GENETIC MONITORING OF CUCURBIT DOWNY MILDEW IN MICHIGAN By Julian Camilo Bello Rodriguez A DISSERTATION Submitted to 2020 Michigan State University in partial fulfillment for the requirements for the degree of Plant Pathology – Doctor of Philosophy ABSTRACT GENETIC MONITORING OF CUCURBIT DOWNY MILDEW IN MICHIGAN By Julian Camilo Bello Rodriguez Cucurbit downy mildew (CDM) caused by the oomycete obligate pathogen, Pseudoperonospora cubensis, incites foliar blighting of several cucurbit genera worldwide. In 2004, the pathogen re-emerged in the U.S. infecting historically resistant cucumber cultivars and requiring the adoption of an intensive fungicide program. Due to an influx of aerially dispersed sporangia from overwinter sources, CDM occurs annually in cucumber growing regions of northern U.S. The genetic monitoring of incoming P. cubensis populations is essential for growers to make informed decisions regarding CDM management strategies. However, the scale and resolution of genetic studies of downy mildews (Peronosporaceae) remains limited due to the logistical constraints involved in the genotyping of these species (e.g. obtaining DNA of sufficient quantity and quality). To gain an evolutionary and ecological perspective of P. cubensis, we describe a targeted enrichment (TE) protocol able to genotype environmental samples of Pseudoperonospora spp. using less than 50 ng of DNA for library preparation. Using the TE protocol, we were able to enrich 736 target genes across 101 samples and identified 2,978 high quality SNP variants. This SNPs resolved the population structure of P. cubensis in Michigan and detected significant (AMOVA, P=0.01) genetic differentiation among the P. cubensis populations from squash (clade I) and cucumber (clade II). No evidence of location-based differentiation was detected within the P. cubensis (clade II) subpopulation of Michigan. Timely alerts of an influx of airborne inoculum of two distinct host-adapted clades of P. cubensis can assist Michigan growers in assessing the need to initiate fungicide sprays. However, the inability to distinguish between the morphologically identical sporangia of P. humuli and P. cubensis has been a significant shortcoming. Using spore traps and qPCR assays, an improved methodology for the aerial monitoring of each Pseudoperonospora taxa was identified. A highly specific qPCR assay differentiated Pseudoperonospora humuli, the causal agent of downy mildew on hop, and the two host-adapted clades of P. cubensis (clade I and II) on spore trap samples. After two years of monitoring, P. cubensis clade II DNA was detected in spore trap samples >2 days before CDM symptoms were first observed in commercial cucumber fields (August), while P. humuli DNA was only detected early in the growing season (May and June). P. cubensis clade I DNA was not detected in air samples before or after the disease onset in cucumber fields. Additionally, the probability for P. cubensis detection in Burkard spore trap samples was higher compared to impaction spore trap samples with approximately the same number of sporangia, suggesting that the efficiency of recovery of sporangia by Burkard spore traps exceeds the recovery of impaction spore traps. The methodology described in this study to monitor the airborne concentrations of Pseudoperonospora spp. sporangia could be used as part of a CDM risk advisory system to time fungicide applications that protect cucurbit crops in Michigan. ACKNOWLEDGMENTS I would like to thank my co-advisor, Dr. Mary Hausbeck, for taking me as her student and for the work-related support she provided to me over the years. It was an honor to be part of her program and I'm thankful for your patience with my stubbornness. I would also like to thank my co-advisor Dr. Monique Sakalidis, for her guidance and questions, they challenged me and encourage me to continue in the graduate program. In addition, Dr. Andy Jarosz and Dr. Timothy Miles, for their invaluable advice as members of my committee. I could not have achieved my degree without the assistance of the technical staff and graduate students of the Hausbeck and Sakalidis Labs. Special thanks to Doug Higgins, my program fellow, for the countless hours of discussions after 5 p.m. Your insights, critics, and suggestions were essential to complete my objectives. I thank my parents for investing in my undergraduate program and for being there for me under any circumstances. I love you both. I also thank my sisters and brother; I made it to this program because you are always challenging me. And last, but not least, I would like to thank my good friends, Marcela Tabares, Oscar Benitez, and Clarissa Strieder, for their emotional support during the hardest times of my time far from home, you guys help me build the best version of myself. iv TABLE OF CONTENTS LIST OF TABLES ........................................................................................................... vii LIST OF FIGURES .......................................................................................................... ix CHAPTER I: LITERATURE REVIEW - PSEUDOPERONOSPORA CUBENSIS THE CAUSAL AGENT OF CUCURBIT DOWNY MILDEW ..................................................... 1 INTRODUCTION .......................................................................................................... 2 TAXONOMY AND DISEASE CYCLE ........................................................................... 5 SURVIVAL AND DISPERSAL OF P. CUBENSIS ........................................................ 8 GENETIC STRUCTURE ............................................................................................. 10 PATHOGENIC VARIATION ........................................................................................ 12 FUNGICIDE RESISTANCE ........................................................................................ 13 BREEDING EFFORTS ............................................................................................... 18 EARLY DETECTION OF P. CUBENSIS ..................................................................... 21 LITERATURE CITED ..................................................................................................... 25 CHAPTER II: DETECTION OF AIRBORNE SPORANGIA OF PSEUDOPERONOSPORA CUBENSIS AND P. HUMULI IN MICHIGAN USING BURKARD SPORE TRAPS COUPLED TO QPCR ....................................................... 39 ABSTRACT ................................................................................................................. 40 INTRODUCTION ........................................................................................................ 41 MATERIALS AND METHODS .................................................................................... 43 RESULTS ................................................................................................................... 48 DISCUSSION .............................................................................................................. 54 APPENDIX ..................................................................................................................... 60 LITERATURE CITED ..................................................................................................... 71 CHAPTER III: OPTIMIZING SPORE TRAPS AND QUANTITATIVE PCR ASSAYS FOR THE MONITORING OF CUCURBIT DOWNY MILDEW. ...................................... 77 ABSTRACT ................................................................................................................. 78 INTRODUCTION ........................................................................................................ 79 MATERIALS AND METHODS .................................................................................... 82 RESULTS ................................................................................................................... 86 DISCUSSION .............................................................................................................. 93 APPENDIX ................................................................................................................... 101 LITERATURE CITED ................................................................................................... 127 v CHAPTER IV: GENOTYPING OF THE OBLIGATE PLANT PATHOGENS PSEUDOPERONOSPORA CUBENSIS AND P. HUMULI USING TARGET ENRICHMENT SEQUENCING .................................................................................... 133 ABSTRACT ............................................................................................................... 134 INTRODUCTION ...................................................................................................... 135 MATERIALS AND METHODS .................................................................................. 138 RESULTS ................................................................................................................. 148 DISCUSSION ............................................................................................................ 156 APPENDIX ................................................................................................................... 164 LITERATURE CITED ................................................................................................... 194 CHAPTER V ................................................................................................................. 202 CONCLUSIONS ........................................................................................................ 203 vi LIST OF TABLES Table 1-1. List of fungicides registered for the control of cucumber downy mildew in the U.S. ................................................................................................................................. 16 Table 2-1. Primers and locked nucleic acids (LNA) probes for the qPCR assay differentiating Pseudoperonospora cubensis and P. humuli using the 105 SNP in the mitochondrial Cox2 gene. ............................................................................................... 61 Table 2-2. Threshold cycle (Cq) values of the qPCR assays using LNA probes and varying concentrations of genomic DNA. ....................................................................... 62 Table 3-1. qPCR assays designed for the differentiation of Pseudoperonospora humuli and Pseudoperonospora cubensis clade 1 and 2 targeting mitochondrial DNA regions. ...................................................................................................................................... 101 Table 3-2. Threshold cycle (Cq) values of the qPCR protocol B using varying concentrations of genomic DNA from Pseudoperonospora cubensis clade 1 and 2 and P. humuli....................................................................................................................... 104 Table 3-3. Percentage of qPCR positive samples collected by Burkard and impaction spore traps ................................................................................................................... 106 Table 3-4. Linear regression analysis of Cq values as a function of the number Pseudoperonospora spp. sporangia ............................................................................ 107 Table 3-5. b estimates of logistical models developed to predict the probability of Pseudoperonospora cubensis detection in Burkard and impaction spore trap samples ...................................................................................................................................... 108 Table 3-6. Detection of Pseudoperonospora cubensis using Burkard spore traps coupled with qPCR and detection of symptoms ........................................................... 109 Table 4-1. Sequencing and alignment results from libraries sequenced using target enrichment (TE), low coverage whole genome sequencing (Lc-WGS) and whole- genome sequencing (WGS) ......................................................................................... 164 Table 4-2. Plant hosts and the location of the 101 samples of Pseudoperonospora spp. samples used for the population analysesa .................................................................. 166 Table 4-3. Pairwise FST comparisons among Pseudoperonospora cubensis (clade I and II) and Pseudoperonospora humuli. ............................................................................. 167 vii Table 4-4. Analysis of molecular variance (AMOVA) for Pseudoperonospora spp. grouped by host. The significance of variance was tested from 999 permutations of the dataa. ............................................................................................................................ 168 Table 4-5. Genotypic diversity estimates for Pseudoperonospora spp. samples grouped by host (clade). ............................................................................................................. 169 Table 4-6. Analysis of molecular variance (AMOVA) of Midwest subpopulationsa of Pseudoperonospora cubensis collected from Cucumis sativus. The significance of variance was tested from 999 permutations of the dataa. ............................................ 170 Table 4-7. Pairwise FST comparisons among subpopulations of Pseudoperonospora cubensis collected from C. sativus in the Midwesta. ..................................................... 171 Table 4-8. Genotypic diversity estimates and index of association of Pseudoperonospora cubensis subpopulations collected from C. sativus in the Midwest. ...................................................................................................................................... 172 Table 4-9. Analysis of molecular variance (AMOVA) of Michigan subpopulations of Pseudoperonospora humuli. The significance of variance was tested from 999 permutations of the dataz. ............................................................................................ 173 Table 4-10. Pairwise FST comparisons among subpopulations of Pseudoperonospora humuli collected from hop yards in Michigana. ............................................................. 174 Table 4-11.Genotypic diversity estimates and index of association of Michigan subpopulations of Pseudoperonospora humuli. ........................................................... 175 viii LIST OF FIGURES Figure 2-1. Regression and amplification curves of Pseudoperonospora cubensis and P. humuli DNA using qPCR. ............................................................................................... 63 Figure 2-2. Standard curves for the quantification of Pseudoperonospora cubensis and P. humuli sporangia using qPCR. .................................................................................. 64 Figure 2-3. Monitoring of Pseudoperonospora cubensis and P. humuli sporangia using Burkard spore traps in Ingham (A) and Muskegon (B) in 2018. ..................................... 65 Figure 2-4. Monitoring of Pseudoperonospora cubensis and P. humuli sporangia using Burkard spore traps in Berrien (A), Ingham (B) and Muskegon (C) in 2019. ................. 66 Figure 3-1. Location of spore traps by county in 2018 (A) and 2019 (B). Green and blue dots indicate Burkard and impaction spore traps, respectively. ................................... 110 Figure 3-2. Standard curves for the quantification of Pseudoperonospora cubensis clade I, P. cubensis clade II and P. humuli DNA using the qPCR protocol A. ....................... 111 Figure 3-3. Standard curves for the quantification of Pseudoperonospora cubensis clade I, P. cubensis clade II and P. humuli DNA using the qPCR protocol B ........................ 112 Figure 3-4. Linear regression of Cq values as a function of sporangial numbers. ....... 113 Figure 3-5. Logistic regression of qPCR results from Burkard (A) and impaction (B) spore traps as a function of sporangial numbers. ........................................................ 114 Figure 3-6. Monitoring of Pseudoperonospora cubensis and P. humuli sporangia using spore traps in Muskegon (A) and Bay (B) counties in 2018 (Location B). .................... 115 Figure 3-7. Monitoring of Pseudoperonospora cubensis and P. humuli sporangia using spore traps in Allegan (A) and Saginaw (B) counties in 2018. ..................................... 117 Figure 3-8. Monitoring of Pseudoperonospora cubensis and P. humuli sporangia using spore traps in Muskegon (A) and Bay (B) counties in 2019. ........................................ 119 Figure 3-9. Monitoring of Pseudoperonospora cubensis and P. humuli sporangia using spore traps in Allegan (A) and Saginaw (B) counties in 2019. ..................................... 121 Figure 3-10. Monitoring of P cubensis and P. humuli sporangia using spore traps in Ingham (A) and Berrien (B) counties in 2019. .............................................................. 123 Figure 3-11. Standard curves for the quantification of Pseudoperonospora cubensis and P. humuli sporangia using the qPCR protocol A. ......................................................... 125 ix Figure 3-12. Standard curves for the quantification of Pseudoperonospora cubensis clade I, P. cubensis clade II and P. humuli sporangia using the qPCR protocol B. ..... 126 Figure 4-1. Cucumis sativus planted acreage and number Humulus lupulus planted in Michigan by county (Adapted from Neufeld, 2017). ..................................................... 176 Figure 4-2. Depth coverage of high-quality SNPs across libraries. .............................. 177 Figure 4-3. Characterization of sequencing error among technical replicates. ............ 178 Figure 4-4. Genetic differentiation of Pseudoperonospora cubensis and Pseudoperonospora humuli samples. .......................................................................... 179 Figure 4-5. Ordination plots of Pseudoperonospora spp. based on 2,978 SNPs. ........ 180 Figure 4-6. Frequency and geographic distribution of Pseudoperonospora cubensis and Pseudoperonospora humuli genotypes (MLG). ............................................................ 181 Figure 4-7. The relationship between genetic differences among samples and geographic distances of the locations from which samples originated. ........................ 182 Figure 4-8. Estimation of the degree of linkage disequilibrium within Pseudoperonospora cubensis and Pseudoperonospora humuli of Michigan. ............. 183 x CHAPTER I: LITERATURE REVIEW - PSEUDOPERONOSPORA CUBENSIS THE CAUSAL AGENT OF CUCURBIT DOWNY MILDEW 1 INTRODUCTION Pseudoperonospora cubensis (Berk. & M. A. Curtis) Rostovzev, the causal agent of cucurbit downy mildew (CDM), is a highly destructive pathogen recognized as one of the greatest contemporary disease threats to cucumber production worldwide (Brzozowski et al., 2016). In the U.S., two distinct evolutionary clades of P. cubensis (clade I and II) affect the production of cucurbit species including agronomically important crops such as cucumber, pumpkin, watermelon, and squash (Wallace et al., 2020). The disease only affects foliage creating small, irregular to angular chlorotic areas, and in most cases, sporulation on the lower leaf surface (Cohen et al., 2015). Leaf lesions coalesce and become necrotic leading to leaf blighting and premature defoliation which results in stunted plants and yield reduction, especially in cucumber (Cucumis sativus L.) and squash (Cucurbita moschata) (Reuveni et al., 1980; Adams et al., 2019; Hausbeck et al., 2019; Perla et al., 2019). In the U.S., the disease was only of minor concern prior to 2004 due to the deployment of resistant cucumber cultivars starting in 1960 (Thomas et al., 2017c). However, in 2004, a highly virulent strain of P. cubensis was introduced into U.S. that overcame this host resistance and stunned the cucumber industry (Holmes et al., 2014; Thomas et al., 2017a). In the absence of fully resistant cucurbit cultivars, chemical control is currently the most effective strategy to control CDM (Hausbeck and Goldenhar, 2017) but the evolution of resistant P. cubensis isolates to multiple fungicides has created an urgent need for alternative disease management strategies (Thomas et al., 2017). 2 Public breeding programs have made progress in the development of new cultivars to respond to the rising virulence of CDM, and have released slicing cucumbers with good levels of resistance (Brzozowski et al., 2016). However, the resistance of these varieties to the downy mildew pathogen has not yet been incorporated into processing (pickling) cucumbers that represent a significant portion of the cucurbit crops grown in the country. The U.S. is the seventh-largest producer of cucumber in the world (Keinath et al., 2017) and in 2019, 756,000 metric tons of fresh market (20%) and pickling cucumbers (80%) were grown on 42000 ha for a combined value of $279 million (USDA, 2020). Michigan is the second-largest producer of cucumbers in the country where approximately half of the national production of cucumbers (300,000 metric tons fresh and pickling) is harvested every year (USDA, 2020). Similarly, Michigan is also a big producer of other susceptible crops to CDM such as squash and pumpkin. In 2019, 85,600 and 41,200 metric tons of squash and pumpkins, respectively, were grown on a combined area of 5700 ha (USDA, 2020). P. cubensis clade II is especially destructive on cucumbers (Holmes et al., 2014). The pathogen cannot overwinter in northern regions of the U.S. that experience frost, but the disease reoccurs yearly due to the influx of airborne sporangia from overwintering sites with mild winters (Ojiambo and Holmes, 2010; Quesada-Ocampo et al., 2012). CDM has changed the way cucumber growers manage their fields in Northern states such as Michigan. The pickling cucumber growers of Michigan apply fungicides frequently to limit CDM, with an estimated cost of more than $6 million annually (Goldenhar and Hausbeck, 2019). Oxathiapiprolin (FRAC 49), ethaboxam (FRAC 22), fluazinam (FRAC group 29), cyazofamid (FRAC 21), 3 ametoctradin/dimethomorph (FRAC 45/40), and the broad-spectrum fungicides mancozeb (FRAC M03) and chlorothalonil (FRAC M05) are among the best fungicide alternatives for CDM control in different regions of the U.S. (Goldenhar and Hausbeck, 2019; Adams et al., 2020; Dutta, 2020; Miller et al., 2020). Application timing of fungicides is crucial when weather conditions are especially favorable for CDM and to reduce the risk of the pathogen developing resistance, fungicides must be applied preventively (Hollomon, 2007). To optimize application timing, the airborne concentration of P. cubensis sporangia can be monitored to coordinate the initiation of fungicide applications with the arrival of the pathogen into cucumber production fields (Granke et al., 2013; Goldenhar and Hausbeck, 2019). The accurate detection and quantification of sporangia can enable a more efficient application timing of fungicides and possibly, delay the development of pathogen resistant isolates to fungicides. Spore trapping can provide quantitative data on airborne spore numbers (Dung et al., 2018); however, the processing and microscopic examination of spore trap samples is time consuming and can result in misidentification due to morphological similarities between Pseudoperonospora taxa (e.g. P. cubensis and P. humuli). The combination of spore traps and PCR technologies is an ongoing area of investigation that seeks to improve and accelerate the detection of P. cubensis sporangia from air samples. Next generation sequencing technologies have recently facilitated the identification of new diagnostics markers for pathogen detection (Withers et al., 2016; Rahman et al., 2019) and the development of qPCR assays that could improve the monitoring of airborne P. cubensis sporangia (Summers et al., 2015a). 4 Similarly, these technologies have also facilitated the implementation of new tools to monitor CDM populations (Summers et al., 2015; Wallace & Quesada-Ocampo, 2017; Withers et al., 2016). This important progress has made possible the execution of more comprehensive genetic studies to better understand the epidemiology of CDM in the U.S. (Summers et al., 2015; Thomas et al., 2017). This review briefly summarizes our current understanding of P. cubensis biology including taxonomy, dispersal, management, fungicide resistance and population genetics. Additionally, we also elaborate on future directions for the effective control of CDM including plant breeding and early pathogen detection TAXONOMY AND DISEASE CYCLE Pseudoperonospora cubensis (Berk. & Curt.) Rost., the causal agent of cucurbit downy mildew (CDM), is an oomycete plant pathogen belonging to the family Peronosporaceae (Oomycota, Oomycetes, Peronosporales). This family is made up of an extensive number of plant pathogens that threaten natural and managed ecosystems including all downy mildews (DM) and other genera such as Phytopythium spp., Halophytophthora spp. and Phytophthora spp. (Thines and Choi, 2015). All DM are considered obligate biotrophic plant parasites and as such, they can only grow in association with living host tissue (Thines, 2014). Phytophthora spp., on the other hand, are hemibiotrophic or necrotrophic and only a small group of species are obligate biotrophs (Thines, 2014; Bourret et al., 2018). So far, 19 downy mildew genera have been described that contain more than 700 species (Thines and Choi, 2015). Most DM genera are present in three major monophyletic groups: 1) downy mildews with colored conidia (Peronospora and 5 Pseudoperonospora), 2) downy mildews with pyriform haustoria (Basidiophora, Benua, Bremia, Novotelnova, Paraperonospora, Plasmopara, Plasmoverna, and Protobremia), and the brassicolous downy mildews (Hyaloperonospora and Perofascia). Apart from these three groups, several grass-parasitic downy mildew genera have been described, of which the graminicolous downy mildews with lasting sporangiophores (Graminivora, Poakatesthia, and Viennotia) seem to be monophyletic (Thines, 2014; Thines and Choi, 2015; Bourret et al., 2018). DM were thought to form a single monophyletic group, however, a recent multigene phylogenetic analysis showed that Graminicolous downy mildews (GDM), brassicolous downy mildews (BDM) and downy mildews with colored conidia (DMCC) form a monophyletic clade with the Phytophthora taxon totara; while downy mildews with pyriform haustoria (DMPH) were placed in their own clade (Bourret et al., 2018). The genus Pseudoperonospora has been place alongside Peronospora in the group of downy mildews with colored conidia. Six species have been described within the Peronosporaceae genus, including two economically important species: P. cubensis and P. humuli (causal agent of hop downy mildew). P. cubensis causes disease on approximately 20 cucurbits genera (approximately 40 – 60 different species) including cucumber (Cucumis sativus), cantaloupe (C. melo), pumpkin (Cucurbita maxima), watermelon (Citrullus lanatus), squash (Cucurbita pepo), gourd (C. moschata), and wild cucurbit species such as balsam apple (Momordica balsamina ), bitter melon (M. charantia) and Buffalo Gourd (Cucurbita foetidissima) (Quesada-Ocampo et al., 2012; Savory et al., 2011; Wallace et al., 2014; Wallace et al., 2015). 6 P. cubensis requires a living host to complete its life cycle which begins when sporangia (2n) land on the adaxial surface of susceptible host. Under the right environmental conditions (15-20 °C, 1-5h of leaf wetness), sporangia release zoospores (2n) capable of swimming and encysting in host stomata (Granke et al., 2013; Cohen et al., 2015). From the encysted zoospore, a germ tube forms, penetrating the host tissue through an appressorium. Hyphae (2n) grow into the intracellular space where nutrient acquisition occurs. Sporangiophores form clonally from the intracellular growing hyphae holding sporangia produced mitotically in their tips (2n). Sporangia dislodge from Sporangiophores by a twisting mechanism that occurs when relative humidity (RH) decreases, and then fly through the air until they reach a new susceptible host re- initiating the cycle. Several oomycete pathogens reproduce in a mixed mode (asexual and sexual reproduction), however, it is not clear if P. cubensis undergoes a sexual phase in U.S. fields. The formation of oospores by P. cubensis seems to be rare but it has been reported several times in other countries such as China, Israel and India (Cohen et al., 2003; Savory et al., 2011; Zhang et al., 2012). A recent study confirmed the presence of two different mating types in the U.S. (A1 and A2) as well as its ability to produce viable oospores in vitro. (Thomas, A. Carbone, I. Ojiambo, 2013). This suggests that P. cubensis could potentially reproduce sexually in cucurbits within U.S. but the actual frequency of oospore formation could be very low due to the strict association of each mating type to a particular host (i.e. the A1 mating type isolates were uniquely found in cucumbers while the A2 mating type was mainly found in squash) (Thomas et al., 2016; Cohen et al, 2011; Cohen et al., 2015). The incidence of sexual reproduction and the 7 formation of oospores could have an important role in the epidemiology of CDM in the country. Monitoring of sexual reproduction in U.S. fields is key to identify new sources of genetic variation. SURVIVAL AND DISPERSAL OF P. CUBENSIS P. cubensis sporangia do not survive for a long time on non-living or necrotic tissue and its ability to infect cucumbers is reduced greatly under dry conditions, surviving only for 22 hours at temperature between 35 and 40 Celsius degrees (Cohen and Rotem, 1971). Additionally, overwintering-sexual structures such as oospores have not been detected in soil or host tissue in the U.S., and are not considered a source of inoculum (Naegele et al., 2016). Thus, P. cubensis survival in the U.S. depends on the availability of susceptible hosts (Savory et al., 2011) and the pathogen cannot survive in fields of regions where cucurbits cannot be grown year-round due to long winters with frost. In northern U.S., CDM reoccurs yearly due to the influx of sporangia that originated from warm weather plantings along the eastern seaboard and/or greenhouses in colder locales where the pathogen can survive on living host tissue (Ojiambo and Holmes, 2010; Quesada-Ocampo et al., 2012). CDM outbreaks in the great lakes region of the U.S. and Ontario, Canada are thought to result from the arrival of sporangia from southern states (e.g. Florida) that migrate north by wind currents using plantings of susceptible crops (Ojiambo et al., 2015). However, migration inferences made at a genetic level suggest a more restricted movement of sporangia in the country (Quesada-Ocampo et al., 2012). Although P. cubensis sporangia can travel up to 1,000 kilometers via air currents and migrate between fields in different states (Ojiambo and Holmes, 2010; Ojiambo et 8 al., 2015; Naegele et al., 2016), the CDM population from southern U.S. (i.e. Florida, Georgia, North Carolina) are highly dissimilar to the pathogen population in the Upper Midwest (Quesada-Ocampo et al., 2012). Evidence has been found that support the exchange of migrants among states in the upper Midwest and Canada (Naegele et al., 2016) but evidence of the movement of sporangia between northern and southern states has not yet been found. The progressive movement of the pathogen between states depends mainly on three factors: the asexual production of sporangia, its passive atmospheric transportation and the availability of new susceptible host (Ojiambo & Holmes, 2010). These factors in combination with environmental variables are correlated with the occurrence of CDM (Granke et al., 2013) and provide the basis of the CDM forecasting system (CDM ipmPIPE) that estimates the risk of outbreaks at any particular area (Ojiambo et al., 2015). Under field conditions, airborne sporangial concentrations, time post-planting, temperature, and leaf wetness are positively associated with disease occurrence, while solar radiation is the only factor negatively associated with disease (Granke and Hausbeck, 2011; Granke et al., 2013). Recent studies indicate that the P. cubensis can also be transmitted by seeds (Cohen et al., 2014) and infect wild cucurbits (Wallace et al., 2014; Wallace et al., 2015; Wallace and Quesada-Ocampo, 2016). CDM symptoms and sporulation have been observed on the leaves of wild species such as Balsam apple (Momordica balsamina), bitter melon (Momordica charantia), buffalo gourd (Cucurbita foetidissima), and bottle gourd (Lagenaria siceraria) (Wallace and Quesada-Ocampo, 2016), however, it is still unknown if the pathogen can overwinter in the dormant tissue of these species (Wallace 9 et al., 2014; Wallace et al., 2015). Further research is needed to establish the role of non-commercial cucurbits in the yearly CDM epidemic, which will aid the efforts of the CDM ipmPIPE to predict disease outbreaks. GENETIC STRUCTURE Research on P. cubensis populations from Europe and the U.S. have identified 6 distinct genetic clusters among 465 samples (Quesada-Ocampo et al., 2012; Kitner et al., 2015). Some clusters were more frequently associated with particular geographical regions, however, all of them were detected in Europe and in the U.S. This suggests that some genotypes are persistent and widely dispersed and/or have migrated from one population to others (Quesada-Ocampo et al., 2012). Initial