGAC CAATCACGAC ---------CAATCACGAC CAATCACGAC ------------------TACTGGCGTT TACTGGCGTT TACTGGCGTT ---------TACTGGCGTT TACTGGCGTT ------------------AAG------AAG------AAG---------------AAG------AAGACGCGCT 781_4 exon 1(last 100 bp) 781_4 exon1 F primer 4 dpi #6-8 8 dpi #4-12 SP #5-9 INTRON 1 781_4 intron 1 span R primer 781_4 exon 2 (first 100 bp) ---------------------------------------------------------------TGAAGAAAGA ---------------------------------------------------------------GTATTATCGC ---------------------------------------------------------------CCAATCGTGC ---------------------------------------------------------------TTCGATCATT ---------------------------------------------------------------CACGGGTTGC 781_4 exon 1(last 100 bp) 781_4 exon1 F primer 4 dpi #6-8 8 dpi #4-12 SP #5-9 INTRON 1 781_4 intron 1 span R primer 781_4 exon 2 (first 100 bp) ----------- --------------- --------------- --------------- --------------- --------------- --------------- -----ATCGCCACGA CACTTA 83 Figure S2.4 (cont'd) B 781_4 exon1 (last 100 bp) 781_4 exon1 F primer 4 dpi #3-1 8 dpi #4-7 SP #2-3 781_4 intron 1 span R primer 781_4 exon 2 (first 100 bp) TGCCAGCTGC ------------------------------------------------------- ATCGCGACGT ------------------------------------------------------- TGGTGAAAGA ------------------------------------------------------- GGACGCCGTT ------------------------------------------------------- AAATTAGCAC ------------------------------------------------------- 781_4 exon1 (last 100 bp) 781_4 exon1 F primer 4 dpi #3-1 8 dpi #4-7 SP #2-3 781_4 intron 1 span R primer 781_4 exon 2 (first 100 bp) CCGTGGAGAT ------------------------------------------------------- TTTGTTTTGG ------TTGG ------TTGG ------TTGG ------TTGG ------------------- CGCTCACTCG CGCTCACTCG CGCTCACTCG CGCTCACTCG CGCTCACTCG ------------------- TGTCTTGGCT TGTCTTG--TGTCTTGGCT TGTCTTGGCT TGTCTTGGCT ------------------- GCTAACGCTT ---------GCTAACGCTT GCTAACGCTT GCTAACGCTT ------------------- 781_4 exon1 (last 100 bp) 781_4 exon1 F primer 4 dpi #3-1 8 dpi #4-7 SP #2-3 781_4 intron 1 span R primer 781_4 exon 2 (first 100 bp) ------------------GTTGCAATCA GTTGCAATCA GTTGCAATCA ---GCAATCA GTTGCAATCA ------------------CGACTACTGG CGACTACTGG CGACTACTGG CGACTACTGG CGACTACTGG ------------------CGTTAAG--CGTTAAG--CGTTAAG--CGTTAAG--CGTTAAGACG ------------------------------------------------------CGCTTGAAGA ------------------------------------------------------AAGAGTATTA 84 Figure S2.4 (cont'd) 781_4 exon1 (last 100 bp) 781_4 exon1 F primer 4 dpi #3-1 8 dpi #4-7 SP #2-3 781_4 intron 1 span R primer 781_4 exon 2 (first 100 bp) ------------------------------------------------------TCGCCCAATC ------------------------------------------------------GTGCTTCGAT ------------------------------------------------------- ------------------------------------------------------- ------------------------------------------------------- Figure S2.4. Multiple sequence alignments of splice variant isoforms. (A) Alignment representing non-spliced, PscRXLR1 isoform. (B) Alignment representing spliced isoform indicative of Psc_781.4. 85 Figure S2.5 Figure S2.5 Heterologous expression of Psc_781.4 in Nicotiana benthamiana. Infiltration and expression of Psc_781.4 does not result in any observable phenotype in N. benthamiana leaves at 4 days post-inoculation (dpi). Circles mark the infiltration zones, visible at 9 dpi. 86 REFERENCES 87 REFERENCES 1. De Wit PJGM, Mehrabi R, Van Den B, Harrold A, Stergiopoulos I (2009) Fungal effector proteins: past, present and future. Mol Plant Pathol 10: 735-747 2. Mansfield JW (2009) From bacterial avirulence genes to effector functions via the hrp delivery system: an overview of 25 years of progress in our understanding of plant innate immunity. Mol Plant Pathol 10: 721-734 3. Schornack S, Huitema E, Cano LM, Bozkurt TO, Oliva R, et al. (2009) Ten things to know about oomycete effectors. Mol Plant Pathol 10: 795-803 4. Phytopathology TPCRi (2006) A catalogue of the effector secretome of plant pathogenic oomycetes. Ann Rev Phytopathol 44: 41-60. 5. Chisholm ST, Coaker G, Day B, Staskawicz BJ (2006) Host-microbe interactions: shaping the evolution of the plant immune response. Cell 124: 803-814. 6. Jones JD, Dangl JL (2006) The plant immune system. Nature 444: 323-329. 7. Haas BJ, Kamoun S, Zody MC, Jiang RH, Handsaker RE, et al. (2009) Genome sequence and analysis of the Irish potato famine pathogen Phytophthora infestans. Nature 461: 393-398. 8. Tyler BM (2006) Phytophthora genome sequences uncover evolutionary origins and mechanisms of pathogenesis. Science 313: 1261-1266 9. Levesque CA, Brouwer H, Cano L, Hamilton JP, Holt C, et al. (2010) Genome sequence of the necrotrophic plant pathogen Pythium ultimum reveals original pathogenicity mechanisms and effector repertoire. Genome Biol 11: R73. 10. Baxter L, Tripathy S, Ishaque N, Boot N, Cabral A, et al. (2010) Signatures of adaptation to obligate biotrophy in the Hyaloperonospora arabidopsidis genome. Science 330: 1549-1551. 11. Rehmany AP, Gordon A, Rose LE, Allen RL, Armstrong MR, et al. (2005) Differential recognition of highly divergent downy mildew avirulence gene alleles by RPP1 resistance genes from two Arabidopsis lines. Plant Cell 17: 1839-1850. 88 12. Whisson SC, Boevink PC, Moleleki L, Avrova AO, Morales JG, et al. (2007) A translocation signal for delivery of oomycete effector proteins into host plant cells. Nature 450: 115-118 13. Bos JI, Kanneganti TD, Young C, Cakir C, Huitema E, et al. (2006) The C-terminal half of Phytophthora infestans RXLR effector AVR3a is sufficient to trigger R3amediated hypersensitivity and suppress INF1-induced cell death in Nicotiana benthamiana. Plant J 48: 165-176. 14. Allen RL (2004) Host-parasite coevolutionary conflict between Arabidopsis and downy mildew. Science 306: 1957-1960 15. Armstrong MR, Whisson SC, Pritchard L, Bos JI, Venter E, et al. (2005) An ancestral oomycete locus contains late blight avirulence gene Avr3a, encoding a protein that is recognized in the host cytoplasm. Proc Natl Acad Sci U S A 102: 7766-7771. 16. Shan W, Cao M, Leung D, Tyler BM (2004) The Avr1b locus of Phytophthora sojae encodes an elicitor and a regulator required for avirulence on soybean plants carrying resistance gene Rps 1b. Mol Plant-Microbe Interact 17: 394-403. 17. Qutob D, Tedman-Jones J, Dong S, Kuflu K, Pham H, et al. (2009) Copy number variation and transcriptional polymorphisms of Phytophthora sojae RXLR effector genes Avr1a and Avr3a. PLoS ONE 4: e5066 18. Dong S, Qutob D, Tedman-Jones J, Kuflu K, Wang Y, et al. (2009) The Phytophthora sojae avirulence locus Avr3c encodes a multi-copy RXLR effector with sequence polymorphisms among pathogen strains. PLoS ONE 4: e5556 19. Matlin AJ, Clark F, Smith CWJ (2005) Understanding alternative splicing: towards a cellular code. Nat Rev Mol Cell Biol 6: 386-398. 20. Win J, Kanneganti TD, Torto-Alalibo T, Kamoun S (2006) Computational and comparative analyses of 150 full-length cDNA sequences from the oomycete plant pathogen Phytophthora infestans. Fungal Genet Biol 43: 20-33. 21. Costanzo S, Ospina-Giraldo M, Deahl K, Baker C, Jones R (2007) Alternate intron processing of family 5 endoglucanase transcripts from the genus Phytophthora. Current Genetics 52: 115-123. 89 22. Shen D, Ye W, Dong S, Wang Y, Dou D (2011) Characterization of intronic structures and alternative splicing in Phytophthora sojae by comparative analysis of expressed sequence tags and genomic sequences. Can J Microbiol 57: 84-90. 23. Savory EA, Granke LL, Quesada-Ocampo LM, Varbanova M, Hausbeck MK, et al. (2011) The cucurbit downy mildew pathogen Pseudoperonospora cubensis. Mol Plant Pathol 12: 217-226. 24. Tian M, Win J, Savory E, Burkhardt A, Held M, et al. (2011) 454 Genome sequencing of Pseudoperonospora cubensis reveals effector proteins with a QXLR translocation motif. Mol Plant-Microbe Interact 24: 543-553. 25. Zerbino DR, Birney E (2008) Velvet: Algorithms for de novo short read assembly using de Bruijn graphs. Genome Res 18: 821-829. 26. Cantarel BL, Korf I, Robb SMC, Parra G, Ross E, et al. (2008) MAKER: An easy-touse annotation pipeline designed for emerging model organism genomes. Genome Res 18: 188-196. 27. Jiang RHY, Tripathy S, Govers F, Tyler BM (2008) RXLR effector reservoir in two Phytophthora species is dominated by a single rapidly evolving superfamily with more than 700 members. Proc Natl Acad Sci U S A 105: 4874-4879. 28. Savory EA, Adhikari BN, Hamilton JP, Vaillancourt B, Buell CR, et al. (2012) mRNASeq Analysis of the Pseudoperonospora cubensis Transcriptome During Cucumber (Cucumis sativus L.) Infection. PLoS ONE 7: e35796. 29. Win J, Morgan W, Bos J, Krasileva KV, Cano LM, et al. (2007) Adaptive evolution has targeted the C-terminal domain of the RXLR effectors of plant pathogenic oomycetes. Plant Cell 19: 2349-2369. 30. Lynch M, Conery JS (2000) The evolutionary fate and consequences of duplicate genes. Science 290: 1151-1155. 31. Jack DL, Yang NM, Saier MH, Jr. (2001) The drug/metabolite transporter superfamily. Eur J Biochem 268: 3620-3639. 90 32. Jacobs KA, Collins-Racie LA, Colbert M, Duckett M, Golden-Fleet M, et al. (1997) A genetic selection for isolating cDNAs encoding secreted proteins. Gene 198: 289-296. 33. Oh SK, Young C, Lee M, Oliva R, Bozkurt TO, et al. (2009) In planta expression screens of Phytophthora infestans RXLR effectors reveal diverse phenotypes, including activation of the Solanum bulbocastanum disease resistance protein Rpi-blb2. Plant Cell 21: 2928. 34. Schornack S, van Damme M, Bozkurt TO, Cano LM, Smoker M, et al. (2010) Ancient class of translocated oomycete effectors targets the host nucleus. Proc Natl Acad Sci U S A 107: 17421-17426. 35. Gurlebeck D, Jahn S, Gurlebeck N, Szczesny R, Szurek B, et al. (2009) Visualization of novel virulence activities of the Xanthomonas type III effectors AvrBs1, AvrBs3 and AvrBs4. Mol Plant Pathol 10: 175-188. 36. Jin P, Wood MD, Wu Y, Xie Z, Katagiri F (2003) Cleavage of the Pseudomonas syringae type III effector AvrRpt2 requires a host factor(s) common among eukaryotes and is important for AvrRpt2 localization in the host cell. Plant Physiol 133: 1072-1082. 37. Shan L, Thara VK, Martin GB, Zhou JM, 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. 38. Nelson BK, Cai X, Nebenführ A (2007) A multicolored set of in vivo organelle markers for co-localization studies in Arabidopsis and other plants. Plant J 51: 1126-1136. 39. Chaparro-Garcia A, Wilkinson RC, Gimenez-Ibanez S, Findlay K, Coffey MD, et al. (2011) The receptor-like kinase SERK3/BAK1 is required for basal resistance against the late blight pathogen Phytophthora infestans in Nicotiana benthamiana. PLoS ONE 6: e16608. 40. Win J, Kamoun S, Jones AM (2011) Purification of effector-target protein complexes via transient expression in Nicotiana benthamiana. Methods Mol Biol 712: 181194. 91 41. Liu T, Ye W, Ru Y, Yang X, Gu B, et al. (2011) Two host cytoplasmic effectors are required for pathogenesis of Phytophthora sojae by suppression of host defenses. Plant Physiol 155: 490-501. 42. Halterman DA, Chen Y, Sopee J, Berduo-Sandoval J, Sanchez-Perez A (2010) Competition between Phytophthora infestans effectors leads to increased aggressiveness on plants containing broad-spectrum late blight resistance. PLoS ONE 5: e10536. 43. Savory EA, Granke LL, Quesada-Ocampo LM, Varbanova M, Hausbeck MK, et al. (2011) The cucurbit downy mildew pathogen Pseudoperonospora cubensis. Molecular Plant Pathology 12: 217-226. 44. Keren H, Lev-Maor G, Ast G (2010) Alternative splicing and evolution: diversification, exon definition and function. Nat Rev Genet 11: 345-355. 45. Franke I, Resch A, Dassler T, Maier T, Bock A (2003) YfiK from Escherichia coli promotes export of O-acetylserine and cysteine. J Bacteriol 185: 1161-1166. 46. Chandran D, Tai YC, Hather G, Dewdney J, Denoux C, et al. (2009) Temporal global expression data reveal known and novel salicylate-impacted processes and regulators mediating powdery mildew growth and reproduction on Arabidopsis. Plant Physiol 149: 1435-1451. 47. Torto TA (2003) EST mining and functional expression assays identify extracellular effector proteins from the plant pathogen Phytophthora. Genome Res 13: 16751685. 48. Rafiqi M, Gan PH, Ravensdale M, Lawrence GJ, Ellis JG, et al. (2010) Internalization of flax rust avirulence proteins into flax and tobacco cells can occur in the absence of the pathogen. Plant Cell 22: 2017-2032. 49. Huang S, Li R, Zhang Z, Li L, Gu X, et al. (2009) The genome of the cucumber, Cucumis sativus L. Nat Genet 41: 1275-1281. 50. Langmead B, Trapnell C, Pop M, Salzberg SL (2009) Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol 10: R25. 92 51. Salamov AA, Solovyev VV (2000) Ab initio Gene Finding in Drosophila Genomic DNA. Genome Res 10: 516-522. 52. Knepper C, Savory EA, Day B (2011) Arabidopsis NDR1 is an integrin-like protein with a role in fluid loss and plasma membrane-cell wall adhesion. Plant Physiol 156: 286-300. 53. Altschul SF, Madden TL, Schaffer AA, Zhang J, Zhang Z, et al. (1997) Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res 25: 3389-3402. 54. Emanuelsson O, Brunak S, von Heijne G, Nielsen H (2007) Locating proteins in the cell using TargetP, SignalP and related tools. Nature Protocols 2: 953-971. 55. Quevillon E, Silventoinen V, Pillai S, Harte N, Mulder N, et al. (2005) InterProScan: protein domains identifier. Nucleic Acids Res 33: W116-120. 56. Larkin MA, Blackshields G, Brown NP, Chenna R, McGettigan PA, et al. (2007) Clustal W and Clustal X version 2.0. Bioinformatics 23: 2947-2948. 57. Yang Z (1997) PAML: a program package for phylogenetic analysis by maximum likelihood. Comp App Biosci 13: 555-556. 58. Tang H, Wang X, Bowers JE, Ming R, Alam M, et al. (2008) Unraveling ancient hexaploidy through multiply-aligned angiosperm gene maps. Genome Res 18: 1944-1954. 59. Edgar RC (2004) MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res 32: 1792-1797. 60. Jones DT, Taylor WR, Thornton JM (1992) The rapid generation of mutation data matrices from protein sequences. Comp App Biosci 8: 275-282. 61. Tamura K, Peterson D, Peterson N, Stecher G, Nei M, et al. (2011) MEGA5: Molecular Evolutionary Genetics Analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol Biol Evol doi: 10.1093/molbev/msr121. 93 CHAPTER 3 mRNA-Seq Analysis of the Pseudoperonospora cubensis transcriptome during cucumber (Cucumis sativus L.) infection This chapter was originally published in PLoS ONE. Savory EA*, Adhikari, BN*, Hamilton JP, Vaillancourt, B, Buell CR, and Day B (2012) mRNA-Seq Analysis of the Pseudoperonospora cubensis transcriptome during cucumber (Cucumis sativus L.) infection. PLoS ONE 7(4): e35796. doi:10.1371/journal.pone.0035796. © 2012 Savory et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited. * These authors contributed equally to this work Author Contributions: Conceived and designed the experiments: EAS, BNA, CRB, and BD. Performed the experiments: EAS, BNA, JPH, and BV. Analyzed the data: EAS, BNA, JPH, CRB, and BD. Contributed reagents/materials/analysis tools: EAS, BNA, JPH, CRB, and BD. Wrote the paper: EAS, BNA, CRB, and BD. 94 ABSTRACT Pseudoperonospora cubensis, an oomycete, is the causal agent of cucurbit downy mildew, and is responsible for significant losses on cucurbit crops worldwide. While other oomycete plant pathogens have been extensively studied at the molecular level, Ps. cubensis and the molecular basis of its interaction with cucurbit hosts has not been well examined. Here, we present the first large-scale global gene expression analysis of Ps. cubensis infection of a susceptible Cucumis sativus cultivar, ‘Vlaspik’, and identification of genes with putative roles in infection, growth, and pathogenicity. Using high throughput whole transcriptome sequencing, we captured differential expression of 2383 Ps. cubensis genes in sporangia and at 1, 2, 3, 4, 6, and 8 days post-inoculation (dpi). Additionally, comparison of Ps. cubensis expression profiles with expression profiles from an infection time course of the oomycete pathogen Phytophthora infestans on Solanum tuberosum revealed similarities in expression patterns of 1,576-6,806 orthologous genes suggesting a substantial degree of overlap in molecular events in virulence between the biotrophic Ps. cubensis and the hemi-biotrophic P. infestans. Coexpression analyses identified distinct modules of Ps. cubensis genes that were representative of early, intermediate, and late infection stages. Collectively, these expression data have advanced our understanding of key molecular and genetic events in the virulence of Ps. cubensis and thus, provides a foundation for identifying mechanism(s) by which to engineer or effect resistance in the host. 95 INTRODUCTION The phytopathogenic oomycete Pseudoperonospora cubensis, the causative agent of cucurbit downy mildew [1,2], infects a wide range of cucurbits, including cucumber (Cucumis sativus L.), squash (Cucurbita spp.), and melon (Cucumis melo L.). As an obligate biotroph, Ps. cubensis is dependent on its host for both reproduction and dispersal, and as such, has evolved a highly specialized host range limited to members of the Cucurbitaceae. At present, downy mildew is the most important foliar disease of cucurbits, affecting cucurbit production throughout the world [1,2]. Under favorable conditions, Ps. cubensis is capable of infecting and defoliating a field in less than two weeks, and as a result, is responsible for devastating economic losses. For more than 50 years, control of downy mildew on cucumber in the U.S. was maintained through genetic resistance; however, since 2004, the likely introduction of a new pathotype into U.S. pathogen populations has resulted in a loss of this resistance [1]. While minimal knowledge of the genetic variation within Ps. cubensis exists - specifically related to virulence, pathogenicity, and host specificity among physiological races - the genetic basis of these processes, and the underlying mechanism(s) associated with infection have not been elucidated [1,2,3,4]. To date, analyses of the Ps. cubensis-C. sativus interaction have been limited to the identification of the aforementioned physiological races, and have largely focused on the utilization of variation in host specificity for the identification and classification of pathotypes [3,5]. To this end, six physiological pathotypes, or races, have been identified within populations in the U.S., Israel, and Japan, as well as additional races 96 throughout Europe [1,2,4]. In the U.S., increased disease pressure on cucumber production since 2004 is hypothesized to be the result of the introduction of a new, more virulent pathotype, capable of overcoming the downy mildew resistance gene dm-1, that has been widely incorporated into commercial cucumber varieties since the 1940’s [6]. While genetic analyses such as Amplified Fragment Length Polymorphism have been used to differentiate these physiological races [4] and some effort has been made to refine the species within Pseudoperonospora [6,7], there is limited information available about pathogenicity or virulence genes in Ps. cubensis or the molecular-genetic basis of resistance to this pathogen in the cucurbits. Recent work generated the first sequence assembly of the Ps. cubensis genome and subsequent in silico analysis has identified candidate effector proteins that may have either virulence or avirulence roles in Ps. cubensis infection [8,9]. Structurally, oomycete effector proteins display a modular organization, consisting of a N-terminal signal peptide, a conserved RXLR (Arg-X-Leu-Arg, where “X” is any amino acid) translocation motif, followed by a variable C-terminal effector domain [10]. In short, it is the function and activity of the variable C-terminal effector domain that drives the activity of these molecules [10,11]. A set of 61 candidate effectors were identified in the first draft of the Ps. cubensis genome [8] and included a large class of variants with sequence similarity to the canonical RXLR motif. Specifically, the function of a QXLR-containing effector, designated PcQNE, was characterized and shown to be a member of a large family of Ps. cubensis QXLR nuclear-localized effectors, which was up-regulated during infection of cucumber [8]. Additionally, internalization of PcQNE into the host cell was shown to 97 require the QXLR-EER motif, thereby establishing a basic functional homology with the well-characterized Phytophthora spp. effector proteins [8]. While this work serves as a substantial development in understanding the genetic basis for pathogenicity in Ps. cubensis, additional work is needed to identify and characterize additional effectors and other proteins involved in establishment of infection and pathogen proliferation. The accessibility of oomycete pathogen genome sequences, combined with gene expression data from both pathogen and host throughout the course of infection, can serve as a basis for identification and curation of genes that may have important roles in both virulence and avirulence [12,13,14,15]. To date, oomycete RXLR effectors have been demonstrated to suppress basal host resistance [16,17], as well as to activate effector-triggered immunity (ETI) [18,19,20,21]. In addition to the RXLR class, other cytoplasmically-localized effectors have been identified in Phytophthora spp. [11]. The Crinkler (CRN) family, for example, has a conserved LXLFLAK motif necessary for translocation into the host cytoplasm and subsequent import into plant nuclei where they elicit a rapid cell death response [22,23]. Finally, oomycete effectors have also been shown to function within the host apoplast, including functions as enzyme inhibitors [24,25,26,27,28], small cysteine-rich proteins [22,29,30], the Nep1-like family of proteins [31,32], and CBEL (Cellulose Binding, Elicitor, and Lectin-like) proteins [33,34]. The initial stages of pathogen infection of a plant host involve adhesion, penetration, and invasive growth within the host cell tissue. As such, cell wall degrading enzymes 98 (CWDE), such as endoglucanases, cutinases, cellulases, and β-glucanases have evolved as essential components of an oomycete’s repertoire for cell wall penetration [20]. Numerous CWDE have been identified computationally from the genomic sequences of several plant pathogenic oomycetes, including Phytophthora sojae, Phytophthora ramorum, Hyaloperonospora arabidopsidis, and Pythium ultimum [30,35,36]. In P. sojae, members of the family 5 and family 12 endoglucanases have been shown to be up-regulated during early stages of infection [37,38]. However, in H. arabidopsidis, which causes downy mildew of Arabidopsis thaliana, CWDE-encoding mRNAs are reduced [36]. This could indicate an adaptation in downy mildew pathogens for evasion of recognition by their host, as break-down products from plant cell wall components can function as elicitors of defense responses [39]. Recent advancements in sequencing technologies have led to an explosive growth in the analysis of in planta-expressed genes of biotrophic plant pathogens [12,40,41,42,43,44]. In the current study, we present the first global gene expression analysis of the infection stages of cucumber by the obligate oomycete pathogen Ps. cubensis, the causal agent of cucurbit downy mildew. Through the analysis of a susceptible cucumber cultivar interaction, we describe the identification of genes with putative roles in infection, growth and pathogenicity. Using next-generation sequencing technology, we assessed gene expression in Ps. cubensis in sporangia and at six time points of infection. By combining visual assessment of symptoms with light microscopy to monitor infection stages as well as minimizing collection of non-inoculated tissues, we were able to capture expression of 7,821 Ps. cubensis genes ranging from 159 genes at 99 1 days post inoculation (dpi) to 7,698 at 8 dpi. In total, this work represents a comprehensive examination of the key infection stages of Ps. cubensis growth and development. In total, the work described herein provides a foundation for further dissection of genes relevant to virulence in this obligate phytopathogen. RESULTS AND DISCUSSION Characterization and sampling of Ps. cubensis infection stages While Ps. cubensis is a major pathogen of cucumber and other cucurbits, limited resources describing the infection process and/or virulence determinants of this obligate oomycete are available. In the current study, we sought to identify Ps. cubensis gene expression from both purified sporangia, as well as from a time course of infected cucumber tissues, representing a wide range of infection stages from 1 to 8 dpi. In total, our goal was to gain a broad perspective of in planta gene expression during infection of a susceptible cucumber host and to correlate this expression with the development of outwardly visible symptoms, as well as the development of microscopic pathogen infection structures. Like other phytopathogenic downy mildews and biotrophic fungi, Ps. cubensis is non-culturable, and proliferates and reproduces only on a susceptible cucurbit host. As with previously published reports on analyzing gene expression in biotrophic phytopathogens, optimization of sampling techniques is key to maximize pathogen tissue compared to host, particularly at early stages of infection (Figure 3.1) [12,43,44,45]. 100 Plants were inoculated on the abaxial leaf surface with purified Ps. cubensis sporangia, and samples were collected using a cork borer, minimizing the amount of non-infected tissue in each sample (Figure 3.1). Initial symptoms of downy mildew infection can be observed on the abaxial leaf surface at 1-3 dpi as water soaking at the site of inoculation, while no visual symptoms are apparent on the upper leaf surface (Figure 3.2). At 1 dpi, zoospores were encysted upon stomata on the lower leaf surface, and by 2 dpi, appressoria and initial penetration hyphae were visible beneath stomata. The yellow angular lesions typical of cucurbit downy mildew were apparent on the upper leaf surface by 4 dpi, and over time, became more chlorotic and necrotic as the infection progressed. By 3 to 4 dpi, multiple haustoria formed within the mesophyll layer. mRNA-Seq data analyses Expression profiling of Ps. cubensis sporangia, as well as infection stages at six time points of cucumber infection, were performed using mRNA-Seq. For each time point, two biological replicates were sequenced. The total number of reads produced for each time point ranged from 55 to 59 million reads, with a median of 57 million reads. Reads were mapped to the Ps. cubensis genome which was generated by assembly of Illumina next generation reads; in total the Ps. cubensis genome encompasses 67.9 Mb, with 23,519 protein coding genes and 23,522 gene models (AHJF00000000; 101 Figure 3.1 102 Figure 3.1 (cont'd) Experimental design and sample collection. A 1 x 10e5 sporangia/ml solution of Pseudoperonospora cubensis was used to inoculate the abaxial leaf surface of cucumber cultivar ‘Vlaspik’. Samples were collected using a #3 cork borer to minimize uninfected tissue (black circles) at 1, 2, 3, 4, 6, and 8 days post-inoculation (dpi). Leaf disks were used for microscopic analysis of infection stages or pooled for RNA extraction. RNA-Seq libraries were made for each time point from 2 biological replicates. Within a biological replicate, libraries were barcoded and sequenced in multiple lanes. The sporangia-only library (SP) was not barcoded and was sequenced on its own. [9]). Of the total reads generated, for each time point, approximately 1.6 to 6.4 million (3-12% of the total; Figure 3.3A) mapped to the Ps. cubensis genome. In turn, a majority of reads in each sample were of host origin, and mapped to the cucumber genome (see accompanying paper, [46]) (Figure S3.1). Through this analysis, we found that there was no significant difference in the total number of combined reads from different time points (p > 0.80); however, the number of Ps. cubensis genes expressed at each time point was significantly different for all time point comparisons (p < 0.05; Figure 3.3A). To assess the experimental variation attributable to biological variation, we compared the gene expression pattern of the genes expressed in both of our biological replicates. In total, our experiments showed very high levels of correlation for biological replicates (in all cases examined, Pearson’s Correlation Coefficient (PCC) > 0.94; Figure S3.2), indicating that our sampling, assay, and analysis methods are robust. To evaluate the effect of sampling depth on gene expression detection and to assess whether we have adequately sampled the mixed mRNA-Seq read pool for Ps. cubensis transcripts, subsets of five to 30 million reads were randomly selected from the total 103 Figure 3.2 Figure 3.2 Symptoms and microscopy images of Ps. cubensis infected Cucumis sativus cultivar 'Vlaspik' of time points used for transcriptome analysis. Symptom images were collected of the adaxial (top row) and abaxial (middle row) at 1, 2, 3, 4, 6, and 8 days post-inoculation (dpi). Microscopy (bottom row) to assess stages of Ps. cubensis invasion were collected from the same time points using ethanol-cleared, trypan blue stained samples. Scale bars at 1-4 dpi are 25 um. Scale bars at 6 and 8 dpi are 50 um. Dotted lines represent position of stomata relative to the pathogen structure. e = encysted zoospore. s = stomate. h = haustorium. 104 read pool from each time point and mapped to the Ps. cubensis genome. The simulation experiment showed a clear positive relationship between sampling depth and number of expressed genes at the lower to medium sequencing depth (5 to 20 million reads; Figure 3.3B). With the exception of 1 dpi in which few genes are expressed, after 20-25 million reads, the number of expressed genes begin to plateau, corresponding to the minimum sampling depth of all libraries sequenced in this study. mRNA-Seq transcriptome profiles In concordance with our visual assessment (Figure 3.2) of pathogen growth throughout the time course, our analyses showed a diversity of transcriptional changes in Ps. cubensis, as well as a correlation between gene expression levels and similar stages of pathogen growth. In support of this, we identified 7,821 genes expressed at different time points of infection (Table S3.1, available at www.plosone.org, e35796) and 129 of those genes (Table S3.2, available at www.plosone.org, e35796), mostly housekeeping, were expressed throughout all time points. Analyses of the top 20 highly expressed genes showed that genes expressed at earlier time points have substantially higher FPKM (fragments per kilobase pair of exon model per million fragments mapped) values than the genes expressed at later time points, consistent with the fewer numbers of genes expressed in the early stages of expression and saturation of detection of Ps. cubensis expression with our sampling depth (Table S3.3, available at www.plosone.org, e35796). For all time points analyzed, the minimum FPKM value was zero but the maximum FPKM values ranged from 8,528 at 8 dpi to 270,121 at 1 dpi 105 Figure 3.3 106 Figure 3.3 (cont'd) Figure 3.3 Number of total RNA-seq reads, reads mapped, and number of genes expressed. (A). Total number of reads, number of reads mapped to the Ps. cubensis genome, and number of genes expressed by Ps. cubensis at different time points are shown. Reads were mapped using Bowtie version 0.12.5 and TopHat version 1.2.0. Fragments per kilobase pair of exon model per million fragments mapped (FPKM) values were calculated using Cufflinks version 0.9.3. Genes were considered expressed if the FPKM and 95% confidence interval lower boundary FPKM value was greater than 0.001 and zero, respectively. dpi = days post-inoculation. (B) Comparison between number of expressed genes detected and sampling depth. For all time points 5, 10, 15, 20, 25, and 30 million reads were randomly selected from the total pool of reads from different time points. Read mapping and expression abundances are as described in panel 3A. 107 (Table S3.3, available at www.plosone.org, e35796). The differences in transcriptome profiles are clearly visible in correlation and cluster analyses between the sampled time points. The Pearson Correlation Coefficient (PCC) values for comparisons of different time points ranged from 0.26 (1 dpi vs. 8 dpi) to 0.79 (3 dpi vs. 4 dpi) (Figure 3.4). Corresponding with our visual assessment of pathogen infection stages showing similar growth at 2 and 3 dpi (i.e., penetration and initial hyphal growth, Figure 3.2), gene expression patterns at 2 dpi are strongly associated with that of 3 dpi. Similarly, genes expressed at 6 dpi showed high correlation with the genes from 8 dpi, which represent comparable stages of pathogen growth and proliferation in the mesophyll (Figure 3.2). Additionally, at 1 dpi when encystment of zoospores is occurring, we observed a poor correlation (PCC ranged from 0.26 to 0.45) between expression at that time point with any other time point, likely due to the unique set of genes that would be involved in this process. Interestingly, gene expression at 3 dpi was highly correlated with other time points (PCC ranged from 0.45 to 0.79), suggesting that events occurring at 3 dpi may represent a transition phase between early and late infection. Differential gene expression To identify genes specifically involved in distinct stages of pathogen infection and development, and to assess gene expression pattern changes over the course of infection, we next evaluated differentially expressed genes between all time points. To provide context for the differential expression, we included expression data from mRNASeq analysis of sporangia in which 8,254 Ps. cubensis genes were expressed (Table 108 Figure 3.4 Figure 3.4 Correlation matrix of Pseudoperonospora cubensis expression profiles throughout a time course of Cucumis sativus infection. Normalized transcript abundances for 7,821 genes were calculated in units of fragments per kilobase pair of exon model per million fragments mapped (FPKM) with Cufflinks version 0.9.3. Pearson product-moment correlations (PCC) of log2 FPKM values were calculated for all pair-wise combinations using R. PCCs were clustered using hierarchical clustering with a Pearson correlation distance metric and average linkage using Multiple Experiment Viewer Software version 4.5. The bootstrap support values shown on tree nodes were obtained from 1000 bootstrap replicates. dpi = days postinoculation. 109 S3.4, available at www.plosone.org, e35796). In concordance with the similarities observed during our visualization of pathogen growth, comparisons with the least number of differentially expressed genes are those between early time points, with only 147 (2%) genes differentially expressed between 1 and 2 dpi (Table 3.1). Additionally, 1 and 2 dpi had fewer differentially expressed genes compared to sporangia than the later time points. Of all the combinations tested, 1 dpi had the highest percentage of differentially expressed genes across all pair-wise comparisons, despite having the lowest number of genes expressed, indicating that the events occurring at this stage of infection are unique among the time course. This corresponds both with our cluster analysis above (Figure 3.4) and our microscopic analysis of infection (Figure 3.2). Interestingly, despite the high correlation between expression patterns at 2 and 3 dpi seen using cluster analysis, there were a large number of genes differentially expressed between 2 and 3 dpi. This is additionally supportive of our hypothesis that 3 dpi is a transition phase between early and late infection. Not surprisingly, the highest number of differentially expressed genes were observed in comparison of all other time points and sporangia with 8 dpi, suggestive of an advanced stage of the infection process and a likely transition to processes involved in sporulation. Expression of genes involved in virulence and pathogenicity As an obligate biotroph, Ps. cubensis must evade and/or overcome basal plant defense responses, as well as effector-triggered immunity, in order to establish growth, proliferate and reproduce within its host. This, as in other phytopathogens, is likely 110 Table 3.1 Number of differentially expressed genes between each combination of time points and sporangia. Differential expression analysis was conducted using the CuffDiff program in Cufflinks version 0.9.3 using the Pseudoperonospora cubensis annotation with a false discovery rate of 0.01. 