studies may have underestimated the diversity of the populations due to the low number of markers used (Quesada-Ocampo et al., 2012; Kitner et al., 2015), however, the general structure patterns have also been observed in analyses using larger numbers of genetic markers (Summers et al., 2015b; Thomas et al., 2017a; Wallace and Quesada-Ocampo, 2017). At the genetic and phenotypic level (i.e. host preference), P. cubensis is structured by host in the U.S. (Thomas et al., 2017a; Wallace et al., 2020). In fact, it was recently shown that P. cubensis in the U.S. can be divided into two host-specific clades (Wallace et al., 2020). Further genetic studies have confirmed that these two clades are host-adapted at the cucurbit species level (Summers et al., 2015b; Thomas et al., 2017a; Wallace et al., 2020), with clade I isolates recovered more frequently from commercial varieties of Cucurbita pepo, C. moschata, C. maxima, and Citrullus lanatus and clade II isolates associated more frequently with commercial varieties of the Cucumis sativus and Cucumis melo (Wallace et al., 2020). Additionally, clade II isolates 10 were also found infecting the wild cucurbit species Lagenaria siceraria while clade I was also isolated from the wild species Momordica balsamina and Momordica charantia (Thomas, A. Carbone, I. Ojiambo, 2013; Wallace et al., 2020). Both clades were only found with low frequency in Cucumis melo and Cucurbita foetidissima (Wallace et al., 2020). It is still unclear if sexual reproduction occurs under field conditions within the P. cubensis population of the U.S. However, signs of recombination were found using genetic markers in clade I consistent with a sexually reproducing population, while no evidence of random mating was found for clade II (Wallace et al., 2020). This suggests that only clade I could be heterothallic while clade II may only reproduce clonally. Thomas et al., (2017) confirmed the presence of two mating types (A1 and A2) in the U.S. able to form oospores in vitro but information on the clade membership of the isolates used is not available. The existence of two different mating types within each clade has not been confirmed and it is unknown if recombination between isolates of different clades can occur. However, both clades are rarely detected within a single host (Cucumis melo) suggesting the possibility of two different mating-types from each clade encountering each other is low. As well as clustering by host, clustering between P. cubensis isolates by geographic location has also been reported in the U.S. (Quesada-Ocampo et al., 2012; Naegele et al., 2016) but it is unclear if this geographic structure is real or is an artifact of the population differentiation driven by the host or a temporal effect of the sampling. Population studies of P. cubensis have are limited by the sporadic occurrence of the disease due to the obligate nature of the pathogen. This makes the collection of isolates 11 highly dependent on the availability of susceptible hosts, whose production is regionally and temporally affected. The cucurbit growing season along the states in the eastern seaboard of the U.S has minimal overlap making comparisons among P. cubensis populations from northern and southern states difficult without considering a temporal factor. In Northern U.S., cucurbits are mainly grown during the summer while southern states such as Florida and Georgia produce cucurbits only during the spring and fall (Aerts and Mossler, 2003). Thus, the genetic differentiation in space (e.g. region, state) detected previously (Quesada-Ocampo et al., 2012) might be an artifact of the host driven structure and/or a biased sampling. A population study performed in Czech Republic over two years of sampling revealed no clustering based on geographical origin (Kitner et al., 2015). PATHOGENIC VARIATION The change in the host range of CDM that occurred in Europe in 2009, when P. cubensis became a significant problem for species like Cucurbita moschata, C. pepo, C. maxima and Citrullus lanatus, was attributed to a significant change in the structure of pathogen population (Kitner et al., 2015). Pre-epidemic samples were different significantly from samples collected after 2009 and they clustered in completely different clades. Only a limited number of heterozygous samples collected after 2009 clustered in the pre-epidemic clade, suggesting the occurrence of rare recombination events between populations (Cohen et al., 2015). In the same way, the emergence of new pathotypes (physiological races) has been proposed as an explanation for the breaking of host resistance in the United States. However, this hypothesis has not yet been proved due to the small number of 12 samples collected before 2004, and the limited resolution of the studies performed so far (Quesada-Ocampo et al., 2012). A better understanding of P. cubensis populations at the local level is key to control and prevent the emergence of more virulent pathogen populations with additional levels of fungicide resistance. FUNGICIDE RESISTANCE In the absence of fully resistant cucumber varieties, fungicide use is currently the most effective method to control CDM (Hausbeck and Goldenhar, 2017). However, P. cubensis is a high-risk pathogen in terms of evolving fungicide resistance both because of its shorth generation time on the genetically uniform monocultures of its hosts and the huge population size during outbreaks that offer many opportunities for mutations (Kitner et al., 2015). In fact, P. cubensis, was the first pathogen to be reported as resistant to the phenylamide fungicide mefenoxam (FRAC 4), a widely used fungicide against most of the oomycete plant pathogens (Reuveni et al., 1980). Over 17 different fungicides (representing 15 FRAC codes) are registered to control CDM (Table 1-1), however, complete resistance or reduction in the sensitivity to multiple fungicides has been reported within P. cubensis populations (Urban and Lebeda, 2006; Goldenhar and Hausbeck, 2019). Complete resistance of P. cubensis to fungicides in the FRAC groups 4 (phenylamides), 11 (quinone outside inhibitors), and 40 (carboxylic acid amides) (Gisi and Sierotzki, 2015; Ojiambo et al., 2015) has been documented. In the U.S., single-site fungicides including mefenoxam (FRAC 4) and azoxystrobin (FRAC 11) were ineffective when the pathogen reemerged in 2004 (Ernest et al., 2005; Gevens and Hausbeck, 2005; Thornton et al., 2006). Since that time, P. cubensis resistance to dimethomorph 13 (FRAC 40) (Zhu et al., 2007) and mandipropamid (FRAC 40) (Hausbeck and Cortright, 2010; Blum et al., 2011) has been reported in the U.S. (Holmes et al., 2014; Keinath, 2015). Similarly, reduced efficacy of fluopicolide and cymoxanil (FRAC 43) against CDM has been observed in field trials in Michigan (Hausbeck and Linderman, 2014; Goldenhar and Hausbeck, 2019), Georgia (Langston and Sanders, 2013), and North Carolina (Adams and Quesada-Ocampo, 2014; Keinath, 2015). Propamomcarb (FRAC 28) was effective against CDM for several years, but in 2013 its efficacy was compromised in field trials in North Carolina (Keinath, 2015; Thomas et al., 2018), Pennsylvania (Gugino and Grove, 2016), and Michigan (Hausbeck and Linderman, 2014; Hausbeck et al., 2017). The number of effective fungicides against CDM is limited, and fungicide efficacy can vary between years (Goldenhar and Hausbeck, 2019) due to changes in environmental conditions and the pathogen population. However, oxathiapiprolin (FRAC 49), ethaboxam (FRAC 22), fluazinam (FRAC group 29), cyazofamid (FRAC 21), ametoctradin/dimethomorph (FRAC 45/40), and the broad-spectrum fungicides mancozeb (FRAC M03) and chlorothalonil (FRAC M05) have shown a good control levels of CDM in field trials in Ohio (Miller et al., 2020), Georgia (Dutta, 2020), North Carolina (Adams et al., 2020) and Michigan (Goldenhar and Hausbeck, 2019; Hausbeck et al., 2019) Oxathiapiprolin is a relatively new active ingredient with proven efficacy against P. cubensis (Cohen 2015; Goldenhar and Hausbeck 2016). However, it is a single site fungicide inhibitor and as such, it is considered at high risk of pathogen resistance (Cohen 2015; FRAC 2018). Similarly, ethaboxam had demonstrated efficacy against 14 CDM (Quesada-Ocampo and Adams, 2014; Gugino and Grove, T.L., 2020) and was classified as a fungicide for which the development of resistance is at low to medium risk (FRAC, 2020); still, in 2017 its efficacy was compromised in field trials in Michigan (Goldenhar and Hausbeck, 2019). Fluazinam, cyazofamid and ametoctradin/dimethomorph have shown consistent good control of CDM (Adams and Quesada-Ocampo, 2014; Keinath, 2015; Goldenhar and Hausbeck, 2019) and only Fluazinam is not widely used by cucurbit growers due to its high cost (Neufeld et al., 2017). Resistance to broad spectrum fungicide is rare in P. cubensis populations, but metalaxyl-resistant isolates found in Israel also exhibited moderate levels of resistance against active ingredients such as mancozeb (Reuveni et al., 1980) It seems inevitable to avoid hastening the emergence of resistance given the limited amount of options against CDM and the high rates at which fungicides are applied (every 7 to 10 days). Thus, resistance management is key to maintaining the efficacy of single-site fungicides, and growers are encouraged to rotate among fungicides from different FRAC groups (FRAC, 2020). The continuous monitoring of fungicide efficiency at the local level is essential to provide the best recommendations to cucumber growers in the U.S. and ensure that P. cubensis is effectively managed. 15 Table 1-1. List of fungicides registered for the control of cucumber downy mildew in the U.S. Product name Registrant FRAC code a Reduce efficacy FRAC chemical group Target site Yes No PhenylAmides (PA) - fungicides RNA polymerase I Multi-site contact Chloronitrites activity Yes Benzamides Yes Carbamates Delocalization of spectrin-like proteins Cell membrane permeability, fatty acids (proposed) No No No Yes Quinone inside inhibitor (Qil) - fungicides Dithio- carbamates (M03) / thiazole carboxamide (22) Cyanoacetamide- oxime No Unestablished Complex III: cytochrome bc1 at Qi site Multi-site contact activity (M03) / B- tubulin assembly in mitosis (22) Unknown Uncouplers of oxidative phosphorylation Active ingredient Metalaxyl/ mefenoxam Ridomil Syngenta 4 Chlorothalonil Bravo Weather Stik Syngenta M05 Fluopicolide Presidio Valent Propamomcarb Previcur Flex Bayer Cyazofamid Ranman FMC 43 28 21 Zoxamide/ mancozeb Mancozeb Cymoxanil Gavel Dithane Curzate Gowan 22/M03 BASF M03 DuPont 27 29 Fluazinam Omega Syngenta aReduce efficacy or completed resistance reported in the U.S. 16 Table 1-1. (cont’d) Active ingredient Mandipropamid Dimethomorph Ametoctradin/ dimethomorph Fluxapyroxad/ pyraclostrobin Pyraclostrobin Famoxadone/ cymoxanil Product name Revus Forum Zampro Priaxor Cabrio Tanos Registrant Syngenta BASF BASF BASF BASF 40 40 45/40 11/7. 11 DuPont 11/27. Oxathiapiprolin Orondis Syngenta 49 Ethaboxam Elumin Valent 22 aReduce efficacy or completed resistance reported in the U.S 17 FRAC code a Reduce efficacy FRAC chemical group Target site Yes Yes No Yes Yes Yes No No Carboxylic Acid Amides (CAA) - fungicides (40) Cellulose synthase (40)/ complex III: cytochrome bc1 (45) Quinone outside Inhibitors (QoI) - fungicides (11) OSBPI oxysterol binding protein homologue inhibition Complex III: cytochrome bc1 at Qo site (11) Lipid homeostasis and transfer/storage Thiazole carboxamide ß-tubulin assembly in mitosis BREEDING EFFORTS Cucumber (Cucumis aestivus) is a widely cultivated plant of the Cucurbitaceae family, with annual production above 71 million tons globally (FAO, 2013). The U.S. is the fifth largest producer of cucumber, and in 2019, 756,000 metric tons of fresh market and pickling cucumbers were grown on approximately 42000 ha for a combined value of $279 million (USDA, 2020). Several diseases affect cucumber production including target spot, powdery mildew and Phytophthora crown and root rot (Tuttle McGrath, 2004; Keinath et al., 2017), however, cucurbit downy mildew (CDM) caused by the oomycete, Pseudoperonospora cubensis is probably the most important disease of cucumber in the country (Thomas et al., 2017a). CDM is a highly destructive foliar disease able to cause 50% reduction in cucumber yield even after fungicides are applied one-week post-symptom appearance (Cohen et al., 2015). Currently no cultivar has robust resistance to the disease, but for decades, downy mildew on cucumbers was effectively managed with genetic host resistance. (Sitterly, 1972). The introgression of the dm-1 gene from PI 197087 into the cultivars ‘Polaris’, ‘Poinsett’, ‘Pixie’, and ‘Chipper’ provided cucumber growers with genetic control of downy mildew for more than 40 years (Call et al., 2012a; Cohen et al., 2015; Thomas et al., 2017a). From 1961 to 2003, downy mildew was only a moderate problem in North America and was easily controlled with fungicides (Cohen et al., 2015). However, the resistance of commercial cultivars in the U.S. was defeated in 2004 when a highly virulent strain of P. cubensis was introduced in the country (Thomas et al., 2017a). Since then, P. cubensis has reoccurred yearly in cucumber production areas of 18 eastern U.S., costing cucumber growers millions annually in fungicide sprays to protect their crop (Granke and Hausbeck, 2011; Savory et al., 2011). Over the last ten years, a number of new downy mildew-resistant cucumber lines have been released by public and private breeding programs including the slicing varieties SV3462CS, SV4719CS, and Bristol, and the pickling varieties Citadel and Peacemaker from Seminis (Holdsworth et al., 2014; Brzozowski et al., 2016). These varieties have shown an intermediate level of resistance to P. cubensis and reached comparable yields to commercial standards (e.g. Vlaspik) but still required the utilization of fungicides to provide full disease control (Hausbeck and Goldenhar, 2017). The Cornell breeding program have made significant progress in the development of new cultivars to control CDM releasing slicing cucumbers with exceptional levels of resistance (Brzozowski et al., 2016). The released line “DMR-NY401” retained the disease resistance of its predecessor line “DMR-NY264” while showing significantly higher fruit length, yield, and earliness of initial harvest (Brzozowski et al., 2016). However, the resistance of these varieties has not yet been incorporated in pickling cucumber varieties which are highly important in the country. Race-specific R gene breeding: Plant pathogens such as P. cubensis secrete effector proteins during infection that modulate host innate immunity (Goss et al., 2013). Many of these effectors function only as virulence factors, but others can be recognized by plant R proteins resulting in the activation of effector triggered immunity. In such cases, the effectors are known as avirulence (AVR) factors (Brzozowski et al., 2016). In most cases, the response induced by AVR factors involves the hypersensitive response (HR) followed by restriction of the invading pathogen (Vleeshouwers et al., 2008). 19 Virulence factor can be identified throughout the phenotypic characterization of the P. cubensis population and can then be functionally profiled on cucurbit species to detect cognate R genes. R genes have been successfully identified using virulence factors in the model system P. infestans-potato plant and other fungal species such as Cryptococcus neoformans (Vleeshouwers et al., 2008; Desjardins et al., 2017). However, the identification of R genes in the cucurbit downy-mildew system using this strategy faces two important obstacles. First, as an obligate biotroph, P. cubensis requires a living host tissue for reproduction and dispersal, complicating the maintenance, identification and phenotyping of isolates. Secondly, the specific spectrum of novel CDM R-genes cannot be assessed in the absence of a diverse panel of isolates or pathotypes (isolates with the same pathogenicity). Due to these difficulties almost every phenotypic or genotypic study performed so far includes only a limited number of isolates. An effort to identify diagnostic pathotypes of P. cubensis was recently performed by Thomas, et al. (2017). In this study, thirteen different pathotypes were identified based on a set of 15 different cucurbit species. The authors suggest the existence of 10 different avirulence factors (Avr genes), assuming that disease resistance is only expressed when an R gene product in the host can recognize the pathogen’s corresponding effector. Interestingly, specific pathotypes form subgroups according to mating type which suggests the association between virulence and mating type (Thomas et al., 2017c). The ways in which virulence factors evolve will be highly valuable in breeding of specific R genes and the designing of effective management strategies in agricultural 20 systems. Unfortunately, these concepts are poorly understood for obligate plant pathogen (Thomas et al., 2017c). In order to understand the evolution of Avr genes, particularly in the cucurbit downy-mildew system, substantial efforts need to be made to identify virulence factors and their genes. This could help to survey the temporal and spatial distribution of virulence factors and race structure of the P. cubensis population. EARLY DETECTION OF P. CUBENSIS In the absence of fully resistant cucurbit cultivars, chemical control is currently the most effective method for controlling CDM (Wu et al., 2016). However, the risk for the evolution of resistance by P. cubensis is high due to its fast mode of reproduction, and the low abundance of multi-site inhibitors to control the disease. It seems inevitable to avoid hastening the emergence of resistant P. cubensis isolates given the limited amount of options for CDM control. Therefore, the reduction of fungicide use and the alternation between fungicides in different FRAC groups is key to maintaining the efficacy of single-site inhibitors (Goldenhar and Hausbeck, 2019). Application timing is crucial to optimize fungicide utilization and preventive application of fungicides can reduce the risk of the pathogen developing resistance (Hollomon, 2007). Thus, coordinating the initiation of fungicide applications with the arrival of pathogens into production fields could result in more efficient fungicide use. In California, measurements of aerial spore loads obtained from spore traps have been used to schedule the timing of fungicide applications against the lettuce pathogen, Bremia lactucae. This strategy resulted in the reduction of sprays without a significant increase in disease incidence (Dhar et al., 2019). 21 Monitoring the airborne concentration of P. cubensis sporangia using spore traps could also help to improve the scheduling of fungicides applications to control CDM in regions where disease occurrence depends on the influx of P. cubensis sporangia (Granke and Hausbeck, 2011; Granke et al., 2013; Dung et al., 2018). However, the processing and examination of spore trap samples need improvement to reduce the processing time and avoid misidentification. The screening of spore trap samples using qPCR can significantly improve the detection of airborne plant pathogens. qPCR has significantly improve the sensitivity and specificity for the detection of Peronospora effusa (Klosterman et al., 2014), Bremia lactucae (Kunjeti et al., 2016), and Botrytis cinerea (Kunjeti et al., 2016) in spore trap samples. The use of molecular markers to differentiate between P. cubensis and P. humuli sporangia could aid to improve the airborne monitoring of CDM. Previous PCR assays have been used to monitor airborne Pseudoperonospora humuli sporangia near hop yards to inform the timing of fungicide sprays, but since this assay also detects P. cubensis, its accuracy decreases in areas where hops and cucurbits are grown in close proximity (Gent et al., 2009). The specificity for the detection of P. cubensis was improved with the development of a qPCR assay able to differentiate between P. cubensis and P. humuli based on the recognition of a single SNP in the cox2 gene (Summers et al., 2015a). However, this assay is unable to differentiated between the two host adapted clades of P. cubensis and its specificity was compromise in samples containing samples from both species (P. cubensis and P. humuli). A new multiplex qPCR assay has been designed based on the recognition of a more polymorphic mitochondrial loci that also allow the differentiation between P. cubensis and P. humuli 22 (Crandall, 2020). These loci have less similarity between species and also allow the differentiation between two distinct clades of P. cubensis (clade I and II) (Thomas et al., 2017a; Wallace et al., 2020). In northern regions of the U.S., CDM occurs annually due to an influx of airborne sporangia from overwinter sources. Burkard spore traps couple with light microscopy have been used to monitor the influx of Pseudoperonospora spp. sporangia and inform growers that the disease is likely to occur. Burkard spore traps collect spores by vacuuming air into a collection chamber that contains a reel mounted in a clockwork mechanism were spores and other particles are impacted onto a greased tape that covers the reel (Burkard Manufacturing Co. Ltd., U.K.). Burkard spore traps can continuously collect data for up to 7 days and the collecting tape can be split at different intervals allowing the precise quantification of spores per unit of time. Although Burkard spore traps are the most common devices used to monitor the air concentration of plant pathogens, the impaction spore traps have become popular in the last decade (Jackson and Bayliss, 2011). These devices possess rods coated with adhesive material, which spin at a standard rate to impact and collect the airborne inoculum (TSE Systems, Chesterfield, MO), however, impaction traps require constant monitoring to allow accurate estimations per unit of time (i.e. an hour or day). Burkard spore traps are robust and highly autonomous but impaction spore traps can be more cost-effective and easier to use by growers (Jackson and Bayliss, 2011; Choudhury et al., 2016a). There are numerous comparative studies investigating the efficiency of these two spore traps, and in most cases, Burkard spore traps perform better than impaction spore traps. However, impaction spore traps have been used successfully for airborne 23 detection of pathogens such as Bremia lactucae, Peronospora effusa and Peronospora schachtii (Jenkyn, 1974; Li and Lin, 1999; Evenhuis et al., 2003), and could also be used for early detection P. cubensis sporangia in cucumber growing regions. 24 LITERATURE CITED 25 LITERATURE CITED Aerts, M., Mossler, M. (2003). Crop profile for cucumbers in Florida. In the National IPM Database [Data file]. Available from https://ipmdata.ipmcenters.org/documents/cropprofiles/FLcucumber2003.pdf Adams, M. L., and Quesada-Ocampo, L. M. (2014) Evaluation of fungicides for control of downy mildew on cucumber. Kinston 2013. Plant Dis Manag Rep 8: V240 Adams, M., Collins, H., Quesada-Ocampo, L.M. (2020) Evaluation of fungicides and cultivars for control of downy mildew on cucumber, Kinston, NC 2019. Plant Dis Manag Rep Report 14: V107 Adams, M., D’Arcangelo, K, Collins, H, Quesada-Ocampo, L (2019) Evaluation of cultivars for control of downy mildew on squash, Kinston 2018. Plant Dis Manag Reports 13: V124 Ali, S., Gautier, A., Leconte, M., Enjalbert, J., de Vallavieille-Pope, C. (2011) A rapid genotyping method for an obligate fungal pathogen, Puccinia striiformis f.sp. tritici, based on DNA extraction from infected leaf and multiplex PCR genotyping. BMC Res Notes 4: 240 Billard, A., Laval, V., Fillinger, S., Leroux, P., Lachaise, H., Beffa, R., and Debieu, D. (2012) The allele-specific probe and primer amplification assay, a new real-time PCR method for fine quantification of single-nucleotide polymorphisms in pooled DNA. Appl Environ Microbiol 78: 1063–1068 Blum, M., Waldner, M., Olaya, G., Cohen, Y., Gisi, U., and Sierotzki, H. (2011) Resistance mechanism to carboxylic acid amide fungicides in the cucurbit downy mildew pathogen Pseudoperonospora cubensis. Pest Manag Sci 67: 1211–1214 Bolger, A.M., Lohse, M., Usadel, B. (2014) Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30: 2114–2120 Bourret, T.B., Choudhury, R.A., Mehl, H.K., Blomquist, C.L., McRoberts, N., Rizzo, D.M. (2018) Multiple origins of downy mildews and mito-nuclear discordance within the paraphyletic genus Phytophthora. PLoS One 13: e0192502 Brown, A.H.D., Feldman, M.W., Nevo, E. (1980) Multilocus structure of natural populations of Hordeum spontaneum. Genetics 96: 523-536. Brzozowski, L., Holdsworth, W., Mazourek, M. (2016) ‘DMR-NY401’: A new downy mildew–resistant slicing cucumber. HortScience 51: 1294–1296 26 Burkhardt, A., Buchanan, A., Cumbie, J.S, Savory, E.A., Chang, J.H., Day, B. (2014) Alternative splicing in the obligate biotrophic oomycete pathogen Pseudoperonospora cubensis. Mol Plant-Microbe Interact 28: 298–309 Burkhardt, A., Day, B. (2016) Transcriptome and small RNAome dynamics during a resistant and susceptible interaction between cucumber and downy mildew. The Plant Genome: 9:1-19. Bustin, S. A., Benes, V., Garson, J.A., Hellemans, J., Huggett, J., Kubista, M., Mueller, R., Nolan, T., Pfaffl, M.W., Shipley, G.L., Vandesompele, J., and Wittwer, C. T (2009) The MIQE guidelines: minimum information for publication of quantitative real-time PCR experiments. Clin Chem 55: 611–622 Call, A.D., Criswell, A.D., Wehner, T.C., Ando, K., Grumet, R. (2012a) Resistance of cucumber cultivars to a new strain of cucurbit downy mildew. HortScience 47: 171–178 Call, A.D., Criswell, A.D., Wehner, T.C., Klosinska, U., Kozik, E.U. (2012b) Screening cucumber for resistance to downy mildew caused by Pseudoperonospora cubensis (Berk. and Curt.) Rostov. Crop Sci 52: 577–592 Carisse, O., McNealis, V., Kriss, A. (2017) Association between weather variables, airborne inoculum concentration, and raspberry fruit rot caused by Botrytis cinerea. Phytopathology 108: 70–82 Carisse, O., Tremblay, D.M., Lévesque, C.A., Gindro, K., Ward, P., and Houde, A. (2009) Development of a Taqman real-time PCR assay for quantification of airborne conidia of Botrytis squamosa and management of Botrytis leaf blight of onion. Phytopathology 99: 1273–1280 Carleson NC, Fieland VJ, Scagel CF, Weiland JE, Grünwald NJ (2019) Population structure of Phytophthora plurivora on Rhododendron in Oregon nurseries. Plant Dis 103: 1923–1930 Chee, H.Y., Nelson, M.E., Grov,e G.G., Eastwell, K.C., Kenny, S.T., Klein, R.E. (2006) Population biology of Pseudoperonospora humuli in Oregon and Washington. Plant Dis 90: 1283–1286 Choi, Y-J., Hong, S-B., Shin, H-D. (2005) A re-consideration of Pseudoperonospora cubensis and P. humuli based on molecular and morphological data. Mycol Res 109: 841–848 Choi, Y-J., Thines, M. (2015) Host jumps and radiation, not co-divergence drives diversification of obligate pathogens. a case study in downy mildews and Asteraceae. PLoS One 10: e0133655 27 Choudhury, R.A., Koike, S.T., Fox, A.D., Anchieta, A., Subbarao, K.V., Klosterman, S.J., McRoberts, N. (2016b) Spatiotemporal patterns in the airborne dispersal of spinach downy mildew. Phytopathology 107: 50–58 Clark, M.J., Chen, R, Lam, H.Y.K., Karczewski, K.J., Chen, R., Euskirchen, G., Butte, A.J., Snyder, M. (2011) Performance comparison of exome DNA sequencing technologies. Nat Biotechnol 29: 908–914 Cohen, Y., Van den Langenberg, K.M., Wehner, T.C, Ojiambo, P.S., Hausbeck, M, Quesada-Ocampo, L.M., Lebeda, A., Sierotzki, H., and Gisi, U. (2015) Resurgence of Pseudoperonospora cubensis: the causal agent of cucurbit downy mildew. Phytopathology 105: 998–1012 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 Cohen, Y., and Rotem, J. (1971) Dispersal and viability of sporangia of Pseudoperonospora cubensis. Trans Br Mycol Soc 57: 67–74 Cohen, Y., Rubin, A.E., Galperin, M. (2011) Formation and Infectivity of oospores of Pseudoperonospora cubensis, the causal agent of downy mildew in cucurbits. Plant Dis 95: 874 Cohen, Y., Rubin, A.E., Galperin, M, Ploch, S., Runge, F., Thines, M. (2014) Seed transmission of Pseudoperonospora cubensis. PLoS One 9: e109766 Cohen, Y., Samoucha, Y. (1983) Cross-resistance to four systemic fungicides in metalaxyl-resistant strains of Phytophthora infestans and Pseudoperonospora cubensis. Plant Dis 68: 137-139. Coley-Smith, J.R. (1962) Overwintering of hop downy mildew Pseudoperonospora humuli (Miy. and Tak.) Wilson. Ann Appl Biol 50: 235–243 Cruz-Dávalos, D.I., Llamas, B., Gaunitz, C., Fages, A., Gamba, C., Soubrier, J., Librado, P., Seguin-Orlando, A., Pruvost, M., Alfarhan, A.H., et al (2017) Experimental conditions improving in-solution target enrichment for ancient DNA. Mol Ecol Resour 17: 508–522 Cui, C., Herlihy, J.H., Bombarely, A., McDowell, J.M., Haak, D.C. (2019) Draft assembly of Phytophthora capsici from long-read sequencing uncovers complexity. Mol Plant- Microbe Interact 32: 1559–1563 Desjardins, C.A., Giamberardino, C., Sykes, S.M., Yu, C-H., Tenor, J.L., Chen, Y., Yang, T., Jones, A.M., Sun, S., Haverkamp, M.R., et al (2017) Population genomics and the evolution of virulence in the fungal pathogen Cryptococcus neoformans. Genome Res 27: 1207-1219 28 Dhar, N., Mamo, B.E., Subbarao, K., Koike, S., Fox, A., Anchieta, A., and Klosterman, S.J. (2019) Measurements of aerial spore load by qPCR facilitates lettuce downy mildew risk advisement. Plant Dis 104: 82–93 Dung, J.K.S, Scott, J.C., Cheng, Q., Alderman, S.C., Kaur, N., Walenta, D.L., Frost, K.E., and Hamm, P.B. (2018) Detection and quantification of airborne Claviceps purpurea sensu lato ascospores from Hirst-type spore traps using real-time quantitative PCR. Plant Dis 102: 2487–2493 Dutta, B. (2020) Evaluation of individual fungicides for downy mildew control in cucumber in Tift County, Georgia, 2019. Plant Dis Manag Rep Report 14: V086 Ernest, E.G., Kee, W.E., and Everts, K.L. (2005) Evaluation of fungicides for control of downy mildew on pickling cucumber, Trial 2, 2005. Plant Dis Manag Rep 2: V066 Evenhuis, A., Verdam, B., Zadoks, J.C. (2003) Splash dispersal of conidia of Mycocentrospora acerina in the field. Plant Pathol 46: 459–469 Faircloth, B.C., Branstetter, M.G., White, N.D., Brady, S.G. (2015) Target enrichment of ultraconserved elements from arthropods provides a genomic perspective on relationships among Hymenoptera. Mol Ecol Resour 15: 489–501 Fletcher, K., Gil, J., Bertier, L.D., Kenefick, A., Wood, K.J., Zhang, L., Reyes-Chin-Wo, S., Cavanaugh, K., Tsuchida, C., Wong, J., et al (2019) Genomic signatures of heterokaryosis in the oomycete pathogen Bremia lactucae. Nat Commun 10: 2645 Fungicide Resistance Action Committee (FRAC) (2018) ode List 2018: Fungicides sorted by mode of action (including FRAC Code numbering). Gent, D.H., Adair, N., Knaus, B.J., Grünwald, N.J. (2019) Genotyping-by-sequencing reveals fine-scale population differentiation in populations of Pseudoperonospora humuli. Phytopathology. doi: 10.1094/PHYTO-12-18-0485-R Gent, D.H., Nelson, M.E., Farnsworth, J.L., Grove, G.G. (2009) PCR detection of Pseudoperonospora humuli in air samples from hop yards. Plant Pathol 58: 1081– 1091 Gent, D.H., Ocamb, C.M. (2009) Predicting Infection Risk of Hop by Pseudoperonspora humuli. Phytopathology 99: 1190–1198 Gent DH, Ocamb CM, Farnsworth JL (2010) Forecasting and Management of hop downy mildew. Plant Dis 94: 425–431 Gevens, A.J., and Hausbeck, M.K. (2005) Control of downy mildew of cucumbers with fungicides, 2005. Plant Dis Manag Rep 2: V066 29 Gisi U, Sierotzki H (2015) Oomycete fungicides: phenylamides, quinone outside inhibitors, and carboxylic acid amide. In H Ishii, DW Hollomon (eds) Fungicide resistance in plant pathogens: Principles and a guide to practical management. Springer Japan, Tokyo, pp 145–174 Gisi, U., Sierotzki, H. (2008) Fungicide modes of action and resistance in downy mildews. Eur J Plant Pathol 122: 157–167 Gnirke, A., Melnikov, A., Maguire, J., Rogov, P., LeProust., E.M., Brockman, W., Fennell, T., Giannoukos, G., Fisher, S., Russ, C., et al (2009) Solution hybrid selection with ultra-long oligonucleotides for massively parallel targeted sequencing. Nat Biotechnol 27: 182–189 Goldenhar, K.E., Hausbeck, M.K. (2019) Fungicides for control of downy mildew on pickling cucumber in Michigan. Plant Heal Prog 20: 165–169 Goss, E.M., Press, C.M., Grünwald, N.J. (2013) Evolution of RXLR-class effectors in the oomycete plant pathogen Phytophthora ramorum. PLoS One 8: e79347 Goudet, J. (2005) hierfstat, a package for r to compute and test hierarchical F-statistics. Mol Ecol Notes 5: 184–186 Granke, L.L., and Hausbeck, M.K. (2011) Dynamics of Pseudoperonospora cubensis sporangia in commercial cucurbit fields in Michigan. Plant Dis 95: 1392–1400 Granke, L.L., Morrice, J.J., and Hausbeck, M.K. (2013) Relationships between airborne Pseudoperonospora cubensis sporangia, environmental conditions, and cucumber downy mildew severity. Plant Dis 98: 674–681 Grünwald, N.J., Everhart, S.E., Knaus, B.J., Kamvar, Z.N. (2017) Best practices for population genetic analyses. Phytopathology 107: 1000–1010 Grünwald, N.J., McDonald, B.A., Milgroom, M.G. (2016) Population genomics of fungal and oomycete pathogens. Annu Rev Phytopathol 54: 323–346 Gugino, B.K., Grove, T.L. (2020) Evaluation of select fungicides and fungicide programs for managing downy mildew of cucumber, 2019. Plant Dis Manag Rep Report 14: V005 Gugino, B.K., and Grove, T.L. (2016) Evaluation of selected fungicides and fungicide programs for managing downy mildew of cucumber, 2015. Plant Dis. Manag. Rep. 10:V072: Hartl, D., Clark, A. (1997) Principles of population genetics. Sinauer Associates, Inc, Sunderland, Massachusetts 30 Hausbeck, M. K., and Linderman, S.D. (2014) Evaluation of fungicides for control of downy mildew of cucumber, 2013. Plant Dis Manag Rep 8: V304 Hausbeck, M., Brisco-McCann, E., Goldenhar, K., and Cook, A. (2017) Evaluation of fungicide programs for control of downy mildew of cucumber, 2016. Plant Dis Manag Reports 11: V063 Hausbeck, M., Goldenhar, K. (2017) Downy mildew recommendations for cucumbers. MSU Ext., https://www.canr.msu.edu/news/downy_mildew_recommendations_for_cucumbers Hausbeck, M.K., and Cortright, B. (2010) Evaluation of fungicides and application methods for control of downy mildew of cucumber, 2010. Plant Dis. Manag. Rep. V6: Hausbeck, M.K., Perla, D.E., and Cook, A.J. (2019) Evaluation of single fungicide products for control of downy mildew of cucumber, 2018. Plant Dis Manag Reports 139: V13 Holdsworth, W.L., Summers, C.F., Glos, M., Smart, C.D., Mazourek, M. (2014) Development of downy mildew-resistant cucumbers for late-season production in the Northeastern United States. HortScience horts 49: 10–17 Hollomon, B. (2007) Fungicide resistance in crop pathogens: How can it be managed?. 