1 dpi 2 dpi 3 dpi 4 dpi 2 dpi† 147 (50%)§ 3 dpi 193 (58%) 848 (28%) 4 dpi 189 (57%) 329 (10%) 560 (14%) 6 dpi 8 dpi † 6 dpi 175 (60%) 306 (7%) 301 (7%) 342 (7%) 8 dpi 192 (59%) 898 (19%) 891 (17%) 820 (16%) 644 (10%) Sporangia 177 (50%) 246 (16%) 391 (13%) 425 (14%) 559 (15%) 1,301 (32%) dpi = days post-inoculation. Numbers in parenthesis indicate the percent of significantly different tests out of the total number of tests that could be performed for each pairwise comparison. § achieved through the secretion of specific effector proteins that function within the host apoplast to interfere with extracellular plant defense responses, such as the activity of glucanases and proteases or cytoplasmically to suppress defense responses. Thus, the identification and characterization of the temporal expression of pathogen-associated genes throughout the course of infection can assist in the identification of secreted effectors that allow for both the promotion of disease, as well as the avoidance of host recognition. In support of this, we identified a suite of 271 candidate RXLR-type effectors within the Ps. cubensis genome [9] with 20 possible amino acid substitutions at position R1, including R and Q. In the current study, we analyzed the expression distribution of all 271-candidate effectors, as well as predicted Crinkler (CRN) effectors, 111 over our time course of infection. As shown in Figure 3.5, the greatest number of expressed candidate effectors have an ASLR (Ala-Ser-Leu-Arg) motif, and are expressed at 2-8 dpi; candidate effectors with RXLR or QXLR motifs were expressed at every time point. Finally, as noted above, the expanded repertoire at the conserved RXLR motif in Ps. cubensis, with a total of 20 amino acids represented at the R1 position, represents a diversity in RXLR-type effectors previously undescribed. The simplest explanation for this expansion at R1 is supported by the hypothesis that RXLRtype effectors may play a role in host range, and that an expanded effector repertoire may impart plasticity. Moving forward, an extensive functional characterization of these RXLR-type effectors will provide insight into both pathogen virulence and host range specificity. Nonetheless, our data suggest that Ps. cubensis possesses a potentially highly expanded virulence capacity, of which, we have determined the expression of 271 RXLR-type effectors over an extensive time-course of susceptibility and disease elicitation in cucumber. Gene families encoding host-targeted hydrolytic enzymes acting on plant proteinases, lipases, and several sugar-cleaving enzymes (carbohydrate active enzymes; CAZymes) were highly expressed in Ps. cubensis at 4 to 8 dpi, suggesting a possible role during infection and proliferation (Figure 3.6). Comparison of glycoside hydrolase (GH), glycosyltransferases (GT), polysaccharide lyase (PL), pectin esterase (PE), and carbohydrate esterase (CE) encoding genes revealed significant differences in number that were expressed as well as diversity across different time points. In total, 178 GH, 135 GT, 2 CE, and 15 PE were expressed throughout all the time points sampled 112 Figure 3.5 Figure 3.5 Candidate effectors expressed at different timepoints. From inner to outermost circles; 2 dpi, 3 dpi, 4 dpi, 6 dpi and 8 dpi. CRN = Crinkler effectors. dpi = days post-inoculation. (Figure 3.6). GH was the most represented family, with expression of 30-78 members followed by GT (17-27 members expressed). The most represented GH families identified were GH3 and GH5, while GT20 and GT48 were the most represented among all GTs. Additionally, substantial differences were observed in the types of CAZymes expressed across different time points. For example, several members of GH (GH family 7, 12, and 31), GT (GT family 1), CE (CE family 5), and PL family were absent in 113 Figure 3.6 Figure 3.6 CAZymes in Pseudoperonospora cubensis expressed during infection on Cucumis sativus. The CAZymes coding genes in the Ps. cubensis genome were annotated using CAZymes Analysis Toolkit- CAT according to the CAZy database in combination with protein family domain analyses. Gene families absent in at least one time point are underlined. CBM = carbohydrate binding module. CE = carbohydrate esterase. GH = glycoside hydrolase. GT = glycosyl transferase. PE = pectin esterase. PL = polysaccharide lyase. dpi = days postinoculation. 114 early time points (i.e., 2 and 3 dpi), yet were expressed at 4 to 8 dpi, suggesting a possible role during the later stage of infection. GH family 12 endoglucanases as well as CE family 5 cutinases have been previously implicated as having a role in infection by Phytophthora spp. [38,47,48]. Comparison to genes induced during P. infestans infection of potato The comparison of gene expression patterns between pathogens during infection of their susceptible hosts can allow for identification of common genes that are specifically involved in pathogenesis, as well as enable the discovery of genes unique to either species. To this end, we chose to compare the gene expression pattern of Ps. cubensis during infection to that of another economically important oomycete pathogen, P. infestans, during the infection of potato, Solanum tuberosum. Using clustering analysis of protein coding genes from both pathogens, we identified 7,374 single copy orthologous genes between these two oomycetes. We then compared the gene expression values obtained from our study (PCU) with those from microarray-based expression profiling [10] experiments with P. infestans-S. tuberosum (PITG). Spearman rank correlation coefficients (SCCs) of log2 expression values were calculated between the single copy orthologs at all time points in the two datasets; between 1,576 and 5,581 genes were included in the pair-wise comparisons (Figure 3.7). The SCC values among all comparisons ranged from 0.12 to 0.76 (Table S3.5, available at www.plosone.org, e35796). Comparisons between time points reflecting similar stages of pathogen infection showed higher overall correlations (0.29 to 0.76) as compared to 115 Figure 3.7 Figure 3.7 Comparison of ribonucleic acid sequencing (RNA-seq) and microarray expression patterns. Microarray expression profiles were obtained from time course analyses of genes expressed in P. infestans (PITG) during infection on S. tuberosum. Single copy orthologous genes between P. infestans and Ps. cubensis (PCU) were identified using OrthoMCL. Log2 transformed expression values of single copy orthologous genes present in both the Ps. cubensis (log2 of fragments per kilobase pair of exon model per million fragments mapped [FPKM]) and P. infestans (log2 intensity) datasets are shown as scatter plots. SCC Spearman correlation coefficient. dpi = days post-inoculation 116 comparisons between dissimilar time points (0.12 to 0.47). The most highly correlated comparisons were those between genes expressed in Ps. cubensis at 4 dpi and P. infestans at 4 dpi (SCC = 0.76). In P. infestans infection on potato, days 2-4 correspond to haustoria formation [10]; likewise, extensive formation of haustoria by Ps. cubensis was observed at 4 dpi (Figure 3.2). Correspondingly, genes expressed at 4 dpi include a haustorium-specific membrane protein, secreted RXLR proteins, as well as an amino acid transporter, which could possibly be involved in nutrient uptake via the haustorium to Ps. cubensis. Gene expression was also highly correlated (SCC = 0.64) between Ps. cubensis 6 dpi samples and P. infestans 5 dpi samples. At 6 dpi, symptoms of Ps. cubensis infection on cucumber manifest as chlorotic yellow lesions (Figure 3.2). Similarly, at 5 dpi, Ph. infestans has entered the mycelial necrotrophic growth stage, showing a similar chlorotic phenotype on its host [10]. Gene co-expression network analyses Correlation analyses in which associations between gene expression patterns are identified are valuable for inferring common function and/or regulatory relationships [49]. In this study, we were primarily interested in identifying genes that are involved in both establishment and maintenance of Ps. cubensis infection, as well as those specifically involved in virulence. To this end, we constructed gene modules to identify highly coexpressed genes, where all members of a module are more highly correlated with each other than to genes outside the module. Using a Coefficient of Variation (CV) cutoff of 1.0, 4,195 genes from an initial total of 7,821 expressed genes were retained for 117 downstream analyses. Using Weighted Gene Correlation Network Analysis (WGCNA) [50], 3,146 genes were assigned to six different gene modules (Modules 1 to 6) containing 107 to 1,312 genes; 1,049 genes were not assigned to any module (Table S3.6, available at www.plosone.org, e35796; Figure S3.3). Eigengenes [51] were calculated for each module and displayed in a heat map (Figure 3.8) revealing discrete gene expression patterns across different time points. As described above, the 1 dpi sample represents both an important initial stage in the infection process, as well as a unique gene expression profile among the infection time points analyzed. This is additionally reflected in Module 1, which contains 146 genes that are highly expressed at 1 dpi, including genes involved in pathogenesis and transport (Figure 3.8, Figure S3.3, Table S3.6, available at www.plosone.org, e35796). The genes in this module are also expressed at 3 dpi, indicating that there may be some similarities between processes involved in zoospore adhesion and encystment, and initiation of haustoria formation. Genes expressed in Module 2 could represent those processes involved in the transition from early to late stages of infection and that are involved in the initial suppression of host defenses and establishment. This module, which contains 508 genes, expressed at 2, 3, and 4 dpi, represents gene expression occurring during initial penetration through the stomata into the host tissue, hyphal growth, and initiation of haustoria formation. It includes genes such as candidate RXLRtype effectors, glucanase inhibitors, CRNs, and a haustorium-specific membrane protein (Table S3.6, available at www.plosone.org, e35796), similar to what has been observed 118 Figure 3.8 Figure 3.8 Heat map of the eigengenes representing each gene module. The columns in the heat map represent time points, and the rows represent eigengenes for each of the six identified co-expression modules. The numbers of genes in each module are given in parentheses. The cells in the heat map show eigengene values between 0 and 1, indicators of relative expression levels of all genes in the module (see Materials and Methods). dpi = days post-inoculation. to be up-regulated during P. infestans infection on potato [10]. Our WGCNA analyses additionally identified genes that are co-expressed during the later stages of infection, specifically in Modules 4, 5, and 6 (Figure 3.3, Figure S3.2). Module 6, in particular, 119 represents genes most highly expressed at 8 dpi possibly indicative of those involved in a shift to the reproductive phase and sporulation. Transcription factors (TFs) are reported to play a key role in the regulation of many biological processes including roles in oomycete pathogenesis [52,53] and within the predicted Ps. cubensis proteome 27 transcription factor-related domains in 440 genes were identified (Table S3.7, S3.8, available at www.plosone.org, e35796). A total of 247 of these were expressed throughout the infection process (Table S3.7, available at www.plosone.org, e35796). We also identified genes encoding transcription factorrelated Pfam domains in all six co-expression modules. Two modules, 2 and 4, with genes co-expressed across different time contained the majority of the transcription factors. The transcription factor-related genes within those modules could play important role in regulation of genes involved in pathogenesis. The bZIP and Myb, DNAbinding transcription factor, which play an important role in oomycete pathogenesis [54,55], were the most abundant transcription factor-related domains expressed during infection. CONCLUSIONS In this study, we present an extensive characterization of the gene expression analysis of the obligate oomycete cucurbit pathogen Ps. cubensis during a compatible interaction. This data set represents the first global gene expression profile of a cucurbit pathogen. Using mRNA-Seq, we analyzed the differential expression of pathogen genes 120 across a time course of infection of cucumber, correlating expression with pathogen infection structures, development, and the onset of disease symptoms. Our study provides a comprehensive examination of the key infection stages of Ps. cubensis growth and development and through clustering and co-expression network analyses, describes genes that are specifically expressed during these stages. In addition, our work has identified an expanded effector repertoire, represented by a unique diversity at the canonical RXLR motif. Overall, the work described herein will significantly enhance our understanding of the regulation of infection of oomycete phytopathogens, as well as a baseline for identifying important virulence determinants in Ps. cubensis. MATERIALS AND METHODS Ps. cubensis inoculation and sample collection Ps. cubensis MSU-1 was maintained on Cucumis sativus cultivar 'Vlaspik' as described previously [8]. Four-week-old cucumber plants were inoculated on the abaxial surface of the first fully-expanded leaf with a 1 x 105 sporangia/ml solution with 20-30 10 μl droplets. Inoculated plants were maintained at 100% relative humidity in the dark for 24 hours and then transferred to growth chambers maintained at 22 °C with a 12 h light/dark photoperiod. Samples were collected at 1, 2, 3, 4, 6, and 8 dpi with a #3 cork borer to collect tissue at the site of inoculation. Samples for RNA extraction were frozen in liquid nitrogen and stored at - 80 °C until use. Samples collected for microscopy were cleared in 95 % ethanol and stored at room temperature. 121 Histological assessment of Ps. cubensis growth Cleared infected leaf discs were stained in a solution of 250 μg/ml trypan blue in equal parts lactic acid, water, and glycerol to visualize infection structures. Microscopy was performed using an Olympus IX71 inverted light microscope. Images were captured using an Olympus DC71 camera and were processed for contrast using Canvas X (ACD Systems International, Inc., Seattle, WA). Library preparation and sequencing Sporangia were washed from the abaxial surface of heavily sporulating leaves, filtered through a 40 μm nylon cell strainer, and pelleted via centrifugation. For RNA extraction (RNeasy Mini Kit, Qiagen, Valencia, CA), sporangia were resuspended in 450 μl RLT buffer with ~50 μl 425-600 μm acid-washed beads and vortexed for 3 minutes to break cells. Additional extraction steps were followed according to the manufacturer's instructions. RNA concentration and quality was determined using the Bioanalyzer 2100 (Agilent Technologies, San Diego CA). The sporangia library was sequenced in two lanes at the UC DNA Sequencing Facility at University of California, Davis (Davis, CA). RNA samples from the infection time course were processed as described in Adhikari et al. (accompanying paper [46]). In brief, RNA was isolated using the RNeasy Mini Kit (Qiagen, Germantown, MD), treated with DNase (Promega, Madison, WI) and barcoded libraries constructed with the Illumina mRNA-seq kit (Illumina, San Diego CA). Libraries 122 were sequenced with the Illumina Genome Analyzer II platform generating 35-42 bp single-end reads. Reads from biological replicates were pooled prior to expression abundance measurements. Reads were deposited in the National Center for Biotechnology Information Sequence Read Archive under accession number SRP009350. mRNA-Seq read mapping and transcript abundance estimation The assembled and annotated Ps. cubensis MSU-1 genome sequence (AHJF00000000; [9]) was used to estimate transcript abundances. mRNA-Seq reads for each time point and control (sporangia) were mapped to the 67.9 Mb Ps. cubensis reference genome using the quality aware alignment algorithms, Bowtie version 0.12.7 [56] and TopHat version 1.2.0 [57]. The single-end reads from different time points were aligned in single-end mode while the paired-end reads from the control were aligned in paired-end mode. The minimum and maximum intron length was set to 5 and 50,000 bp, respectively and the insert size for paired-end mode was set to 140 bp. The aligned read files produced by TopHat were processed by Cufflinks v0.9.3 [58]. A reference annotation of the Ps. cubensis genome (23,519 loci and 23,522 gene models) was provided and the maximum intron length was set to 50,000 bp. Normalized gene expression levels were calculated and reported as FPKM. The quartile normalization option was used to improve differential expression calculations of lowly expressed genes [58]; all other parameters were used at the default settings. A gene was 123 considered expressed in a specific sample if the FPKM value and FPKM 95% confidence interval lower boundary was greater than 0.001 and zero, respectively. Pearson product-moment correlation analyses of log2 FPKM values among mRNA-Seq libraries were performed using R (http://cran.r-project.org/), with all log2 FPKM values less than zero set to zero. Only tests significant at p = 0.05 are shown. Correlation values depicted as a heat map were clustered with hierarchical clustering using a Pearson correlation distance metric and average linkage. The bootstrap support values were calculated from 1000 replicates using Multiple Experiment Viewer Software (MeV) v4.5 [59]. To understand variability among biological replicates, Pearson correlation coefficients were calculated for the log2 transformed FPKM values of the genes expressed in both replicates at a particular time point. Identification of differentially expressed genes Once transcript abundance estimation was calculated, differential expression analysis was conducted using the Cuffdiff program within Cufflinks version 0.9.3 [58] utilizing the read alignment files described above. The expression testing was done at the level of genes. Quartile normalization [60] and a false discovery rate of 0.01 after BenjaminiHochberg correction for multiple testing were used. The Ps. cubensis genome and the annotation files were provided as input parameters. All other parameters were used at the default levels. Cuffdiff was used to perform pairwise comparisons of six time points and sporangia. 