2nd ed. FRAC, Brussels Holmes, G.J., Ojiambo, P.S., Hausbeck, M.K., Quesada-Ocampo, L., and Keinath, A.P. (2014) Resurgence of cucurbit downy mildew in the United States: a watershed event for research and extension. Plant Dis 99: 428–441 Jackson, S.L., Bayliss, K.L. (2011) Spore traps need improvement to fulfil plant biosecurity requirements. Plant Pathol 60: 801–810 Jenkyn, J. (1974) A comparison of seasonal changes in deposition of spores of Erysiphe graminis on different trapping surfaces. Ann Appl Biol 76: 257–267 Johnson, D.A, Alldredge, J.R, Allen, J.R, Allwine, R. (1991) Spatial pattern of downy mildew in hop yards during severe and mild disease epidemics. Phytopathology 81: 1369–1374 Jombart, T., and Ahmed, I. (2011) adegenet 1.3-1: new tools for the analysis of genome-wide SNP data. Bioinformatics 27: 3070–3071 Jones, L., Riaz, S., Morales-Cruz, A., Amrine, K.C.H., McGuire, B., Gubler, W.D., Walker, M.A., Cantu, D. (2014) Adaptive genomic structural variation in the grape powdery mildew pathogen, Erysiphe necator. BMC Genomics 15: 1081 31 Kamvar, Z.N., Brooks, J.C., Grünwald, N.J. (2015) Novel R tools for analysis of genome-wide population genetic data with emphasis on clonality. Front Genet 6: 208 Kamvar, Z.N., Tabima, J.F., Grünwald, N.J. (2014) Poppr: an R package for genetic analysis of populations with clonal, partially clonal, and/or sexual reproduction. PeerJ 2: e281 Kanetis, L., Holmes, G.J, and Ojiambo, P.S. (2010) Survival of Pseudoperonospora cubensis sporangia exposed to solar radiation. Plant Pathol 59: 313–323 Kanetis, L., Vasiliou, A., Neophytou, G., Samouel, S., Tsaltas, D. (2013) First report of downy mildew caused by Peronospora belbahrii on sweet basil (Ocimum basilicum) in Cyprus. Plant Dis 98: 283 Keinath, A. P. (2015) Utility of a cucumber plant bioassay to assess fungicide efficacy against Pseudoperonospora cubensis. Plant Dis 100: 490–499 Keinath, A.P., Wintermantel WM, Zitter TA, eds (2017) Compendium of cucurbit diseases and pests, 2nd ed. St. Paul, Minnesota, USA Kitner, M., Lebeda, A., Sharma, R., Runge, F., Dvořák, P., Tahir, A., Choi, Y-J., Sedláková, B., Thines, M. (2015) Coincidence of virulence shifts and population genetic changes of Pseudoperonospora cubensis in the Czech Republic. Plant Pathol 64: 1461–1470 Klosterman, S.J., Anchieta, A., McRoberts, N., Koike, S.T., Subbarao, K. V., Voglmayr, H., Choi, Y-J., Thines, M., and Martin, F.N. (2014) Coupling spore traps and quantitative PCR assays for detection of the downy mildew pathogens of spinach (Peronospora effusa) and beet (P. schachtii). Phytopathology 104: 1349–1359 Knaus, B.J., Grünwald, N.J. (2017) vcfr: a package to manipulate and visualize variant call format data in R. Mol Ecol Resour 17: 44–53 Kozarewa, I., Armisen, J., Gardner, A.F., Slatko, B.E., Hendrickson, C.L. (2015) Overview of target enrichment strategies. Curr Protoc Mol Biol 112: 7.21.1-7.21.23 Kunjeti, S.G., Anchieta, A., Martin ,F.N., Choi, Y-J., Thines, M., Michelmore, R.W., Koike, S.T., Tsuchida, C., Mahaffee, W., Subbarao, K. V., Klosterman, S. J. (2016) Detection and quantification of Bremia lactucae by spore trapping and quantitative PCR. Phytopathology 106: 1426–1437 Lane, C.R., Beales, P.A., O’Neill, T.M., McPherson, G.M., Finlay, A.R., David, J., Constantinescu, O., Henricot, B. (2005) First report of Impatiens downy mildew (Plasmopara obducens) in the UK. Plant Pathol 54: 243 32 Langston, D., Sanders, F.H. (2013) Evaluation of fungicides for control of downy mildew on cucumber in Georgia I, 2012. Plant Dis. Manag. Rep. 7: V108. Lebeda, A., Urban, J. (2010) Screening for resistance to cucurbit downy mildew (Pseudoperonospora cubensis). MASS Screen. Tech. Sel. Crop. Resist. TO Dis. International Atomic Energy Agency, Viena, pp 284–295 Li, C-S., Lin, Y-C. (1999) Sampling performance of impactors for fungal spores and yeast cells. Aerosol Sci Technol 31: 226–230 Li, H., Durbin, R. (2009) Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics. doi: 10.1093/bioinformatics/btp324 Li, H., Handsaker, B., Wysoker, A., Fennell, T., Ruan J., Homer, N., Marth, G., Abecasis, G., Durbin, R., Subgroup 1000 Genome Project Data Processing (2009) The sequence alignment/map format and SAMtools. Bioinformatics 25: 2078–2079 Lim, H.C., Braun, M.J. (2016) High-throughput SNP genotyping of historical and modern samples of five bird species via sequence capture of ultraconserved elements. Mol Ecol Resour 16: 1204–1223 Lizotte, E., and Miles, T.D. (2020) Hop disease management in Michigan for 2020. MSU Ext., https://www.canr.msu.edu/news/managing-hop-downy-mildew-in-michigan Ma, X., Shao, Y., Tian, L., Flasch, D.A., Mulder, H.L., Edmonson, M.N., Liu, Y., Chen, X., Newman, S., Nakitandwe, J., et al (2019) Analysis of error profiles in deep next- generation sequencing data. Genome Biol 20: 50 McCormack JE, Tsai WLE, Faircloth BC (2016) Sequence capture of ultraconserved elements from bird museum specimens. Mol Ecol Resour 16: 1189–1203 Michigan Department of Agriculture & Rural Development (2018) Michigan Agriculture Facts & Figures. https://www.michigan.gov/documents/mdard/MI_Ag_Facts__Figures_474011_7.pdf Milgroom, M.G. (2015) Population biology of plant pathogens: genetics, ecology, and evolution. The American Phytopathological Society, St. Paul, Minnesota Milgroom, M.G. (1996) Recombination and the multilocus structure of fungal populations. Annu Rev Phytopathol 34: 457–477 Miller SA, Vargas A, Amrhein J (2020) Bioassay for sensitivity of Pseudoperonospora cubensis to fungicides in OH, 2019. Plant Dis Manag Rep Report 14: V024 33 Mitchell, M.N., Ocamb, C.M., Grünwald, N.J., Mancino, L.E., and Gent, D.H. (2011) Genetic and pathogenic relatedness of Pseudoperonospora cubensis and P. humuli. Phytopathology 101: 805–818 Naegele, R. P., Quesada-Ocampo, L. M. , Kurjan, J. D., Saude, C., and Hausbeck, M. K. (2016) Regional and temporal population structure of Pseudoperonospora cubensis in Michigan and Ontario. Phytopathology 106: 372–379 Nei, M (1972) Genetic distance between populations. Am Nat 106: 283–292 Neufeld, K.N., Keinath, A.P., Ojiambo, P.S. (2017) Evaluation of a model for predicting the infection risk of squash and cantaloupe by Pseudoperonospora cubensis. Plant Dis 102: 855–862 Ojiambo, P.S., Gent, D.H., Quesada-Ocampo, L.M., Hausbeck, M.K., and Holmes, G.J. (2015) Epidemiology and population biology of Pseudoperonospora cubensis: a model system for management of downy mildews. Annu Rev Phytopathol 53: 223–246 Ojiambo, P.S., and Holmes, G.J. (2010) Spatiotemporal spread of cucurbit downy mildew in the Eastern United States. Phytopathology 101: 451–461 Ojiambo, P.S., Holmes, G.J., Britton, W., Babadoost, M., Bost, S.C., Boyles, R., Brooks, M., Damicone, J., Draper, M.A., Egel, D.S., et al (2011) Cucurbit Downy Mildew ipmPIPE: A next generation web-based interactive tool for disease management and extension outreach. Plant Heal Prog 12: 26 Oyola, S.O., Gu, Y., Manske, M., Otto, T.D., O’Brien, J., Alcock, D., MacInnis, B., Berriman, M., Newbold, C.I., Kwiatkowski, D.P, et al (2013) Efficient depletion of host DNA contamination in malaria clinical sequencing. J Clin Microbiol 51: 745 LP – 751 Perea, C., De La Hoz, J.F., Cruz, D.F., Lobaton, J.D., Izquierdo, P., Quintero, J.C., Raatz, B., Duitama, J. (2016) Bioinformatic analysis of genotype by sequencing (GBS) data with NGSEP. BMC Genomics 17: 498 Perla, D., Hausbeck, M.K., Cook, A. (2019) Evaluation of fungicide programs for control of downy mildew of cucumber, 2018. Plant Dis Manag Reports 147: V13 Potapov, V., Ong, J.L. (2017) Examining sources of error in PCR by single-molecule sequencing. PLoS One 12: e0169774 Quesada-Ocampo L.M., Granke, L.L., Olsen, J., Gutting, H.C., Runge, F., Thines, M., Lebeda, A., and Hausbeck, M.K. (2012) The genetic structure of Pseudoperonospora cubensis populations. Plant Dis 96: 1459–1470 Quinlan, AR, Hall, IM (2010) BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26: 841–842 34 Rahman, A., Góngora-Castillo, E., Bowman, M.J., Childs, K.L., Gent, D.H., Martin, F.N., Quesada-Ocampo, L.M. (2019) Genome sequencing and transcriptome analysis of the hop downy mildew pathogen Pseudoperonospora humuli reveal species-specific genes for molecular detection. Phytopathology 109: 1354–1366 Reuveni, M., Eyal, H., Cohen, Y. (1980) Development of resistance to metalaxyl in Pseudoperonospora cubensis. Plant Dis 64: 1108–1109 Rivera, Y., Salgado-Salazar, C., Gulya, T.J., Crouch, J.A. (2016) Newly emerged populations of Plasmopara halstedii infecting rudbeckia exhibit unique genotypic profiles and are distinct from sunflower-infecting strains. Phytopathology 106: 752–761 Rogers, S.L., Atkins, S.D., and West, J.S. (2009) Detection and quantification of airborne inoculum of Sclerotinia sclerotiorum using quantitative PCR. Plant Pathol 58: 324–331 Royle, D.J. (1973) Quantitative relationships between infection by the hop downy mildew pathogen, Pseudoperonospora hutnuli, and weather and inoculum factors. Ann Appl Biol 73: 19–30 Runge, F., and Thines, M. (2011) Reevaluation of host specificity of the closely related species Pseudoperonospora humuli and P. cubensis. Plant Dis 96: 55–61 Savory, E.A., Adhikari, B.N., Hamilton, J.P., Vaillancourt, B., Buell, C.R., Day, B. (2012a) mRNA-Seq analysis of the Pseudoperonospora cubensis transcriptome during cucumber (Cucumis sativus L.) infection. PLoS One 7: e35796 Savory, E.A., Granke, L.L., Quesada-Ocampo, L.M., Varbanova, M., Hausbeck, M.K., Day, B. (2011) The cucurbit downy mildew pathogen Pseudoperonospora cubensis. Mol Plant Pathol 12: 217–26 Savory, E.A., Zou, C., Adhikari, B.N., Hamilton, J.P., Buell, C.R., Shiu, S-H., Day, B. (2012b) 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 Sharma, R., Xia, X., Cano, L.M., Evangelisti, E., Kemen, E., Judelson, H., Oome, S., Sambles, C., van den Hoogen, D.J., Kitner, M., et al (2015) Genome analyses of the sunflower pathogen Plasmopara halstedii provide insights into effector evolution in downy mildews and Phytophthora. BMC Genomics 16: 741 Sitterly, W.R. (1972) Breeding for disease resistance in cucurbits. Annu Rev Phytopathol 10: 471–490 35 Starrett, J., Derkarabetian, S., Hedin, M., Bryson, Jr. R.W., McCormack, J.E., Faircloth, B.C. (2017) High phylogenetic utility of an ultraconserved element probe set designed for Arachnida. Mol Ecol Resour 17: 812–823 Stassen, J., Seidl, M., Vergeer, P., Nijman, I., Snel, B., Cuppen, E., Van den ackerveken, G. (2012) Effector identification in the lettuce downy mildew Bremia lactucae by massively parallel transcriptome sequencing. Mol Plant Pathol 13: 719–731 Summers, C., Adair, N., Gent, D., Mcgrath, M., and Smart, C. (2015a) Pseudoperonospora cubensis and P. humuli detection using species-specific probes and high definition melt curve analysis. Can J Plant Pathol 37: 150605144651006 Summers C.F., Gulliford C.M., Carlson C.H., Lillis J.A., Carlson M.O., Cadle-Davidson L., Gent, D.H., Smart, C.D. (2015b) Identification of genetic variation between obligate plant pathogens Pseudoperonospora cubensis and P. humuli using RNA sequencing and genotyping-by-sequencing. PLoS One 10: e0143665 Tabima, J.F., Coffey, M.D., Zazada, I.A., Grünwald, N.J. (2018) Populations of Phytophthora rubi show little differentiation and high rates of migration among states in the western United States. Mol Plant-Microbe Interact 31: 614–622 Thines, M. (2014) Phylogeny and evolution of plant pathogenic oomycetes—a global overview. Eur J Plant Pathol 138: 431–447 Thines, M., Choi, Y-J. (2015) Evolution, Diversity, and Taxonomy of the Peronosporaceae, with Focus on the Genus Peronospora. Phytopathology 106: 6–18 Thomas, A., Carbone, I., Ojiambo, P. (2013) Comparative genomic analysis of Pseudoperonospora cubensis to elucidate the genetic basis of host specialization. Phytopathology 103: S2.145 Thomas, A., Carbone, I., Choe, K., Quesada-Ocampo, L.M., and Ojiambo, P.S. (2017a) Resurgence of cucurbit downy mildew in the United States: Insights from comparative genomic analysis of Pseudoperonospora cubensis. Ecol Evol 7: 6231–6246 Thomas, A., Carbone, I., Cohen, Y., Ojiambo, P.S. (2017b) Occurrence and Distribution of Mating Types of Pseudoperonospora cubensis in the United States. Phytopathology 107: 313–321 Thomas, A., Carbone, I., Lebeda, A., Ojiambo, P.S. (2017c) Virulence structure within populations of Pseudoperonospora cubensis in the United States. Phytopathology 107: 777–785 Thomas, A., Neufeld, K.N., Seebold, K.W., Braun, C.A., Schwarz, M.R., and Ojiambo, P.S. (2018) Resistance to fluopicolide and propamocarb and baseline sensitivity to 36 ethaboxam among isolates of Pseudoperonospora cubensis from the Eastern United States. Plant Dis 102: 1619–1626 Thornton, A.C, Adams, M.L., and Holmes, G.J. (2006) Evaluation of fungicides for control of downy mildew of cucumber I, 2006. Plant Dis Manag Rep 1: V122 McGrath, M (2004) Diseases of cucurbits and their management BT - Diseases of fruits and vegetables volume I: diagnosis and management. In SAMH Naqvi (ed) Springer Netherlands, Dordrecht, pp 455–510 Urban, J., Lebeda, A. (2006) Fungicide resistance in cucurbit downy mildew – methodological, biological and population aspects. Ann Appl Biol 149: 63–75 USDA (2020) United States Department of Agriculture. 2019. Vegetables Annual Summary. https://usda-library-cornell- edu.proxy1.cl.msu.edu/concern/publications/02870v86p?locale=en Vleeshouwers, V.G., Rietman, H, Krenek, P., Champouret, N., Young, C., Oh, S-K., Wang, M., Bouwmeester, K., Vosman, B., Visser, R.G.F., et al (2008) Effector genomics accelerates discovery and functional profiling of potato disease resistance and Phytophthora Infestans avirulence genes. PLoS One 3: e2875 Wallace, E., Adams, M., Ivors, K., Ojiambo, P.S., Quesada-Ocampo, L.M. (2014) First report of Pseudoperonospora cubensis causing downy mildew on Momordica balsamina and M. charantia in North Carolina. Plant Dis 98: 1279 Wallace, E., Adams, M., Quesada-Ocampo, L.M. (2015) First report of downy mildew on buffalo gourd (Cucurbita foetidissima) caused by Pseudoperonospora cubensis in North Carolina. Plant Dis 99: 1861 Wallace, E., Quesada-Ocampo, L. (2016) Pseudoperonospora cubensis on commercial and non-commercial cucurbits in North Carolina: Population structure determined by simple sequence repeats. Phythopathology 106: (Suppl) S412 Wallace, E.C., D’Arcangelo, K.N., Quesada-Ocampo, L.M. (2020) Population analyses reveal two host-adapted clades of Pseudoperonospora cubensis, the causal agent of cucurbit downy mildew, on commercial and wild cucurbits. Phytopathology. doi: 10.1094/PHYTO-01-20-0009-R Wallace, E.C., Quesada-Ocampo, L.M. (2017) Analysis of microsatellites from the transcriptome of downy mildew pathogens and their application for characterization of Pseudoperonospora populations. PeerJ 5: e3266 Withers, S., Gongora-Castillo, E., Gent, D., Thomas, A., Ojiambo, P.S., and Quesada- Ocampo, L.M. (2016) Using Next-Generation sequencing to develop molecular diagnostics for Pseudoperonospora cubensis, the cucurbit downy mildew pathogen. Phytopathology 106: 1105–1116 37 Wong, F.P, and Wilcox, W.F (2001) Comparative physical modes of action of azoxystrobin, mancozeb, and metalaxyl gainst Plasmopara viticola (grapevine downy mildew). Plant Dis 85: 649–656 Wu, E-J., Yang, L-N., Zhu, W., Chen, X-M., Shang, L-P., Zhan, J. (2016) Diverse mechanisms shape the evolution of virulence factors in the potato late blight pathogen Phytophthora infestans sampled from China. Sci Rep 6: V26182 Zhang, Y., Pu, Z., Qin, Z., Zhou, X., Liu, D., Dai, L., Wang, W. (2012) A study on the overwintering of cucumber downy mildew oospores in China. J Phytopathol 160: 469– 474 Zhu, S.S., Liu, X.L., Wang, Y., Wu, X.H., Liu, P.F., Li, J.Q., Yuan, S.K., and Si, N.G. (2007) Resistance of Pseudoperonospora cubensis to flumorph on cucumber in plastic houses. Plant Pathol 56: 967–975 38 CHAPTER II: DETECTION OF AIRBORNE SPORANGIA OF PSEUDOPERONOSPORA CUBENSIS AND P. HUMULI IN MICHIGAN USING BURKARD SPORE TRAPS COUPLED TO QPCR 39 ABSTRACT Cucurbit downy mildew (CDM), caused by the oomycete pathogen Pseudoperonospora cubensis, is a devastating foliar disease on cucumber resulting in reduced yields. In 2004, the pathogen re-emerged in the U.S., infecting historically resistant cucumber cultivars and requiring the adoption of an intensive fungicide program. The pathogen cannot overwinter in Michigan fields but due to an influx of airborne sporangia cucurbit downy mildew occurs annually. In Michigan, spore traps are used to monitor the presence of airborne P. cubensis sporangia in cucumber growing regions to guide the initiation of a fungicide program. However, Pseudoperonospora humuli sporangia, the causal agent of downy mildew on hop, are morphologically indistinguishable from P. cubensis sporangia. This morphological similarity reduces the ability to accurately detect P. cubensis from spore trap samples when examined with the aid of light microscopy. To improve P. cubensis detection, we adapted a qPCR- based assay to allow the differentiation between P. cubensis and P. humuli on Burkard spore trap samples collected in the field. Specifically, we evaluated the specificity and sensitivity of P. cubensis detection on Burkard spore trap tapes using a morphological based and qPCR-based identification assay and determined whether sporangia of P. cubensis and P. humuli on Burkard samples could be distinguished using qPCR. We found that the qPCR assay was able to detect a single sporangium of each species on spore trap samples collected in the field with Cq values below 35.5. The qPCR assay also allowed the detection of P. cubensis and P. humuli in samples containing sporangia from both species. However, the number of sporangia quantified using light microscopy explained only 54% and 10% of the variation in the Cq values of P. cubensis 40 and P. humuli, respectively, suggesting a limited capacity of the qPCR assay for the absolute quantification of sporangia in field samples. After two years of monitoring using Burkard spore traps coupled with the qPCR in cucumber fields, P. humuli sporangia were detected more frequently than P. cubensis early in the growing season (May and June). P. cubensis sporangia were detected approximately 5 -10 days before cucurbit downy mildew symptoms were first observed in cucumber fields during both years. This research describes an improved sporangial detection system that is key for the monitoring and management of P. cubensis in Michigan. INTRODUCTION Pseudoperonospora cubensis (Berk. & M. A. Curtis) Rostovzev, the causal agent of cucurbit downy mildew (CDM), infects approximately 20 cucurbit genera including the economically important crops of cucumber, cantaloupe, squash, watermelon and pumpkin (Cohen et al., 2015; Lebeda and Cohen, 2011). CDM symptoms include angular, chlorotic lesions that coalesce and become necrotic, resulting in leaf blight and death; pathogen sporulation occurs on the abaxial side of the leaf (Salcedo et al., 2020). In cucumber, foliar blighting resulting from CDM can result in yield reduction (Hausbeck et al., 2019; Perla et al., 2019; Reuveni et al., 1980). In the U.S., Michigan is the largest producer of pickling cucumber and the second-largest producer of cucumber for the fresh market, with approximately 4.1 million cwt of cucumbers sold in 2018 (USDA, 2020). For nearly 40 years, resistant cucumber cultivars had been used successfully to mitigate CDM (Brzozowski et al., 2016). In 2004, a highly virulent strain of P. cubensis emerged in the U.S. overcoming this host resistance (Thomas et al., 2017) and since then fungicides have been relied on 41 for control (Blum et al., 2011; Holmes et al., 2014). However, fungicide-resistant P. cubensis isolates have presented crop protection challenges. Single-site fungicides including mefenoxam and azoxystrobin were ineffective when the pathogen emerged in 2004 (Ernest et al., 2005; Gevens and Hausbeck, 2005; Thornton et al., 2006). Since that time, P. cubensis resistance to dimethomorph (Zhu et al., 2007) and mandipropamid (Hausbeck and Cortright, 2010; Blum et al., 2011) has been reported in the U.S. (Holmes et al., 2014; Keinath, 2015). Similarly, reduced efficacy of fluopicolide against CDM has been observed in field trials in Michigan (Hausbeck and Linderman, 2014), Georgia (Langston and Sanders, 2013), and North Carolina (Adams and Quesada-Ocampo, 2014). While propamomcarb was effective against CDM for several years, since 2013 its efficacy appeared to be compromised in field trials in North Carolina (Keinath, 2015; Thomas et al., 2018), Pennsylvania (Gugino and Grove, 2016), and Michigan (Hausbeck and Linderman, 2014; Hausbeck et al., 2017). P. cubensis is an obligate pathogen and its survival depends on the availability of susceptible hosts (Cohen et al., 2015). The pathogen does not survive in regions that experience frost, instead its sporangia are dispersed to northern latitudes from overwintering sites (Ojiambo and Holmes, 2010; Quesada-Ocampo et al., 2012). Airborne sporangia concentrations influence CDM onset (Granke et al., 2013) and under conducive weather conditions, P. cubensis sporangia can spread rapidly within and between fields (Ojiambo et al, 2015). Airborne concentrations of P. cubensis sporangia in Michigan’s cucumber fields have been monitored using Burkard spore traps (Burkard Manufacturing Co Ltd, U.K.) with light microscopy used to identify and enumerate the pathogen sporangia based on morphology (Granke and Hausbeck, 42 2011; Granke et al., 2013). Pseudoperonospora humuli, the causal agent of hop downy mildew (HDM) is nearly identical to P. cubensis morphologically (Runge and Thines, 2011) but rarely infects cucurbits in the U.S. (Mitchell et al., 2011). Approximately 400 ha of hops are planted in Michigan (Michigan Department of Agriculture & Rural Development, 2018) and P. humuli is prevalent (Lizotte et al., 2020). Thus, relying on morphological identification, alone, to monitor airborne sporangial concentrations of P. cubensis could result in inaccurate estimations of the pathogen's presence and concentration. PCR-based methods have been used successfully to detect and quantify airborne plant pathogens such as Peronospora effusa (Klosterman et al., 2014), Peronospora schachtii (Klosterman et al., 2014) Claviceps purpurea (Dung et al., 2018) and P. humuli, infecting spinach, beet, grass-seed and hop (Gent et al., 2009), respectively. A qPCR assay was developed that differentiates between P. cubensis and P. humuli sporangia (Summers et al., 2015). This assay, with or without microscopic visualization of spore trap tapes, could accelerate the speed and accuracy of P. cubensis detection and inform the initiation of fungicide sprays. The objective of our study was to improve the detection of airborne concentrations of P. cubensis sporangia by adapting a qPCR-based assay (Summers et al., 2015) that distinguishes between P. cubensis and P. humuli using Burkard spore trap samples collected in the field. MATERIALS AND METHODS In vitro evaluations to assess the sensitivity of the qPCR-based assay were performed using isolates of P. cubensis (CDM23) and P. humuli (HDM19) obtained in 2017 using methods similar to those described by Thomas et al. (2017). Briefly, 43 diseased tissue was placed in a moist chamber overnight to induce sporulation. Sporangia from a single cucumber leaf lesion or an infected hop basal shoots (spike) were suspended in 1 ml of distilled water and the resulting inoculum (1,000-10,000 sporangia/ml) applied to the abaxial side of detached leaves of ‘Vlaspik’ cucumber or ‘Centennial’ hop, respectively, contained in Petri dishes (100 X 15 mm). Inoculated leaves were then incubated in a growth chamber at 18°C under a 12/12-hr light/dark cycle. Seven to 10 days post inoculation, sporangia were gently rinsed from infected leaves using a Preval spray power unit (Preval, Chicago) filled with distilled water. A new set of leaves were inoculated with the resulting sporangia. Collection of sporangia and extraction of genomic DNA. P. cubensis (CDM23) and P. humuli sporangia (HDM19) were gently rinsed from host tissue into centrifuge tubes (50 ml) using a Preval spray power unit filled with distilled water. The sporangial suspension was concentrated by centrifugation (5424R centrifuge, Eppendorf) at 14,000 rpm for 5 min and homogenized in impact-resistant 2mL tubes (Lysing Matrix H, MP Biomedicals) using a TissueLyser II (Qiagen, Valencia, CA) for 4min at 30 Hz. DNA was extracted using a NucleoSpin Plant II isolation kit (Macherey- Nagel, Bethlehem, PA) following manufacturer’s instructions and DNA concentration was determined using the Qubit double-stranded DNA High Sensitivity Assay Kit (Life Technologies, Carlsbad, CA). Competitive qPCR internal control. A competitive internal positive control (IC) was designed in this study and incorporated into every qPCR reaction to monitor for the presence of PCR inhibitors in each sample. The IC consisted of a single-stranded linear synthetic DNA that utilizes the same primers of the target mitochondrial cox2 gene, and 44 an additional fluorogenic probe (ICprobeJ2: /5CYS/A+GCATTATT+GTTTAT+CATATATACA/3IABkFQ/) for amplification and detection. The sequence of the internal control (Table 2-1) showed no significant nucleotide identity to any known naturally occurring PCR-amplifiable nucleotide sequences reported in the NCBI database. qPCR protocol with purified DNA. All qPCR experiments were conducted using a modified version of the protocol described by Summers et al. (2015) in accordance with the Minimum Information for Publication of Quantitative Real-Time Experiments (MIQE) guidelines (Bustin et al., 2009). The Summers et al. (2015) assay was modified by changing the commercial master mix IQ Supermix (Bio-rad, Hercules, CA) to the Prime-Time Gene Expression Master Mix (IDT, Skokie, IL). This new master mix reduced the variation between technical replicates and increased the amplification efficiency of qPCR reactions. Additionally, we also added an internal-positive control to identify any alterations in amplification efficiency in field air samples. qPCR reactions with a final volume of 20 µl were manually assembled in 96-well white plates (Bio-rad MLL9651) containing 10 µl of the Prime-Time Gene Expression Master Mix, 2 µl of template DNA, 600 nM of each primer (RT33F and RT182R), 500 nM of the LNA probe HUMprobeSNP105, 250 nM of the LNA probe CUBprobeSNP105, 250 nM of the LNA probe ICprobeJ2, and 7.5 x10-10 nM (0.75 aM) of our internal positive control (IC) (Table 2-1). The IC was set to this concentration to obtain a Cq value of 29 without affecting the sensitivity or specificity of the other probes. Negative control reactions lacking the DNA template were included in every plate run. The qPCR protocol run on a CFX 96 Touch qPCR system (Bio-rad) and included an initial denaturation step at 95°C for 3 min 45 followed by 38 cycles of 95°C for 10 s and 65°C for 45 s. Two technical replicates of each sample were run and the average Cq and standard deviation were calculated using Bio-rad CFX Manager software (version 3.1) (Bio-rad 1845000). Sensitivity and specificity of qPCR with LNA probes. The Burkard spore traps use a vacuum pump to draw air (approximately 10 liters/min) into a collection chamber containing a reel, covered with a melinex tape; the reel was mounted on a clockwork mechanism (Hirst, 1952). The melinex tape (Burkard Manufacturing Co. Ltd., U.K.) was coated with an adhesive of petroleum jelly and paraffin (9:1 wt/wt) dissolved in sufficient toluene to provide adequate coverage of tape at the desired thickness (Granke et al., 2013). The sensitivity and specificity of the qPCR assay was evaluated using three different experiments; the first generated a standard curve using pure DNA of P. humuli and P. cubensis, the second used samples that contained a mixture of P. humuli and P. cubensis DNA and the third used samples of P. cubensis sporangia that also contained the melinex tape and the adhesive used on the Burkard spore trap tapes. The first procedure included ten-fold dilutions of genomic DNA from two independent DNA extractions from each pathogen isolate (CDM23 and HDM19) which were used to generate standard curves ranging from 10 to 100,000 fg. Three technical replicates of each sample dilution were tested using qPCR and the average Cq values with the standard deviation were calculated using the Bio-rad CFX Manager software (Bio-rad). Mean Cq values were plotted against the log10 of template DNA concentrations and used to generate standard curves. The second procedure included the evaluation of mixed-DNA samples to assess the specificity of the qPCR assay. An in vitro 46 assessment was used to determine whether the assay could detect P. cubensis and P. humuli in mixed samples containing DNA from both pathogens. Ten-fold dilutions of genomic DNA from each species were made from 100 to 100,000 fg and used as templates, both separately and mixed in varying concentrations (Table 2-2), for the qPCR assay described above. Three technical replicates of each concentration and mixture were run and the average Cq and standard deviations were calculated using Bio-rad CFX Manager software. The third procedure included five independent extractions from solutions containing 10, 100, or 1000 sporangia of P. cubensis prepared using a hemocytometer counting cell chamber. Sporangia were homogenized in impact-resistant 2 mL tubes using a TissueLyser (4 min at 30 Hz) and DNA was extracted using the NucleoSpin Plant II isolation kit (Macherey-Nagel, Bethlehem, PA). Subsequently, 2 µl of the extraction product were evaluated using qPCR. Finally, 9 x 48 mm sections (representing a 24 h-sampling period) of melinex tape with adhesive were spiked with 10, 20, 50, 100 or 300 P. cubensis sporangia. DNA was extracted and evaluated using qPCR as previously described. Three technical replicates of an average of four independent extractions of each sample dilution were tested and the average Cq values with standard deviation were calculated using the Bio-rad CFX Manager software. Collection of field samples for screening using qPCR. Airborne sporangial concentrations were monitored during the cucumber growing season (May to September) in 2018 and 2019 using Burkard spore traps. Each year, a spore trap was placed 20 m from a commercial cucumber field located in Muskegon County in northwest Michigan and a cucumber research plot at the Michigan State University 47 (MSU) Plant Pathology Farm located in Ingham County in southcentral Michigan. The MSU cucumber research plot (0.25 ha) was direct seeded during the last week of July and was located 200 m from an abandoned hop research yard (0.25 ha) where systemically infected basal shoots (spikes) were observed beginning in late April 2019. An additional Burkard spore trap was placed in a commercial hop yard in Berrien County in 2019. The reel in each Burkard trap was covered with a melinex tape coated with an adhesive as described previously. The tape was removed weekly and cut longitudinally along the center line into two subsections of 9 x 336 mm each (Rogers et al., 2009). The first section was then cut into 48-mm lengths (equivalent to a monitoring period of 24 h), scored at hourly intervals (2 mm) and stained to facilitate counting according to the protocol described by Granke et al. (2013). The second section was also cut into 48- mm lengths, placed into impact-resistant 2mL tubes (Lysing Matrix H, MP Biomedicals) and subjected to DNA extraction as previously described using a NucleoSpin Plant II isolation kit (Macherey-Nagel, Bethlehem, PA). Subsequently, 2 µl of the extraction product was evaluated using qPCR. Fields were scouted weekly for signs and/or symptoms of P. cubensis. Leaf samples with lesions resembling CDM symptoms and signs of the pathogen were returned to the laboratory and examined using light microscopy to verify the presence of sporangia. . RESULTS Sensitivity and specificity of qPCR. Using 10-fold dilutions of P. cubensis and P. humuli DNA, the qPCR assay exhibited a significant linear response with an efficiency of 