124 mRNA-Seq and microarray comparative analyses Gene expression data from a P. infestans-S. tuberosum time course experiment [10] was used to assess gene expression pattern similarities/differences in two ooymcete pathogens. The data set included P. infestans gene expression over a five-day (2-5 days) time course of a potato infection. Raw data was downloaded from the Gene Expression Omnibus (http://www.ncbi.nlm.nih.gov/geo/) (GSE14480) [61]. The probe intensities were normalized using Robust Multichip Analyses method [62]. For the mRNA-Seq to microarray comparative analysis, single copy orthologous genes were identified using OrthoMCL [63,64] with default parameters. Clustering of 23,522 and 18,140 protein-coding genes from Ps. cubensis and P. infestans, respectively, yielded 7,374 clusters with single copy genes from both species. Only single copy orthologous genes were used for the analyses. FPKM and probe intensity values were log2 transformed, and Spearman correlation coefficients were calculated using R (http://cran.r-project.org/). Functional analysis Functional annotation for all Ps. cubensis genes were generated from searches of the UniProt databases (Uuniref100) [65] with BLAST and combined with Pfam [66] protein families assignment performed using HMMER3 [67]. Functional annotations for Ps. cubensis sequences were taken from the best possible UniRef sequence match, but if there was no UniRef sequence match, functional annotations were made based on the best Pfam domain alignment. Transcription factors were identified based on PFAM 125 domains. Gene co-expression network analysis Gene co-expression network analysis was done according to the methods described by Childs et al. [68] with some modifications. First, the FPKM gene expression values were log2 transformed and FPKM values less than 1 were transformed to zero. Second, genes showing no variation across time points were filtered out using a coefficient of variance (CV) cutoff (1.0). Third, the β and treecut parameters were 7 and 0.6, respectively. Eigengenes were calculated using the WGCNA package [51]. The heat map of eigengenes for each gene module was constructed using R (http://cran.rproject.org/). Genes assigned to co-expression modules were annotated based on the Ps. cubensis functional annotation. 126 APPENDIX 127 Figure S3.1 Figure S3.1 Number of total RNA-seq reads, reads mapped, and number of genes expressed at different timepoints. Total number of reads, number of reads mapped and number of genes expressed in Cucumis sativus mapped, (Cusa) and Pseudoperonospora cubensis (Pcu) at different time-points are shown. Reads were mapped using Bowtie version 0.12.5 and TopHat version 1.2.0. Fragments per kilobase pair of exon model per million fragments mapped (FPKM) values were calculated using Cufflinks version 0.9.3. Genes were considered expressed if the 95% confidence interval lower boundary FPKM value was greater than zero. dpi = days post-inoculation. 128 Figure S3.2 129 Figure S3.2 (cont'd) Concordance of FPKM values of the genes expressed in two biological replicates of the Pseudoperonospora cubensis transcriptome. Reads from different timepoints were mapped to Ps. cubensis genome using Bowtie version 0.12.5 and TopHat version 1.2.0. Fragments per kilobase pair of exon model per million fragments mapped (FPKM) values were calculated using Cufflinks version 0.9.3 and Ps. cubensis genome annotations. For each time point Log2 transformed FPKM values of equal number of genes from both replicates are plotted. Pearson Correlation Coefficient (PCC) was calculated using R. dpi, days post-inoculation. 130 Figure S3.3 Figure S3.3 Trend plots of the normalized gene expression values for each gene from six identified gene coexpression modules. Modules consisting of genes expressed modules 1, 2, 3, 4, 5, and 6 are shown. 131 REFERENCES 132 REFERENCES 1. Savory EA, Granke LL, Quesada-Ocampo LM, Varbanova M, Hausbeck MK, et al. (2011) The cucurbit downy mildew pathogen Pseudoperonospora cubensis. Mol Plant Pathol 12: 217-226. 2. Lebeda A, Cohen Y (2011) Cucurbit downy mildew (Pseudoperonospora cubensis)— biology, ecology, epidemiology, host-pathogen interaction and control. Eur J Plant Pathol 129: 157-192. 3. Thomas C, Inaba T, Cohen Y (1987) Physiological Pseudoperonospora cubensis. Phytopathol 77: 1621-1624. specialization in 4. Sarris P, Abdelhalim M, Kitner M, Skandalis N, Panopoulos N, et al. (2009) Molecular polymorphisms between populations of Pseudoperonospora cubensis from Greece and the Czech Republic and the phytopathological and phylogenetic implications. Plant Pathol 58: 933-943. 5. Lebeda A, Widrlechner MP (2003) A set of Cucurbitaceae taxa for differentiation of Pseudoperonespora cubensis pathotypes. J Plant Dis Prot 110: 337–349. 6. Runge F, Choi Y-J, Thines M (2011) Phylogenetic investigations in the genus Pseudoperonospora reveal overlooked species and cryptic diversity in the P. cubensis species cluster. Eur J Plant Pathol 129: 135-146. 7. Choi Y (2005) A re-consideration of Pseudoperonospora cubensis and P. humuli based on molecular and morphological data. Mycological Res 109: 841-848. 8. Tian M, Win J, Savory E, Burkhardt A, Held M, et al. (2011) 454 Genome sequencing of Pseudoperonospora cubensis reveals effector proteins with a QXLR translocation motif. Mol Plant-Microbe Interact 24: 543-553. 9. Savory EA, Zou C, Adhikari BN, Hamilton JP, Buell CR, et al. (2012) Alternative Splicing of a Multi-Drug Transporter from Pseudoperonospora cubensis Generates an RXLR Effector Protein That Elicits a Rapid Cell Death. PLoS ONE 7: e34701. 133 10. Haas BJ, Kamoun S, Zody MC, Jiang RH, Handsaker RE, et al. (2009) Genome sequence and analysis of the Irish potato famine pathogen Phytophthora infestans. Nature 461: 393-398. 11. Kamoun S (2006) A catalogue of the effector secretome of plant pathogenic oomycetes. Ann Rev Phytopathol 44: 41-60. 12. Cabral A, Stassen JH, Seidl MF, Bautor J, Parker JE, et al. (2011) Identification of Hyaloperonospora arabidopsidis transcript sequences expressed during Infection reveals isolate-specific effectors. PLoS ONE 6: e19328. 13. Sierra R, Rodriguez RL, Chaves D, Pinzon A, Grajales A, et al. (2010) Discovery of Phytophthora infestans genes expressed in planta through mining of cDNA libraries. PLoS ONE 5: e9847. 14. Torto-Alalibo TA, Tripathy S, Smith BM, Arredondo FD, Zhou L, et al. (2007) Expressed sequence tags from Phytophthora sojae reveal genes specific to development and infection. Mol Plant-Microbe Interact 20: 781-793. 15. Randall TA, Dwyer RA, Huitema E, Beyer K, Cvitanich C, et al. (2005) Large-scale gene discovery in the oomycete Phytophthora infestans reveals likely components of phytopathogenicity shared with true fungi. Mol Plant-Microbe Interact 18: 229–243. 16. Bos JI, Kanneganti TD, Young C, Cakir C, Huitema E, et al. (2006) The C-terminal half of Phytophthora infestans RXLR effector AVR3a is sufficient to trigger R3amediated hypersensitivity and suppress INF1-induced cell death in Nicotiana benthamiana. Plant J 48: 165-176. 17. Fabro G, Steinbrenner J, Coates M, Ishaque N, Baxter L, et al. (2011) Multiple Candidate Effectors from the Oomycete Pathogen Hyaloperonospora arabidopsidis Suppress Host Plant Immunity. PLoS Pathog 7: e1002348. 18. Armstrong MR, Whisson SC, Pritchard L, Bos JI, Venter E, et al. (2005) An ancestral oomycete locus contains late blight avirulence gene Avr3a, encoding a protein that is recognized in the host cytoplasm. Proc Natl Acad Sci U S A 102: 7766-7771. 19. Allen RL (2004) Host-Parasite Coevolutionary Conflict Between Arabidopsis and Downy Mildew. Science 306: 1957-1960 134 20. Money NP, Davis CM, Ravishankar JP (2004) Biomechanical evidence for convergent evolution of the invasive growth process among fungi and oomycete water molds. Fung Gen Biol 41: 872-876. 21. Dong S, Qutob D, Tedman-Jones J, Kuflu K, Wang Y, et al. (2009) The Phytophthora sojae avirulence locus Avr3c encodes a multi-copy RXLR effector with sequence polymorphisms among pathogen strains. PLoS ONE 4: e5556 22. Torto TA (2003) EST mining and functional expression fssays identify extracellular effector proteins from the plant pathogen Phytophthora. Genome Res 13: 16751685. 23. Schornack S, van Damme M, Bozkurt TO, Cano LM, Smoker M, et al. (2010) Ancient class of translocated oomycete effectors targets the host nucleus. Proc Natl Acad Sci U S A 107: 17421-17426. 24. Damasceno CM, Bishop JG, Ripoll DR, Win J, Kamoun S, et al. (2008) Structure of the glucanase inhibitor protein (GIP) family from Phytophthora species suggests coevolution with plant endo-beta-1,3-glucanases. Mol Plant-Microbe Interact 21: 820-830. 25. Rose JK, Ham KS, Darvill AG, Albersheim P (2002) Molecular cloning and characterization of glucanase inhibitor proteins: coevolution of a counterdefense mechanism by plant pathogens. Plant Cell 14: 1329-1345. 26. Tian M, Benedetti B, Kamoun S (2005) A Second Kazal-like protease inhibitor from Phytophthora infestans inhibits and interacts with the apoplastic pathogenesisrelated protease P69B of tomato. Plant Physiol 138: 1785-1793. 27. Tian M, Huitema E, Da Cunha L, Torto-Alalibo T, Kamoun S (2004) A Kazal-like extracellular serine protease inhibitor from Phytophthora infestans targets the tomato pathogenesis-related protease P69B. J Biol Chem 279: 26370-26377. 28. Tian M, Win J, Song J, van der Hoorn R, van der Knaap E, et al. (2007) A Phytophthora infestans cystatin-like protein targets a novel tomato papain-like apoplastic protease. Plant Physiol 143: 364-377. 29. Liu Z, Bos JI, Armstrong M, Whisson SC, da Cunha L, et al. (2005) Patterns of diversifying selection in the phytotoxin-like scr74 gene family of Phytophthora infestans. Mol Biol Evol 22: 659-672. 135 30. Levesque CA, Brouwer H, Cano L, Hamilton JP, Holt C, et al. (2010) Genome sequence of the necrotrophic plant pathogen Pythium ultimum reveals original pathogenicity mechanisms and effector repertoire. Genome Biol 11: R73. 31. Fellbrich G, Romanski A, Varet A, Blume B, Brunner F, et al. (2002) NPP1, a Phytophthora-associated trigger of plant defense in parsley and Arabidopsis. Plant J 32: 375-390. 32. Cabral A, Oome S, Sander N, Kuefner I, Nürnberger T, et al. (2012) Non-toxic Nep1-like proteins of the downy mildew pathogen Hyaloperonospora arabidopsidis; repression of necrosis-inducing activity by a surface-exposed region. Mol Plant-Microbe Interact. 33. Gaulin E, Drame N, Lafitte C, Torto-Alalibo T, Martinez Y, et al. (2006) Cellulose binding domains of a Phytophthora cell wall protein are novel pathogenassociated molecular patterns. Plant Cell 18: 1766-1777. 34. Gaulin E, Jauneau A, Villalba F, Rickauer M, Esquerré-Tugayé MT, et al. (2002) The CBEL glycoprotein of Phytophthora parasitica var-nicotianae is involved in cell wall deposition and adhesion to cellulosic substrates. J Cell Sci 115: 4565. 35. Tyler BM (2006) Phytophthora genome sequences uncover evolutionary origins and mechanisms of pathogenesis. Science 313: 1261-1266 36. Baxter L, Tripathy S, Ishaque N, Boot N, Cabral A, et al. (2010) Signatures of adaptation to obligate biotrophy in the Hyaloperonospora arabidopsidis genome. Science 330: 1549-1551. 37. Moy P, Qutob D, Chapman BP, Atkinson I, Gijzen M (2004) Patterns of gene expression upon infection of soybean plants by Phytophthora sojae. Mol PlantMicrobe Interact 17: 1051-1062. 38. Costanzo S, Ospina-Giraldo MD, Deahl KL, Baker CJ, Jones RW (2006) Gene duplication event in family 12 glycosyl hydrolase from Phytophthora spp. Fung Gen Biol 43: 707-714. 39. Yamaguchi Y, Huffaker A (2011) Endogenous peptide elicitors in higher plants. Curr Opinion Plant Biol 14: 351-357. 136 40. Fernandez D, Tisserant E, Talhinhas P, Azinheira H, Vieira ANA, et al. (2011) 454pyrosequencing of Coffea arabica leaves infected by the rust fungus Hemileia vastatrix reveals in planta-expressed pathogen-secreted proteins and plant functions in a late compatible plant–rust interaction. Mol Plant Pathol 13: 17-37. 41. Joly DL, Feau N, Tanguay P, Hamelin RC (2010) Comparative analysis of secreted protein evolution using expressed sequence tags from four poplar leaf rusts (Melampsora spp.). BMC Gen 11: 422. 42. Miranda M, Ralph SG, Mellway R, White R, Heath MC, et al. (2007) The transcriptional response of hybrid poplar (Populus trichocarpa x P. deltoides) to infection by Melampsora medusae leaf rust involves induction of flavonoid pathway genes leading to the accumulation of proanthocyanidins. Mol PlantMicrobe Interact 20: 816-931. 43. Duplessis S, Hacquard S, Delaruelle C, Tisserant E, Frey P, et al. (2011) Melampsora larici-populina tanscript profiling during germination and timecourse infection of poplar leaves reveals dynamic expression patterns associated with virulence and biotrophy. Mol Plant-Microbe Interact 24: 808-818. 44. Mosquera G, Giraldo M, Khang CH, Coughlan S, Valent B (2009) Interaction transcriptome analysis identifies Magnaporthe oryzae BAS1-4 as biotrophyassociated secreted proteins in rice blast disease. Plant Cell 21: 1273-1290. 45. Polesani M, Desario F, Ferrarini A, Zamboni A, Pezzotti M, et al. (2008) cDNA-AFLP analysis of plant and pathogen genes expressed in grapevine infected with Plasmopara viticola. BMC Gen 9: 142 - 156. 46. Adhikari BN, Savory EA, Vaillancourt B, Childs KL, Hamilton JP, et al. (2012) Expression Profiling of Cucumis sativus in Response to Infection by Pseudoperonospora cubensis. PLoS ONE 7: e34954. 47. Ospina-Giraldo M, Griffith J, Laird E, Mingora C (2010) The CAZyome of Phytophthora spp.: A comprehensive analysis of the gene complement coding for carbohydrate-active enzymes in species of the genus Phytophthora. BMC Gen 11: 1-16. 48. Ospina-Giraldo M, McWalters J, Seyer L (2010) Structural and functional profile of the carbohydrate esterase gene complement in Phytophthora infestans. Curr Gen 56: 495-506. 137 49. Ihmels J, Bergmann S, Berman J, Barkai N (2005) Comparative gene expression analysis by a differential clustering approach: Application to the Candida albicans transcription program. PLoS Genet 1: e39. 50. Langfelder P, Zhang B, Horvath S (2008) Defining clusters from a hierarchical cluster tree: the Dynamic Tree Cut package for R. Bioinformatics 24: 719 - 720. 51. Langfelder P, Horvath S (2007) Eigengene networks for studying the relationships between co-expression modules. BMC Sys Biol 1: 54. 52. Iyer LM, Anantharaman V, Wolf MY, Aravind L (2008) Comparative genomics of transcription factors and chromatin proteins in parasitic protists and other eukaryotes. In J Parasitol 38: 1-31. 53. Wang Y, Dou D, Wang X, Li A, Sheng Y, et al. (2009) The PsCZF1 gene encoding a C2H2 zinc finger protein is required for growth, development and pathogenesis in Phytophthora sojae. Micro Path 47: 78-86. 54. Blanco FA, Judelson HS (2005) A bZIP transcription factor from Phytophthora interacts with a protein kinase and is required for zoospore motility and plant infection. Mol Micro 56: 638-648. 55. Judelson H, Ah-Fong A (2010) The kinome of Phytophthora infestans reveals oomycete-specific innovations and links to other taxonomic groups. BMC Gen 11: 700. 56. Langmead B, Trapnell C, Pop M, Salzberg SL (2009) Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol 10: R25. 57. Trapnell C, Pachter L, Salzberg SL (2009) TopHat: discovering splice junctions with RNA-Seq. Bioinformatics 25: 1105-1111. 58. Trapnell C, Williams BA, Pertea G, Mortazavi A, Kwan G, et al. (2010) Transcript assembly and quantification by RNA-Seq reveals unannotated transcripts and isoform switching during cell differentiation. Nat Biotechnol 28: 511-515. 59. Saeed AI, Bhagabati NK, Braisted JC, Liang W, Sharov V, et al. (2006) TM4 microarray software suite. Methods Enz 411: 134-193. 138 60. Bullard J, Purdom E, Hansen K, Dudoit S (2010) Evaluation of statistical methods for normalization and differential expression in mRNA-Seq experiments. BMC Bioinformatics 11: 94. 61. Edgar R, Domrachev M, Lash AE (2002) Gene Expression Omnibus: NCBI gene expression and hybridization array data repository. Nucleic Acids Res 30: 207210. 62. Irizarry R, Hobbs B, Collin F, Beazer-Barclay Y, Antonellis K, et al. (2003) Exploration, normalization, and summaries of high density oligonucleotide array probe level data. Biostatistics 4: 249-264. 63. Li L, Stoeckert CJ, Jr., Roos DS (2003) OrthoMCL: identification of ortholog groups for eukaryotic genomes. Genome Res 13: 2178-2189. 64. Chen F, Mackey AJ, Vermunt JK, Roos DS (2007) Assessing performance of orthology detection strategies applied to eukaryotic genomes. PLoS ONE 2: e383. 65. Suzek BE, Huang H, McGarvey P, Mazumder R, Wu CH (2007) UniRef: comprehensive and non-redundant UniProt reference clusters. Bioinformatics 23: 1282-1288. 66. Bateman A, Birney E, Durbin R, Eddy SR, Howe KL, et al. (2000) The Pfam protein families database. Nucleic Acids Res 28: 263-266. 67. Eddy SR (2009) A new generation of homology search tools based on probabilistic inference. Genome Inform 23: 205-211. 68. Childs KL, Davidson RM, Buell CR (2011) Gene coexpression network analysis as a source of functional annotation for rice genes. PLoS ONE 6: e22196. 139 CHAPTER 4 Expression profiling of Cucumis sativus in response to infection by Pseudoperonospora cubensis This chapter was originally published in PLoS ONE. Adhikari, BN*, Savory EA*, Vaillancourt, B, Childs KL, Hamilton JP, Day B and Buell CR (2012) Expression profiling of Cucumis sativus in response to infection by Pseudoperonospora cubensis. PLoS ONE 7(4): e34954. doi:10.1371/journal.pone.0034954 © 2012 Adhikari et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited. * These authors contributed equally to this work. Author Contributions: Conceived and designed the experiments: EAS, BNA, KLC, BD and CRB. Performed the experiments: BNA, EAS, KLC, and JPH. Analyzed the data: BNA, EAS, KLC, and JPH. Contributed reagents/materials/analysis tools: BNA, EAS, BV, KLC, JPH, BD and CRB. Wrote the paper: BNA, EAS, BD and CRB. 140 ABSTRACT The oomycete pathogen, Pseudoperonospora cubensis, is the causal agent of downy mildew on cucurbits, and at present, no effective resistance to this pathogen is available in cultivated cucumber (Cucumis sativus). To better understand the host response to a virulent pathogen, we performed expression profiling throughout a time course of a compatible interaction using whole transcriptome sequencing. As described herein, we were able to detect the expression of 15,286 cucumber genes, of which 14,476 were expressed throughout the infection process from 1 day post-inoculation (1 dpi) to 8 dpi. A large number of genes, 1,612 to 3,286, were differentially expressed in pair-wise comparisons between time points. We observed the rapid induction of key defense related genes, including catalases, chitinases, lipoxygenases, peroxidases, and protease inhibitors within 1 dpi, suggesting detection of the pathogen by the host. Coexpression network analyses revealed transcriptional networks with distinct patterns of expression including down-regulation at 2 dpi of known defense response genes suggesting coordinated suppression of host responses by the pathogen. Comparative analyses of cucumber gene expression patterns with that of orthologous Arabidopsis thaliana genes following challenge with Hyaloperonospora arabidopsidis revealed correlated expression patterns of single copy orthologs suggesting that these two dicot hosts have similar transcriptional responses in response to related pathogens. In total, the work described herein presents an in-depth analysis of the interplay between host susceptibility and pathogen virulence in an agriculturally important pathosystem. 141 INTRODUCTION Cucumber (Cucumis sativus L.) is an economically important vegetable crop cultivated in over 80 countries, with more than 66 million tons produced annually for both fresh use and processing (http://faostats.fao.org). Cucumber has been utilized extensively as a model system to study sex determination [1], vascular biology [2], and induced defense responses [3,4]. Despite its extensive use as a model system in research, as well as its obvious economic importance, genetic and genomic resources remain limited. In recent years, the publication of both a genetic map [5] and genome sequences [6,7] of cucumber, as well as generation of large-scale expression data sets [8,9], have provided the first comprehensive resources for genetic and genomic based inquiries into cucumber biology. Cucumber has limited genetic diversity, few wild relatives, and only 7 pairs of chromosomes (2n = 14), whereas other Cucumis spp., such as melon (Cucumis melo), have 12 chromosomes, making interspecific breeding difficult, if not impossible. As such, advances in breeding for important agronomic traits such as increased yield, fruit quality, and disease resistance are slow. Cucumber production is hindered by diseases caused by bacterial (e.g., Pseudomonas syringae pv. lachrymans), viral (e.g., Cucumber mosaic virus), fungal (e.g., Sphaerotheca fulginea and Erysiphe cichoracearum), and oomycete (e.g., Phytophthora capsici and Pseudoperonospora cubensis) pathogens [10,11]. Of these, the most destructive is Ps. cubensis, the causal agent of cucurbit downy mildew, which threatens cucumber production in nearly 80 countries and causes severe economic losses 142 [11,12,13]. Ps. cubensis is an obligate, biotrophic oomycete pathogen with a host range limited to the Cucurbitaceae [11]. In recent years, a foundation has been established to support advances in this area, including studies on epidemiology [14], host specificity [15,16,17,18], pathogenic variation [19,20,21], and more recently, generation of a draft genome sequence of Ps. cubensis [22,23]. The molecular and biochemical mechanisms associated with host resistance have been investigated to a limited extent in cucumber and other cucurbits. In large part, signaling of resistance is primarily associated with systemic acquired resistance (SAR) [24], for which cucumber has historically been a model system [25,26]. In addition to SAR, structural modifications (e.g., callose deposition), as well as the induction of defenserelated genes (e.g., peroxidases chitinases, and glucanases) are often associated with the onset of resistance signaling following pathogen infection. Moreover, like other wellcharacterized plant-pathogen systems, the presence of nucleotide-binding site (NBS)containing genes encoding protein products that recognize cognate pathogen effector proteins or their perturbations [27] have been postulated to play a role in disease resistance in cucumber. To this end, analysis of the cucumber genome sequence has identified 61 NBS-Resistance genes, considerably less than have been identified in other plant genomes, such as Arabidopsis (200) or rice (600) [6]. However, of the 15 genes known to control disease resistance in cucumber, none have been cloned, nor have they been associated through linkage maps with the 61 predicted NBS-Resistance genes [28]. It is hypothesized that cucumber has an expanded lipoxygenase (LOX) 143 pathway which may provide an additional mechanism(s) to facilitate responses to biotic stress [6]. While genetically conferred host resistance is the ideal means of disease control in crop species, it has become ineffective in controlling cucurbit downy mildew, particularly in the U.S., where introduction of a new, more virulent pathotype of Ps. cubensis is responsible for economic losses in recent years [12,29]. To this end, control methods for cucurbit downy mildew in both Europe and the U.S. require the use of fungicides, coupled with a single host resistance locus, the recessive dm1 gene, which has been incorporated into most commercial cucumber germplasm [11]. However, the identification of the dm1 locus, as well as its functional role in resistance signaling remains undefined. In addition to widespread incorporation of dm1, breeding of Ps. cubensis resistance has focused mainly on genes from melon [30], as limited diversity for resistance is available in cucumber or its wild relative, Cucumis hardwickii. Largescale screening trials to identify tolerant germplasm are in progress, but have yet to identify a source of complete resistance to Ps. cubensis [16,31]. As such, new resources must be explored to support development of improved cultivars and identify new sources of resistance for breeding programs. Next generation sequencing of the transcriptome (mRNA-Seq) permits deep, robust assessments of transcript abundances and transcript structure [32]. When gene expression profiling is applied to host-pathogen interactions of economically important crops, insights into the mechanisms these pathogens use to suppress and subvert the 144 host defense response can be made [33,34,35]. In the current study, we performed expression profiling, using the susceptible cucumber cultivar 'Vlaspik', over a time course of infection with the downy mildew pathogen Ps. cubensis to identify genes, pathways, and systems that are altered during a compatible interaction. Using this technology, deep profiling of both the host and pathogen transcriptome (see accompanying manuscript, [36]) was achieved, providing the first in-depth analysis of this important plant-pathogen interaction. In this study, we cataloged the expression of 14,476 genes from cucumber through an 8-day time course of infection with a virulent Ps. cubensis isolate. In total, this work identified major changes in gene expression in cucumber at 1 days post-inoculation (dpi) continuing through 8 dpi, with up to 3,286 genes differentially expressed between time points. Comparative analyses revealed correlated gene expression patterns in cucumber and Arabidopsis thaliana leaves infected with downy mildew suggesting orthologous host responses in these two dicotyledonous hosts. Through co-expression network analyses, modules of temporalspecific transcriptional networks were constructed that provide a framework to connect transcription factors with defense response genes. RESULTS AND DISCUSSION Response of C. sativus leaves to pathogen infection To correlate gene expression and host responses with observable disease symptoms and pathogen invasion, the progression of infection in susceptible C. sativus cv. 145 'Vlaspik' was monitored at 1, 2, 3, 4, 6, and 8 dpi. As shown in Figure 4.1, the first visible symptoms of Ps. cubensis infection were apparent at 1 dpi, in the form of water soaking on the abaxial leaf surface at the inoculation site (Figure 4.1A). These symptoms correspond to zoospore encystment and initial penetration through the stomata into the host (see accompanying paper, [36]). In similar pathosystems systems such as Hyaloperonospora arabidopsidis, the causal agent of downy mildew on A. thaliana, analogous processes occur in the early stages of infection, with the exception that H. arabidopsidis penetrates between anticlinal walls of epidermal cells rather than utilizing natural openings like stomata [37]. While no symptoms are visible on the upper leaf surface in cucumber, water soaking on the lower leaf surface can be seen as early as 1 dpi, and remains present through 4 dpi, during which time hyphal growth through the mesophyll of the host tissue occurs and haustoria formation begins (see accompanying paper, [36]). Yellow angular lesions bound by leaf veins characteristic of cucurbit downy mildew were first visible on the upper leaf surface at 4 dpi (Figure 4.1D), and became more chlorotic and slightly necrotic at the centers by 8 dpi. These symptoms are associated with extensive growth of hyphae through the plant mesophyll (see accompanying paper, [36]). Gene expression profiling in C. sativus We performed mRNA-Seq profiling of C. sativus leaves following infection with Ps. 146 Figure 4.1 Figure 4.1 Symptoms of Pseudoperonospora cubensis infection on susceptible Cucumis sativus cv. ‘Vlaspik’. Images were collected of the adaxial (left column) and abaxial (right column) leaf surfaces at 1, 2, 3, 4, 6, and 8 days post-inoculation (dpi). (A) 1 dpi, (B) 2 dpi, (C) 3 dpi, (D) 4 dpi, (E) 6 dpi, (F) 8 dpi. cubensis over an 8-day period. Leaf disc samples were collected using a cork borer to 147 maximize the amount of infected tissue in each sample (Figure 4.2, black circles), pooled within a given time point, and RNA was isolated. Two biological replicates of each time point were sequenced, yielding 55 to 59 million reads from both replicates at each time point. Additionally, a mock-inoculated C. sativus sample was collected and sequenced, yielding 5.8 million reads. The number of reads that mapped to the C. sativus genome ranged from 48 to 53 million (Figure 4.3A, 84-93% of the total reads) per time point while the number of genes expressed at different time points ranged from 12,257 to 13,048 (Figure 4.3A). The libraries were constructed from inoculated leaves and therefore, the reads represent transcripts from the host (C. sativus) and the pathogen (Ps. cubensis). At the early time points, nearly all of the reads obtained were of host origin, which is consistent with our microscopy analysis revealing limited pathogen biomass. However, as we are surveying a susceptible interaction, pathogen biomass increases throughout the time course and consequentially, pathogen transcripts increase in the total read pool in the later time points (see accompanying paper, [36]). However, even with the increased percentage of pathogen reads in the later time points, we have saturated sampling of the C. sativus transcriptome with our deep read coverage of the libraries. Randomly selected subsets of reads, 5 to 30 million, from the total read pool were used to evaluate the effect of sampling depth on gene expression detection. The simulation demonstrates a positive relationship between sampling depth and numbers of expressed genes at lower sequencing depths (5 to 20 million reads) 148 Figure 4.2 149 Figure 4.2 (cont'd) Experimental design and sample collection. Cucumber cv. ‘Vlaspik’ leaves were inoculated on the abaxial leaf surface with 10-30 10 l droplets of a 1 x 105 sporangia/ml solution. Samples were collected at 1, 2, 3, 4, 6, and 8 days post-inoculation (dpi) using a #3 cork borer to isolate tissue immediately around the infection point (black circles). Samples from cucumber leaves mock-inoculated with 10 l droplets of dH2O were collected in the same manner at 1 dpi. Leaf disks from each time point were pooled for RNA extraction. mRNA-Seq libraries were made for each time point from two separate biological replicates. Within a biological replicate, libraries were barcoded and sequenced in multiple lanes with the exception of the mock-inoculated library, which was not barcoded. (Figure 4.3B). The number of expressed genes, however, begins to plateau at approximately 20 million reads, corresponding to the minimum sampling depth of all libraries in this study. To study the repeatability of two biological replicates, pair-wise scatter plots of gene expression values were generated. For biological replicates of each time point, nearly all genes fell along the diagonal of plots, indicating no major variation with correlation coefficients (R2) ranging from 0.97 to 0.98, thus providing evidence for high reproducibility of biological replicates (Figure S4.1). Host transcriptional changes in response to infection Over the infection period, a total of 14,476 C. sativus genes were expressed (Table S4.1, available at www.plosone.org, e34954), with 10,350 genes common to all time points. For all data points analyzed, the minimum fragments per kilobase exon model per million mapped reads (FPKM) value was zero, yet the maximum FPKM values ranged from 8,869 at 1 dpi to 34,017 at 8 dpi (Table S4.1, available at www.plosone.org, e34954). Interestingly, the highest up-regulated gene at 1 dpi was a putative galactinol synthase (Csa6M000080.1), 150 which has Figure 4.3 151 Figure 4.3 (cont'd) Figure 4.3 Comparison of total mRNA-Seq reads, reads mapped and number of genes expressed. (A) Total number of reads, total reads mapped, and number of genes expressed as determined from pooling of both biological replicates are shown. Reads were mapped to the C. sativus genome [6] using Bowtie version 0.12.5 [55] and TopHat version 1.2.0 [53]. Fragments per kilobase pair of exon model per million fragments mapped (FPKM) values for the annotated C. sativus genes were calculated using Cufflinks version 0.9.3 [55]. Genes were considered expressed if the FPKM value and 95% confidence interval lower boundary FPKM value was greater than 0.001 and zero, respectively. (B) Effect of sampling depth on detection of expressed genes. For all time points 5, 10, 15, 20, 25, and 30 million reads were randomly selected from the total pool of reads. Read mapping and expression abundance estimates were performed as describe above. Solid lines indicate number of genes expressed and the dashed lines indicate number of reads mapped at different time points. dpi, days post-inoculation. 