93.6% (R2=0.99) and 90.7% (R2=1), respectively (Figs. 2-1A, 2-1C). Both 48 species-specific LNA probes detected each pathogen within total DNA template amounts ranging between 100 to 100,000 fg per reaction (Figs. 2-1A, 2-1C). The average Cq values for samples containing 100 fg of P. humuli and P. cubensis DNA was < 35.5. Most samples with concentrations below 100 fg were either not detected or detected without reasonable certainty (<95% of the times tested); thus, 100 fg of template DNA was considered as the lower limit of detection (LOD) of the qPCR assay for both species. Although the LNA probes were specifically designed to detect either P. humuli or P. cubensis based on the recognition of a SNP at the 105-base of the cox2 gene (Summers et al., 2015), the HEX-labelled LNA probe CUBprobeSNP105 for P. cubensis detection showed nonspecific amplification of P. humuli DNA. However, the amplification curves of P. humuli DNA with this probe did not show the same shape as those generated using the P. cubensis DNA (Fig. 2-1B). P. cubensis samples were classified as positive based on the shape of the amplification curve using the probe CUBprobeSNP105 and no amplification of the FAM-labelled LNA probe HUMprobeSNP105. This probe (HUMprobeSNP105) designed to recognize P. humuli was highly specific and no background signal was observed when P. cubensis DNA was analyzed (Figs. 2-2D). Additionally, when mixed-samples containing DNA from both species were assessed using the qPCR assay, no significant changes were observed in the Cq values of the samples containing P. cubensis or P. humuli DNA in a 1:1 ratio (Table 2- 2). However, when 10,000 fg of P. cubensis DNA was mixed with 100 fg of P. humuli DNA, no detection of P. humuli DNA was observed (Table 2-2). Detection of P. cubensis occurred in all mixtures, but when 10,000 fg of P. humuli DNA was mixed with 49 1,000 or 100 fg of P. cubensis DNA, the detection of P. cubensis occurred at significant lower Cq values than in reactions including only P. cubensis DNA (Table 2-2). Mixed- samples containing DNA from both species generated amplification curves with both probes (CUBprobeSNP105 and HUMprobeSNP105). This clearly differentiated them from samples containing only P. cubensis DNA for which there was only amplification with the CUBprobeSNP105 probe (Supplementary Fig. 2-S1 A, C). Although the samples containing only P. humuli DNA also generated amplification curves with both probes (Supplementary Fig. 2-S1 B), the amplification curves generated with the CUBprobeSNP105 probe for mixed samples showed faster growth in the exponential phase of the curves. This was the case for mixed-samples containing DNA from both species in a 1:1 ratio and samples containing P. cubensis and P. humuli DNA in 10:1 ratio (Supplementary Fig. 2-S1 C, D). Only the amplification curves of the samples containing P. cubensis and P. humuli DNA in 1:10 and 1:100 ratio were not clearly differentiated from the amplification curves containing only P. humuli DNA (Supplementary Fig. 2-S1 B, E). Both LNA probes detected DNA from extractions containing 10, 100 or 1000 sporangia of P. cubensis and P. humuli. Upon regression analysis, a linear relationship between Cq values and DNA extracted from purified sporangia was observed for both species (P. cubensis, R2= 0.99, Pvalue = 0.03 and P. humuli, R2= 0.98, Pvalue = 0.084) (Fig. 2-2A). The average Cq value for detecting 10 P. cubensis sporangia was < 35.5 (Fig. 2-2A) and samples with less than 10 sporangia could not be detected with reasonable certainty (>95% of the times tested). Cq values ≤ 35.5 were classified as 50 specific to P. humuli and P. cubensis and were used as a threshold to evaluate field samples. The sensitivity of the assay to detect P. cubensis was minimally affected by the adhesive applied to the melinex tape (Fig. 2-2B). DNA was detected from extractions of tapes spiked with 20, 50,100, or 300 sporangia and a reduced number of samples (4/10) containing 10 sporangia had average Cq values below 35.5. All the extractions showed Cq values that were significantly different from the background signal observed in the negative controls. The relationship between sporangial numbers and Cq values was significant (p = 0.003) and the assay exhibited a linear response with a R2 value of 0.99 (Fig. 2-2B). The average Cq values of the different sporangial dilutions were within the 95% confidence interval. However, a high standard error of the mean was observed among biological replicates of extractions with the same number of sporangia (Fig. 2- 2B) indicating that the extraction affects the precision of quantifying sporangia using the qPCR-based assay. Assessment of field samples using light microscopy and qPCR. A total of 560 samples collected from May to August in 2018 and 2019 were assessed using qPCR. P. cubensis or P. humuli DNA was detected using qPCR on field samples with fewer than 10 sporangia (Fig. 2-2C). Approximately 90% of all samples that tested positive for either pathogen using qPCR (204 out of 227) had one or more sporangia on the corresponding half of the tape analyzed using light microscopy. The average Cq value of the IC remained relatively constant and had an average of 28.8 ± 1.7 (SD) among all the field samples evaluated. Regression analysis indicated that the number of sporangia on the second half of the tape of the samples (quantified using light 51 microscopy) explained 54% (R2 = 0.54) and 10% (R2 = 0.10) of the variation in the Cq values of P. cubensis and P. humuli, respectively (Fig. 2-2C). Using light microscopy, Pseudoperonospora spp. sporangia were first detected in 2018 during May (Muskegon Co. commercial field, 15 May) and June (Ingham Co. research field, June 13) (Fig. 2-3). From May to July at the research field, fewer than 5 sporangia/day were observed via light microscopy with P. cubensis DNA confirmed using qPCR on 13 June and 10, 23 July; while P. humuli was confirmed on 19 June and 7 July (Fig. 2-3A). At the commercial field, fewer than 10 sporangia/day were detected using light microscopy from May to July, except for 21 June (Fig. 2-3B). Using qPCR, we confirmed the presence of P. humuli DNA on 29 May to 4 June and 21 to 26 June (Fig. 2-3B) while P. cubensis DNA was detected on 5 and 13 June (Fig. 2-3B). Airborne sporangial concentrations increased during the first week of August and reached the maximum during the third week of the month in both locations monitored in 2018 (Figs. 2-3A, 2-3B). CDM symptoms were observed at the research field on 15 August, a peak of 16 sporangia was observed via light microscopy 10 days prior (5 August) and P. cubensis DNA was confirmed by qPCR on 1 to 6 August and 10 August (Fig. 2-3A). After CDM symptoms were observed in the research field, daily sporangial counts via light microscopy exceeded 10 sporangia/day with P. cubensis DNA detected nearly every day using qPCR (Figs. 2-3A, 2-3B). CDM symptoms were observed at the commercial field on 7 August, a peak of 22 sporangia was observed via light microscopy four days prior (3 August) and P. cubensis DNA was confirmed by qPCR on 3, 5 August, and 28, 30, and 31 July (Fig. 2-3B). The day following the detection of 52 CDM symptoms, more than 80 sporangia/day were captured by the spore traps and P. cubensis DNA was detected every day using qPCR (Fig. 2-3B). In 2019, Pseudoperonospora spp. sporangia were first detected in May across all locations using light microscopy (Fig. 2-4). During May and June, concentrations of airborne sporangia exceeded 10 sporangia/day in the commercial hop yard (Fig. 2-4A) and the research field (Fig. 2-4B). Using qPCR, P. humuli DNA was detected several times in the commercial hop yard from May through August while P. cubensis DNA was only detected on 12, 14, and 18 August (Fig. 2-4A). Based on data from light microscopy, the research field which was in proximity to a non-treated hop yard, had more than 10 sporangia/day during the last week of May, the first and fourth week of June and the first week of July (Fig. 2-4B). At this location, P. humuli DNA was detected from May to July using qPCR (Fig. 2-4B). During August, fewer than 10 sporangia/day were observed at the research field. CDM symptoms were confirmed at this site on 21 August and P. cubensis DNA was verified by qPCR on 11, 19, and 20 August. Following CDM symptoms, P. cubensis DNA was detected from 22 to 31 August (Fig. 2-4B). In the commercial cucumber field, fewer than 10 sporangia/day were observed using light microscopy from May through July with the exception of 18, 21 and 23 June (Fig. 2-4C). P. humuli DNA was confirmed with qPCR on 22 and 23 June (Fig. 2-4C). At this location, the sporangial counts increased from the third to the last week of August. CDM symptoms were confirmed on 16 August and P. cubensis DNA was detected using qPCR in air samples every day from 12 to 31 August (Fig. 2-4C). 53 DISCUSSION The ability to detect and differentiate between P. cubensis and P. humuli in field air samples using qPCR represents an important advance for CDM monitoring and management. The qPCR detection of airborne sporangia could be used as a decision- making tool to initiate fungicide sprays (Dhar et al., 2019) or as a complementary variable to forecast the risk (Carisse et al., 2009) of CDM outbreaks in Michigan. Early and specific detection of P. cubensis sporangia could ensure timely crop protection and avoid unnecessary fungicide applications. P. cubensis does not overwinter in Michigan and for disease to occur, the pathogen must be introduced into the state’s growing regions annually (Naegele et al., 2016). Burkard spore traps coupled with light microscopy have been used since 2007 to alert Michigan growers to an influx of P. cubensis sporangia into their growing region. However, HDM is prevalent in the state (Lizotte et al., 2020) where approximately 400 ha of hops have been planted (Hop Growers of America, 2019). Using a qPCR assay, we were able to distinguish between the morphologically similar sporangia of P. cubensis and P. humuli collected from Burkard spore traps. During the two years of monitoring using Burkard spore traps coupled with the qPCR assay in cucumber fields, P. cubensis sporangia were detected approximately 5-10 days before CDM symptoms were observed in monitored cucumber fields. We adapted the qPCR assay developed by Summers et al. (2015) to a high degree of sensitivity for use with the Burkard spore trap samples. Using DNA extracted from purified sporangial suspensions of P. cubensis and P. humuli, we were able to 54 detect DNA concentrations ranging from 100 fg to 100,000 fg. This sensitivity was validated with the detection of the two downy mildew pathogens in field samples containing less than 10 sporangia (Cq <35.5). We split the tape of Burkard spore traps to facilitate the comparison between light microscopy and qPCR and observed that the number of sporangia deposited onto one half of the tape was linearly correlated with the Cq values obtained after the assessment of the other half using qPCR. However, the change in the number of sporangia on field samples quantified using light microscopy explained only 54% and 10% of the variation in the Cq values of P. cubensis and P. humuli, respectively, suggesting a limited capacity of the qPCR assay for the absolute quantification of sporangia in field samples. The low correlation between Cq values and sporangial numbers of field samples may be explained by the high variation in the yield of DNA extraction among samples (Summers et al., 2015), inaccurate visual quantification, the low specificity of one of our probes, and possibly, the multicopy nature of the target sequence (Klosterman et al., 2014; Kunjeti et al., 2016; Dung et al., 2018). Similarly, the yield variation of DNA extractions among samples may also explain the variation observed in the Cq values of samples with the same number of sporangia in vitro. This variation is introduced in all the samples collected in the field and may reduce the precision for the quantification of sporangia using the extraction protocol and qPCR assay described in this study. However, assessing the first half of the spore trap tape using qPCR could reduce the number of samples that require microscopic analysis for spore quantification, accelerating the turn-around time associated with monitoring airborne P. cubensis sporangia. The reduction of variation in the yield of DNA extractions and the utilization 55 of a qPCR assay based on a single-copy marker may be more appropriate for quantification (Rahman et al., 2020) but may result in a system with reduced sensitivity when compared to the multi-copy system that we used. The nonspecific amplification of P. humuli DNA affected the quantification capacity of the assay when both species were present in the same reaction. However, the inclusion of a second probe ensured that the detection of each species was possible even when P. cubensis and P. humuli were present in the same sample. The nonspecific amplification of P. humuli did not occur under the conditions described by Summers et al. (2015) and was a consequence of the change in the commercial master mix used for the qPCR reactions (Supplementary Fig. 2-S2). The master mix used in this study reduced the variation among technical replicates (data not shown) and increased the amplification efficiency of the qPCR reactions (exponential phase) but affected the specificity of the assay. Different qPCR master mixes influences how oligonucleotides (primers and probes) bind to target regions (Morinha et al., 2020), thus the suitability of new reagents must be carefully evaluated as they may condition the results of the qPCR. The detection of both species using our qPCR assay was hindered only in samples with a significantly higher amount of P. humuli compared to P. cubensis (i.e. samples containing P. cubensis and P. humuli DNA in 1:10 and 1:100 ratio). In these cases, or in locations where a higher number of P. humuli sporangia relative to the number of P. cubensis sporangia is expected (e.g. hop yards) the use of the IQ Supermix (Bio-rad, Hercules, CA) as described by Summers et al. (2015) for qPCR reactions should allow a more accurate evaluation of the samples. 56 During two years of monitoring in commercial cucumber fields, we did not detect any periods when both pathogens were detected simultaneously, however, overlapping periods may have occurred at the commercial hop yard during late August. The identification of genetic regions with a higher number of polymorphisms has allowed the design of more specific primer and probes for P. cubensis detection (Rahman et al., 2020). Using this new set of primers and probes in combination with the probe HUMprobeSNP105 (for detection of P. humuli) could ensure both specific detection and quantification of P. cubensis and P. humuli sporangia using qPCR even during periods when both species are present. Despite the limitations of the qPCR assay described in this study, we were able to detect low atmospheric concentrations of P. cubensis and P. humuli (<10 sporangia/day). Detection of P. cubensis before symptoms developed in the field was linked to a sporangial concentration below 10 sporangia/day as estimated using light microscopy. In other crops including lettuce and onion, measurements of aerial spore load (sporangia/day) have been used to guide fungicide application to control Bremia lactucae (Dhar et al., 2019) and Botrytis squamosa (Carisse et al., 2009), respectively. In these systems, fungicide applications began once spore loads reached a critical level between 300-500 spores/day (10 sporangia/m3). In the cucumber fields monitored in 2018, CDM symptoms were observed after airborne P. cubensis sporangial concentrations were greater than 15 sporangia/day as determined via light microscopy, suggesting that the critical concentration for CDM could be close to this number depending on the coincident environmental conditions. Using the qPCR assay, P. cubensis sporangia were detected before concentrations reached >15 sporangia/day. 57 More than 15 sporangia/day were also observed one month before CDM was detected in the cucumber fields (June 2018 and 2019), however, these sporangia were identified as P. humuli using qPCR. In Michigan, information on the airborne concentration of sporangia is used to provide an early warning for growers that CDM is likely to occur (Hausbeck, 2020) and prompt the application of fungicides. Using light microscopy only, P. humuli sporangia could have triggered unnecessary fungicide applications, highlighting the importance of a qPCR assay system that reliably distinguishes between P. cubensis and P. humuli. Using light microscopy and qPCR, differences in the airborne sporangial concentrations of P. humuli and P. cubensis were detected between the two years of monitoring. From June to August 2018 at the commercial cucumber field, we detected higher airborne sporangial concentrations of P. cubensis compared to 2019. A relatively cold and rainy spring delayed the planting of cucumbers for pickling in 2019 (USDA, 2020). This reduced host availability may have also resulted in reduced infection and P. cubensis sporangia production. Similarly, from May to July 2018, low concentrations of P. humuli (<10 sporangia/day) were detected in the two monitored fields whereas a higher concentration (>10 sporangia/day) was observed at the three locations monitored in 2019. P. humuli overwinters in dormant hop buds or crowns, growing into expanding basal shoots in spring and early summer (Coley-Smith, 1962). Extended periods of wetness, high RH, and temperatures below 20°C (Royle, 1973; Gent and Ocamb, 2009) occurred during the cold and rainy spring of 2019 (NOAA, 2019) and may have favored the pathogen’s reproduction and infection. 58 These results suggest that the qPCR-based assay allowed for precise monitoring of airborne P. cubensis and P. humuli sporangia over two different years; the specific detection of these two species was not possible using light microscopy only. Cucumber growers in Michigan desire to know when sporangia of P. cubensis have arrived in their production region/field so that scouting efforts can be intensified, and costly fungicide programs initiated. The information on P. cubensis detection derived from spore traps coupled with qPCR could be used by growers to make informed decisions regarding fungicide usage leading to increased efficiency. Judicious use of fungicides may slow the development of pathogen resistance and decrease the cost associated with CDM control. The deployment of a broader network of spore traps and the evaluation of air samples using qPCR could also improve the risk assessment of CDM epidemics. Future evaluation of more cost-effective spore traps such as impaction traps for the monitoring of P. cubensis is essential to increase the geographic coverage of the spore trapping network in Michigan. The use of more spore traps at a local level could make the monitoring more geographically precise and trigger the execution of disease management practices only in fields at high risk of infection based on the qPCR detection of P. cubensis and the local environmental conditions. 59 APPENDIX 60 Table 2-1. Primers and locked nucleic acids (LNA) probes for the qPCR assay differentiating Pseudoperonospora cubensis and P. humuli using the 105 SNP in the mitochondrial Cox2 gene. APPENDIX Code name [Conc]a Sequence 5'->3' Primer RT33Fb Primer RT182Rb 600 nM 600 nM AACTCCCGTTATGGAAGGTATT CCATGTACAACAGTAGCTGGA Probe CUBprobeSNP105b 250 nM HEX/A+C+AAA+C+G+AATA+CT/BHQ c Probe HUMprobeSNP105b 500 nM FAM/AA+C+AAA+C+A+AATA+CTG/BHQ c Probe ICprobeJ2 250 nM CYS/A+GCATTATT+GTTTAT+CATATATACA/BHQc Internal Control IC 7.5 x10-10 nM GATTTGTATATATGATAAACAATAATGCTATAAC (0.75 aMd) ATAGAGTCTCTTTCATGAATAATCCAGCTACTGT AACTCCCGTTATGGAAGGTATTATCATTAATCAT TGTACATGG aConcentrations used in a 20 µl qPCR reaction. bPrimers and LNA probes were adapted from Summers et al. (2015). cLocked nucleic acids in the probes are followed by a plus (+) sign dAttomole (aM) = 10-18 moles per liter 61 Table 2-2. Threshold cycle (Cq) values of the qPCR assays using LNA probes and varying concentrations of genomic DNA. DNA Probe CUBprobeSNP105 HUMprobeSNP105 Cq-HEX SD Cq-FAM P. cubensis P. humuli 10,000 fg 1,000 fg Un- 100 fg mixed -- -- -- -- -- -- 28.39a 32.12b 35.49c 10,000 fg 29.88nc 1,000 fg 32.96nc 100 fg NA 10,000 fg 10,000 fg 28.67a 1,000 fg 1,000 fg 31.44b 100 fg 100 fg 35.32c Mixed 10,000 fg 1,000 fg 28.37a 10,000 fg 100 fg 28.39a 1,000 fg 10,000 fg 29.78x 100 fg 10,000 fg 30.58y 0.10 0.21 0.74 0.34 0.17 NA 0.23 0.37 0.98 0.15 0.26 0.37 0.24 NA NA NA 28.29d 31.70e 36.61f 28.52d 32.42e 35.11f 32.07e NAz 27.96d 24.99d SD NA NA NA 0.08 0.30 0.14 0.60 0.08 0.59 0.33 NA 0.34 5.91 The HEX-labelled probe (CUBprobeSNP105) was designed to detect only DNA from P. cubensis and the FAM-labelled probe (HUMprobeSNP105) was designed to detect only DNA from P. humuli. Cq values with the same letter are not significantly different (t-Test; P=0.05). --: not DNA added. NC: not used for comparison in the t-Test. NA: not defined. 62 Figure 2-1. Regression and amplification curves of Pseudoperonospora cubensis and P. humuli DNA using qPCR. A. Standard curve for the quantification of P. cubensis and P. humuli DNA using the LNA probe CUBprobeSNP105. The log10 of DNA (100fg, 1,000fg, 10,000fg, and 100,000fg) is plotted against the quantification cycle (Cq) values. Each curve was plotted separately using DNA from each pathogen. The data points below 100fg were not included in the regression analysis. B. Amplification curves of P. cubensis and P. humuli DNA with different concentrations using the LNA probe CUBprobeSNP105. Each curve was plotted separately using DNA from each pathogen C. Standard curve for the quantification of P. cubensis and P. humuli DNA using the LNA probe HUMprobeSNP105. The log10 of DNA (100fg, 1,000fg, 10,000fg, and 100,000fg) is plotted against the quantification cycle (Cq) values. Each curve was plotted separately using DNA from each pathogen. D. Amplification curves of P. cubensis and P. humuli DNA with different concentrations using the LNA probe HUMprobeSNP105. Each curve was plotted separately using DNA from each pathogen 63 Figure 2-2. Standard curves for the quantification of Pseudoperonospora cubensis and P. humuli sporangia using qPCR. A. The log10 of the number of sporangia is plotted against the quantification cycle values (Cq). The centerline represents the line of fit and error bars represent standard error of the mean. Each curve was plotted separately using the LNA probes specific to each pathogen. Data points represent three technical replicates from two DNA extractions. B. Standard curves based on the qPCR assays of DNA extractions from P. cubensis sporangia (20, 50, 100 and 300) in the presence and absence of the adhesive mix used on the melinex tape. All data points are from three technical replicates from 4 independent DNA extractions. C. Linear regression of Pseudoperonospora spp. sporangia counted using light microscopy against corresponding mean Cq values. All the samples used in this regression were collected using spore traps in the field. 64 Figure 2-3. Monitoring of Pseudoperonospora cubensis and P. humuli sporangia using Burkard spore traps in Ingham (A) and Muskegon (B) in 2018. The data from each county was divided into two panels. The first top panel represents the daily sporangia numbers estimated through the analysis of Burkard spore trap samples using light microscopy (blue bars). The y-axis was trimmed to 40 sporangia to facilitate the visualization of low counts. The middle panel represents the qPCR results for the detection of P. cubensis (red bars) and P. humuli (green bars) in the tape of Burkard spore traps. Yellow triangles denote the monitoring starting date. Red triangles denote the first confirmed report of cucurbit downy mildew in the state. The dashed line denotes the date of cucurbit downy mildew detection in the field. Scouting efforts to detect CDM symptoms in growing cucumber regions are intensified once sporangial loads exceed 10 sporangia/day. 65 Figure 2-4. Monitoring of Pseudoperonospora cubensis and P. humuli sporangia using Burkard spore traps in Berrien (A), Ingham (B) and Muskegon (C) in 2019. The data from each county was divided into two panels. The first top panel represents the daily sporangia numbers estimated through the analysis of Burkard spore trap samples using light microscopy (blue bars). The y-axis was trimmed to 40 sporangia to 66 Figure 2-4. (cont’d) facilitate the visualization of low counts. The middle panel represents the qPCR results for the detection of P. cubensis (red bars) and P. humuli (green bars) in the tape of Burkard spore traps. Yellow triangles denote the monitoring starting date. Red triangles denote the first confirmed report of cucurbit downy mildew in the state. The dashed line denotes the date of cucurbit downy mildew detection in the field. Scouting efforts to detect CDM symptoms in growing cucumber regions are intensified once sporangial loads exceed 10 sporangia/day. 67 A M A F U F R 20000 10000 0 30 Cycles B 20000 M A F U F R 10000 0 30 Cycles C 20000 M A F U F R 10000 0 35 30 Cycles D 20000 10000 M A F U F R 0 35 30 Cycles E 20000 10000 M A F U F R 0 35 30 Cycles A M A F U F R 20000 10000 0 B 20000 10000 M A F U F R 0 C M A F U F R D M A F U F R E M A F U F R 20000 10000 0 20000 10000 0 20000 10000 0 25 25 25 25 25 Sample Phu_0fg__Pcu_10000fg Phu_0fg__Pcu_1000fg Phu_0fg__Pcu_100fg X E H U F R Phu_0fg__Pcu_10000fg Phu_0fg__Pcu_1000fg Phu_0fg__Pcu_100fg 1000 750 Sample 500 250 35 25 40 30 Cycles 35 40 0 1000 25 750 Sample Sample Phu_10000fg__Pcu_0fg Phu_1000fg__Pcu_0fg Phu_100fg__Pcu_0fg X E H U F R 500 Phu_10000fg__Pcu_0fg Phu_1000fg__Pcu_0fg Phu_100fg__Pcu_0fg 250 35 25 40 30 Cycles 35 40 0 1000 25 Sample Phu_10000fg__Pcu_10000fg Phu_1000fg__Pcu_1000fg Phu_100fg__Pcu_100fg 750 Sample X E H U F R 500 Phu_10000fg__Pcu_10000fg Phu_1000fg__Pcu_1000fg Phu_100fg__Pcu_100fg 250 25 40 30 Cycles 35 40 0 1000 25 Sample Phu_1000fg__Pcu_10000fg Phu_100fg__Pcu_10000fg 750 Sample 500 Phu_1000fg__Pcu_10000fg Phu_100fg__Pcu_10000fg 250 X E H U F R 25 40 30 Cycles 35 40 0 1000 25 Sample Phu_10000fg__Pcu_1000fg Phu_10000fg__Pcu_100fg 750 Sample 500 Phu_10000fg__Pcu_1000fg Phu_10000fg__Pcu_100fg 250 X E H U F R 0 1000 750 500 250 0 X E H U F R 30 Cycles 1000 X E H U F R 750 500 250 0 30 Cycles 1000 X E H U F R 750 500 250 0 30 Cycles 1000 X E H U F R 750 500 250 0 30 Cycles 1000 X E H U F R 750 500 250 0 Sample Sample P. cub Sample P. hum 0 fg 0 fg 0 fg 10,000 fg Phu_0fg__Pcu_10000fg Phu_0fg__Pcu_1000fg 1,000 fg Phu_0fg__Pcu_100fg 100fg Phu_0fg__Pcu_10000fg Phu_0fg__Pcu_1000fg Phu_0fg__Pcu_100fg 35 25 40 30 Cycles 35 40 Sample Sample Phu_10000fg__Pcu_0fg Phu_1000fg__Pcu_0fg Phu_100fg__Pcu_0fg Sample P. hum 10,000 fg 1,000 fg 100 fg Phu_10000fg__Pcu_0fg Phu_1000fg__Pcu_0fg Phu_100fg__Pcu_0fg P. cub 0 fg 0 fg 0fg 35 25 40 30 Cycles 35 40 Sample Sample Phu_10000fg__Pcu_10000fg Phu_1000fg__Pcu_1000fg Phu_100fg__Pcu_100fg P. cub P. hum Sample 10,000 fg 1,000 fg 100 fg 10,000 fg Phu_10000fg__Pcu_10000fg Phu_1000fg__Pcu_1000fg 1,000 fg Phu_100fg__Pcu_100fg 100fg 35 25 40 30 Cycles 35 40 Sample Sample P. cub P. hum Sample 1,000 fg 100 fg 10,000 fg Phu_1000fg__Pcu_10000fg 1,0000 fg Phu_100fg__Pcu_10000fg Phu_1000fg__Pcu_10000fg Phu_100fg__Pcu_10000fg 35 25 40 30 Cycles 35 40 Sample Sample P. hum Sample 10,000 fg 10,000 fg P. cub 1,000 fg Phu_10000fg__Pcu_1000fg 100 fg Phu_10000fg__Pcu_100fg Phu_10000fg__Pcu_1000fg Phu_10000fg__Pcu_100fg 25 40 30 Cycles 35 40 25 30 Cycles 35 25 40 30 Cycles 35 40 Supplementary Figure S1. Amplification curves of Pseudoperonospora cubensis (Pcu) and P. humuli Supplementary Figure 2-S1. Amplification curves of Pseudoperonospora cubensis (Phu) DNA using qPCR. A. Amplification curves of P. cubensis DNA with different concentrations using the LNA probes HUMprobeSNP105 (left, FAM) and CUBprobeSNP105 (right. HEX). B. Amplification (Pcu) and P. humuli (Phu) DNA using qPCR. A. Amplification curves of P. cubensis curves of P. humuli DNA with different concentrations using the LNA probes HUMprobeSNP105 (left, DNA with different concentrations using the LNA probes HUMprobeSNP105 (left, FAM) FAM) and CUBprobeSNP105 (right, HEX). C. Amplification curves of P. humuli and P. cubensis DNA mix and CUBprobeSNP105 (right. HEX). B. Amplification curves of P. humuli DNA with in a 1:1 ratio using the LNA probes HUMprobeSNP105 (left, FAM) and CUBprobeSNP105 (right, HEX). different concentrations using the LNA probes HUMprobeSNP105 (left, FAM) and D. Amplification curves of P. humuli and P. cubensis DNA mix in a 1:10 and 1:100 ratio using the LNA probes HUMprobeSNP105 (left, FAM) and CUBprobeSNP105 (right, HEX). E. Amplification curves of P. CUBprobeSNP105 (right, HEX). cubensis and P. humuli DNA mix in a 1:10 and 1:100 ratio using the LNA probes HUMprobeSNP105 (left, FAM) and CUBprobeSNP105 (right, HEX) 68 Supplementary Figure 2-S1. (cont’d) C. Amplification curves of P. humuli and P. cubensis DNA mix in a 1:1 ratio using the LNA probes HUMprobeSNP105 (left, FAM) and CUBprobeSNP105 (right, HEX). D. Amplification curves of P. humuli and P. cubensis DNA mix in a 1:10 and 1:100 ratio using the LNA probes HUMprobeSNP105 (left, FAM) and CUBprobeSNP105 (right, HEX). E. Amplification curves of P. cubensis and P. humuli DNA mix in a 1:10 and 1:100 ratio using the LNA probes HUMprobeSNP105 (left, FAM) and CUBprobeSNP105 (right, HEX) 69 A 25000 M A F U F R 20000 15000 10000 5000 0 C 3000 M A F U F R 2000 1000 0 20 25 30 Cycles 35 40 Species P_cubensis P_humuli fg 10000fg 1000fg 100fg Species P_cubensis P_humuli fg 10000fg 1000fg 100fg B 600 X E H U F R 400 200 0 D 200 X E H U F R 100 0 20 25 30 Cycles 35 40 Species P_cubensis P_humuli fg 10000fg 1000fg 100fg Species P_cubensis P_humuli fg 10000fg 1000fg 100fg 20 25 35 40 30 Cycles 20 25 35 40 30 Cycles Supplementary Figure S2. Amplification curves of Pseudoperonospora cubensis and P. humuli DNA using qPCR. A. Amplification curves of P. cubensis and P. humuli DNA with different concentrations using the LNA probe HUMprobeSNP105 and the Prime-Time Gene Expression Master Mix (IDT, Skokie, IL). B. Supplementary Figure 2-S2. Amplification curves of Pseudoperonospora cubensis and Amplification curves of P. cubensis and P. humuli DNA with different concentrations using the LNA probe P. humuli DNA using qPCR. A. Amplification curves of P. cubensis and P. humuli DNA CUBprobeSNP105 and the Prime-Time Gene Expression Master Mix. C. Amplification curves of P. with different concentrations using the LNA probe HUMprobeSNP105 and the Prime- cubensis and P. humuli DNA with different concentrations using the LNA probe HUMprobeSNP105 and Time Gene Expression Master Mix (IDT, Skokie, IL). B. Amplification curves of P. the IQ Supermix (Bio-rad, Hercules, CA). D. Amplification curves of P. cubensis and P. humuli DNA with different concentrations using the LNA probe CUBprobeSNP105 and the IQ Supermix. cubensis and P. humuli DNA with different concentrations using the LNA probe CUBprobeSNP105 and the Prime-Time Gene Expression Master Mix. C. Amplification curves of P. cubensis and P. humuli DNA with different concentrations using the LNA probe HUMprobeSNP105 and the IQ Supermix (Bio-rad, Hercules, CA). D. Amplification curves of P. cubensis and P. humuli DNA with different concentrations using the LNA probe CUBprobeSNP105 and the IQ Supermix. 70 LITERATURE CITED 71 LITERATURE CITED Adams, M. L., and Quesada-Ocampo, L. M. (2014) Evaluation of fungicides for control of downy mildew on cucumber. Kinston 2013. Plant Dis Manag Rep 8: V240 Blum, M., Waldner, M., Olaya, G., Cohen, Y., Gisi, U., and Sierotzki, H. (2011) Resistance mechanism to carboxylic acid amide fungicides in the cucurbit downy mildew pathogen Pseudoperonospora cubensis. Pest Manag Sci 67: 1211–1214 Brzozowski, L., Holdsworth, W., and Mazourek, M. (2016) ‘DMR-NY401’: A new downy mildew–resistant slicing cucumber. HortScience 51: 1294–1296 Bustin, S. A., Benes, V., Garson, J.A., Hellemans, J., Huggett, J., Kubista, M., Mueller, R., Nolan, T., Pfaffl, M.W., Shipley, G.L., Vandesompele, J., and Wittwer, C. T (2009) The MIQE guidelines: minimum information for publication of quantitative real-time PCR experiments. Clin Chem 55: 611–622 Carisse, O., Tremblay, D.M., Lévesque, C.A., Gindro, K., Ward, P., and Houde, A. (2009) Development of a Taqman real-time PCR assay for quantification of airborne conidia of Botrytis squamosa and management of Botrytis leaf blight of onion. Phytopathology 99: 1273–1280 Cohen, Y., Van den Langenberg, K.M., Wehner, T.C, Ojiambo, P.S., Hausbeck, M, Quesada-Ocampo, L.M., Lebeda, A., Sierotzki, H., and Gisi, U. (2015) Resurgence of Pseudoperonospora cubensis: the causal agent of cucurbit downy mildew. Phytopathology 105: 998–1012 Cohen, Y., and Rotem, J. (1971) Dispersal and viability of sporangia of Pseudoperonospora cubensis. Trans Br Mycol Soc 57: 67–74 Coley-Smith, J.R. (1962) Overwintering of hop downy mildew Pseudoperonospora humuli (Miy. and Tak.) Wilson. Ann Appl Biol 50: 235–243 Dhar, N., Mamo, B.E., Subbarao, K., Koike, S., Fox, A., Anchieta, A., and Klosterman, S.J. (2019) Measurements of aerial spore load by qPCR facilitates lettuce downy mildew risk advisement. Plant Dis 104: 82–93 Dung, J.K.S, Scott, J.C., Cheng, Q., Alderman, S.C., Kaur, N., Walenta, D.L., Frost, K.E., and Hamm, P.B. (2018) Detection and quantification of airborne Claviceps purpurea sensu lato ascospores from Hirst-type spore traps using real-time quantitative PCR. Plant Dis 102: 2487–2493 Ernest, E.G., Kee, W.E., and Everts, K.L. (2005) Evaluation of fungicides for control of downy mildew on pickling cucumber, Trial 2, 2005. Plant Dis Manag Rep 2: V066 72 Gent, D.H., Ocamb, C.M. (2009) Predicting infection risk of hop by Pseudoperonospora humuli. PhytopathologyTM 99: 1190–1198 Gent, D.H., Nelson, M.E., Farnsworth, J.L. and Grove, G.G. (2009), PCR detection of Pseudoperonospora humuli in air samples from hop yards. Plant Pathology, 58: 1081- 1091. Gevens, A.J., and Hausbeck, M.K. (2005) Control of downy mildew of cucumbers with fungicides, 2005. Plant Dis Manag Rep 2: V066 Granke, L.L., and Hausbeck, M.K. (2011) Dynamics of Pseudoperonospora cubensis sporangia in commercial cucurbit fields in Michigan. Plant Dis 95: 1392–1400 Granke, L.L., Morrice, J.J., and Hausbeck, M.K. (2013) Relationships between airborne Pseudoperonospora cubensis sporangia, environmental conditions, and cucumber downy mildew severity. Plant Dis 98: 674–681 Gugino, B.K., and Grove, T.L. (2016) Evaluation of selected fungicides and fungicide programs for managing downy mildew of cucumber, 2015. Plant Dis. Manag. Rep. 10: V072: Hausbeck, M. K., and Linderman, S.D. (2014) Evaluation of fungicides for control of downy mildew of cucumber, 2013. Plant Dis Manag Rep 8: V304 Hausbeck, M., Brisco-McCann, E., Goldenhar, K., and Cook, A. (2017) Evaluation of fungicide programs for control of downy mildew of cucumber, 2016. Plant Dis Manag Reports 11: V063 Hausbeck, M.K., and Cortright, B. (2010) Evaluation of fungicides and application methods for control of downy mildew of cucumber, 2010. Plant Dis. Manag. Rep. V6: Hausbeck, M.K., Perla, D.E., and Cook, A.J. (2019) Evaluation of single fungicide products for control of downy mildew of cucumber, 2018. Plant Dis Manag Reports 139: V13 Hausbeck, M.K. (2020) First cucumber downy mildew outbreak of 2020 confirmed in Berrien County. MSU Ext., https://www.canr.msu.edu/news/cucumber-downy-mildew- outbreak-of-2020 Hirst, J.M. (1952) An automatic volumetric spore trap. Annals of Applied Biology, 39, 259–265. Holmes, G.J., Ojiambo, P.S., Hausbeck, M.K., Quesada-Ocampo, L., and Keinath, A.P. (2014) Resurgence of cucurbit downy mildew in the united states: a watershed event for research and extension. Plant Dis 99: 428–441 73 Hop growers of America (2019) 2018 Statistical report - hops. HGA Stat. Pack. - Revis., https://www.usahops.org/img/blog_pdf/168.pdf Kanetis, L., Holmes, G.J, and Ojiambo, P.S. (2010) Survival of Pseudoperonospora cubensis sporangia exposed to solar radiation. Plant Pathol 59: 313–323 Keinath, A. P (2015) Utility of a cucumber plant bioassay to assess fungicide efficacy against Pseudoperonospora cubensis. Plant Dis 100: 490–499 Klosterman, S.J., Anchieta, A., McRoberts, N., Koike, S.T., Subbarao, K. V., Voglmayr, H., Choi, Y-J., Thines, M., and Martin, F.N. (2014) Coupling spore traps and quantitative PCR assays for detection of the downy mildew pathogens of spinach (Peronospora effusa) and beet (P. schachtii). Phytopathology 104: 1349–1359 Kunjeti, S.G., Anchieta, A., Martin, F.N., Choi, Y-J., Thines, M., Michelmore, R.W., Koike, S.T., Tsuchida, C., Mahaffee, W., Subbarao, K. V., Klosterman, S. J. (2016) Detection and quantification of Bremia lactucae by spore trapping and quantitative PCR. Phytopathology 106: 1426–1437 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 Langston, D., Sanders, F.H. (2013) Evaluation of fungicides for control of downy mildew on cucumber in Georgia I, 2012. Plant Dis. Manag. Rep. 7: V108. Lizotte, E., and Miles, T.D. (2020) Hop disease management in Michigan for 2020. MSU Ext., https://www.canr.msu.edu/news/managing-hop-downy-mildew-in-michigan Michigan Department of Agriculture & Rural Development (2018) Michigan Agriculture Facts & Figures. https://www.michigan.gov/documents/mdard/MI_Ag_Facts__Figures_474011_7.pdf Mitchell, M.N., Ocamb, C.M., Grünwald, N.J., Mancino, L.E., and Gent, D.H. (2011) Genetic and pathogenic relatedness of Pseudoperonospora cubensis and P. humuli. Phytopathology 101: 805–818 Morinha, F., Magalhães, P. and Blanco, G. (2020), Different qPCR master mixes influence telomere primer binding within and between bird species. J Avian Biol, 51: doi:10.1111/jav.02352 Naegele, R. P., Quesada-Ocampo, L. M., Kurjan, J. D., Saude, C., and Hausbeck, M. K. (2016) Regional and temporal population structure of Pseudoperonospora cubensis in Michigan and Ontario. Phytopathology 106: 372–379 NOAA (2019) National Climate Report - May to August 2019. https://www.ncdc.noaa.gov/sotc/national/201905 74 Ojiambo, P.S., Gent, D.H., Quesada-Ocampo, L.M., Hausbeck, M.K., and Holmes, G.J. (2015) Epidemiology and Population Biology of Pseudoperonospora cubensis: A Model System for Management of Downy Mildews. Annu Rev Phytopathol 53: 223–246 Ojiambo, P.S., and Holmes, G.J. (2010) Spatiotemporal spread of cucurbit downy mildew in the Eastern United States. Phytopathology 101: 451–461 Perla, D., Hausbeck, M.K., Cook, A. (2019) Evaluation of fungicide programs for control of downy mildew of cucumber, 2018. Plant Dis Manag Reports 147: V13 Quesada-Ocampo L.M., Granke, L.L., Olsen, J., Gutting, H.C., Runge, F., Thines, M., Lebeda, A., and Hausbeck, M.K. (2012) The genetic structure of Pseudoperonospora cubensis populations. Plant Dis 96: 1459–1470 Rahman, A., Standish, J., D’Arcangelo, K.N., Quesada-Ocampo, L.M. (2020) Clade- specific biosurveillance of Pseudoperonospora cubensis using spore traps for precision disease management of cucurbit downy mildew. Phytopathology. doi: 10.1094/PHYTO- 06-20-0231-R Reuveni, M., Eyal, H., Cohen, Y. (1980) Development of resistance to metalaxyl in Pseudoperonospora cubensis. Plant Dis 64: 1108–1109 Rogers, S.L., Atkins, S.D., and West, J.S. (2009) Detection and quantification of airborne inoculum of Sclerotinia sclerotiorum using quantitative PCR. Plant Pathol 58: 324–331 Royle, D.J. (1973) Quantitative relationships between infection by the hop downy mildew pathogen, Pseudoperonospora hutnuli, and weather and inoculum factors. Ann Appl Biol 73: 19–30 Runge, F., and Thines, M. (2011) Reevaluation of host specificity of the closely related species Pseudoperonospora humuli and P. cubensis. Plant Dis 96: 55–61 Salcedo, A., Hausbeck, M., Pigg, S., Quesada-Ocampo, L.M. (2020) Diagnostic Guide for Cucurbit Downy Mildew. Plant Heal Prog 21: 166–172 Summers, C., Adair, N., Gent, D., Mcgrath, M., and Smart, C. (2015) Pseudoperonospora cubensis and P. humuli detection using species-specific probes and high definition melt curve analysis. Can J Plant Pathol 37: 150605144651006 Thomas, A., Carbone, I., Choe, K., Quesada-Ocampo, L.M., and Ojiambo, P.S. (2017) Resurgence of cucurbit downy mildew in the United States: Insights from comparative genomic analysis of Pseudoperonospora cubensis. Ecol Evol 7: 6231–6246 Thomas, A., Neufeld, K.N., Seebold, K.W., Braun, C.A., Schwarz, M.R., and Ojiambo, P.S. (2018) Resistance to Fluopicolide and Propamocarb and baseline sensitivity to 75 Ethaboxam among isolates of Pseudoperonospora cubensis from the Eastern United States. Plant Dis 102: 1619–1626 Thornton, A.C, Adams, M.L., and Holmes, G.J. (2006) Evaluation of fungicides for control of downy mildew of cucumber I, 2006. Plant Dis Manag Rep 1: V122 USDA (2020) United States Department of Agriculture. 2019. Vegetables Annual Summary. https://usda-library-cornell- edu.proxy1.cl.msu.edu/concern/publications/02870v86p?locale=en Withers, S., Gongora-Castillo, E., Gent, D., Thomas, A., Ojiambo, P.S., and Quesada- Ocampo, L.M. (2016) Using Next-Generation sequencing to develop molecular diagnostics for Pseudoperonospora cubensis, the cucurbit downy mildew pathogen. Phytopathology 106: 1105–1116 Zhu, S.S., Liu, X.L., Wang, Y., Wu, X.H., Liu, P.F., Li, J.Q., Yuan, S.K., and Si, N.G. (2007) Resistance of Pseudoperonospora cubensis to flumorph on cucumber in plastic houses. Plant Pathol 56: 967–975 76 CHAPTER III: OPTIMIZING SPORE TRAPS AND QUANTITATIVE PCR ASSAYS FOR THE MONITORING OF CUCURBIT DOWNY MILDEW. 77 ABSTRACT Current management of Pseudoperonospora cubensis, the causal agent of cucurbit downy mildew (CDM), relies on an intensive fungicide program. In Michigan, CDM occurs annually due to an influx of airborne sporangia; timely alerts of airborne inoculum can assist growers in assessing the need to initiate fungicide sprays. The main objective of our research was to improve the detection and quantification of airborne concentrations of P. cubensis sporangia by adapting two qPCR-based assays to distinguish between P. cubensis clade I and II and P. humuli in spore trap samples. We also aim to evaluate the efficiency of Burkard and impaction spore traps for the detection airborne concentrations of P. cubensis sporangia. A new qPCR assay improved the specificity of P. cubensis detection and resulted in a better linear correlation between the number of sporangia observed using light microscopy and Cq values obtained from Burkard spore traps (R2=0.6; p = 0.01). After two years of monitoring, P. cubensis clade II and P. humuli DNA were detected in air samples collected in commercial cucumber fields, while P. cubensis clade I DNA was not detected. P. cubensis clade II DNA was detected in spore trap samples >2 days before CDM symptoms were first observed in cucumber fields (August), while P. humuli DNA was only detected early in the growing season (May and June). P. cubensis clade I DNA was not detected in air samples before or after the disease onset in cucumber fields. Additionally, the probability for P. cubensis detection in Burkard spore trap samples was higher compared to impaction spore trap samples with approximately the same number of sporangia, suggesting that the efficiency of recovery of sporangia by Burkard spore traps exceeds the recovery of impaction spore traps. Our study identified 78 an improved methodology to monitor the airborne concentrations of Pseudoperonospora spp. sporangia using spore traps coupled with qPCR. This methodology could be used as part of a CDM risk advisory system to time fungicide applications that protect cucurbit crops in Michigan. INTRODUCTION Cucurbit downy mildew (CDM), caused by the obligate oomycete Pseudoperonospora cubensis, incites foliar blighting of several Cucurbitaceae species worldwide (Mitchell et al., 2011). Symptoms include irregular to angular chlorotic leaf lesions and pathogen sporulation on the lower leaf surface (Cohen et al., 2015) leading to premature defoliation, stunted plants and reduced yield, especially in cucumber (Reuveni et al., 1980; Adams et al., 2019; Hausbeck et al., 2019; Perla et al., 2019). In the U.S., P. cubensis overwinters on living hosts in regions that do not experience a frost or in heated greenhouses (Ojiambo et al., 2011; Naegele et al., 2016). The pathogen’s airborne sporangia disperse to northern U.S. growing regions annually from overwintering sources (Ojiambo & Holmes, 2010; Ojiambo et al., 2015). Michigan is the number one producer of pickling cucumbers and squash in the U.S. (USDA, 2020), but since 2005 CDM has threatened cucumber production annually in the state. In the absence of CDM resistant cucumber cultivars, growers rely on intensive fungicide spray programs to limit disease at a significant cost (Savory et al., 2011), however, P. cubensis has developed resistance to key fungicides (Zhu et al., 2007; Blum et al., 2011; Holmes et al., 2014; Keinath, 2015). For more than 10 years, information on the sporangial concentrations of P. cubensis derived from Burkard spore 79 traps has been used in Michigan as a decision-making tool to initiate fungicide sprays to control CDM (Granke and Hausbeck, 2011; Granke et al., 2013). Burkard and impaction spore traps are the most widely used devices for atmospheric sampling in plant pathology (Frenz, 1999). The Burkard spore trap has been used in aerobiological studies for over 60 years (West and Kimber, 2015) and operates by drawing air into a collection chamber containing a reel mounted onto a clockwork mechanism. Spores and other air-borne particles are impacted onto a greased tape covering the reel (Burkard Manufacturing Co. Ltd., U.K.) that rotates past the intake orifice at 2 mm/hr for 7 days. The impaction spore trap, also known as a rotating-arm spore sampler or rotorod spore trap, has become popular for the early detection of airborne plant pathogens (Jackson and Bayliss, 2011; Klosterman et al., 2014; Fall et al., 2015; Choudhury et al., 2016a; Kunjeti et al., 2016). This device includes rods coated with adhesive material which spin at a standard rate in a rotating arm, impacting and collecting airborne particles (TSE Systems, Chesterfield, MO). However, frequent monitoring is required to obtain accurate estimations of inoculum per unit of time (i.e. an hour or day). Although Burkard spore traps are robust and highly autonomous, impaction spore traps can be more cost-effective and grower friendly (i.e. easier to use) (Jackson and Bayliss, 2011; Choudhury et al., 2016a). The efficiency of impaction spore traps to monitor P. cubensis sporangia in comparison to Burkard spore traps has not yet been assessed under field conditions. While the spore traps provide quantitative data on airborne sporangial concentrations, the processing and microscopic examination of the tapes and/or rods is time consuming and can result in misidentification due to the morphological similarities between species (Dung et al., 2018). A qPCR assay was designed for molecular 80 differentiation of the morphologically identical sporangia of P. cubensis and Pseudoperonospora humuli, the causal agent of hop downy mildew (HDM) (Summers et al., 2015a). The detection of each species is based on the recognition of a conserved single nucleotide polymorphism (SNP) in the cytochrome oxidase subunit II (cox2) gene. However, quantification of sporangia using this assay was compromised in spore trap samples containing DNA from both species due to the high similarities between them in the region targeted (Bello et al., 2020). To further improve the specificity and facilitate the quantification of P. cubensis and P. humuli sporangia, a new qPCR assay targeting unique sequences in the pathogens’ mitochondrial genome was developed that also enables detection and differentiation of both species in a single reaction (Crandall, 2020). This assay can also differentiate between P. cubensis clades I and II. Isolates belonging to these two clades are host-adapted at the cucurbit species level (Summers et al., 2015b; Thomas et al., 2017a; Wallace et al., 2020). Clade I isolates are recovered more frequently from commercial varieties of Cucurbita pepo, C. moschata, C. maxima, and Citrullus lanatus and clade II isolates are associated more frequently with commercial varieties within the Cucumis genus (Wallace et al., 2020). Accurate identification of airborne sporangia of P. humuli and P. cubensis (clades I and II) is critical to monitor the pathogen’s arrival into Michigan’s growing regions so as to inform fungicide applications. This new multiplex qPCR assay has the potential to improve quantification of airborne sporangia in spore trap samples without sacrificing accuracy. The main objective of our research was to improve the detection and quantification of airborne concentrations of Pseudoperonospora spp. sporangia by adapting two qPCR-based assays to distinguish between P. cubensis and P. humuli in 81 spore trap samples collected in the field. The specific detection and monitoring of sporangia from each clade of P. cubensis (clade I and II) in commercial cucumber fields was of particular interest. Additionally, we also aimed to evaluate Burkard and impaction spore traps for their efficiency in detecting airborne concentrations of P. cubensis sporangia. MATERIALS AND METHODS Pseudoperonospora isolates. To perform in vitro evaluations, a single-lesion isolate of P. cubensis clade II (CDM23 cucumber) and clade I (CDM-YUM squash), and a single-spike isolate of P. humuli (isolate HDM19) were maintained as described by Bello et al. (2020). Sporangia from each isolate were rinsed from the host tissue into falcon tube with distilled water and the sporangial suspensions were concentrated by centrifugation (5424R centrifuge, Eppendorf, New York, NY) at 14.000 rpm for 5 min. The resulting pellet was homogenized in impact-resistant 2mL tubes (Lysing Matrix H, MP Biomedicals, Irvine, CA) using a TissueLyser II (Qiagen, Valencia, CA) for 4 min at 30 Hz. DNA was extracted using a NucleoSpin Plant II isolation kit (Macherey-Nagel, Bethlehem, PA) following manufacturer’s instructions and the DNA concentration was determined using the Qubit double-stranded DNA High Sensitivity Assay Kit (Life Technologies, Carlsbad, CA). Multiplexing of qPCR assays. All qPCR experiments were conducted using the protocols described by Bello et al. (2020) and Crandall et al. (2020) referred to hereafter as protocol A and B, respectively. All qPCR reactions were manually assembled into 96- well white plates (Bio-rad MLL9651) containing 10 µl of the Prime-Time Gene Expression Master Mix (IDT, Skokie, IL), 2µl of template DNA, and 8 µl of a solution 82 containing primers, probes and internal controls as described in Table 3-1. Negative control reactions lacking a DNA template were included in each plate run. The qPCR protocols were run on a CFX 96 Touch qPCR system (Bio-rad); the cycling conditions are summarized in Table 3-1. Two technical replicates of each sample were run and the average Cq and standard deviation were calculated using Bio-rad CFX Manager software (version 3.1). Specificity and sensitivity of qPCR assays. The sensitivity and specificity of the qPCR protocols were tested using ten-fold dilutions of genomic DNA from two independent DNA extractions of each isolate (CDM23, CDM-YUM and HDM19). Three technical replicates of each sample dilution were tested using both qPCR protocols (Table 3-1) and the average Cq values with standard deviation were calculated using Bio-rad CFX Manager software. Mean Cq values were plotted against the log10 of template DNA concentrations and used to generate standard curves. To assess the specificity of the qPCR protocol B, samples of mixed DNA from the three isolates were evaluated to determine whether the assay could detect P. cubensis (clades 1 and 2) and P. humuli. Ten-fold dilutions of genomic DNA from each isolate were mixed in varying concentrations (Table 3-2) and subjected to qPCR. Three technical replicates of each mixture were run and the average Cq and standard deviations were calculated using Bio-rad CFX Manager software. To evaluate the relationship between sporangial concentrations and the Cq values from the multiplexed qPCR assays in vitro, dilution series containing 1, 3, 5, 10, 25 and 50 sporangia were prepared as described by Crandall et al. (2020) and regressed against the corresponding Cq values of each assay. There were fifteen replicates each for the sporangial counts of 1, 3, and 5 and eight replicates each for 83 sporangial counts of 10, 25, and 50. DNA was extracted using a NucleoSpin Plant II isolation kit (Macherey-Nagel, Bethlehem, PA) following manufacturer’s instructions. Collection and qPCR evaluation of field samples. Airborne sporangial concentrations were monitored during the growing season (May to August) in 2018 and 2019 using Burkard and impaction spore traps. Each year, a Burkard spore trap and an impaction spore trap were placed side by side approximately 20 m from commercial cucumber fields in the Michigan counties of Muskegon, Allegan, Bay, and Saginaw (Fig. 3-1). In 2019, a cucumber research plot (0.25 ha) at the Michigan State University (MSU) Plant Pathology Farm located in Ingham County was also monitored using a Burkard spore trap and impaction spore trap. An abandoned hop research yard (0.25 ha) with basal shoots infected by P. humuli was located 200 m from the cucumber research plot. A Burkard spore trap, only, was placed in a commercial hop yard in Berrien County in 2019 (Fig. 3-1). Burkard spore traps were set to an approximate airflow rate of 10 l/min. The reel of each spore trap was covered with a melinex tape coated with an adhesive mixture of petroleum jelly and paraffin (9:1 wt/wt) dissolved in sufficient toluene to provide the desired thickness. The tape was removed weekly and cut longitudinally along the center line in two subsection of 9 x 336 mm each (Rogers et al., 2009). The two resulting sub- section were processed as described by Bello et al. (2020). Briefly, the first section was cut at 24-hr segments (48-mm lengths) and subjected to DNA extraction. The second section was also cut at 24-h segments and screened using light microscopy to estimate the number of sporangia captured per day (Granke et al., 2013). The impaction traps were constructed using a motor (RF-500TB-10750, Mabuchi) that spun at 2700 rpm when powered by a 12-V marine battery. Each 84 impaction trap was mounted onto a 1.5-m tower that holds the collection rods 1 m above the ground. The impaction traps operated continuously throughout each growing season (May to August), and the greased-coated rods (1.2-by-3.5-mm stainless steel) were collected four times per week at intervals of 24, 24, 24, 96 h (Table 3-3) (Choudhury et al., 2016a). Samples from Burkard and impaction traps were carefully placed into impact-resistant 2mL tubes (Lysing Matrix H, MP Biomedicals) containing 100 µl of PL1 buffer (Macherey-Nagel, Bethlehem, PA). The samples were homogenized using a TissueLyser II (Qiagen, Valencia, CA) for 4 min at 30 Hz and DNA was extracted using NucleoSpin Plant II isolation kit (Macherey-Nagel, Bethlehem, PA) following manufacturer’s instructions. Subsequently, 2 µl of the extraction products were evaluated using qPCR. All 2019 samples were assessed using both A and B protocols, however, all the samples collected during 2018 were evaluated using only protocol A. Correlating Cq values with sporangial counts of field samples. Linear regression using R (version 3.6.1) was used to assess the relationship between the Cq values of each qPCR assay and the number of sporangia in trap samples. The Cq values of each qPCR assay were regressed against the number of sporangia quantified using light microscopy in the second half of the tape of the Burkard spore traps. The linear equation obtained from the regression analysis (Table 3-4) was used to determine the corresponding number of sporangia expressed as NCq where N represents the number of sporangia calculated using Cq values. Additionally, binary logistic regression was used to model the relationship between the number of airborne sporangia and the probability of a positive detection in samples collected by Burkard or impaction spore traps. The log10 number of sporangia 85 quantified in the first half of the Burkard spore traps were regressed with the results from protocol A obtained from trap samples as a categorical (binary) variable. The Cq values of protocol A were categorized using a threshold Cq value of 35.5 (Cq>35.5 negative and Cq≤35.5 positive). The regression was performed independently using the qPCR results from each spore trap separately (i.e. Burkard or impaction). We used this approach to compare between traps and estimate the probability of a positive detection of P. cubensis DNA given an approximate number of sporangia in the air. For example, the probability of a positive detection in impaction samples given an estimated number of 10 sporangia in air samples. Assuming that p is the probability of a positive detection by the qPCR, the logistic regression line is described by Equation 1. Equation 1: Binary logistic regression used to model the relationship between the number of airborne sporangia and the probability of a positive detection in spore trap samples. "#$!"% &'− &* = b"+ b!∗"#$!"(/!) & is the probability of a positive detection. b0 to b1 are parameters. RESULTS Sensitivity and specificity of qPCR assays. Using ten-fold dilutions of genomic DNA, the probes of both protocols detected P. humuli and each clade of P. cubensis within total DNA template amounts between 100 to 1000 fg (R2³ 0.99; p < 0.05) (Fig. 3- 2 and 3-3). The assays are intended to be used to quantify sporangia in environmental samples, thus, further testing outside of these template amounts was not done. Both protocols detected 100 fg of each taxon with average Cq values below 37.5 as reported 86 previously (Bello et al., 2020; Crandall et al., 2020; Summers et al. 2015). Most samples with concentrations below 100 fg were not detected reliably (>95% of the times tested), thus, 100 fg of template DNA was considered as the lower limit of detection (LOD) of the qPCR assays for both species (Figs. 3-2 and 3-3). In singleplex reactions with DNA recovered from isolated sporangia of each Pseudoperonospora taxa, unspecific amplification was observed using protocol A (Fig. 3-2 A). On the contrary, singleplex reactions of protocol B using DNA recovered from isolated sporangia allowed the specific detection of each taxon and background amplification of nontarget taxa was not observed (Fig. 3-3 A, B, C). Multiplex reactions containing all the primers and probes of each qPCR protocol yielded the same results. Small, yet significant differences in the Cq values of mixed and unmixed DNA samples assessed using protocol A were previously reported (Bello et al., 2020). Similar results were also observed using protocol B (Table 3-2). The presence of DNA from a second non-target taxa slightly affected the sensitivity of the probes Pcub2 and Pcub_RFLP_qP1 designed for the detection of P. cubensis clades II and I, respectively (Table 3-1). When 100 fg of P. cubensis clade II DNA were mixed in a 1:1 or 1:100 ratio with DNA from a second taxon, no amplification or a significant increase in the Cq value was observed with the Pcub2 probe. Similarly, when 100 fg of P. cubensis clade I DNA were mixed in a 1:100 ratio with DNA from a second taxon, no amplification was observed with the probe Pcub_RFLP_qP1 (Table 3-2). The regression between sporangial concentrations and the Cq values of both assays (A and B) demonstrated reliable pathogen detection with as few as three sporangia for each taxon (Fig. 3-11 and 3-12). Amplification results of DNA from a 87 single sporangium were inconsistent and outside of the linear relationship between sporangial counts and Cq values. Correlation between Cq values and sporangial counts of Burkard spore trap samples. DNA of P. cubensis clade II and P. humuli was detected using both qPCR protocols on field samples containing fewer than 10 sporangia (Fig. 3-4). Regression analysis indicated that the number of sporangia on the second half of the tape of field samples (quantified using light microscopy) explained 37% and 60% of the variation in the Cq values of P. cubensis clade II DNA obtained using protocols A and B, respectively (Fig. 3-4 A). Significant differences between the protocols were detected using regression analysis (p<2.24e-05, Table 3-4) with a more inclined regression line observed for protocol B (Fig. 3-4 A). Similarly, the number of sporangia on the second half of the tape explained 27% and 30% of the variation in the Cq values of P. humuli DNA obtained using protocols A and B, respectively (Fig. 3-4 B). No significant differences between the protocols were detected (P=0.26, Table 3-4) using regression analysis. b estimates of equations 2 and 3 that describe the relationship between the Cq values of each qPCR assay and the sporangial numbers of each species are summarized in Table 3-4. Equation 2: Regression line that describes the relationship between Cq values and sporangial numbers of Pseudoperonospora cubensis clade II. 12#.%&' )= b"− b! "#$!"(3Lj$75)+ b) "#$!"(2819)b* "#$!"(3Lj$75)(2819 ) The values of the b0, b1, b2, b3 parameters are summarized in Table 3-4. The qPCR assay is express as a binary variable: qPCR assay A = 1, qPCR assay B = 0. 88 Equation 3: Regression line that describes the relationship between Cq values and sporangial numbers of Pseudoperonospora humuli. 12#.+&= b"− b! "#$!"(3Lj$75)). The values of the b0 to b1 parameters are summarized in Table 3-4. Logistic regression. After categorizing the qPCR results of both spore traps (positive/negative) a higher number of impaction samples were negative compared to the Burkard samples (Fig. 3-5 C, D). This occurred most frequently when the atmospheric concentration of sporangia was below 100 sporangia/day as estimated using microscopic analysis of the Burkard tape. Using a logistic regression to model the relationship between the number of atmospheric sporangia and the probability of detection, higher estimates were obtained for the Burkard traps (Table 3-5). Therefore, a higher probability of P. cubensis detection given any number of sporangia was estimated for Burkard trap samples (Figs. 3-5 A, B). A probability of detection above 90% was obtained for Burkard and impaction trap samples with an approximate number of sporangia equal to 15 and 120 sporangia, respectively. Assessment of field samples using light microscopy and qPCR. During 2018 and 2019, an average of 15 reels of Burkard spore traps were collected from May to August in all the locations monitored (Table 3-3, Figs. 3-5 to 3-10). The tape from each reel was divided by days for microscopic and qPCR analysis generating a total of 105 to 112 samples of 24 h per location each year (Table 3-3, Figs. 3-5 to 3-10). Similarly, during both years a total of 544 impaction trap samples were collected from May to August among all the locations monitored. An average of 45 and 15 impaction 89 trap samples per location were collected every 24 and 96 h, respectively (Table 3-3, Figs. 3-5 to 3-10). The qPCR assays detected Pseudoperonospora spp. DNA on 41.84% of the 24 h-samples collected using Burkard spore traps (Table 3-3). Similarly, detection of Pseudoperonospora spp. DNA occurred in 23.35 and 29.81% of the impaction trap samples collected every 24 and 96h, respectively (Table 3-3). In 2018, Pseudoperonospora spp. sporangia were first observed in May or June across the monitored cucumber fields (Figs. 3-6 and 3-7). During May, June and July, the number of sporangia observed in the tape of Burkard spore traps remained below 10 sporangia/day and P. cubensis clade II DNA was detected with Cq values above 31 (Fig. 3-6 and 3-7). This Cq value corresponds to <3Cq sporangia according to the regression line that describes the relationship between Cq values and sporangial numbers (Equation 2 and Table 3-4). Using light microscopy, 10 to 40 sporangia/day were detected in June in Bay Co. (22, 27 June) and Allegan Co. (6, 7, 24 June) and were identified as P. humuli using qPCR with Cq values between 31 and 27 (corresponding to 3Cq to 15Cq sporangia) (Figs. 3-6 B and 3-7 A). In all locations, the number of airborne Pseudoperonospora spp. sporangia observed in the tape of the Burkard tape increased during the last week of July and reached a maximum in the second or third week of August. During this time, P. cubensis clade II DNA was regularly detected in all fields with Cq values that range between 31 and 22 (corresponding to 3Cq and 1200 Cq sporangia) but P. humuli was not detected (Figs. 3-6 and 3-7). In 2018, CDM symptoms were detected in August for each monitored cucumber field, after 12 to 40 sporangia/day were observed in the Burkard spore trap tape. Correspondingly, before symptoms were observed in the fields, P. cubensis clade 90 II detection using qPCR occurred one to seven days earlier with Cq values below 31 (equivalent to >3Cq sporangia) (Figs. 3-6 and 3-7). Similar results were also obtained after the evaluation of impaction trap samples using qPCR in 2018. However, when the number of sporangia per day was low (<20 sporangia/day), pathogen detection in impaction trap samples using qPCR was less consistent. For instance, 1 to 20 P. humuli sporangia were observed in the first half of the Burkard spore trap tape in June (Bay and Allegan Counties) and detected in the second half using qPCR, but P. humuli detection did not occur in the impaction trap samples of the same dates (Figs. 3-6B and 3-7A). Similarly, before CDM symptoms were observed, P. cubensis DNA was consistently detected in Burkard samples in all the fields monitored, but it was only detected in impaction trap samples in Muskegon and Allegan Counties. (Figs. 3-6 A and 3-7 A). Using the Burkard spore trap, P. cubensis clade II DNA was detected with Cq values below 31 approximately 10 days before CDM symptoms were observed in the field (Figs. 3-6 and 3-7). On the other hand, P. cubensis clade II DNA was detected in impaction trap samples approximately 7 days before symptoms were observed in the field only in Muskegon and Allegan Counties. (Figs. 3-6 and 3-7, Table 3-6). In 2019, the number of sporangia observed in Burkard samples from May to July were generally below 10 sporangia/day, but exceptions occurred in May and/or June in Muskegon, Allegan, Saginaw, Ingham and Berrien Counties where daily counts reached values between 10 to 40 sporangia (Figs. 3-8, 3-9 and 3-10). From May to July, P. cubensis clade II DNA was occasionally detected with Cq values above 36 (corresponding to <3Cq sporangia) in all fields monitored. When sporangial numbers were above 10 sporangia/day, the Cq values for P. humuli detection reached values 91 between 31 and 25. These Cq values correspond to >3Cq and 50 Cq sporangia according to the regression line that describes the relationship between Cq values and sporangial numbers (Equation 2 and Table 3-4). Most Cq values for P. humuli detection below 31 were detected only during June, when the highest numbers of sporangia (>40) were observed using light microscopy. In the commercial cucumber fields monitored, the concentration of Pseudoperonospora spp. sporangia estimated using light microcopy increased during August but did not reach the numbers observed in 2018 (Figs. 3-8 and 3-9). At these locations, CDM symptoms were detected during the third or fourth week of August. More regular detection of P. cubensis clade II DNA in Burkard trap samples with Cq values between 31 and 20 (corresponding to 3Cq to 400Cq sporangia) occurred after the observation of CDM symptoms. P. cubensis detection using qPCR before CDM symptoms were observed in the cucumber fields occurred one two seven days earlier with Cq values between 36 and 30 (corresponding >3Cq to 10Cq sporangia) (Figs. 3-8 and 3-9). On the other hand, P. humuli DNA was detected almost every day from May to August in the commercial hop yard (Fig. 3-10 B) and from May to July in the cucumber research plot (Fig. 3-10 A). At these two locations, concentrations between 10 and 40 sporangia were observed from the Burkard trap tapes during May and June and Cq values between 31 and 23 (corresponding to >3Cq to 100Cq sporangia) were registered (Fig. 3-10). The monitoring of Pseudoperonospora spp. using impaction trap samples in the commercial cucumber fields had similar results to those of the Burkard spore traps in 2019 (Fig. 3-8, 3-9 and 3-10). However, when atmospheric concentrations of P. cubensis sporangia were below 10 sporangia/day (estimated with the aid of a light 92 microscope), DNA detection occurred less frequently in samples collected by impaction spore traps compared to the Burkard spore traps. Using qPCR, P. cubensis was detected in samples from both traps approximately two weeks before symptoms were observed in the cucumber fields monitored in Muskegon, Allegan and Saginaw counties. Only in Bay county, P. cubensis was not detected two weeks before symptoms developed in impaction trap samples (Table 3-6). Generally, P. cubensis detection in 2019 occurred with higher Cq values compared to 2018. DISCUSSION Early detection and quantification of airborne P. cubensis sporangia could improve the timing of fungicide initiation in Michigan as the pathogen is reintroduced to northern U.S. production regions each year (Bello et al., 2020). The Burkard spore traps have been used in the state for this purpose since 2008, however, the inability to distinguish between the morphologically identical sporangia of P. humuli and P. cubensis has been a significant shortcoming. Using an improved qPCR assay (Crandall et al., 2020), we were able to distinguish between three host-adapted Pseudoperonospora taxa in spore trap samples: P. humuli, P. cubensis clade I, and P. cubensis clade II. During two years of monitoring using Burkard and impaction spore traps coupled with qPCR in commercial cucumber fields, P. cubensis clade II sporangia were detected 2 to 10 days before CDM symptoms were observed. Both spore traps recorded similar trends in the airborne concentration of P. humuli and P. cubensis (clade I and II) sporangia, however, in our study, the Burkard spore trap was a more efficient instrument for detecting P. cubensis sporangia. 