152 been shown to be up-regulated in Cucumis melo (melon) in response to abiotic stress [38] as well as in an inbred C. sativus line 'IL57' with a high level of resistance to downy mildew [39]. The expression patterns of the top 20 highly expressed genes showed expression of genes involved in defense responses including catalases, chitinases, lipoxygenases, peroxidases, and protease inhibitors, beginning at 1 dpi and extending through 8 dpi (Table S4.2, available at www.plosone.org, e34954). The detection of defense-related genes within 1 dpi suggests that there is an active response by C. sativus to early infection stages of Ps. cubensis, including zoospore encystment, appressorium formation, and penetration via stomata (see accompanying paper, [36]). Additionally, no pathogen defense-related genes are present within the top 20 highly expressed genes in the mock-inoculated samples, which mainly consists of housekeeping genes (Table S4.2, available at www.plosone.org, e34954). Correlation and cluster analyses were used to identify similarities in transcriptome profiles among the sampled time points. Pearson Correlation Coefficient (PCC) values for time point comparisons ranged from 0.78 to 0.93, with tight clustering readily apparent, revealing patterns that highlight the extent of transcriptional diversity underlying early (1 dpi), intermediate (2, 3, and 4 dpi), and advanced (6 and 8 dpi) stages of disease progression and the corresponding responses in host gene expression (Figure 4.1). As described above, C. sativus defense-related genes are expressed within 1 dpi of Ps. cubensis inoculation, and based on our correlation analysis (Figure 4.4), these likely represent a distinct cluster of genes responding specifically to initial recognition of sporangia, germination of zoospores, and zoospore 153 Figure 4.4 Figure 4.4 Correlation matrix of Cucumis sativus expression profiles during infection by Pseudoperonospora cubensis. Tissue samples were collected from C. sativus at different time points of infection with Ps. cubensis. Normalized transcript abundances for 14,476 genes were calculated as fragments per kilobase pair of exon model per million fragments mapped (FPKM) with Cufflinks version 0.9.3 [55]. Pearson correlation coefficient of log2 FPKM values were calculated for all pair-wise comparisons using R. Hierarchical clustering was performed using Pearson correlation distance metric and average linkage with the Multiple Experiment Viewer Software version 4.5 [57]. The bootstrap support values shown on tree nodes were obtained from 1000 bootstrap replicates. dpi, days post-inoculation. 154 encystment in stomata. Genes expressed at 2-4 dpi also cluster more with each other than to 1 dpi or to later time points, which also reflects the similar stages of Ps. cubensis infection apparent at days 2-4 dpi, indicating that these genes may be involved in host responses to hyphal penetration, growth and haustorium formation. The clustering of gene expression at later time points (6 and 8 dpi) likely corresponds to similar symptoms observed (Figure 4.1) as well as plant responses to extensive Ps. cubensis hyphal growth that is occurring at those time points (see accompanying paper, [36]). Conserved host responses in C. sativus and A. thaliana in compatible interactions as measured through expression profiling Host responses to pathogen challenge have been well documented in the model species A. thaliana [40], including those to the downy mildew pathogen H. arabidopsidis [34,37]. To identify genes induced in a compatible interaction with a downy mildew pathogen common to both C. sativus and A. thaliana, we identified single copy orthologous genes in both plant genomes and analyzed their expression patterns. A total of 7,396 clusters of single copy genes from both species were identified by clustering 23,248 and 27,416 representative protein coding genes from C. sativus and A. thaliana, respectively. Data from a previous microarray-based expression profiling [41] experiment of a compatible A. thaliana-H. arabidopsidis interaction was compared with our mRNA-Seq-based expression data. The H. arabidopsidis infection time points were 0, 0.5, 2, 4, and 6 dpi, similar to the 1-8 dpi time points assayed in this study. Spearman rank correlation coefficients (SCCs) of log2 expression values were calculated between the single copy orthologs at all time points in the two datasets; 155 between 2,136 and 3,446 gene pairs were included in the pair-wise comparisons. Among the six comparisons between similar time points, the SCC values ranged from 0.10 to 0.72 (Table S4.3, available at www.plosone.org, e34954). The correlations between early time points were the lowest between the two interactions, possibly due to the differences in penetration strategies between the two pathogens. The highest correlations were observed between 6 dpi (SCC = 0.72) followed by 2 dpi (SCC = 0.66) for both host-pathogen interactions (Figure 4.5). Overall, correlation coefficients between analogous time points (0.65 to 0.72) were greater than comparisons between dissimilar time points (0.10 to 0.33) indicating similar expression patterns for single copy orthologous genes in C. sativus and A. thaliana. Differential gene expression throughout the infection process Differences in gene expression patterns between time points can provide insight into the host response following pathogen perception and subsequent infection; thus, the mechanism(s) through which pathogenicity occurs may be inferred [34,42,43,44]. For example, a recent publication by Gaulin et al. [45] used a comparative approach analyzing the transcriptomes of Aphanomyces euteiches and Phytophthora spp. to identify novel pathogenicity factors and expanded repertoires for virulence. Through comparison of gene expression patterns across time points, we identified genes differentially expressed between all time points and the control sample, 1 dpi mock inoculation; between 1,170 and 3,286 genes were differentially expressed in pair-wise 156 Figure 4.5 157 Figure 4.5 (cont'd) Comparison of orthologous gene expression in Cucumis sativus and Arabidopsis thaliana in a compatible interaction. Microarray expression profiles were obtained from time-course analyses of genes expressed in A. thaliana (Arth) during infection by Hyaloperonospora arabidopsidis [39]. Single copy orthologous genes between C. sativus (Cusa) and A. thaliana were identified using OrthoMCL [60]. Log2 transformed expression values of single copy orthologous genes expressed in the C. sativus mRNA-Seq dataset (log2 fragments per kilobase pair of exon model per million fragments mapped [FPKM]) and A. thaliana microarray dataset (log2 intensity) are shown as scatter plots. SCC, Spearman correlation coefficient; dpi, days post-inoculation. comparison between the mock inoculated and/or the inoculated time points (Table 4.1, Table S4.4, available at www.plosone.org, e34954). In general, 12 to 31% of the genes tested under different conditions were differentially expressed in the pair-wise comparisons. In infected samples, comparisons of the three early to intermediate time points (i.e., 1, 2, and 3 dpi) showed a higher number of differentially expressed genes (2,214 to 3,286) than those between three intermediate to later (4, 6 and, 8 dpi) time points (1,612 to 2,074), suggesting more correlated gene expression in later stages of infection. Gene co-expression pattern analyses Using Weighted Gene Correlation Network Analysis (WGCNA) [46], we identified sets of highly correlated genes and constructed modules where all members are more highly correlated with each other than they are to genes outside the module [47]. Out of 15,286 genes expressed in the mock control or throughout the infection time course, 4,410 genes passed the Coefficient of Variance (CV) filter (0.4) and were retained for downstream analyses. Of the 4,410 genes, a total of 2,169 were assigned to 11 gene 158 Table 4.1 Number of genes differentially expressed between different time points. MockControl 1 dpi 2 dpi 3 dpi 1 dpi¥ 1,170 (12%)§ 2 dpi 1,906 (19%) 3 dpi 2,596 (26%) 4 dpi 2,509 (25%) 6 dpi 1,713 (17%) 8 dpi 2,556 (26%) 3,286 (31%) 2,948 (30%) 2,214 (23%) 2,849 (27%) 1,736 (18%) 1,590 (16%) 2,010 (20%) 2,718 (27%) 1,840 (18%) 1,842 (18%) 3,014 (30%) 3,006 (29%) 2,218 (23%) 2,074 (21%) 1,612 (16%) 4 dpi 6 dpi Differential expression analysis was conducted using the cuffdiff program in Cufflinks version 0.9.3 [52], Cucumis sativus v2 annotation, and false discovery rate of 0.01. ¥ dpi, days post-inoculation. § Numbers in parenthesis indicate the percent of significantly different tests out of the total number of tests that could be performed for each pair-wise comparison. modules that contained between 50 and 428 genes (Table S4.5); 2,241 genes were not assigned to any module. To visualize the relationship of the modules to each other with respect to the progression of infection, eigengenes for each module were calculated and displayed in a heat map [48]. As shown in Figure 4.6, some modules are representative of genes with correlated co-expression at primarily a single or two time points (Modules F, G, H, I, J, and K) whereas other modules represent genes that share broader co-expression patterns across multiple time points (Modules A, B, C, D, E). Examination of trend plots for the modules (Figure 4.7, Figure S4.2) reveals the direction and magnitude of gene expression patterns. Genes within Module B are expressed in the mock control and at 1 dpi, but are coordinately down-regulated at 2 159 dpi, remaining less abundant through 8 dpi (Figure 4.7). This set of 272 genes includes a large suite of genes implicated in resistance including six lipoxygenase genes, four cationic peroxidases, two cinnamate 4-hydroxylases, an anthocyanidin 3- glucosyltransferase, an anthocyanin 5-aromatic acyltransferase, a cysteine protease, and the elicitor-inducible protein EIG-J7. All of these genes have been implicated in defense responses in other plants [49,50,51,52,53,54], and in total, their coordinate down-regulation at 2 dpi is suggestive of an active mechanism by the pathogen to alter host defense responses. For example, the lipoxygenase pathway has been hypothesized to be involved in defense responses in cucumber, due to its expansion within the genome [6], and our data presented herein showing a rapid down-regulation of these transcripts during infection supports this hypothesis. In addition, cinnamate 4hydroxylase, one of the primary enzymes in the phenylpropanoid pathway responsible for the conversion of cinnamic acid to p-coumaric acid, was shown to be rapidly induced in C. sativus in response to abiotic stress [51]. In this context, it is reasonable to hypothesize that the down-regulation of its expression (observed in the present study), as well as that of other genes within the phenylpropanoid pathway, are suggestive of an active virulence mechanism to abrogate host responses associated with stress-induced signaling. In addition to the direct correlation among defense gene expression noted above, this module also includes the coordinated expression of 32 transcription factors that may also have critical roles in regulating genes responsible for the induction of defense signaling in the host. A total of 36 genes of no known function are also in this module, and further examination of their roles in defense responses may provide new 160 Figure 4.6 Figure 4.6 Heat map of eigengenes representing each gene module. The mock control and post-infection time points are represented in columns and the eigengenes for each of the 11 identified co-expression modules [46] are presented in rows. The numbers of genes in each module are given in parentheses. The eigengene values, which range from 0 to 1, are a measure of centrality and indicate the relative expression levels of all genes in the module (see Materials and Methods). dpi, days post-inoculation. insight into critical host genes modulated by virulent pathogens. Modules F, G, H, I, J, and K all have discrete time points where genes are up-regulated compared to the other sampled time points. As shown in Figure S4.2, genes in Module F (162 genes total) have a peak of expression solely at 2 dpi. This module contains 20 transcription factors which could be key to regulating genes from Module G that have peak expression at 3 dpi (Figure 4.7). Likewise, Module G (244 genes total) has 21 161 transcription factors that may regulate genes within Module H (89 genes total) that are coordinately up-regulated at 4 dpi (Figure 4.7). Within all of these modules (F, G, H, I, J, and K, 726 genes total) there are 199 genes with no known function and placement of these genes in transcriptional networks provides a new functional annotation and contextual information on their function. CONCLUSIONS The work described herein represents the first genome-scale analysis of the cucumberdowny mildew interaction in which we catalog gene expression throughout an 8 day infection period and identify differentially expressed genes that could be correlated with pathogen growth and development in planta. With expression profiles for nearly 15,000 genes during a compatible interaction, we have new insight into molecular events at the host-pathogen interface including a suite of defense response-related genes that are down-regulated early upon infection and transcriptional networks that respond in a temporal manner throughout the infection cycle. Most intriguingly, these networks include transcription factors and genes of no known function, which may have a role in the host-pathogen interaction. 162 Figure 4.7 163 Figure 4.7 (cont'd) Trend plots of the normalized gene expression values from six identified gene co-expression modules. Modules consisting of genes downregulated at 2 dpi (Module B), genes up-regulated at a single time point (Modules G, H, I, K) and genes up-regulated at two time-points (Module J) are shown. The number of genes in each module is shown in parentheses. MATERIALS AND METHODS Plant materials, inoculum, and pathogen infection C. sativus cv. 'Vlaspik' was grown in growth chambers maintained at 22 °C with a 12 h light/dark photoperiod. Ps. cubensis MSU-1 was maintained as previously described [22]. The first fully expanded leaf of four-week-old cucumber plants was inoculated on the abaxial surface with 20-30 10 μl droplets of a 1 x 105 sporangia/ml solution. After inoculation, plants were kept at 100% relative humidity for 24 hours in the dark. Plants were returned to growth chambers for disease progression. Sampling and experimental design Samples from two biological replicates were collected at 1, 2, 3, 4, 6, and 8 dpi at the site of inoculation using a #3 cork borer. Additionally, a mock-inoculated sample (i.e.,10 μl dH2O), which was inoculated as described above and kept at 100% relative humidity for 24 hours in the dark was collected at 1 dpi. Samples were frozen immediately in liquid nitrogen and stored at -80 C until use. 164 Library preparation and sequencing Total RNA was isolated from infected leaf discs using the RNeasy Mini Kit (Qiagen, Germantown, MD) and treated with DNase (Promega, Madison, WI) per the manufacturer's instructions. RNA concentration and quality was determined using the Bioanalyzer 2100 (Agilent Technologies, Santa Clara, CA). The mRNA-Seq sample preparation was done using the Illumina mRNA-Seq kit (Illumina, San Diego CA) according to the manufacturer’s protocol. Parallel sequencing was performed using an Illumina Genome Analyzer II (Illumina, Inc., San Diego, CA) at the Research Technology Support Facility (RTSF) at Michigan State University. Each library within a biological experiment was barcoded, six different barcodes for 6 time-points, pooled and run on multiple lanes. Two biological replicates of each time-point were sequenced multiple times and single-end reads between 35 and 42 bp were generated. Reads from both biological replicates were pooled for determining expression abundances. The mock-inoculated sample library was made as described above, but not barcoded and run in a single lane. Processing of mRNA-Seq data mRNA-Seq reads obtained from Illumina Pipeline version 1.3 were quality evaluated on the Illumina purity filter, percent low quality reads, and distribution of phred-like scores at each cycle. Reads were deposited in the National Center for Biotechnology Information Sequence Read Archive under accession number SRP009350. Reads in 165 FASTQ formats were aligned to the C. sativus [6] reference genome using TopHat v1.2.0/Bowtie v0.12.5 [55,56]. A reference annotation of C. sativus (version 2) from the Cucurbit Genomics Database (http://www.icugi.org/cgi-bin/ICuGI/misc/download.cgi) was provided in which a representative isoform, the gene model with the longest CDS, was used to estimate expression of the gene; a total of 23,248 gene models from a total of 25,600 gene models were used. All other isoforms were discarded from the annotation set. The minimum and maximum intron length was set to 5 and 50,000 bp, respectively; all other parameters were set to the default values. Normalized gene expression levels were calculated using Cufflinks v0.9.3 [57] using the quartile normalization option [58] to improve differential expression calculations of lowly expressed genes. The maximum intron length parameter was set to 50,000 bp. All other parameters were used at the default settings. Sampling depth was evaluated on expression estimates by randomly selecting 5, 10, 15, 25 and 30 million reads from the total pool of reads at all time points. Pearson correlation coefficient analyses of log2 FPKM values were performed using R (http://cran.r-project.org/), in which all log2 FPKM values less than zero were set to zero. Only tests significant at p = 0.05 are shown. The correlation values were clustered with hierarchical clustering using a Pearson correlation distance metric with average linkage and depicted as a heat map. For each node, bootstrap support values were calculated from 1000 replicates using Multiple Experiment Viewer Software (MeV) v4.5 [59]. To examine biological variation, PCC was calculated for the log2 transformed FPKM values of the genes expressed in both replicates at a particular time point. 