93 The multiplex qPCR assay developed by Crandall et al. (2020) that targets the open reading frames (orf 374, orf 367, orf 329) in the mitochondrial genome of each taxon provided high specificity for detecting DNA of P. humuli and each clade of P. cubensis (Crandall et al., 2020). This qPCR assay allowed us to estimate the concentrations of P. cubensis and P. humuli sporangia in Burkard and impaction spore trap field samples. The new multiplex qPCR assay had the same sensitivity of the qPCR assay developed by Summers et al. (2015), which was also developed to target a mitochondrial DNA region. Both assays detected P. cubensis and P. humuli DNA at amounts ranging from 100-1000 fg in vitro and were equally sensitive to a qPCR assay targeting a nuclear DNA region (Rahman et al., 2020). This high sensitivity was validated with the detection of less than 10 sporangia of each species in field spore trap samples. In vitro, samples containing three sporangia of each clade were reliably detected (Crandall et al., 2020) using these qPCR assays. Comparatively, the qPCR assay that targets the single-copy nuclear gene c255.3e7 showed a detection limit of 10 sporangia (Rahman et al., 2020). Additionally, the detection of each taxon without cross reactivity in samples containing mixed DNA of P. humuli and either clade of P. cubensis validated the high specificity of the assay developed by Crandall et al. (2020). This improved specificity was possible because the primers and probes of this new qPCR assay target more polymorphic regions in the mitochondrial genome compared to the single nucleotide polymorphism used for differentiation in the assay developed by Summer et al. (2015). Improving the specificity also resulted in an increased linear correlation between the number of P. cubensis sporangia quantified with the aid of a light microscope and the Cq values of the qPCR (R2=0.6). We used the equation that describes this 94 relationship to calculate the number of sporangia based on Cq values; however, the standard error of the b estimates used in this equation suggests a low precision for the quantification of sporangia using qPCR in spore trap samples. Multiple factors such as high variation during DNA extraction (Summers et al., 2015a), user variation during qPCR, and the multicopy nature of the target mitochondrial genes (González- Domínguez et al., 2020) could explain these results. The number of mitochondria can vary greatly among cells (O’Hara et al., 2019) reducing the precision for the quantification of cells using qPCR assays that target mitochondrial genes, however, the additional number of mitochondria per cell can increase the ability to detect low concentrations of sporangia. The more precise the assay (i.e. regression model) the closer predictions are to the observed number of sporangia. In this study, the use of Cq values to predict the number of sporangia resulted in misestimates compared to the quantification using light microscopy. Single-copy nuclear genes are thought to offer more precision for cell quantification compared to mitochondrial genes. However, the correlation of Cq values obtained from the amplification of the single-copy nuclear gene c255.3e7 with sporangial numbers revealed a standard deviation across more than three amplification cycles for samples with the same number of sporangia (Rahman et al., 2020). Similar results were obtained using the mitochondrial genes orf 374, orf 367and orf 329 (Crandall et al., 2020) suggesting that the precision for cell quantification using qPCR is not significantly increased with the utilization of nuclear genes as amplification targets. During the two years of monitoring in commercial cucumber fields in Michigan, P. humuli detection occurred early in the growing season. This was expected because hop downy mildew is prevalent in the state (Higgins et al., 2020; Lizotte et al., 2020) and the 95 pathogen overwinters in dormant hop crowns, growing into expanding basal shoots in the spring (Coley-Smith, 1962). Similarly, detection of P. cubensis clade II occurred as expected because airborne sporangia are dispersed every year from overwintering sources resulting in CDM outbreaks in cucumber production regions of the upper Midwest (Ojiambo et al., 2011; Naegele et al., 2016). However, P. cubensis clade I was not detected in the 980 Burkard and 544 impaction spore trap samples that were collected over the two years of sampling from May to August. The absence of P. cubensis clade I sporangia in the air samples collected at the monitored cucumber fields may be due to the reduced number of crops planted in the state that are known to be hosts of this clade (i.e. C. pepo, C. moschata, C. maxima, and C. lanatus) (Wallace et al., 2020). Approximately 6000 ha of C. maxima (pumpkin), C. pepo (squash). C. moschata (butternut squash) and C. lanatus (watermelon) are planted in Michigan, compared to the more than 15000 ha of C. sativus (cucumber) planted in the state every year (USDA, 2020). In North Carolina, the hectares of cucumber planted annually are also greater compared than those planted with other cucurbits; clade I was consistently detected using spore traps in the fall (September and October), while clade II was detected in the summer and fall (Rahman et al., 2020). In our study, we report on the airborne concentration of P. cubensis sporangia through August and most of the cucumber acreage is harvested by this time. In Michigan, the foliage of pumpkins and hard squash during September and October begins to senesce and is compromised by powdery mildew; CDM is rarely reported on these crops in the state. Determining if sporangia from both clades are dispersed to the Great Lakes cucurbit growing regions during September to October remains to be seen by future studies. P. cubensis clade I was not 96 detected in North Carolina and Michigan during the summer probably due to the total area planted to hosts susceptible to this clade. In the absence of airborne P. cubensis clade I sporangia, an intensive fungicide program for CDM may be unnecessary in non- cucumber hosts. Timely regional information on the atmospheric concentrations of each clade of P. cubensis could inform control measures to minimize the negative impact of CDM across different cucurbits. This information could be used as a decision-making tool to initiate fungicide sprays to protect susceptible crops, as it is used in other crop production systems (Carisse et al., 2009; Fall et al., 2015; Dhar et al., 2019; Van der Heyden et al., 2020). In lettuce, a threshold Cq value of 24, equivalent to 324 Bremia lactucae sporangia/day, is used to determine whether fungicides should be applied. This approach reduced the number of fungicide applications to control downy mildew in small lettuce plots without a significant increase in disease incidence (Dhar et al., 2019). Similarly, in our study, CDM symptoms were observed after airborne sporangial concentrations exceeded 10- 15 sporangia/day (estimated using light microscopy) or Cq values between 33 to 30 (qPCR assay B) were detected in Burkard spore trap samples. This suggests that the critical concentration to trigger fungicide sprays against CDM could be close to these numbers. However, further research is required to better understand the interaction between sporangial concentrations, environmental conditions, and symptom development (Fall et al., 2015), which could ultimately lead to the establishment of spore concentration thresholds to trigger fungicide application. Accurate sampling of low inoculum loads and real-time monitoring is critical to develop a biosurveillance system that accurately assesses the risk of CDM in cucurbits. Burkard and impaction spore traps are the most widely used devices for atmospheric 97 sampling in plant pathology (Frenz, 1999) and have played a significant role in epidemiological studies in horticultural and agricultural settings (Granke et al., 2013; Choudhury et al., 2016a; Carisse et al., 2017; Wyka et al., 2017). They have also been used to accelerate the detection of airborne plant pathogens (Jackson and Bayliss, 2011) prior to symptom development (Villari et al., 2016; Thiessen et al., 2017; Dung et al., 2018; Dhar et al., 2019). However, Burkard and impaction spore traps had not been used side by side to monitor Pseudoperonospora spp. sporangia. After two years of monitoring, our results suggest that the Burkard spore traps are a more efficient device detecting airborne sporangia at low concentrations (<100 sporangia/day). This is consistent with theoretical expectations that impaction spore traps are likely to offer lower particle recoveries than Burkard spore traps (Frenz, 1999). We found that the probability of P. cubensis detection is above 90% for Burkard and impaction samples that contain approximately 15 and 120 sporangia, respectively. These results indicate that Burkard spore traps can collect at least eight times (120/15) more sporangia than impaction spore traps at the conditions tested. The difference in the spore recovery between these two spore traps is expected to increase inversely proportional to the size of the particles being collected. According to Aylor (1993), Burkard spore traps can recover three times more particles than impaction spore traps when the particle size is approximately the same as that of P. cubensis sporangia (40 µm). However, differences of up to seven times in particle recovery between these two collection devices have also been reported (Solomon et al., 1980). Wind direction and wind velocity can also affect the particle collection efficiency of both spore traps (West and Kimber, 2015), but the more important factors in the performance of impaction spore traps are sampling surface width and angular velocity 98 (Solomon et al., 1980). We used rods with similar width surface (1.2-by-3.5-mm stainless steel) to rods utilized in previous studies (Klosterman et al., 2014; Rahman et al., 2020) but increasing the collection surface of the impaction rod samplers could result in higher particle recovery and should be considered for future monitoring studies. Other factors to consider for increasing the recovery of P. cubensis sporangia by impaction spore traps include using multiple traps per location and longer deployment times of spore traps before rod collection. In California, at least two impaction spore traps are utilized to monitor B. lactucae spore loads in lettuce fields 50 to 200 times smaller than the commercial fields monitored in this study (Dhar et al., 2019); doubling the sample surface and the amount of air sampled could increase the chances for impaction of airborne sporangia in daily samples. We also observed a higher proportion of positive samples among the impaction spore trap samples collected over a longer period of time (96 h) which indicates that the adhesive medium we used on the rods (High-Vacuum Grease) was not completely saturated. It may not be necessary to change the impaction rods every 24 to 72 hours. Considering the expense of Burkard spore traps and the relatively low cost of impaction spore traps and their ease of use, it is important to improve the efficiency of sporangia detection using impaction spore traps. The combination of Burkard and impaction spore traps with the qPCR assay developed by Crandall et al., 2020 facilitated the sensitive and specific monitoring of P. humuli and two host-adapted clades P. cubensis in Michigan. Using spore traps and qPCR, we detected P. cubensis clade II sporangia three to seven days before disease onset in commercial cucumber fields. During two years of monitoring, P. cubensis clade I was never detected during the summer season (May to August) in the fields 99 monitored. Our data suggest that Burkard spore traps are more efficient than impaction traps for the detection of airborne Pseudoperonospora spp. sporangia at low concentrations (<100 sporangia/day). Impaction spore traps could be modified to increase the probability for the collection of sporangia. In the future, the ability to rapidly detect both clades of P. cubensis using qPCR could be incorporated with environmental data and disease development information as part of a CDM risk advisory system to time fungicide applications that protect cucurbit crops in Michigan. 100 APPENDIX 101 APPENDIX Table 3-1. qPCR assays designed for the differentiation of Pseudoperonospora humuli and Pseudoperonospora cubensis clade I and II. Assay Primers/probes Final concentr ationa 60 µm RT33F RT182R 60 µm CUBprobeSNP105 2.5 µm HUMprobeSNP105 5 µm InCp_J2 5 µm Internal control (IC) 7.5 x10-10 nM PC_RFLP_2F PC_RFLP_3R PH_RFLP_4R PC-4 F PC-4 R Pcub_RFLP_qP1 Pcub2 20 µm 10 µm 10 µm 10 µm 10 µm 5 µm 10 µm Phum_RFLP_qP4 5 µm Sequence 5'->3' Protocolb AACTCCCGTTATGGAAGGTATT CCATGTACAACAGTAGCTGGA HEX/A+C+AAA+C+G+AATA+CT/BHQ FAM/AA+C+AAA+C+A+AATA+CTG/BHQ CYS/A+GCATTATT+GTTTAT+CATATATACA/BHQ AACTCCCGTTATGGAAGGTATTATCATTAATCAT GATTTGTA TATATGATAAACAATAATGCTATAACATAGAGTC TCTTTCAT GAATAATCCAGCTACTGTTGTACATGG. CTGCTTTATCTTTTTCTTTTTG AGAGAAGATTTAGATTATAATTC AGAGACGATTTGGATTATAATT CAAGACCACCATTTTTATGTC TGGAAATTAAAAATTTTCTATTAC FAM/AACAAACTCAAGTAGAACTTCAACAAA/BH Q HEX/AGGATTGATTTTCATTAATTCCTTTTTGTAA TAGAA/BHQ RED/CCAACAGTTATACTTGTAATAAAC ATCAAG/BHQ 95°C for 3 min followed by 40 cycles of 95°C for 10 s and 65°C for 45 s 95⁰C for 3 minutes followed by 45 cycles of 95⁰C for 15 s and 58⁰C for 45 s Ac Bd 102 Table 3-1. (cont’d) a. Final concentrations used in a 20 ul qPCR reaction. b. The amplification protocols run on a CFX 96 Touch qPCR system (Bio-rad). c. This set of primers and probes were adapted from (Summers et al., 2015a). Locked nucleic acids in the probes are followed by a plus (+) sign. d. This set of primers and probes were adapted from (Crandall et al., 2020). 103 Table 3-2. Threshold cycle (Cq) values of the qPCR protocol B using varying concentrations of genomic DNA from Pseudoperonospora cubensis clade 1 and 2 and P. humuli Species/Clade DNA Probe Mix ratio Un- mixed 1/1 Pcub2 Cq±SD Phum_ RFLP_qP4 Cq±SD NA NA NA 28.17±0.47a NA 32.01±1.1b NA 34.11±0.25c NA NA NA NA Pcub_ RFLP_qP1 Cq±SD 27.53±0.35a NA -- 31.69±0.21b NA -- 34.76±0.94c NA -- NA -- NA -- NA -- 10.000 fg NA NA 1.000 fg 100 fg NA 27.75±0.02a 28.15±0.05a NA -- -- 31.91±0.38b 32.82±0.47b NA 34.67±0.41c 38.65±0.15* NA -- 27.21±0.69a 31.11±1.1b 35.64±1.2c P. humuli P. cubensis clade II -- -- -- 10.000 fg 1.000 fg 100 fg -- -- -- P. cubensis clade I 10.000 fg 1.000 fg 100 fg -- -- -- -- -- -- 10.000 fg 10.000 fg 1.000 fg 1.000 fg 100 fg 100 fg 10.000 fg 10.000 fg NA -- -- 1.000 fg NA 100 fg -- NA 10.000 fg -- -- 1.000 fg 100 fg -- 1.000 fg 100 fg 10.000 fg 27.39±0.07a NA 31±0.28* 1.000 fg NA 100 fg 34.73±0.5c NA 28.58±0.15a 26.57±0.31a 33.27±0.12* 31.25±0.05b NA** 34.71±0.7c 27.85±0.22a 31.33±0.64b 35.73±0.34c 104 Table 3-2. (cont’d) Species/Clade DNA Probe Mix ratio 1/10 1/100 Pcub2 P. humuli P. P. cubensis cubensis clade I clade II 10.000 fg 1.000 fg 1.000 fg 10.000 fg 10.000 fg 1.000 fg -- -- 1.000 fg -- 10.000 fg Pcub_ RFLP_qP1 Cq±SD 27.61±0.01a 31.9±0.56b NA NA** 28.21±0.18a NA 28.24±0.52a 29.51±0.13* NA 10.000 fg NA 33.24±1.77b 26.32±0.32a 31.06±1.11b NA 1.000 fg Phum_ RFLP_qP4 Cq±SD Cq±SD -- -- 27.78±0.33 a -- 1.000 fg 10.000 fg 100 fg 100 fg 10.000 fg 100 fg -- -- 10.000 fg -- -- 10.000 fg 100 fg 100 fg 10.000 fg 34.25±0.48* NA -- -- 100 fg 10.000 fg NA** 27.87±0.25a NA 28.11±0.05a NA 27.66±0.06a NA** NA** 35.76±0.26c 27.61±0.17a NA NA 27.92±0.55a 28.08±0.78a 33.04±0.9* NA** 26.65±0.52* NA 10.000 fg NA Cq values with the same letter are not significantly different. Cq values with the symbol * are significantly different from values obtained for un-mixed samples with the same concentration of DNA (t-Test; P=0.05). --: not DNA added. NC: not used for comparison in the t-Test. NA: not defined. 105 Table 3-3. Percentage of qPCR positive samples collected by Burkard and impaction spore traps Location Year Burkard 24 ha Reel Nd Positivee Impaction 96 hc Muskegon Bay Allegan Saginaw Ingham Average 2018 15 2019 16 2018 15 2018 16 2018 15 2019 16 2018 15 2019 16 2018 -- 2019 16 105 40.95% 112 33.04% 105 39.05% 112 43.75% 105 48.57% 112 33.93% 105 24.76% 112 41.96% -- 112 70.54% 15.6 108 41.84% -- 24 hb Nd 45 45 45 45 45 45 48 45 -- 43 45 Positive Nd 33.33% 16 15 8.89% 16 13.33% 20.00% 15 15 26.67% 15 20.00% 12.50% 16 15 28.89% -- -- 15 46.51% 23.35% 15 Positivee 25.00% 40.00% 18.75% 40.00% 13.33% 33.33% 31.25% 20.00% -- 46.67% 29.81% a. Reels were collected every 7 days and the tape was divided every 24 hours. b. Rod samples collected every 24 hours. c. Rod samples collected every 96 hours. d. Total number of samples collected. e. Percentage of samples collected that tested positive for P. cubensis or P. humuli. 106 Table 3-4. Linear regression analysis of Cq values as a function of the number Pseudoperonospora spp. sporangia Taxon P. humuli P. cubensis clade II b Variablesa b0 32.33 Intercept b1 -4.77 log10(Xb) qPCR(A/B) -0.77 b2 log10(Xb)*qPCR (A/B) c b3 0.71 b0 33.69 Intercept b1 log10(Xb) -5.34 qPCR (A/B) -2.30 b2 log10(Xb)*qPCR(A/B) c b3 2.27 Estimate S.E 0.42 0.47 0.55 0.63 0.59 0.49 0.65 0.53 t value 76.13 -10.05 -1.40 1.12 57.03 -10.89 -3.50 4.28 Pr(>|t|) <2e-16 <2e-16 0.16 0.26 <2e-16 <2e-16 5.1e-04 2.2e-05 Pr(>F) -- <2e-16 0.39 0.26 -- <2.20e-16 0.91 <2.2e-05 a. Linear regression equation: !"!.#$%= b&− b' &'('&(*+',-.(/-)+ b( &'('&("2!3)+ b) &'('&(*+',-.(/-)("2!3 ), !"!.*$= b&− b' &'('&(*+',-.(/-)). b. X = number of sporangia. c. qPCR assay express as a binary variable: qPCR assay A = 1, qPCR assay B = 0. 107 Table 3-5. b estimates of logistical models developed to predict the probability of Pseudoperonospora cubensis detection in Burkard and impaction spore trap samples Spore trap b0 Burkard Impaction -1.57 -2.33 SD 0.10 0.21 Pvalue <2e-16 <2e-17 b1 2.16 1.60 SD 0.16 0.21 Pvalue <2e-16 6.e-14 AIC 1333.8 405.4 Logistic regression models were developed using the spore trapping data collected in Allegan, Muskegon, Saginaw, Berrien and Bay counties in 2018 and 2019. Cq values of the qPCR protocol A were categorized as a binary variable using a threshold value of 35.5 (0>35.5 and 1≤35.5). To predict the probability of P. cubensis detection in Burkard and impaction spore traps as a function of the number of sporangia b estimates should be replaced in equation 1. 108 Table 3-6. Detection of Pseudoperonospora cubensis using Burkard spore traps coupled with qPCR and detection of symptoms Year Location qPCR Spore trap CDM Symptoms qPCR detection of P. cubensis pre- Burkard Aug 7 Jul 23, 24, 26, 28, 29, 30, 31 and Aug 1, Jul 27, 28, 29, 30, 31and Aug 1, 2, 3, 4, 5, symptomsa 2, 3, 4, 5, 6 6 Muskegon A 2018 Bay Allegan Saginaw Muskegon Bay 2019 Allegan Saginaw Ingham A A A B B B B B Impaction Burkard Impaction Burkard Impaction Burkard Impaction Burkard Impaction Burkard Impaction Burkard Impaction Burkard Impaction Burkard Impaction Aug 7 Aug 21 Aug 2 Aug 16 Aug 20 Aug 16 Aug 22 Aug 21 Jul 27, 30 and Aug 5, 6 --- Aug 4, 6, 7, 8, 9, 10, 11, 12, 13, 14 Aug 2, 3, 4, 5, 6, 7, 14 Jul 27, 31 --- Aug 5, 6, 7, 9, 11, 12, 13, 14, 15 Aug 8, 9, 10, 11, 12, 13, 14, 15 Aug 12 --- Aug 5, 8, 11, 12, 13, 14, 15 Aug 8, 9, 10, 11, 15 Aug 11, 19, 21 Aug 15, 16, 17, 18, 19 Aug 1, 19, 20 --- qPCR detectionb 16 days 12 days 11 days --- 17 days 19 days 6 days --- 11 days 9 days 8 days --- 11 days 9 days 11 days 9 days 21 days --- a. qPCR detection within a window of three weeks before symptoms were observed in the field b. Number of days between the first detection of P. cubensis DNA using qPCR and the detection of CDM symptoms in the field 109 A B n a hig e Mic k a L Muskegon Co. Allegan Co. L a k e H u r o n n a hig e Mic k a L Bay Co. Berrien Co. L a k e H u r o n Saginaw Co. Ingham Co. Figure 3-1. Location of spore traps by county in 2018 (A) and 2019 (B). Green and blue dots indicate Burkard and impaction spore traps, respectively. 110 Figure 3-2. Standard curves for the quantification of Pseudoperonospora cubensis clade I, P. cubensis clade II and P. humuli DNA using the qPCR protocol A. Standard curves of the protocol A using (A) the probe CUBprobeSNP105 for P. cubensis detection and (B) the probe HUMprobeSNP105 for P. humuli detection. The curves were constructed assessing the DNA from all taxa independently. Detection of P. cubensis clade I and II was observed using the probe HUMprobeSNP105. The log10 of DNA fg is plotted against the quantification cycle values (Cq). All data points are from three technical replicates derived from two DNA extractions. The centerline represents the line of fit and error bars represent standard deviation. 111 Figure 3-3. Standard curves for the quantification of Pseudoperonospora cubensis clade I, P. cubensis clade II and P. humuli DNA using the qPCR protocol B Standard curves of the protocol B using the probes Pcub2 (A), Pcub_RFLP_qP1 (B) and Phum_RFLP_qP4 (C) for detection of P. cubensis clade II, P. cubensis clade I and P. humuli, respectively. The curves were constructed assessing the DNA from all taxa independently. Amplification of non-target taxa was not observed. The log10 of DNA fg is plotted against the quantification cycle values (Cq). All data points are from three technical replicates derived from two DNA extractions. The centerline represents the line of fit and error bars represent standard deviation. 112 A q C 35 30 25 20 0 B Y = 31 – 2.9 log10(X) Y = 33 - 5 log10(X) q C qPCR A (SPcHx) R2=0.37 qPCR B (qPc2Hx) R2=0.60 1 3 Log10 (number of sporangia) 2 35 30 25 20 0.0 Y = 32 – 4.29 log10(X) Y = 32 – 4.1 log10(X) qPCR A (qPhRd ) R2=0.27 qPCR B (SPhFa ) R2=0.30 0.5 2.0 Log10 (number of sporangia) 1.0 1.5 2.5 Figure 3-4. Linear regression of Cq values as a function of sporangial numbers. (A) Linear regression of Pseudoperonospora spp. sporangia counted using light microscopy against corresponding mean Cq values of the protocol A (rhomboids) and B (triangles). (B) Linear regression of Pseudoperonospora spp. sporangia counted using light microscopy against corresponding mean Cq values of the protocol A (circles) and B (triangles). The log10 number of sporangia is plotted against the quantification cycle values (Cq). 113 A y t i l i b a b o r P C i ) a g n a r o p s f o r e b m u n ( 0 1 g o L 4 3 2 1 0 1.0 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 0 1 Negative/Positive D 4 3 t s e T 2 1 0 C i ) a g n a r o p s f o r e b m u n ( 0 1 g o L 4 3 2 1 0 0 1 0 1 Negative/Positive Negative/Positive y t i l i b a b o r P B D 4 3 t s e T 2 1 0 0 1 0 1 Negative/Positive Negative/Positive 0 1 Negative/Positive C D i ) a g n a r o p s f o t s r e e T b m u n ( 0 1 g o L 4 4 3 3 2 2 1 1 0 0 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 0.0 2.0 Log10 (number of sporangia) 1.5 0.5 1.0 0.0 2.5 Log10 (number of sporangia) 2.0 1.5 0.5 1.0 Figure 3-5. Logistic regression of qPCR results from Burkard (A) and impaction (B) spore traps as a function of sporangial numbers. 114 Figure 3-6. Monitoring of Pseudoperonospora cubensis and P. humuli sporangia using spore traps in Muskegon (A) and Bay (B) counties in 2018 (Location B). 115 Figure 3-6. (cont’d) The data from each county was divided into three rows of panels. The first row panels represent the daily sporangial numbers estimated through the analysis of Burkard spore trap samples using light microscopy (blue bars). The y-axis was trimmed to 40 sporangia to facilitate the visualization of low counts. The middle panels represent the qPCR results of protocol A for the detection of P. cubensis (red bars) and P. humuli (green circles) in the tape of Burkard spore traps. The bottom panels represent the qPCR results of protocol A for the detection of P. cubensis (red bars) and P. humuli (green circles) in the rods of impaction spore traps. A black arrow denotes the monitoring starting date. The dashed line denotes the date of cucurbit downy mildew symptom detection in the field. Scouting efforts to detect CDM symptoms in growing cucumber regions are increased once sporangial loads exceed 10 sporangia/day. Bars below the x-axis denote the time the reels and rods were changed from each trap. 116 Figure 3-7. Monitoring of Pseudoperonospora cubensis and P. humuli sporangia using spore traps in Allegan (A) and Saginaw (B) counties in 2018. 117 Figure 3-7. (cont’d) The data from each county was divided into three rows of panels. The first row panels represent the daily sporangial numbers estimated through the analysis of Burkard spore trap samples using light microscopy (blue bars). The y-axis was trimmed to 40 sporangia to facilitate the visualization of low counts. The middle panels represent the qPCR results of protocol A for the detection of P. cubensis (red bars) and P. humuli (green circles) in the tape of Burkard spore traps. The bottom panels represent the qPCR results of protocol A for the detection of P. cubensis (red bars) and P. humuli (green circles) in the rods of impaction spore traps. A black arrow denotes the monitoring starting date. The dashed line denotes the date of cucurbit downy mildew symptoms detection in the field. Scouting efforts to detect CDM symptoms in growing cucumber regions are increased once sporangial loads exceed 10 sporangia/day. Bars below the x-axis denote the time the reels and rods were changed from each trap. 118 Figure 3-8. Monitoring of Pseudoperonospora cubensis and P. humuli sporangia using spore traps in Muskegon (A) and Bay (B) counties in 2019. 119 Figure 3-8. (cont’d) The data from each county was divided into three rows of panels. The first row panels represent the daily sporangial numbers estimated through the analysis of Burkard spore trap samples using light microscopy (blue bars). The y-axis was trimmed to 40 sporangia to facilitate the visualization of low counts. The middle panels represent the qPCR results of protocol B for the detection of P. cubensis (red bars) and P. humuli (green circles) in the tape of Burkard spore traps. The bottom panels represent the qPCR results of protocol B for the detection of P. cubensis (red bars) and P. humuli (green circles) in the rods of impaction spore traps. A black arrow denotes the monitoring starting date. The dashed line denotes the date of cucurbit downy mildew symptoms detection in the field. Scouting efforts to detect CDM symptoms in growing cucumber regions are increased once sporangial loads exceed 10 sporangia/day. Bars below the x-axis denote the time the reels and rods were changed from each trap. 120 Figure 3-9. Monitoring of Pseudoperonospora cubensis and P. humuli sporangia using spore traps in Allegan (A) and Saginaw (B) counties in 2019. 121 Figure 3-9. (cont’d) The data from each county was divided into three rows of panels. The first row panels represent the daily sporangial numbers estimated through the analysis of Burkard spore trap samples using light microscopy (blue bars). The y-axis was trimmed to 40 sporangia to facilitate the visualization of low counts. The middle panels represent the qPCR results of protocol B for the detection of P. cubensis (red bars) and P. humuli (green circles) in the tape of Burkard spore traps. The bottom panels represent the qPCR results of protocol B for the detection of P. cubensis (red bars) and P. humuli (green circles) in the rods of impaction spore traps. A black arrow denotes the monitoring starting date. The dashed line denotes the date of cucurbit downy mildew symptoms detection in the field. Scouting efforts to detect CDM symptoms in growing cucumber regions are increased once sporangial loads exceed 10 sporangia/day. Bars below the x-axis denote the time the reels and rods were changed from each trap. 122 Figure 3-10. Monitoring of P cubensis and P. humuli sporangia using spore traps in Ingham (A) and Berrien (B) counties in 2019. 123 Figure 3-10. (cont’d) The data from Ingham county was divided into five rows of panels. The first row panels represent the daily sporangial numbers estimated through the analysis of Burkard spore trap samples using light microscopy (blue bars). The y-axis was trimmed to 40 sporangia to facilitate the visualization of low counts. The middle second and third row panels represent the qPCR results of protocol A (row 2) and B (row 3) for the detection of P. cubensis (red bars) and P. humuli (green circles) in the tape of Burkard spore traps. The fourth and five row panels represent the qPCR results of protocol A (row 4) and B (row 5) for the detection of P. cubensis (red bars) and P. humuli (green circles) in the rods of impaction spore traps. The data from Berrien county was divided into three row panels. The first row panels represent the daily sporangial numbers estimated through the analysis of Burkard spore trap samples using light microscopy (blue bars). The y-axis was trimmed to 40 sporangia to facilitate the visualization of low counts. The middle panels represent the qPCR results of protocol A for the detection of P. cubensis (red bars) and P. humuli (green circles) in the tape of Burkard spore traps. The bottom panels represent the qPCR results of protocol B for the detection of P. cubensis (red bars) and P. humuli (green circles) in the tape of Burkard spore traps. A black arrow denotes the monitoring starting date. The dashed line denotes the date of cucurbit downy mildew symptoms detection in the field. Scouting efforts to detect CDM symptoms in growing cucumber regions are increased once sporangial loads exceed 10 sporangia/day. Bars below the x-axis denote the time the reels and rods were changed from each trap. 124 Figure 3-11. Standard curves for the quantification of Pseudoperonospora cubensis and P. humuli sporangia using the qPCR protocol A. Five-point standard curves for protocol A using (A) the probe CUBprobeSNP105 for P. cubensis detection and (B) the probe HUMprobeSNP105 for P. humuli detection. All data points represent an average of 8 to 15 replicate spore-count samples of each species. Error bars on points represent the standard error of the mean. There was inconsistent amplification for the 1-count sporangium samples and they were excluded from the final linear regression. 125 Figure 3-12. Standard curves for the quantification of Pseudoperonospora cubensis clade I, P. cubensis clade II and P. humuli sporangia using the qPCR protocol B. Five-point standard curves for protocol B using the probes Pcub2 (A), Pcub_RFLP_qP1 (B) and Phum_RFLP_qP4 (C) for detection of P. cubensis clade II, P. cubensis clade I and P. humuli, respectively. All data points represent an average of 8 to 15 replicate spore-count samples of each taxa. Error bars on points represent the standard error of the mean. There was inconsistent amplification for the 1-count sporangium samples and they were excluded from the final linear regression. 126 LITERATURE CITED 127 LITERATURE CITED Adams, M., D’Arcangelo, K, Collins, H, Quesada-Ocampo, L (2019) Evaluation of cultivars for control of downy mildew on squash, Kinston 2018. Plant Dis Manag Reports 13: V124 Aylor, D.E. (2017) Chapter 20: Aerial Sampling. Aerial dispersal of pollen and spores. The American Phytopathological Society, pp 339–360 Aylor, D.E. (1993) Relative collection efficiency of Rotorod and Burkard spore samplers for airborne Venturia inaequalis ascospores. Phythopathology 83: 1116–1119 Bello, J.C., Sakalidis, M.L., Perla, D.E., Hausbeck, M.K. (2020) Detection of airborne sporangia of Pseudoperonospora cubensis in Michigan using Burkard spore traps coupled to qPCR. Accepted in Plant Dis. Blum, M., Waldner, M., Olaya, G., Cohen, Y., Gisi, U., Sierotzki, H. (2011) Resistance mechanism to carboxylic acid amide fungicides in the cucurbit downy mildew pathogen Pseudoperonospora cubensis. Pest Manag Sci 67: 1211–1214 Carisse, O., McNealis, V., Kriss, A. (2017) Association between weather variables, airborne inoculum concentration, and raspberry fruit rot caused by Botrytis cinerea. Phytopathology 108: 70–82 Carisse, O., Tremblay, D.M., Lévesque, C.A., Gindro, K., Ward, P., Houde, A. (2009) Development of a Taqman real-time PCR assay for quantification of airborne conidia of Botrytis squamosa and management of Botrytis leaf blight of onion. Phytopathology 99: 1273–1280 Choudhury, R.A., Koike, S.T., Fox, A.D., Anchieta, A., Subbarao, K.V., Klosterman, S.J., McRoberts, N. (2016a) Season-long dynamics of spinach downy mildew determined by spore trapping and disease incidence. Phytopathology 106: 1311–1318 Choudhury, R.A., Koike, S.T., Fox, A.D., Anchieta, A., Subbarao, K.V., Klosterman, S.J., McRoberts, N. (2016b) Spatiotemporal patterns in the airborne dispersal of spinach downy mildew. Phytopathology 107: 50–58 Coley-Smith, J.R. (1962) Overwintering of hop downy mildew Pseudoperonospora humuli (Miy. and Tak.) Wilson. Ann Appl Biol 50: 235–243 Crandall, S. G., Ramon, M. L., Burkhardt, A. K., Bello J. C., Adair, N., Gent, D. H., Hausbeck, M. K., Quesada-Ocampo, L. M. and Martin, F. N. A multiplex TaqMan qPCR assay for detection and quantification of clade 1 and clade 2 isolates of Pseudoperonospora cubensis and P. humuli. Under review in Plant Dis. 128 Dhar, N., Mamo, B.E., Subbarao, K., Koike, S., Fox, A., Anchieta, A., Klosterman, S.J .(2019) Measurements of aerial spore load by qPCR facilitates lettuce downy mildew risk advisement. Plant Dis 104: 82–93 Dung, J.K.S., Scott, J.C., Cheng, Q., Alderman, S.C., Kaur, N., Walenta, D.L., Frost, K.E., Hamm, P.B. (2018) Detection and quantification of airborne Claviceps purpurea sensu lato ascospores from Hirst-type spore traps using real-time quantitative PCR. Plant Dis 102: 2487–2493 Fall. M.L., Van der Heyden, H., Brodeur, L., Leclerc, Y., Moreau, G., Carisse, O. (2015) Spatiotemporal variation in airborne sporangia of Phytophthora infestans: characterization and initiatives towards improving potato late blight risk estimation. Plant Pathol 64: 178–190 Frenz, D.A. (1999) Comparing pollen and spore counts collected with the Rotorod Sampler and Burkard spore trap. Ann Allergy, Asthma Immunol 83: 341–349 González-Domínguez, E., Berlanas, C., Gramaje, D., Armengol, J., Rossi, V., Berbegal, M. (2020) Temporal dispersal patterns of Phaeomoniella chlamydospora, causal agent of petri disease and esca, in vineyards. Phytopathology 110: 1216–1225 Granke, L.L., Hausbeck, M.K. (2011) Dynamics of Pseudoperonospora cubensis sporangia in commercial cucurbit fields in Michigan. Plant Dis 95: 1392–1400 Granke, L.L., Morrice, J.J., Hausbeck, M.K. (2013) Relationships between airborne Pseudoperonospora cubensis sporangia, environmental conditions, and cucumber downy mildew severity. Plant Dis 98: 674–681 Hausbeck, M., Brisco-McCann, E., Goldenhar, K., Cook, A. (2017) Evaluation of fungicide programs for control of downy mildew of cucumber, 2016. Plant Dis Manag Reports 11: V063 Hausbeck, M.K., Perla, D.E., Cook, A.J. (2019) Evaluation of single fungicide products for control of downy mildew of cucumber, 2018. Plant Dis Manag Reports 139: V13 Hollomon, B. (2007) Fungicide resistance in crop pathogens: How can it be managed?, 2nd ed. FRAC, Brussels Holmes, G.J., Ojiambo, P.S., Hausbeck, M.K., Quesada-Ocampo, L., Keinath, A.P. (2014) Resurgence of cucurbit downy mildew in the united states: a watershed event for research and extension. Plant Dis 99: 428–441 Jackson, S.L., Bayliss, K.L. (2011) Spore traps need improvement to fulfil plant biosecurity requirements. Plant Pathol 60: 801–810 129 Keinath, A.P. (2015) Utility of a cucumber plant bioassay to assess fungicide efficacy against Pseudoperonospora cubensis. Plant Dis 100: 490–499 Klosterman, S.J., Anchieta, A., McRoberts, N., Koike, S.T., Subbarao, K.V., Voglmayr, H., Choi, Y-J., Thines, M., Martin, F.N. (2014) Coupling spore traps and quantitative PCR assays for detection of the downy mildew pathogens of spinach (Peronospora effusa) and beet (P. schachtii). Phytopathology 104: 1349–1359 Kunjeti, S.G., Anchieta, A., Martin, F.N., Choi, Y-J., Thines, M., Michelmore, R.W., Koike, S.T., Tsuchida, C., Mahaffee, W., Subbarao, K. V., et al (2016) Detection and quantification of Bremia lactucae by spore trapping and quantitative PCR. Phytopathology 106: 1426–1437 Lizotte, E., and Miles, T.D. (2020) Hop disease management in Michigan for 2020. MSU Ext., https://www.canr.msu.edu/news/managing-hop-downy-mildew-in-michigan Naegele, R.P., Quesada-Ocampo, L.M., Kurjan, J.D., Saude, C., Hausbeck, M.K. (2016) Regional and temporal population structure of Pseudoperonospora cubensis in Michigan and Ontario. Phytopathology 106: 372–379 O’Hara, R., Tedone, E., Ludlow, A., Huang, E., Arosio, B., Mari, D., Shay, J.W. (2019) Quantitative mitochondrial DNA copy number determination using droplet digital PCR with single-cell resolution. Genome Res 29: 1878–1888 Ojiambo, P.S., Gent, D.H., Quesada-Ocampo, L.M., Hausbeck, M.K., Holmes, G.J. (2015) Epidemiology and population biology of Pseudoperonospora cubensis: a model system for management of downy mildews. Annu Rev Phytopathol 53: 223–246 Ojiambo, P.S., Holmes, G.J. (2010) Spatiotemporal spread of cucurbit downy mildew in the Eastern United States. Phytopathology 101: 451–461 Ojiambo, P.S., Holmes, G.J., Britton, W., Babadoost, M., Bost, S.C., Boyles, R., Brooks, M., Damicone, J., Draper, M.A., Egel, D.S., et al (2011) Cucurbit downy mildew IPMpipe: a next generation web-based interactive tool for disease management and extension outreach. Plant Heal Prog 12: 26 Perla, D., Hausbeck, M.K., Cook, A. (2019) Evaluation of fungicide programs for control of downy mildew of cucumber, 2018. Plant Dis Manag Reports 147: V13 Rahman, A., Standish, J., D’Arcangelo, K.N., Quesada-Ocampo, L.M. (2020) Clade- specific biosurveillance of Pseudoperonospora cubensis using spore traps for precision disease management of cucurbit downy mildew. Phytopathology. Reuveni, M., Eyal, H., Cohen, Y. (1980) Development of resistance to metalaxyl in Pseudoperonospora cubensis. Plant Dis 64: 1108–1109 130 Rogers, S.L., Atkins, S.D., West, J.S. (2009) Detection and quantification of airborne inoculum of Sclerotinia sclerotiorum using quantitative PCR. Plant Pathol 58: 324–331 Savory, E.A., Granke, L.L., Quesada-Ocampo, L.M., Varbanova, M., Hausbeck, M.K., Day, B. (2011) The cucurbit downy mildew pathogen Pseudoperonospora cubensis. Mol Plant Pathol 12: 217–26 Solomon, W.R., Burge, H.A., Boise, J.R., Becker, M. (1980) Comparative particle recoveries by the retracting rotorod, rotoslide and Burkard spore trap sampling in a compact array. Int J Biometeorol 24: 107–116 Summers, C., Adair, N., Gent, D., Mcgrath, M., Christine, S. (2015a) Pseudoperonospora cubensis and P. humuli detection using species-specific probes and high definition melt curve analysis. Can J Plant Pathol 37: 150605144651006 Summers, C.F., Gulliford, C.M., Carlson, C.H., Lillis, J.A., Carlson, M.O., Cadle- Davidson, L., Gent, D.H., Smart, C.D. (2015b) Identification of genetic variation between obligate plant pathogens Pseudoperonospora cubensis and P. humuli using RNA sequencing and genotyping-by-sequencing. PLoS One 10: e0143665 Thiessen, L.D., Neill, T.M., Mahaffee, W.F. (2017) Timing fungicide application intervals based on airborne erysiphe necator concentrations. Plant Dis 101: 1246–1252 Thomas, A., Carbone, I., Choe, K., Quesada-Ocampo, L.M., Ojiambo, P.S. (2017) Resurgence of cucurbit downy mildew in the United States: Insights from comparative genomic analysis of Pseudoperonospora cubensis. Ecol Evol 7: 6231–6246 USDA (2020) United States Department of Agriculture. 2019. Vegetables Annual Summary. Van der Heyden, H., Bilodeau, G., Carisse, O., Charron, J-B. (2020) Monitoring of Peronospora destructor primary and secondary inoculum by Real-Time qPCR. Plant Dis. doi: 10.1094/PDIS-03-20-0687-RE Villari, C., Mahaffee, W.F., Mitchell, T.K., Pedley, K.F., Pieck, M.L., Hand, F.P. (2016) Early detection of airborne inoculum of Magnaporthe oryzae in turfgrass fields using a quantitative LAMP assay. Plant Dis 101: 170–177 Wallace, E.C., D’Arcangelo, K.N., Quesada-Ocampo, L.M. (2020) Population analyses reveal two host-adapted clades of Pseudoperonospora cubensis, the causal agent of cucurbit downy mildew, on commercial and wild cucurbits. Phytopathology. doi: 10.1094/PHYTO-01-20-0009-R West, J., Kimber, R. (2015) Innovations in air sampling to detect plant pathogens. Ann Appl Biol 166: 4–17 131 Wyka, S.A., McIntire, C.D., Smith, C., Munck, I.A., Rock, B.N., Asbjornsen, H., Broders, K.D. (2017). Effect of climatic variables on abundance and dispersal of Lecanosticta acicola spores and their impact on defoliation on eastern white pine. Phytopathology 108: 374–383 Zhu, S.S., Liu, X.L., Wang, Y., Wu, X.H., Liu, P.F., Li, J.Q., Yuan, S.K., Si, N.G. (2007) Resistance of Pseudoperonospora cubensis to flumorph on cucumber in plastic houses. Plant Pathol 56: 967–975 132 CHAPTER IV: GENOTYPING OF THE OBLIGATE PLANT PATHOGENS PSEUDOPERONOSPORA CUBENSIS AND P. HUMULI USING TARGET ENRICHMENT SEQUENCING 133 ABSTRACT Technological advances in genome sequencing have improved our ability to catalog genomic variation and have led to an expansion of the scope and scale of genetic studies over the past decade. Yet, for agronomically important plant pathogens such as the downy mildews (Peronosporaceae) the scale of genetic studies remains limited. This is, in part, due to the difficulties associated with maintaining obligate pathogens, and the logistical constraints involved in the genotyping of these species (e.g. obtaining DNA of sufficient quantity and quality). To gain an evolutionary and ecological perspective of downy mildews, adaptable methods for the genotyping of their populations are required. Here, we describe a targeted enrichment (TE) protocol to genotype isolates from two Pseudoperonospora species (P. cubensis and P. humuli) using less than 50 ng of mixed pathogen and plant DNA for library preparation. We were able to enrich 736 target genes across 101 samples and identified 2,978 high- quality SNP variants. Using these SNPs, we detected significant genetic differentiation (AMOVA, p=0.01) between P. cubensis subpopulations from Cucurbita moschata (clade I) and Cucumis sativus (clade II) in Michigan (U.S.). No evidence of location-based differentiation was detected within the P. cubensis (clade II) subpopulation of Michigan. However, a significant effect of location on the genetic variation of the P. humuli subpopulation was detected in the state (AMOVA, p=0.01). Mantel tests found evidence that the genetic distance among P. humuli samples was associated with the physical distance of the hop yards from which the samples were collected (p=0.005). The differences in the distribution of genetic variation of the P. humuli and P. cubensis subpopulations of Michigan suggest differences in the dispersal of these two species. 134 The TE protocol described here provides an additional tool for genotyping obligate biotrophic plant pathogens and the execution of new genetic studies. INTRODUCTION Downy mildew (DM) pathogens (Peronosporaceae) cause foliar disease in several agronomically important plant species (Gent et al., 2009; Kanetis et al, 2013; Kunjeti et al., 2016; Lane et al., 2005; Rivera et al., 2016; Wallace & Quesada-Ocampo, 2017; Wong & Wilcox, 2001). The group is comprised of at least twenty different genera representing a large portion of the described pathogenic species of oomycetes (Thines, 2014; Choi and Thines, 2015). All DM species are considered obligate biotrophs, growing only in association with living host tissue (Bourret et al., 2018). Over the past decade, several studies have provided key insight into DM biology, including virulence mechanisms (Savory et al., 2012; Baxter et al., 2010; Burkhardt & Day, 2016; Purayannur et al., 2020), host specificity (Choi and Thines, 2015; Summers et al., 2015b; Rivera et al., 2016; Wallace et al., 2020), and fungicide resistance (Gisi and Sierotzki, 2008; Blum et al., 2011). Despite this progress, several unresolved research questions in ecology and evolution remain, many of which could be addressed with emerging genomic and genetic approaches. Recent technological advances in high-throughput sequencing have facilitated the genotyping of dozens of samples and the analysis of whole plant-pathogen genomes (Savory et al., 2012b; Sharma et al., 2015; Rivera et al., 2016; Withers et al., 2016; Cui et al., 2019; Fletcher et al., 2019; Rahman et al., 2019). The recent availability of genome-wide sequence information from multiple individuals of the same species has facilitated population genomic studies of several plant pathogens 135 (Grünwald et al., 2017; Tabima et al., 2018; Carleson et al., 2019; Gent et al., 2019). Whole or reduced representation genome sequencing technologies have become increasingly popular to monitor genetic changes in plant-pathogen populations. However, these technologies have not been broadly used to study DM populations due to difficulties in obtaining sufficient amounts of DNA from single obligate biotrophic individuals (Milgroom, 2015). Traditionally, other genotyping technologies that require significantly lower inputs of DNA such as Sanger sequencing and microsatellites (SSRs) are used with DNA extracted from fresh and stored symptomatic tissue to study DM populations (Quesada- Ocampo et al., 2012; Kitner et al., 2015; Naegele et al., 2016; Rivera et al., 2016; Wallace and Quesada-Ocampo, 2017; Wallace et al., 2020). However, these approaches are limited by the relatively low number of variants that can be obtained when compared to genome-wide sequencing options. DNA from symptomatic tissue can be used for sequencing with high-throughput sequencing technologies but the low concentration of pathogen DNA and large amounts of exogenous material (e.g. plant and bacterial DNA) can add further expense and complexity to the sequencing and bioinformatic analysis (Stassen et al., 2012). The quality of high-throughput sequencing data depends on the quality and purity of the DNA sample, which means that the target pathogen is ideally present in higher amounts when compared to the host plant before DNA extraction (Jouet et al., 2019). A spore propagation approach using detached susceptible leaves can be used to propagate bulk amounts of DM sporangia from which additional pathogen DNA can be recovered (Summers et al., 2015b; Thomas et al., 2017a; Gent et al., 2019); however, 136 multiple growing cycles (7-10 days/cycle) are needed making this approach time consuming and labor-intensive (Ali et al., 2011; Gent et al., 2019). Spore propagation is most successful when using freshly collected symptomatic leaves, but propagating sporangia from samples stored over long periods of time or under poor conditions (e.g. samples stored >-80 C) can be complicated. Additionally, sporangial propagation requires continuous maintenance of fresh sporangia on a highly susceptible host, which may bias the genetic composition of the populations under study due to the selection of genotypes by the host and the propagation conditions (Jones et al., 2014; Thomas et al., 2017a). While sporangial propagation can provide large amounts of high-quality DNA needed for sequencing, this approach is impractical for population studies requiring a large number of samples collected over time (i.e. years). Sequence capture methods may provide a solution for genotyping DM pathogens that excludes non-target DNA while also facilitating high coverage sequencing of several target loci (Kozarewa et al., 2015; Lim & Braun, 2016). These techniques use affinity probes (RNA/DNA) to isolate particular sequences of interest (“target regions”) out of a larger pool of DNA fragments (DNA library). Sequence capture methods such as target enrichment (TE) have been used across a variety of genomic studies with model (Gnirke et al., 2009; Clark et al., 2011) and non-model organisms (Faircloth et al., 2015; McCormack et al., 2016; Starrett et al., 2017) and have also facilitated the study of museum specimens with low amounts of poor quality DNA (Cruz-Dávalos et al., 2017). Additionally, sequence capture methods have also been used to study the genetic variation of other plant pathogens such as Phytophthora infestans, Phytophthora (cid:0)apsica (Thilliez et al., 2019), and the obligate biotroph Albugo candida 137 (Jouet et al., 2019). These techniques could facilitate the study of DM pathogen populations using symptomatic tissue samples, alleviating the need for sporangial propagation and expanding the type and condition of the samples used. In this study, we evaluated the genotypic variation of two DM species of the genus Pseudoperonospora using TE sequencing. Pseudoperonospora cubensis and Pseudoperonospora humuli infect cucurbits and hops, respectively, worldwide and are considered the most economically important species of the genus Pseudoperonospora (Choi et al., 2005; Mitchell et al., 2011). P. cubensis (clades I and II) causes foliar blight (DM) of cucumber (Cucumis sativus), melon (Cucumis melo), pumpkin (Cucurbita maxima), watermelon (Citrullus lanatus), and squash (Cucurbita moschata) (Summers et al., 2015b; Thomas et al., 2017a; Wallace et al., 2020). P. humuli negatively impacts hop (Humulus lupulus) cone yield (Gent et al., 2010). We describe an optimized TE procedure for sequencing DM and a bioinformatic pipeline for population genetic analyses using TE data. Evaluation of the structure, diversity, and reproduction of P. cubensis and P. humuli populations in Michigan was of particular interest. MATERIALS AND METHODS Sample collection and DNA extraction. Over a 12 year period (2007-2009, 2012-2013, 2015, and 2018-2019). P. cubensis sporangia were obtained from symptomatic cucurbit tissue collected in Michigan, other U.S. states (Indiana, Iowa, Ohio, Florida, and Wisconsin) and one Canadian province: Ontario (Fig. 4-1 A; Supplementary Table 4-1). Sporangia were harvested by gently rinsing infected leaves exhibiting multiple lesions and pathogen signs using a Preval spray power unit (Preval, Chicago) filled with distilled water as described by Mitchell et al. (2011). Dislodged 138 sporangia were transferred into a centrifuge tube (2 mL), pelleted (14000 rpm for 5 min, 5424 Centrifuge, Eppendorf) and subjected to DNA extraction. A second group of samples was processed by excising single lesions using a sterile scalpel. Sporangia from each lesion were harvested by vigorously shaking the microcentrifuge tube (2 mL) containing the tissue and 1 mL of distilled H2O for 30 s. Dislodged sporangia were pelleted (14000 rpm for 5 min, 5424 Centrifuge, Eppendorf) and subjected to DNA extraction. P. humuli samples were processed using methods similar to those described by Chee et al. (2006). Sporangial suspensions were obtained from diseased basal shoots collected from six commercial hop yards and a hop research plot located in Michigan. The three hop yards in the north region were located within a radius of approximately 25 km (hop yards A, B and C) (Fig. 4-1 B). The two hop yards in the west region of the state were separated by approximately 100 km and operated by the same producer (hop yards D and E) (Fig. 4-1 B). The sixth commercial hop yard was located in the east region (hop yard F) and was located approximately 80 km from the research plot (hop yard G) located in the central region (Fig. 4-1B). Diseased basal shoots were brought to the laboratory, stems placed into beakers of water, and shoots individually covered with a plastic bag overnight to induce sporulation. The following day, sporangia were washed from the abaxial leaf surface using a Preval spray power unit filled with distilled water. Sporangial suspensions from single shoots were transferred into centrifuge tubes (2 ml) and pelleted by centrifugation (14000 rpm for 5 min). All DNA extractions were performed on the pelleted sporangial suspensions using the NucleoSpin Plant II 139 isolation kit (Macherey-Nagel, Bethlehem, PA, U.S.) following manufacturer’s instructions. DNA library construction. TE libraries: Whole DNA was extracted from 275 samples as described above and quantified using a Qubit 2.0 fluorimeter (Thermo Fisher Scientific). Before fragmentation of DNA using a M220 Focused-ultrasonicator (Covaris, Woburn, MA, U.S.) the presence of P. cubensis and/or P. humuli in the samples was confirmed by sanger sequencing of the ITS region as described by Quesada-Ocampo et al. (2012). A subset of five samples were divided in two individual tubes and submitted independently for library preparation and enrichment. These samples were used to estimate the number of errors introduced into a sample during the TE workflow. Sequencing libraries were prepared using the KAPA HyperPrep library sequencing preparation kit (Kapa/Roche, Pleasanton, CA, U.S.) following the manufacturer’s recommendations and an additional KAPA Pure Bead (Kapa/Roche, Pleasanton, CA, U.S.) cleanup to ensure the libraries were free from any adapters. Insert size and library quality was verified using the Agilent 2100 Bioanalyzer (Agilent Technologies, Santa Clara, CA, U.S.). To optimize the use of the Mybaits custom kit’s reactions (80bp baits, Arbor Biosciences, formerly Mycroarray Inc., Ann Arbor, MI, U.S.), good quality libraries were pooled into groups of four libraries for enrichment. Probes (baits) were designed according to manufacturer protocols (probe compatibility, repeat masking, and melting temperature filters) to cover 118 population informative DNA regions and 712 DNA regions annotated as protein genes at a 2X coverage. Protein gene sequences were obtained using bedtools (Quinlan and Hall, 2010) with the coordinates information derived from the genome annotation of P. 140 cubensis available online in the server of the Oregon State University Libraries (dx.doi.org/10.7267/N9TD9V7M; Burkhardt et al., 2014). All the genes annotated as extracellular toxins, hydrolytic enzymes, enzyme inhibitors, cell-entering RxLR effectors, cell-entering crinkler effectors, secreted proteins and transcription factors were targeted for sequencing. These regions were selected to identify annotated genes containing SNPs that may play a role in controlling host-specificity, and selectively neutral regions for population level analyses (Quesada-Ocampo et al., 2012; Kitner et al., 2015; Wallace and Quesada-Ocampo, 2017). Hybridization of DNA libraries and probes were performed at 65 °C for 24 or 48 h (Supplementary Table 4-1). After hybridization, library pools were bound to Dynabeads MyOne Streptavidin C1 magnetic beads (Life Technologies) for enrichment. Amplification of the enriched libraries was performed using the 2X KAPA HiFi HotStart ReadyMix (Kapa/Roche, Pleasanton, CA, U.S.) and 5 μM of each TruSeq forward and reverse primers. Amplification conditions were 98 °C for 2 min, followed by 6-14 cycles of 98 °C for 20 s, 60 °C for 30 s and 72 °C for 60 s and then a final extension of 72 °C for 5 min. Following PCR, libraries were quantified using a Qubit 2.0 fluorimeter (Invitrogen). Enriched libraries were quality controlled and quantified using a combination of Qubit dsDNA HS and Caliper LabChipGX HS DNA assays. Based on these quantifications the libraries were combined in equimolar amounts into two groups for sequencing. The first set of samples (TES1) included 62 enriched libraries with a hybridization time of 24 h. The second set (TES2) of samples included 72 enriched libraries with a hybridization time of 48 h (Supplementary Table 4-1). Both sets were quantified using the Qubit 2.0 and the Kapa Biosystems Illumina Library Quantification 141 qPCR kit. Each set was loaded onto one line of an Illumina HiSeq 4000 flow cell and sequenced in a 2x150bp paired-end format. Low-coverage whole-genome sequencing (Lc-WGS) libraries: In order to assess the performance of TE in comparison to Lc-WGS and to estimate the amount of exogenous material in the DNA samples, ten samples of genomic DNA directly extracted from sporangial suspensions derived from individual lesions were submitted to the Michigan State University Research Technology Support Facility (MSU RTSF) for DNA library preparation and sequencing. Due to the low DNA concentration of some of the samples, the Rubicon ThruPLEX DNA-Seq library preparation kit (Takara Bio, Mountain View, CA, U.S.) was used to prepare the sequencing libraries. Library preparation was performed following manufacturer’s recommendations with an additional AMPureXP bead cleanup (Beckman Coulter, Indianapolis, IN, U.S.). Completed libraries were quality controlled and quantified using a combination of Qubit dsDNA HS (Thermo Fisher Scientific/Life Technologies, Carlsbad, CA, U.S.) and Caliper LabChipGX HS DNA assays (PerkinElmer, Waltham, MA, U.S.). Based on these quantifications the libraries were grouped in equimolar amounts into two pools that were quantified using the Kapa Biosystems Illumina Library Quantification qPCR kit (Kapa/Roche, Pleasanton, CA, U.S.). Each pool was loaded into one lane of an Illumina MiSeq standard flow cell (v2) (Illumina, San Diego, CA, U.S.) and sequencing was performed in a 2x250bp paired end format using a v2 500 cycle MiSeq reagent cartridge (Illumina, San Diego, CA, U.S.). Genotyping and variant calling. In order to assess the performance of TE for the genotyping of P. cubensis and P. humuli, we analyzed the data in combination with 142 previously published whole-genome sequencing (WGS) data generated for a comparative genomic analyses of P. cubensis. This data set included a total of 10 samples collected from different cucurbits in the U.S. states of Alabama, South Carolina (N=2), Florida, North Carolina (N=2), Georgia, California, New York and Oregon. The datasets can be found in the bioprojects PRJNA360426 (Thomas et al., 2017) and PRJNA675756 (this project) available online through the National Center for Biotechnology Information (NCBI). Sequencing raw reads of all libraries were trimmed and quality filtered using Trimmomatic 0.32 (Bolger et al., 2014). The publicly available P. cubensis (Savory et al., 2012b) and C. sativus genome sequences were used as references for alignment of the reads. To estimate plant contamination, high-quality reads were initially aligned to the C. sativus genome using the Burrows-Wheeler aligner (BWA-0.7.17) (Li and Durbin, 2009). All reads were then aligned to a reference genome of P. cubensis containing only the contigs where the target genes are located. Clean reads of the Bioproject PRJNA360426 were also aligned to this subset of the genome. Samtools-0.1.19 (Li et al., 2009) was used to sort the alignments and PCR duplicates were removed with Picard (https://broadinstitute.github.io/picard/). Coverage depth was calculated using BEDTools-2.27.1. The Next Generation Sequencing Plugin (NGSEP) for analysis of high- throughput sequencing data was used for calling SNPs and producing the variant call format file (VCF) (Perea et al., 2016). The VCF file generated was filtered using the NGSEP plugin in order to retain the most informative SNPs of sufficient quality. Only biallelic SNPs, present in 70% of samples, with a minimum quality of 40 were retained. The resulting VCF file was imported into R using the package vcfR (version 1.5.0) 143 (Knaus and Grünwald, 2017). Missing genotype data were deleted before further analysis (Tabima et al., 2018). Samples with more than 40% of missing variants were removed from the analysis to retain the maximum number of SNPs. Variants with more than 75% of missing information across samples were also removed. To increase the stringency for the genotyping, we excluded all the samples with an average coverage below 50X. A nonparametric Kruskal-Wallis rank-sum test was used to assess the differences among the mean coverage depth across libraries. Population Assignment. P. cubensis samples were assigned to several different subpopulations depending on the analyses. To study the distribution of the genetic variation by host, P. cubensis samples were assigned to subpopulations according to its host species. To study the distribution of the genetic variation by geographic location, only the P. cubensis samples collected from Cucumis spp. were used. The samples collected in the U.S. states of Michigan, Wisconsin, Indiana, Ohio, and the Canadian province of Ontario, CA, were assigned to the Midwest subpopulation (Fig, 4-1 C). Midwest samples excluding those from Michigan were assigned to the Midwest-w/o-MI subpopulation (Fig. 4-1 C). Samples from Michigan were further divided into three subpopulations according to the region of collection (east, west and central) (Fig. 4-1 C). Samples collected from Iowa, Alabama, Oregon, California, South Carolina, Florida, New York, North Carolina, and Georgia, were assigned to their respective state of origin. P. humuli samples were assigned to either the north or west subpopulation based on their location of collection in Michigan (Fig. 4-1 B, C; Table 4-2; Supplementary Table 4-1). Three hop yards (A, B and C) were located within a radius of 144 approximately 25 km and comprised the north subpopulation (Fig. 4-1 C). Hop yards D and E in southwest Michigan are operated by one producer and make up the west subpopulation (Fig. 4-1 C). Representative samples from central, east Michigan and Oregon were also included in our analysis. Genetic differentiation and population structure. The filtered SNPs were used to construct a genetic tree based on Nei’s genetic distance (Nei, 1972) and visualize the genetic relationship among all samples, 1000 bootstrap replicates were performed to obtain branch support (Kamvar et al., 2015). Clone-correction was performed by collapsing samples to the average genetic distance detected between technical replicates of sequencing. For this, the “mlg.filter” function of the R package Poppr (version 2.4.1) was used with a bitwise distance equal to 0.02 and the farthest neighbor algorithm. Ordination plots based on principal component analysis (PCA) were also constructed to visualize the differentiation of samples within and among subpopulations from different host or regions across the Midwest (Fig. 4-1 C). Three distinct PCAs were constructed to visualize the relationship among i) all P. cubensis and P. humuli samples by host ii) all P. cubensis samples collected from C. sativus and iii) all P. humuli samples collected in Michigan and Oregon. The PCAs were performed using the “glPca” function of R package Adegenet (Jombart and Ahmed, 2011) and the ordination plots were constructed using the R package Ggplot2. Samples were colored according to host or subpopulations and ellipses of 90% confidence intervals were drawn on the ordination plots. Genetic differentiation between and within subpopulations was quantified using hierarchical F-statistics (hierfstat version 0.04.