166 Microarray data acquisition and processing Comparative analyses of host gene expression responses during a compatible interaction with the model species A. thaliana utilized microarry-based gene expression data from an H. arabidopsidis-A. thaliana time course experiment [41]. The dataset was comprised of A. thaliana genes expressed in response to H. arabidopsidis infection at 0, 0.5, 2, 4, and 6 dpi using the ATH1 Affymetrix platform. The probe intensity values were downloaded from Gene Expression Omnibus (http://www.ncbi.nlm.nih.gov/geo/; GSE22274) [60], normalized using Robust Multichip Analysis method [61]. OrthoMCL [62] was used to identify clusters of orthologs and close paralogs in cucumber and A. thaliana. For the mRNA-Seq to microarray comparative analysis only single copy orthologous genes were considered for further analyses. The FPKM and intensity values were log2 transformed, and Spearman rank correlations of the single copy orthologs in both hosts were calculated in R (http://cran.r-project.org/). Only tests significant at p = 0.05 are shown. Identification of differentially expressed genes The Cuffdiff program within Cufflinks version 0.9.3 [57] was used to identify differentially expressed genes using pair-wise comparisons of the six time points and the control. The minimum number of alignments at a gene required to test was set to 100. Quartile normalization and a false discovery rate of 0.01 after Benjamini-Hochberg correction for multiple testing were used. Significance in the numbers of differentially expressed 167 genes between time points was tested using the two-sample t-test. Functional analyses of differentially expressed genes Functional annotation for all C. sativus genes were generated from BLAST searches of the UniProt databases (Uniref100) [63] and combined with Pfam (version 24.0, [64]) protein families assignment performed using HMMER3 [65]. C. sativus sequences were functionally annotated based on the best possible UniRef sequence match using a minimum E value cutoff of 1E-5. If there was no UniRef sequence match, functional annotations were assigned using Pfam domains. Transcription factors were annotated based on PFAM domains assignment. Gene co-expression network analysis Gene modules of highly correlated genes were identified using the WGCNA method [47] implemented in R. All gene FPKM expression values were log2 transformed and any transformed FPKM value less than 1 was converted to zero. Genes without variation across the mock sample and 6 time points were filtered out using a Coefficient of Variation (CV = σ/μ) cutoff of 0.4. The β and treecut parameters for WGCNA were 13 and 0.9, respectively. All other parameters were used with their default values. Eigengenes for each gene module [48] were calculated and presented as a heat map. 168 ACKNOWLEDGEMENTS We thank members of the Day lab for critical reading of the manuscript. 169 APPENDIX 170 Figure S4.1 171 Figure S4.1 (cont'd) Concordance of expression values in two biological replicates of Cucumis sativus during infection by Pseudoperonospora cubensis. Reads from different time points were mapped to the C. sativus genome using Bowtie version 0.12.5 and TopHat version 1.2.0. Fragments per kilobase pair of exon model per million fragments mapped (FPKM) values were calculated using Cufflinks version 0.9.3 and the C. sativus genome annotations. For each time point, log2 transformed FPKM values of equal number of genes from both replicates are plotted. R2, correlation coefficient; dpi, day post-inoculation. 172 Figure S4.2 173 Figure S4.2 (cont'd) 174 Figure S4.2 (cont'd) 175 Figure S4.2 (cont'd) Figure S4.2 Trend plots for all 11 modules. All 11 modules generated using WGCNA are shown (Modules A through K). The number of genes in each module is shown in parentheses. 176 REFERENCES 177 REFERENCES 1. Tanurdzic M, Banks JA (2004) Sex-determining mechanisms in land plants. Plant Cell 16 Suppl: S61-71. 2. Lough TJ, Lucas WJ (2006) Integrative plant biology: role of phloem long-distance macromolecular trafficking. Ann Rev Plant Biol 57: 203-232. 3. Hammerschmidt R (1999) PHYTOALEXINS: What have we learned after 60 years? Ann Rev Phytopathol 37: 285-306. 4. Hammerschmidt R (1999) Induced disease resistance: how do induced plants stop pathogens? Physiol Mol Plant Pathol 55: 77-84. 5. Ren Y, Zhang Z, Liu J, Staub JE, Han Y, et al. (2009) An integrated genetic and cytogenetic map of the cucumber genome. PLoS ONE 4: e5795. 6. Huang S, Li R, Zhang Z, Li L, Gu X, et al. (2009) The genome of the cucumber, Cucumis sativus L. Nat Genet 41: 1275-1281. 7. Wóycicki R, Witkowicz J, Gawroński P, Dąbrowska J, Lomsadze A, et al. (2011) The genome sequence of the North-European cucumber (Cucumis sativus L.) unravels evolutionary adaptation mechanisms in plants. PLoS ONE 6: e22728. 8. Guo S, Zheng Y, Joung JG, Liu S, Zhang Z, et al. (2010) Transcriptome sequencing and comparative analysis of cucumber flowers with different sex types. BMC Gen 11: 384. 9. Ando K, Grumet R (2010) Transcriptional profiling of rapidly growing cucumber fruit by 454-pyrosequencing analysis. J Am Soc Hort Sci 135: 291-302. 10. Olczak-Woltman H, Schollenberger M, Niemirowicz-Szczytt K (2009) Genetic background of host-pathogen interaction between Cucumis sativus L. and Pseudomonas syringae pv. lachrymans. J Appl Genet 50: 1-7. 11. Savory EA, Granke LL, Quesada-Ocampo LM, Varbanova M, Hausbeck MK, et al. (2011) The cucurbit downy mildew pathogen Pseudoperonospora cubensis. Mol Plant Pathol 12: 217-226. 178 12. Colucci SJ, Wehner TC, Holmes GJ (2006) The downy mildew epidemic of 2004 and 2005 in the Eastern United States. In: Holmes G, editor. Proc Cucurbitaceae 2006. Raleigh, NC: Universal Press. 13. Lebeda A, Pavelková J, Urban J, Sedláková B (2011) Distribution, host range and disease severity of Pseudoperonospora cubensis on cucurbits in the Czech Republic. J Phytopathol 159: 589-596. 14. Arauz LF, Neufeld KN, Lloyd AL, Ojiambo PS (2010) Quantitative models for germination and infection of Pseudoperonospora cubensis in response to temperature and duration of leaf wetness. Phytopathol 100: 959-367. 15. Thomas C, Inaba T, Cohen Y (1987) Physiological Pseudoperonospora cubensis. Phytopathol 77: 1621-1624. Specialization in 16. Shetty NV, Wehner TC, Thomas CE, Doruchowski RW, Vasanth Shetty KP (2002) Evidence for downy mildew races in cucumber tested in Asia, Europe, and North America. Scientia Hort 94: 231–239. 17. Lebeda A, Widrlechner MP (2003) A set of Cucurbitaceae taxa for differentiation of Pseudoperonespora cubensis pathotypes. J Plant Dis Prot 110: 337–349. 18. Cohen Y, Meron I, Mor N, Zuriel S (2003) A new pathotype of Pseudoperonospora cubensis causing downy mildew in cucurbits in Israel. Phytoparasitica 31: 458– 466. 19. Lebeda A (2002) Pathogenic variation of Pseudoperonospora cubensis in the Czech Republic and some other European countries. II International Symposium on Cucurbits 588. pp. 137–141 20. Lebeda A, Urban J (2004) Distribution, harmfulness and pathogenic variability of cucurbit downy mildew in the Czech Republic. Acta Fytotech Zootech 7: 170-173. 21. Sarris P, Abdelhalim M, Kitner M, Skandalis N, Panopoulos N, et al. (2009) Molecular polymorphisms between populations of Pseudoperonospora cubensis from Greece and the Czech Republic and the phytopathological and phylogenetic implications. Plant Pathol 58: 933–943. 179 22. Tian M, Win J, Savory E, Burkhardt A, Held M, et al. (2011) 454 Genome sequencing of Pseudoperonospora cubensis reveals effector proteins with a QXLR translocation motif. Mol Plant-Microbe Interact 24: 543-553. 23. Savory EA, Zou C, Adhikari BN, Hamilton JP, Buell CR, et al. (2012) Alternative Splicing of a Multi-Drug Transporter from Pseudoperonospora cubensis Generates an RXLR Effector Protein That Elicits a Rapid Cell Death. PLoS One 7: e34701. 24. Durrant WE, Dong X (2004) Systemic acquired resistance. Ann Rev Phytopathol 42: 185-209. 25. Phuntumart V, Marro P, Métraux J-P, Sticher L (2006) A novel cucumber gene associated with systemic acquired resistance. Plant Sci 171: 555-564. 26. Sticher L, Mauch-Mani B, Metraux JP (1997) Systemic acquired resistance. Ann Rev Phytopathol 35: 235-270. 27. Bent AF, Mackey D (2007) Elicitors, effectors, and R genes: The new paradigm and a lifetime supply of questions. Ann Rev Phytopathol 45: 399-436. 28. Call AD, Wehner T (2010) Gene list 2010 for cucumber. Cucurbi Genet Coop Rep 28-29: 105-141. 29. Runge F, Choi Y-J, Thines M (2011) Phylogenetic investigations in the genus Pseudoperonospora reveal overlooked species and cryptic diversity in the P. cubensis species cluster. Eur J Plant Pathol 129: 135-146. 30. Taler D, Galperin M, Benjamin I, Cohen Y, Kenigsbuch D (2004) Plant eR genes that encode photorespiratory enzymes confer resistance against disease. Plant Cell 16: 172-184. 31. Criswell A (2008) Screening cucumber (Cucumis sativus) for resistance to downy mildew (Pseudoperonospora cubensis). NCSU Thesis. 32. Wang ET, Sandberg R, Luo S, Khrebtukova I, Zhang L, et al. (2008) Alternative isoform regulation in human tissue transcriptomes. Nature 456: 470 - 476. 180 33. Boddu J, Cho S, Muehlbauer GJ (2007) Transcriptome analysis of trichotheceneinduced gene expression in barley. Mol Plant-Microbe Interact 20: 1364-1375. 34. Huibers RP, de Jong M, Dekter RW, Van den Ackerveken G (2009) Diseasespecific expression of host genes during downy mildew infection of Arabidopsis. Mol Plant-Microbe Interact 22: 1104-1115. 35. Gupta S, Chakraborti D, Sengupta A, Basu D, Das S (2010) Primary metabolism of chickpea is the initial target of wound inducing early sensed Fusarium oxysporum f. sp. ciceri race I. PLoS ONE 5: e9030. 36. Savory EA, Adhikari BN, Hamilton JP, Vaillancourt B, Buell CR, et al. (2012) mRNASeq Analysis of the Pseudoperonospora cubensis Transcriptome During Cucumber (Cucumis sativus L.) Infection. PLoS One 7: e35796. 37. Coates ME, Beynon JL (2010) Hyaloperonospora arabidopsidis as a pathogen model. Ann Rev Phytopathol 48: 329-345. 38. Volk GM, Haritatos EE, Turgeon R (2003) Galactinol synthase gene expression in melon. J Am Soc Hort Sci 128: 8-15. 39. Li JW, Liu J, Zhang H, Xie CH (2011) Identification and transcriptional profiling of differentially expressed genes associated with resistance to Pseudoperonospora cubensis in cucumber. Plant Cell Rep 30: 345-357. 40. Knepper C, Day B (2010) From perception to activation: The molecular-genetic and biochemical landscape of disease resistance signaling in plants. The Arabidopsis Book: e012. 41. Wang W, Barnaby JY, Tada Y, Li H, Tor M, et al. (2011) Timing of plant immune responses by a central circadian regulator. Nature 470: 110-114. 42. Miranda M, Ralph SG, Mellway R, White R, Heath MC, et al. (2007) The transcriptional response of hybrid poplar (Populus trichocarpa x P. deltoides) to infection by Melampsora medusae leaf rust involves induction of flavonoid pathway genes leading to the accumulation of proanthocyanidins. Mol PlantMicrobe Interact 20: 816-831. 181 43. Polesani M, Bortesi L, Ferrarini A, Zamboni A, Fasoli M, et al. (2010) General and species-specific transcriptional responses to downy mildew infection in a susceptible (Vitis vinifera) and a resistant (V. riparia) grapevine species. BMC Gen 11: 117. 44. Polesani M, Desario F, Ferrarini A, Zamboni A, Pezzotti M, et al. (2008) cDNA-AFLP analysis of plant and pathogen genes expressed in grapevine infected with Plasmopara viticola. BMC Gen 9: 142. 45. Gaulin E, Madoui MA, Bottin A, Jacquet C, Mathe C, et al. (2008) Transcriptome of Aphanomyces euteiches: new oomycete putative pathogenicity factors and metabolic pathways. PLoS ONE 3: e1723. 46. Langfelder P, Zhang B, Horvath S (2008) Defining clusters from a hierarchical cluster tree: the Dynamic Tree Cut package for R. Bioinformatics 24: 719 - 720. 47. Zhang B, Horvath S (2005) A general framework for weighted gene co-expression network analysis. Stat App Gen Mol Biol 4: Article 17. 48. Langfelder P, Horvath S (2007) Eigengene networks for studying the relationships between co-expression modules. BMC Sys Biol 1: 54. 49. Takemoto D, Demirdöven A (2007) Lipoxygenase in fruits and vegetables: A review. Enz Micro Tech 40: 491-496. 50. Hilaire E, Young SA, Willard LH, McGee JD, Sweat T, et al. (2001) Vascular defense responses in rice: peroxidase accumulation in xylem parenchyma cells and xylem wall thickening. Mol Plant-Microbe Interact 14: 1411-1419. 51. Varbanova M, Porter K, Lu F, Ralph J, Hammerschmidt R, et al. (2011) Molecular and biochemical basis for stress-induced accumulation of free and bound pcoumaraldehyde in cucumber. Plant Physiol 157: 1056-1066. 52. Brader G, Djamei A, Teige M, Palva ET, Hirt H (2007) The MAP kinase kinase MKK2 affects disease resistance in Arabidopsis. Mol Plant-Microbe Interact 20: 589-596. 53. Bernoux M, Timmers T, Jauneau A, Brière C, de Wit PJGM, et al. (2008) RD19, an Arabidopsis cysteine protease required for RRS1-R-mediated resistance, Is 182 relocalized to the nucleus by the Ralstonia solanacearum PopP2 effector. Plant Cell 20: 2252-2264. 54. Takemoto D, Doke N, Kawakita K (2001) Characterization of elicitor-inducible tobacco genes isolated by differential hybridization. J Gen Plant Pathol 67: 8996. 55. Trapnell C, Pachter L, Salzberg SL (2009) TopHat: discovering splice junctions with RNA-Seq. Bioinformatics 25: 1105-1111. 56. Langmead B, Trapnell C, Pop M, Salzberg S (2009) Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol 10: R25. 57. Trapnell C, Williams BA, Pertea G, Mortazavi A, Kwan G, et al. (2010) Transcript assembly and quantification by RNA-Seq reveals unannotated transcripts and isoform switching during cell differentiation. Nat Biotechnol 28: 511-515. 58. Bullard J, Purdom E, Hansen K, Dudoit S (2010) Evaluation of statistical methods for normalization and differential expression in mRNA-Seq experiments. BMC Bioinformatics 11: 94. 59. Saeed AI, Bhagabati NK, Braisted JC, Liang W, Sharov V, et al. (2006) TM4 microarray software suite. Methods Enzymol 411: 134-193. 60. Edgar R, Domrachev M, Lash AE (2002) Gene Expression Omnibus: NCBI gene expression and hybridization array data repository. Nucleic Acids Res 30: 207210. 61. Irizarry RA, Hobbs B, Collin F, Beazer-Barclay YD, Antonellis KJ, et al. (2003) Exploration, normalization, and summaries of high density oligonucleotide array probe level data. Biostat 4: 249 - 264. 62. Li L, Stoeckert CJ, Jr., Roos DS (2003) OrthoMCL: identification of ortholog groups for eukaryotic genomes. Genome Res 13: 2178-2189. 63. Suzek BE, Huang H, McGarvey P, Mazumder R, Wu CH (2007) UniRef: comprehensive and non-redundant UniProt reference clusters. Bioinformatics 23: 1282-1288. 183 64. Bateman A, Birney E, Durbin R, Eddy SR, Howe KL, et al. (2000) The Pfam Protein Families Database. Nucleic Acids Res 28: 263-266. 65. Finn RD, Clements J, Eddy SR (2011) HMMER web server: interactive sequence similarity searching. Nucleic Acids Res 39: W29-37. 184 CHAPTER 5 Conclusions and Future Directions 184 CONCLUSIONS Cucurbit downy mildew, caused by the oomycete pathogen Pseudoperonospora cubensis, emerged as a significant economic threat to cucumber production in the United States during the 2004 and 2005 growing seasons and is currently the major limiting factor of cucumber production [1]. Although downy mildew had been a major issue in Europe since the mid-1980's, in the U.S., the disease on cucumber had been successfully controlled since the 1940's through host resistance. As such, only limited research was conducted to understand the biology, genetics, and virulence of the pathogen. The focus of this dissertation was to expand knowledge of this economically important oomycete pathogen using cell biology, genetic, genomic, and transcriptomic approaches to better understand the molecular-genetic basis for the interaction between cucumber (Cucumis sativus) and Ps. cubensis. The establishment of genetic resources in the form of a draft genome assembly laid the groundwork for additional analyses, including identification and characterization of candidate effector proteins (Chapter 2) as well as providing a framework for mapping mRNA-Seq reads from a large-scale gene expression study of infection (Chapter 3). One RXLR-type effector protein, PscRXLR1 was characterized in detail and shown to be the result of alternative splicing of a non-effector protein (Chapter 2). Additionally, due to the obligate nature of Ps. cubensis and our sampling methods for the mRNA-Seq study, we were able to measure host gene expression changes during C. sativus infection with Ps. cubensis, providing the first in-depth analysis of this important plantpathogen interaction (Chapter 4). Overall, the research presented in this dissertation 185 greatly expands our knowledge of the C. sativus-Ps. cubensis interaction and provides much-needed insight into the virulence of Ps. cubensis and the defense response of C. sativus. The establishment of genomic resources for major oomycete pathogens such as Phytophthora infestans [2], Phytophthora sojae, Phytophthora ramorum [3], Pythium ultimum [4], as well as for the Arabidopsis thaliana pathogen, Hyaloperonospora arabidopsidis [5], and subsequent identification of RXLR and RXLR-like effector proteins has anchored the recent evolution of molecular plant pathology [6]. The study of oomycete effector proteins has been central to the study of oomycete pathogenicity and virulence and the understanding of the interactions between these pathogens and their hosts [7,8,9,10,11,12,13,14,15,16]. Therefore, the first step in identifying these important pathogenicity and virulence determinants in Ps. cubensis was the generation of a 64.4 Mbp genomic assembly of the MSU-1 isolate and prediction of 23,519 loci and 23,522 gene models (Chapter 2). Due to their modular structure, including an N- terminal signal peptide, an RXLR translocation motif, and a C-terminal functional domain, RXLR effectors are readily identified via bioinformatics [17,18]. We identified 271 candidate effector proteins within the Ps. cubensis genome with variable RXLR motifs, including the previously identified R and Q, predicting 20 possible amino acids at position R1. Additionally, mRNA-Seq analysis of infection provided expression support for 19 of the 20 possible R1 substitutions (except Y, Tyr; Table S2.1, Figure 3.5). While candidate effectors with these R1 amino acid substitutions, with the exception of QXLR [19], have yet to be functionally validated, their expression during a compatible 186 interaction supports their putative virulence function and role as bona fide effector proteins. This variation may provide clues to the diversity among Ps. cubensis pathotypes in terms of their virulence and host specificity. In-depth functional analysis of one Ps. cubensis effector protein, RXLR protein 1 (PscRXLR1) led to a surprising and interesting result - it was demonstrated that PscRXLR1 arises as a product of alternative splicing (Figure 2.5), making this the first example of an alternative splicing event in plant pathogenic oomycetes transforming a non-effector gene into a functional effector protein. Using a set of experiments designed to validate effector protein function, it was shown that PscRXLR1 was upregulated during the early stages of infection and that heterologous expression of PscRXLR1 in Nicotiana benthamiana elicits a rapid cell death phenotype (Figure 2.3, Figure 2.4). Additionally, characterization of the closest P. infestans ortholog, PITG_17484, a member of the Drug/Metabolite Transporter (DMT) superfamily to Psc_RXLR1 confirmed that it did not elicit the same cell death response (Figure 2.3). The orthologous relationship between PscRXLR1 and PITG_17484 led us to question if such effector-non-effector ortholog pairs are common among oomycete plant pathogens. We subsequently examined the relationship(s) among putative ortholog pairs in Ps. cubensis and P. infestans. Of 271 predicted Ps. cubensis effector proteins, only 109 (41%) had a putative ortholog in P. infestans and evolutionary rate analysis of these orthologs showed an evolutionary rate significantly faster than most other genes (Figure 2.1). In total, these data provide a basis for comparative analysis of candidate effector proteins and their non-effector orthologs as a means of understanding function 187 and evolutionary history of pathogen effectors. In addition, the discovery of an effector protein arising from an alternative splicing event may indicate an adaptive mechanism utilized by Ps. cubensis and potentially other pathogens to generate proteome complexity during infection. Next generation sequencing of the transcriptome (mRNA-Seq) permits deep, robust assessments of transcript abundance and structure, and when applied to host-pathogen interactions, enables insight into pathogen mechanisms to suppress and subvert host defense responses [20,21,22,23]. In this dissertation, the first large-scale global transcriptome analysis of Ps. cubensis infection of a susceptible C. sativus cultivar, ‘Vlaspik’ is presented, yielding new information about both pathogen and host gene expression during a compatible interaction (Chapter 3, 4). Sampling Ps. cubensis- infected C. sativus tissue during an 8-day time course and then separating speciesspecific reads in silico enabled simultaneously collection and analysis of infection transcriptome dynamics from both pathogen and host. As described herein, we were able to detect the expression of 7,821 Ps. cubensis genes (Chapter 3) and 14,476 cucumber genes (Chapter 4) throughout the infection process from 1 day postinoculation (1 dpi) to 8 dpi, providing the first genome-scale analysis of the cucurbit downy mildew interaction. In Ps. cubensis, genes involved in virulence, including RXLR and RXLR-like effectors, Crinkler (CRN) effectors, and host-targeted hydrolytic enzymes acting on plant proteinases, lipases, and several sugar-cleaving enzymes were all differentially 188 expressed throughout the infection time course. Corresponding with our visual assessment of symptoms and infection structures, clustering and co-expression analyses identified distinct modules of Ps. cubensis genes that were representative of early, intermediate, and late infection stages. The 1 dpi time point, when encystment of zoospores is occurring, represents a unique pattern of gene expression supported by multiple modes of analysis. Gene expression at this time point poorly correlated with all other time points (Figure 3.4), had the highest percentage of differentially expressed genes across all pair-wise comparisons (Table 3.1), and was represented by a unique module, Module 1 (Figure 3.8), when analyzed using Weighted Gene Correlation Network Analysis (WGCNA). The genes represented here are likely those involved in zoospore encystment, appressorium production, and the initial penetration of stomata. Our data also supports an intermediate, transition stage of infection, represented mainly by overlapping gene expression between the 2, 3, and 4 dpi time points. Module 2, containing 508 genes including candidate RXLR-type effectors, CRNs, and haustoriumspecific proteins, represents expression during hyphal penetration, growth, and initiation of haustoria formation, which occurs between 2-4 dpi. The late stage of infection, at 6 and 8 dpi, is characterized by extensive hyphal growth within the mesophyll (Figure 3.1) and a transition to the expression of genes likely involved in sporulation (Module 6, Figure 3.8). Overall, these expression data and analyses and their correlation with pathogen growth have advanced our understanding of molecular and genetic events in the infection of Ps. cubensis. 189 During Ps. cubensis infection, expression of defense-related genes, including catalases, chitinases, lipoxygenases, peroxidases, and protease inhibitors, was highly upregulated in C. sativus within 1 dpi (Table S4.2), suggesting an active host defense response to early infection by Ps. cubensis. These genes are coordinately downregulated at 2 dpi (Module B, Figure 4.7) and remain lowly expressed for the continuation of the time course, suggesting that they are suppressed by the pathogen. With expression profiles for nearly 15,000 genes during a compatible interaction, we have new insight into molecular events at the host-pathogen interface including a suite of defense response-related genes that are down-regulated early upon infection and transcriptional networks that respond in a temporal manner throughout the infection cycle. Most intriguingly, these networks include transcription factors and genes of no known function, which may have a role in the host-pathogen interaction. Overall, the work presented in this dissertation represents a substantial advance in the understanding of the Ps. cubensis-C. sativus interaction. Using a variety of methods ranging from cell biology to genetics to large-scale transcriptomic analysis, it serves to advance knowledge of Ps. cubensis biology, genomics, virulence determinants, and gene expression during infection, as well as corresponding host responses. 190 FUTURE DIRECTIONS While substantial progress has been made over the course of this dissertation project to establish genetic, genomic, and transcriptomic resources for the study of the Ps. cubensis-C. sativus interaction, it only represents a start to the study of this important and fascinating pathosystem. To build on the work herein, further functional characterization of additional candidate effector proteins, exploration of alternative splicing events over the course of infection, and a more detailed analysis of the Ps. cubensis transcriptome data set is needed to provide additional insight into both the lifestyle and virulence of Ps. cubensis. For C. sativus, additional analysis of the mRNASeq data would increase our knowledge of host defense responses in cucumber, and more transcriptome analyses of lines with resistance to Ps. cubensis could potentially lead to discoveries of new sources of resistance which could be utilized in breeding of resistant varieties for growers. This dissertation and recent work [19] have identified 271 candidate effector proteins with variable RXLR motifs and established a pipeline for their characterization using both bioinformatics and molecular biology tools. The prediction of 20 possible amino acid substitutions at the R1 position has expanded our previous knowledge of the distribution of R1 domains in putative Ps. cubensis effector proteins, previously thought to be equally distributed between RXLR and QXLR motifs [19]. While expression support exists for 19 of these 20 substitutions, additional functional validation is necessary and would add to both general knowledge of oomycete and fungal effector 191 proteins and provide insight into the possible mechanisms utilized by Ps. cubensis to cause infection. Functional characterization of the effector protein PscRXLR1 demonstrated that it arises as a product of alternative splicing from Psc_781.4, making it the first example of an alternative splicing event in plant pathogenic oomycetes transforming a non-effector gene into a functional effector protein. This was substantiated using a combination of RT-PCR and transcriptome analysis of Ps. cubensis gene expression during infection. To build upon this work, analysis to quantify changes in expression levels of PscRXLR1 and Psc_781.4 over the course of infection should be completed. Using qRT-PCR, an up-regulation of mRNA was identified at 1 dpi, extending through 4 dpi. However, due to limitations with this technique, we were unable to distinguish between PscRXLR1 and Psc_781.4. More in depth analysis of alternative splicing events in Ps. cubensis during infection will allow us to distinguish between these isoforms, as well as enable us to identify additional changes in splicing over the course of infection. In addition, such analyses would enable the identification of global splicing events in Ps. cubensis, which would expand general knowledge about alternative splicing changes during infection, an as-yet understudied phenomenon. The large-scale mRNA-Seq data sets generated in this dissertation provide a wealth of knowledge about gene expression during Ps. cubensis infection on C. sativus, and while extensive, the analyses and data presented herein only represent a small fraction of what could be gleaned from them. The analyses for this project, focused on genes 192 involved in pathogenicity and infection, such as effector proteins, pathogenicity-related genes, and CAZymes. Ps. cubensis is an obligate biotroph and is dependent on C. sativus for completion of its life cycle. As such, this mRNA-Seq dataset would be ideal for analyses focused on identifying genes involved in obligate biotrophy and the establishment of this intricate relationship between pathogen and host. This would not only yield insight into the nature of obligate biotrophic relationships which could be extrapolated to other systems, but could also be useful in understanding the pathogenicity of Ps. cubensis, potentially leading to improved control strategies. Monitoring the expression of both pathogen and host genes during infection can provide insight into the interplay between resistance and susceptibility. Initial analysis of the mRNA-Seq data focused on studying genes and expression patterns that were either Ps. cubensis or C. sativus specific. This data set is unique in that expression data from both host and pathogen were collected from the same samples and as such, ideal for correlative analyses that could look at the co-regulation of gene expression between pathogen and host. This could potentially aide in the identification of pathogen effectors that are down-regulating host defense responses or specific host genes that are upregulated in response to pathogen virulence determinants. Looking specifically at the C. sativus mRNA-Seq data in more detail could increase our knowledge of host defense responses in cucumber. We identified a suite of genes upregulated at 1 dpi and subsequently down-regulated at 2 dpi (Figure 4.7) - functional characterization of a set of these genes could provide insight into the host defense response of C. sativus, as one could hypothesize that they could potentially play a role in defense if pathogen infection results in their down-regulation. Finally, the mRNA-Seq 193 data presented herein represents a compatible interaction between a virulent pathotype of Ps. cubensis and a susceptible C. sativus cv., 'Vlaspik'. While ideal for studying pathogen virulence determinants, to better study the host defense response, additional transcriptome analyses of C. sativus lines with resistance to Ps. cubensis should be done. This could potentially lead to discoveries of new sources of resistance and breeding of resistant varieties for growers. The Ps. cubensis-C. sativus interaction, while in its infancy in terms of our understanding of the genetic and biochemical processes that regulate pathogenicity and resistance, presents an opportunity to develop a crop-based model system in which to study oomycete-plant interactions, ultimately focused on identifying mechanisms of host resistance that can be translated into field application. The work presented in this dissertation lays a substantial groundwork to achieve that goal by establishing genomic, genetic and transcriptomic resources for both Ps. cubensis and C. sativus that can be built upon to better understand the interplay between pathogen and host. 194 REFERENCES 195 REFERENCES 1. Savory EA, Granke LL, Quesada-Ocampo LM, Varbanova M, Hausbeck MK, et al. (2011) The cucurbit downy mildew pathogen Pseudoperonospora cubensis. Mol Plant Pathol 12: 217-226. 2. Haas BJ, Kamoun S, Zody MC, Jiang RH, Handsaker RE, et al. (2009) Genome sequence and analysis of the Irish potato famine pathogen Phytophthora infestans. Nature 461: 393-398. 3. Tyler BM (2006) Phytophthora Genome Sequences Uncover Evolutionary Origins and Mechanisms of Pathogenesis. Science 313: 1261-1266. 4. Levesque CA, Brouwer H, Cano L, Hamilton JP, Holt C, et al. (2010) Genome sequence of the necrotrophic plant pathogen Pythium ultimum reveals original pathogenicity mechanisms and effector repertoire. Genome Biol 11: R73. 5. Baxter L, Tripathy S, Ishaque N, Boot N, Cabral A, et al. (2010) Signatures of adaptation to obligate biotrophy in the Hyaloperonospora arabidopsidis genome. Science 330: 1549-1551. 6. Birch PR, Armstrong M, Bos J, Boevink P, Gilroy EM, et al. (2009) Towards understanding the virulence functions of RXLR effectors of the oomycete plant pathogen Phytophthora infestans. J Exp Bot 60: 1133-1140. 7. Allen RL (2004) Host-Parasite Coevolutionary Conflict Between Arabidopsis and Downy Mildew. Science 306: 1957-1960. 8. Shan W, Cao M, Leung D, Tyler BM (2004) The Avr1b locus of Phytophthora sojae encodes an elicitor and a regulator required for avirulence on soybean plants carrying resistance gene Rps 1b. Mol Plant-Microbe Interact 17: 394–403. 9. Rehmany AP, Gordon A, Rose LE, Allen RL, Armstrong MR, et al. (2005) Differential recognition of highly divergent downy mildew avirulence gene alleles by RPP1 resistance genes from two Arabidopsis lines. Plant Cell 17: 1839-1850. 10. Armstrong MR, Whisson SC, Pritchard L, Bos JI, Venter E, et al. (2005) An ancestral oomycete locus contains late blight avirulence gene Avr3a, encoding a protein that is recognized in the host cytoplasm. Proc Natl Acad Sci U S A 102: 7766-7771. 196 11. Bos JI, Kanneganti TD, Young C, Cakir C, Huitema E, et al. (2006) The C-terminal half of Phytophthora infestans RXLR effector AVR3a is sufficient to trigger R3amediated hypersensitivity and suppress INF1-induced cell death in Nicotiana benthamiana. Plant J 48: 165-176. 12. Whisson SC, Boevink PC, Moleleki L, Avrova AO, Morales JG, et al. (2007) A translocation signal for delivery of oomycete effector proteins into host plant cells. Nature 450: 115-118. 13. Qutob D, Tedman-Jones J, Dong S, Kuflu K, Pham H, et al. (2009) Copy Number Variation and Transcriptional Polymorphisms of Phytophthora sojae RXLR Effector Genes Avr1a and Avr3a. PLoS ONE 4: e5066. 14. Dong S, Qutob D, Tedman-Jones J, Kuflu K, Wang Y, et al. (2009) The Phytophthora sojae Avirulence Locus Avr3c Encodes a Multi-Copy RXLR Effector with Sequence Polymorphisms among Pathogen Strains. PLoS ONE 4: e5556. 15. Bos JI, Armstrong MR, Gilroy EM, Boevink PC, Hein I, et al. (2010) Phytophthora infestans effector AVR3a is essential for virulence and manipulates plant immunity by stabilizing host E3 ligase CMPG1. Proc Natl Acad Sci U S A 107: 9909-9914. 16. Cabral A, Stassen JH, Seidl MF, Bautor J, Parker JE, et al. (2011) Identification of Hyaloperonospora arabidopsidis Transcript Sequences Expressed during Infection Reveals Isolate-Specific Effectors. PLoS ONE 6: e19328. 17. Schornack S, Huitema E, Cano LM, Bozkurt TO, Oliva R, et al. (2009) Ten things to know about oomycete effectors. Mol Plant Pathol 10: 795–803. 18. Kamoun S (2006) A catalogue of the effector secretome of plant pathogenic oomycetes. Ann Rev Phytopathol 44: 41-60. 19. Tian M, Win J, Savory E, Burkhardt A, Held M, et al. (2011) 454 Genome sequencing of Pseudoperonospora cubensis reveals effector proteins with a QXLR translocation motif. Mol Plant-Microbe Interact 24: 543-553. 20. Wang E, Sandberg R, Luo S, Khrebtukova I, Zhang L, et al. (2008) Alternative isoform regulation in human tissue transcriptomes. Nature 456: 470 - 476. 197 21. Boddu J, Cho S, Muehlbauer GJ (2007) Transcriptome analysis of trichotheceneinduced gene expression in barley. Mol Plant-Microbe Interact 20: 1364-1375. 22. Huibers RP, de Jong M, Dekter RW, Van den Ackerveken G (2009) Diseasespecific expression of host genes during downy mildew infection of Arabidopsis. Mol Plant-Microbe Interact 22: 1104-1115. 23. Gupta S, Chakraborti D, Sengupta A, Basu D, Das S (2010) Primary metabolism of chickpea is the initial target of wound inducing early sensed Fusarium oxysporum f. sp. ciceri race I. PLoS ONE 5: e9030. 198