-22, Goudet, 2005) and analysis of 145 molecular variance (AMOVA) (poppr version 2.4.1, Kamvar et al., 2014) in R version 3.6.1. To test the hypothesis that the Pseudoperonospora spp. populations are differentiated by host, we performed an AMOVA and calculated FST statistics using all the samples collected by P. cubensis and P. humuli grouped by host. Similarly, to test the hypothesis that the P. cubensis subpopulation from C. sativus in Michigan is differentiated by geographic location, we performed an AMOVA and calculated FST statistics using the samples from the Michigan’s subpopulations (east, west and central) and the Midwest-w/o-Mi subpopulation. Population differentiation among the P. humuli subpopulations of Michigan (north and west) was analyzed in the same way. A Mantel test (Mantel, 1967) was performed to evaluate the relationship between the genetic distance of the samples and the physical distance of the location where the samples originated. This analysis was performed for each species independently (P. cubensis and P. humuli) using only the samples collected in Michigan. The Mantel test was performed as described by Gent et al. (2019) using the “mantel.rtest“ function of the ade4 package (Dray and Dufour, 2007). The genetic and Euclidean-geographic distance matrices were created with the function “bitwise.dist” (Kamvar et al., 2014) and “dist”, respectively. Physical distances were calculated using the coordinates of each sample location. Genotypic diversity. Genotypic diversity was determined using the filtered SNP data (not clone corrected) using the R package Poppr (Kamvar et al., 2014). The diversity estimates calculated included genotypic richness (MLG), expected number of genotypes based on rarefaction (eMLG), evenness (E5) and the indices of Shannon- Wiener (Shannon, 1948), Stoddart and Taylor’s (Stoddart and Taylor, 1988), and 146 Simpson’s (Simpson, 1949). The within-population gene diversity (Hexp) (Nei, 1978) that describes the proportion of heterozygous genotypes was calculated using the function “basic.stats” of the R package hierfstat. The diversity estimates were calculated using i) all P. cubensis and P. humuli samples grouped by host ii) all P. cubensis samples collected from C. sativus in each Michigan subpopulation and the Midwest-w/o- Michigan subpopulation iii) the P. humuli samples from the north and west subpopulations of Michigan. Only categories or subpopulations with at least five individuals were included for this analysis. Reproductive Mode. To infer the predominant reproductive mode of P. cubensis and P. humuli in Michigan (e.g., sexual, clonal, or mixed) we calculated the index of association IA (Brown et al., 1980; Milgroom, 1996) according to Tabima et al. (2018) using only the P. cubensis and P. humuli samples collected in the state from C. sativus and H. lupulus, respectively. For this, we calculated the index of association (IA) using a subset of 1000 random high-quality SNP and compared to the observed value of simulated populations with 0, 25, 50, 75, and 100% linkage. The mean IA values of each species were compared independently against the mean IA values of the simulated populations. Simulations were conducted with a dataset consisting of 2,978 loci analogous to the observed P. cubensis and P. humuli data. A nonparametric Kruskal- Wallis rank-sum test was used to assess the differences among means for all IA distributions. Multiple comparisons between IA distribution were performed using the nonparametric Kruskal- Wallis test and Tukey’s honest significant difference (HSD) test. 147 RESULTS Library sequencing and read mapping. The sequencing and alignment results of all libraries used in this study are summarized in Table 4-1. A total of 270 samples containing 2 – 500 ng of DNA were used for library preparation and enrichment. However, we could not generate enough enrichment products from 136 libraries, including 51 libraries prepared from samples with less than 50 ng of DNA. The remaining libraries (N=134) were enriched successfully and were prepared from samples containing 2 – 500 ng of DNA; these libraries included 92 libraries of P. cubensis and 42 libraries of P. humuli that were sequenced using two lanes of HiSeq4000. TE libraries: 510 million paired end reads (2 X 150bp) were obtained from the sequencing of these libraries (134) generating approximately 3.9 ± 2.6 SD (standard deviation) million reads per library. A range of 35 to 54% of the reads aligned to the reference genome of P. cubensis. The total number of reads obtained per library was influenced by the hybridization time of the libraries with the capture probes for enrichment. The libraries that hybridized for 48 h generated almost two times more aligned reads (2.64 ± 2.6 SD million) than the libraries that hybridized for 24 h (1.34 ± 1.0 SD million) (Table 4-1). Less than 20,000 sequencing reads were generated for only eight libraries that were eliminated (six samples were hybridized for 24 h and two were hybridized for 48 h), leaving 126 enriched libraries for downstream analysis. Five percent of the reads per library that did not align to the reference genome of P. cubensis aligned to the C. sativus reference genome. 148 Lc-WGS libraries: Seven libraries of P. cubensis and two libraries of P. humuli were sequenced without enrichment using MiSeq (Lc-WGS) generating a total of 41 million sequencing reads (2 X 150 bp). Approximately 0.55 ± 0.37 SD million reads were obtained per library, but only 25% of them aligned to the reference genome of P. cubensis. Twenty-five percent of the remaining reads aligned to the C. sativus reference genome. WGS libraries: A total of 350 million paired end reads (2 x 150 bp) were obtained from the bioproject PRJNA360426, that included WGS data of 9 libraries of P. cubensis and one library of P. humuli. An average of 15.6 ± 6 SD million reads per sample we obtained after sequencing. Eighty-nine percent of the total reads aligned to the P. cubensis reference genome and approximately 2% of the reads per sample aligned to the reference genome of C. sativus. Genotyping and variant calling. SNPs were retrieved from the sequencing reads generated from 126 enriched DNA libraries and the reads obtained from the 10 libraries sequenced using whole genome sequencing (Bioproject PRJNA360426). Ninety percent of the enriched libraries were prepared from input DNA amounts between 2 to 100 ng (Fig. 4-2 B). After mapping the reads to the reference genome of P. cubensis, 5,957 biallelic SNPs were identified across 70% of the 136 samples. To retain the maximum number of SNPs, we excluded 31 of the samples prepared using TE as they contained more than 40% missing data (24 of these were hybridized for 24 h, with the remaining libraries hybridized for 48 h) (Table 4-1). SNPs with more than 75% of the missing data across the entire data set were also excluded. This resulted in a total of 2,978 high-quality (HQ) SNPs with <10% missing data for downstream 149 analysis. All HQ SNPs were located within 127 out of the 812 genes targeted, 66% of the SNPs were found within secreted and effectors proteins. The remaining 44% of the SNPs were contained within esterases (16.3%), glucanasas (3%), lyases (3%), SSRs (4%), and other (7%) genes. The coverage depth (CD) across the 2,978 HQ SNPs was variable among samples and significant differences in the CD were observed between the libraries genotyped using TE and the libraries genotyped using WGS (Kruskal-Wallis p = 2.2e- 16; Fig. 4-2 A). The enriched libraries with a hybridization time of 48 h (TES3) exhibited a significantly higher CD than the enriched libraries generated with 24 h of hybridization time (TES2) or the libraries genotyped using WGS (Kruskal-Wallis p = 2.2e-16; Fig. 4-2 A). The CD across all HQ SNPs had an average of 262 ± 546 SD and 358 ± 546 SD for the libraries with hybridization times of 24 and 48 h, respectively (Table 4-1; Fig. 4-2 A). Among the samples sequenced using WGS, the CD across HQ SNPs had an average of 180 ± 116 SD (Table 4-1). No correlation was observed between the amount of input DNA used for library preparation and the CD across HQ SNPs (p > 0.05, Fig. 4-2 B). Characterization of technical error. Technical replicates were used to characterize the number of errors introduced into a sample during TE. Of the five samples with two technical replicates submitted for enrichment, only three samples (18- 35, L32 and L682) and their corresponding replicates were retained after the quality filtering of the SNPs. Differences in the average coverage between technical replicates of the same sample were observed (Fig. 4-3 A), however, only a small proportion of loci were different between replicates (Fig. 4-3 B). On average, 30 SNPs (out of 2,978 total 150 SNPs) differed between technical replicates, suggesting an error rate of roughly 1% within the high-quality variants. Genetic differentiation and population structure . In the genetic distance tree of HQ SNPs (2,978) all 101 samples genotyped fell into one of two clades (100% bootstrap support); either P. cubensis (67 samples) or P. humuli (34 samples) (Fig. 4-4). Within the P. cubensis clade, samples were either clade I or II of P. cubensis (Thomas et al., 2017). The majority of samples from Cucumis spp. aligned with clade II of P. cubensis, while the remaining samples collected from C. moschata and C. lanatus aligned with clade I (Fig. 4-4). Samples in the ordination plot formed 3 clusters corresponding to P. humuli, P. cubensis clade I, and P. cubensis clade II; these clusters were consistent with the stratification by species and within P. cubensis by host in the genetic distance tree (Fig. 4-5 A). F-statistics (FST > 0.25) and AMOVA detected significant genetic differentiation (P = 0.01) between pathogen species, supporting the clustering of the ordination plot (Table 4-3). Stratification by host was supported by a high level of genetic differentiation among samples collected from Cucumis spp. and samples collected from C. moschata (FST = 0.11, Table 4-3). AMOVA revealed that 78% of the total genetic variance was significantly associated with differences between host within P. cubensis (P = 0.01, Table 4-4). All P. cubensis samples collected from C. sativus in the Midwest including Michigan and the reference sample from Florida clustered loosely together across all four quadrants of the ordination plot (Fig. 4-5 B). However, the reference samples from North Carolina and California clustered separately from each other and were not 151 contained within the 90% confidence ellipses of the other subpopulations (Fig. 4-5 B). More than 70% of the samples from the subpopulations of the central and east Michigan were contained within the 90% confidence ellipse of the Midwest-w/o-MI subpopulation (Fig. 4-5 B); however, only 40% of the samples from west Michigan were contained within the ellipse of the of the Midwest-w/o-MI subpopulation. AMOVA did not support significant genetic differentiation among the subpopulations of P. cubensis from C. sativus in Michigan and the Midwest-w/o-MI subpopulation (P = 0.21). Only 1% of the total genetic variance was associated with difference among subpopulations (or regions) (Tables 4-6 and 4-7). The clone correction of all P. cubensis samples collected from C. sativus (N=57) resulted in 40 MLGs, 23% of which were detected in multiple subpopulations (Fig. 4-6 A, B). MLG 100 was detected in west Michigan and the Midwest-w/o-MI subpopulations (Fig. 4-6 A). The subpopulation of west Michigan also shared three MLG with the subpopulation of east Michigan (MLG 27, 36 and 46) (Fig. 4-6 A). The subpopulations of east and central Michigan shared MLG 33 and MLG 91 and two identical samples corresponding to MLG 47 were detected in two distant areas of the Midwest-w/o-MI subpopulation (i.e. Indiana and Ontario) (Fig. 4-6 A). Additionally, the MLG 29 from Florida was also detected in the east Michigan subpopulation (Fig. 4-6 A). Supporting these findings, the Mantel test did not detect a significant relationship between the genetic and physical distances of the P. cubensis samples collected from C. sativus in the Midwest (P = 0.454, Fig. 4-7 A). The level of genetic differentiation among P. humuli samples from Michigan and Oregon (Fst = 0.0168) was higher when compared to the genetic differentiation 152 detected among P. cubensis samples from C. sativus originating in the Midwest, California, Florida and North Carolina (Fst = 0.0024). In the ordination plot of all the P. humuli samples, most samples from the subpopulation of west Michigan were loosely dispersed in the left quadrants while most samples from the subpopulation of north Michigan (75%) were clustered tightly in the right quadrants (Fig. 4-5 C). Only two samples from the north subpopulation were contained within the 90% confidence ellipse of the west subpopulation (Fig. 4-5 C). The samples from central and east Michigan were either contained in the ellipses from the north or west subpopulations. The reference sample from Oregon was not contained within the 90% confidence ellipse of Michigan subpopulations (Fig. 4-5 C). These patterns were supported by AMOVA in which significant genetic differentiation between the subpopulations of north and west Michigan was detected (P=0.01, Table 4-9). A total of 11.25% of the genetic variance was significantly associated with differences among subpopulations and only 2.5% of the variance was associated with differences among hop yards within a subpopulation (Table 4-10). This suggests that the geographic region has a significant effect on the structure of the P. humuli population in Michigan. Clone correction of the 34 P. humuli samples resulted in 23 MLGs, only one of them (MLG 4) was detected in two different subpopulations (Fig. 4-6 C, D). Only two different MLGs were shared among hop yards within regions. MLG 3 was detected in two different years at the three yards in north Michigan and MLG 59 was detected once in the two hop yards sampled in west Michigan (Fig. 4-6 C). The Mantel test supported the hypothesis that the genetic distance among P. humuli samples collected in Michigan 153 was significantly associated (P= 0.0042) with the geographic distance among hop yards (Fig. 4-7 B). Genotypic diversity. The highest level of genotypic diversity was observed among the samples of P. humuli followed by P. cubensis clade I and P. cubensis clade II (Table 4-5). Each species or clade consisted of different MLGs spread relatively evenly. This was reflected in the estimates of genotypic richness (eMLG= 6 to 9.54) and evenness (E5 > 0.8) (Table 4-5). The Shannon-Wiener index, Stoddart and Taylor’s index, and Simpson’s index were highest among P. cubensis clade II samples, followed by P. humuli samples, and then P. cubensis clade I samples (Table 4-5). The greatest expected heterozygosity or proportion of heterozygous genotypes expected was observed among P. humuli samples followed by P. cubensis clade I samples, and then P. cubensis clade II samples (Table 4-5). The Michigan subpopulations and the Midwest-w/o-MI subpopulation of P. cubensis collected from C. sativus exhibited similarly low levels of expected heterozygosity and genotypic diversity (Table 4-8). Expected heterozygosity ranged between 0.178 and 0.197 while genotypic richness (eMLG) ranged between 6 to 9.57. The east Michigan subpopulation had the greatest expected heterozygosity (Hexp= 0.197) followed by Midwest-w/o-MI (Hexp= 0.190), west Michigan (Hexp= 0.189), and central Michigan subpopulations (Hexp= 0.178) (Table 4-8). The east Michigan subpopulation had the highest genotypic richness (eMLG), while west Michigan and the Midwest-w/o-MI subpopulations had larger sample size and lower number of MLGs (Table 4-8). The Shannon-Wiener index, Stoddart and Taylor’s index, and Simpson’s index were all highest for the west Michigan subpopulation, followed by the 154 subpopulations of east Michigan, Midwest-w/o-MI, and then central Michigan (lowest indices) (Table 4-8). The highest evenness (E5) was detected for the central Michigan subpopulation followed by the subpopulations of east Michigan, west Michigan and Midwest-w/o-MI (lowest E5) (Table 4-8). The genotypic diversity varied more widely across the P. humuli subpopulations of Michigan (eMLG = 5 to 11) compared to the variation detected among P. cubensis subpopulations (eMLG = 6 to 9.57). The expected heterozygosity among the subpopulations of P. humuli was higher and ranged between 0.206 to 0.220. The subpopulations of north (Hexp= 0.224) and east-central Michigan (Hexp= 0.220) had the greatest expected heterozygosity followed by the subpopulation of west Michigan (Hexp= 0.206) (Table 4-11). The subpopulation of west Michigan had the highest genotypic richness, while the north Michigan subpopulation had the largest sample size and the most MLGs. The Shannon-Wiener index, Stoddart and Taylor’s index, and Simpson’s index were all highest for the north Michigan subpopulation, followed by the subpopulations of west and east-central Michigan (lowest indices) (Table 4-11). The highest evenness (E5) was detected in the subpopulations of west and east-central Michigan followed by the north Michigan subpopulation (lowest E5) (Table 4-11). Reproductive Mode. To determine whether allelic variants were randomly associated as expected in populations with a sexual mode of reproduction (low linkage), we calculated the mean index of association across loci for each species in Michigan and compared to the estimated index of association of simulated populations under strong linkage (100% linked loci), moderate linkage (75 and 50% linked loci), and low linkage (25% and 0 loci under linkage) (Fig. 4-8). Upon comparison, significant 155 differences were observed between the mean index of association (IA) of each species and the mean values estimated for the simulated populations (P. cubensis, P < 2.2e-16 and P. humuli, P < 2.2e-16). The IA mean value of the P. cubensis samples was situated between the IA mean values of the simulated data with 50 and 75% linkage (Fig. 4-8 A). This indicated that populations of P. cubensis have a mixed-mode of reproduction with a predominantly clonal phase. The IA mean value of the P. humuli samples was not significantly different to the simulated data with 25% linkage (Fig. 4-8 B), suggesting that the populations of P. humuli may have a mixed mode of reproduction that could be predominantly sexual. DISCUSSION Technological advances in genome sequencing have accelerated the cataloging of genomic variation of plant pathogens, however, extracting large amounts of high- quality DNA from obligate pathogens responsible for DM is challenging, and hampers their genotyping using next-generation sequencing. We adapted a TE method that facilitated the sequencing of 736 genes annotated as virulence factors in the genome of P. cubensis. We used this method to sequence P. cubensis and P. humuli DNA extracted from sporangia collected from plant tissue with signs of the pathogen. This approach facilitated the genotyping of samples that contained very low amounts of pathogen DNA mixed with other environmental contaminants (i.e. plant DNA, bacteria DNA). After aligning the sequenced DNA, we identified 2,978 SNPs and resolved the population structure of P. cubensis and P. humuli in Michigan. A significant effect of location on the genetic variation of P. humuli was detected and the genetic distance among samples was associated with the physical distance of the hop yards. Evidence 156 of location-based differentiation within Michigan was not detected for the P. cubensis population. By using affinity probes and several amplification cycles, our TE protocol was designed to provide high coverage sequencing of specific loci in the P. cubensis and P. humuli genome. This facilitated the sequencing of a low amount of target (pathogen) DNA from environmental samples containing a significant amount of contaminant DNA from plant tissues and other microorganisms. Sequencing DNA directly from environmental samples reduces time, labor, and DNA input required for other sequencing approaches such as genotyping by sequencing (GBS) or WGS (Summers et al., 2015b; Thomas et al., 2017a; Gent et al., 2019). This is particularly advantageous for DM, as other sequencing approaches (GBS, WGS, Rad-Seq) rely on large amounts of high-quality DNA that can only be obtained using a laborious propagation procedure (Thomas et al., 2017a; Gent et al., 2019). Using TE, several samples with less than 50 ng of DNA were successfully genotyped. Most of the samples that failed were leaf lesions containing few or no sporangia (P. cubensis had been previously confirmed via qPCR). The TE protocol also enabled the genotyping of samples that were no longer viable for propagation, expanding the number of samples that could be genotyped including those collected more than ten years ago. Using a reduced amount of DNA (2 to 100x less), the samples sequenced after TE reached similar sequencing coverage when compared to the samples sequenced using a WGS approach, at the region’s target (250x). However, the enriched samples required 6 to 7 times less space for sequencing (60 to 70 samples per lane of HiSeq) compared to the space used for the sequencing of libraries using WGS (10 samples per 157 lane of HiSeq). This was possible because TE reduced the complexity of the P. cubensis genome from 88.22 MB to < 1MB (i.e. 1.13% of the genome). This reduction in the genome complexity and the enrichment of up to four samples in a single enrichment experiment (MyBaits) make TE a cost-effective alternative to WGS. However, we did find that our TE protocol can result in a higher number of sequencing errors per sample compared to WGS. The high number of amplification cycles used for the enrichment of samples with low amounts of DNA may have led to the introduction of PCR errors and a subsequent reduction in genotyping accuracy. Typically, high-throughput sequencing (e.g. WGS and GBS) without enrichment of samples with low amounts of DNA (<250ng) requires an average of 14 PCR cycles, however, 6-14 extra cycles were used after library construction in our TE protocol (for a total of 20-28 cycles). Thus, possibly due to the high number of extra amplification cycles, the samples genotyped using TE were subject to an error rate of approximately 1%, an error rate 10 times higher than the estimated error rate for high-throughput sequencing data of >0.1% (Grünwald et al., 2017; Ma et al., 2019). Despite an error rate of 1%, the genetic distance between samples in 95% of cases was greater than the distance generated due to PCR errors, providing confidence that our findings reflect patterns linked to the pathogen biology than an artifact of sequencing. Using TE, we detected a significant effect of host in the population structure of P. cubensis, and the samples collected from C. sativus were genetically different from the samples collected from C. moschata. These results support previous studies that have shown that the structure of the P. cubensis population is driven by host preference, with samples from Cucumis spp. and C. moschata belonging to two 158 distinct evolutionary clades (Thomas et al., 2017a; Wallace et al., 2020). Thomas et al. (2017) identified two P. cubensis clades in the U.S.; clade I occurs on C. pepo, C. moschata, C. maxima, C. lanatus, and clade II occurs on Cucumis spp. In our analyses, all P. cubensis samples clustered by clades according to that distribution. Annual CDM infections in the northern U.S. are driven by an influx of airborne P. cubensis sporangia from overwintering sites (Bello et al., 2020; Naegele et al., 2016; Goldenhar and Hausbeck, 2019). CDM limits cucumber yield in Michigan where 15000 ha of cucumbers are planted every year (USDA, 2020). Small, yet significant genetic differences among cucumber production regions in Michigan were previously reported (Naegele et al., 2016). However, a significant effect of location in the subpopulation structure of P. cubensis was not detected in the current study, suggesting no differentiation within and among Michigan’s regions. An exchange of migrants has a homogenizing effect on subpopulations (Milgroom, 2015) and may occur in Michigan due to the availability of susceptible crops across the state. The detection of the same multi-locus genotypes (MLG) in multiple locations supports this hypothesis and also suggests that P. cubensis sporangia may be disseminated unrestricted in the state as there is no geographical barrier (e.g. mountain range or water body) that may limit the homogenizing of geographically distant populations. The wide host availability in Michigan may also facilitate the establishment of incoming MLG from other subpopulations in the Midwest. The exchange of migrants may occur among Midwestern states outside of Michigan but the absence of susceptible crops between geographically distance populations may result in lower rates of exchange and 159 subsequently more genetically differentiated populations. Additional sampling is needed to test this hypothesis. Significant differences were detected between the P. humuli populations from the north and west Michigan regions, but we did not find any evidence of genetic differentiation among hop yards within the same region. In Michigan, approximately 400 ha of hops are planted across more than 50 commercial hop yards (Michigan Department of Agriculture & Rural Development, 2018). A lack of differentiation among hop yards within regions could also be a consequence of the continuous exchange of P. humuli genotypes among them. However, despite the potential for airborne dispersal of hop DM (Bello et al, 2020; Gent et al., 2009) only one MLG was detected in multiple hop yards within the same region. This is consistent with the restricted pattern of dispersal suggested for this pathogen; new infections of hop plants by P. humuli are less likely to occur far from their inoculum source (Johnson et al., 1991). We detected a significant correlation between the genetic and geographic distance of P. humuli samples, providing circumstantial evidence of a higher probability for new P. humuli infections to occur close to their inoculum source. In Oregon, Gent et al. (2019) found significant genetic differences between hop yards planted within 10 km of each other and a low amount of population differentiation between hop yards established from the same planting material. This suggests that infected hop plant material may be a more important source of primary inoculum than airborne migration for DM. Most of the hop yards sampled for this study were among the first yards establish from rhizomes in the mid-2000s when commercial hop production returned to Michigan and very few sources of propagation material were available 160 (Sirrine et al., 2014). As the hop industry matured in the state, new sources of propagation material became available, but still in limited numbers. The lack of genetic differentiation among hop yards within regions in Michigan may be a result of the introduction of a reduced number of genotypes via planting material from the same origin source into multiple hop yards. The detection of the same MLGs in multiple hop yards and relatively low genetic diversity of north Michigan partially support this hypothesis but further sampling and more information on sources of plant material are needed to verify this hypothesis in Michigan. The restricted airborne dispersal of P. humuli sporangia compare to P. cubensis could be attributed to the cultivation practices and geographic distribution of hops. Generally, diseased basal shoots are close to the ground and during most of the season sheltered within a canopy of healthy basal shoots that hamper dispersal (Johnson et al., 1991). Additionally, the area planted with hops in the state is equivalent to only 2% of the area planted with cucumbers. This lower host availability may result in reduced P. humuli sporangia production compared to P. cubensis and lower aerial exchange of MLG among subpopulations. The restricted exchange of MLG can result in genetically differentiated subpopulations (Milgroom, 2015). The relatively low level of genetic differentiation and genetic diversity detected within the Pseudoperonospora spp. populations of Michigan is consistent with the clonal reproductive mode of P. cubensis and inbreeding reproductive mode of P. humuli (Naegele et al., 2016; Gent et al., 2019). Although we estimated relatively low indices (65%) of association for both species that suggest they experience a sexual phase, especially P. humuli, we believe these values are an artifact of our sampling strategy. 161 Our samples were collected from a bulked inoculum of multiple leaf lesions (cucurbits) or diseased shoots (hops), so it is likely that our sample units may contain multiple genotypes creating an effect of random mating (no linkage-disequilibrium). An earlier study of P. humuli populations in the Pacific Northwest that used a similar sampling strategy compared to our study and suggested that the P. humuli population of Oregon was reproducing sexually (Chee et al., 2006). However, in a genetic study of P. humuli using more precise sample units in the same region, GBS revealed strong evidence of linkage disequilibrium (Gent et al., 2019). This is consistent with the nonrandom mating expected from species with an asexually reproducing or highly inbreed mode (Gent et al., 2019). Similarly, Wallace et al. (2020) provided evidence of non-random mating or recombination consistent with selfing or asexual reproduction for P. cubensis clade II. Future studies using TE should consider the use of a single lesion as a sample unit and sequencing selectively neutral regions (Grünwald et al., 2016). We included 94 regions (genes) containing small sequence repeats (SSRs) markers among the regions targeted for sequencing, as SSRs can provide less biased estimates of population differentiation due to their neutrality. However, the sequencing of these regions was of poor quality and we could not detect SSRs within them. Instead, we found 115 SNPs within these regions that were not significantly differentiated among or within subpopulations (data not shown). Most of the signal of population differentiation was found within 66% of the HQ SNPs used, these SNPs were contained within secreted proteins or effector genes that corresponded to 10% of the genes targeted. Thus, we encourage the use of neutral polymorphic regions more evenly distributed across the genome. 162 We also recommend the use of a higher number of baits for enrichment when working with low concentrations of target DNA and the quantification of plant and pathogen DNA using qPCR before library preparation. More baits should increase the sequencing coverage across samples, facilitating the calling of variants. Similarly, the utilization of samples with a higher amount of pathogen DNA relative to the amount of contaminating DNA could increase the chances for the successful preparation of libraries using TE. Future studies should also include more technical replicates to assess error due to batch effects as the high number of amplification cycles resulted in lower genotyping accuracy. The introduction of technical replicates from multiple generations of clones is also advised. This is critical when working with high-throughput sequencing data due to the introduction of false mutations in the data that could create additional multi-locus genotypes (Gent et al., 2019; Potapov and Ong, 2017). In summary, our results reveal the key strengths of TE for the genotyping of DM. This approach provides a solution for the genotyping of obligate biotrophic pathogens; for which high-throughput sequencing is typically constrained by the low amounts of target DNA and high amounts of non-target (contaminating) DNA. TE provides a cost- effective approach for the genotyping of unpurified field samples and the assessment of sequence polymorphisms across a large number of individuals. This method could be adapted to a diverse group of pathogens even without a reference genome. Future population studies using TE should carefully consider the sampling strategy and the size and complexity of the genomic regions targeted. Including technical replicates will also be important to ensure the accurate genotyping of the samples after enrichment and the reproducibility of the experiments. 163 APPENDIX 164 Table 4-1. Sequencing and alignment results from libraries sequenced using target enrichment (TE), low coverage whole genome sequencing (Lc-WGS) and whole-genome sequencing (WGS) APPENDIX Sequencing characteristics Sequencing platform Sequencing format (bp) Sequencing lines Expected sequencing output per lane (Gbp) Input amount of DNA for library preparation Hybridization time with probes (baits) Total number of libraries sequenced Total number of clean reads (millions) % Reads aligned to reference genome of P. cubensis % Reads aligned to reference genome of C. sativus Aligned reads per samples ± SD (millions) c % Libraries with high quality SNPs d Average coverage among high quality SNPs ± SD e Lc-WGS WGSb TE seta 1 TE set 2 HiSeq4000 HiSeq4000 MiSeq.v2.300 HiSeq 2x100 2x150 1 1 105 105 5-400 ng > 1 µg -- 24 h 10 62 211 350 89% 35% < 5% 2% 16.6 ± 18 1.34 ± 1.0 100% 61% 262 ± 546 180 ± 116 2x150 2 3.6 – 4.5 < 50 ng -- 9 41 25% > 25% 0.55 ± 0.37 -- -- 2x150 1 105 5-400 ng 48 h 72 290 54% < 5% 2.64 ± 2.6 84% 358 ± 546 aTarget enrichment set with hybridization times of 24h and 48h bAll the sample genotyped using whole-genome sequencing were retrieved from the bioproject PRJNA360426. cMillions of reads aligned to the reference genome of Pseudoperonospora cubensis per sample. dLibraries (%) retained containing 2,978 high quality SNPs. eAverage coverage among high quality SNPs within the libraries retained. 165 Table 4-2. Plant hosts and the location of the 101 samples of Pseudoperonospora spp. samples used for the population analysesa West Michigan East Michigan North Michigan Central Michigan Other states within the U.S Total Pathogen and host species P. cubensis clade I Cucurbita spp. and Citrullus lanatus P. cubensis clade II Cucumis spp. P. humuli Humulus lupulus Totala 1 20 8 34 1 15 0 25 0 0 20 20 0 6 5 11 6 18 1 11 8 59 34 101 aThe total of samples reflects the sum of the samples retained after quality filtering. 166 Table 4-3. Pairwise FST comparisons among Pseudoperonospora cubensis (clade I and II) and Pseudoperonospora humuli. FST 0.11 0.27 0.25 Pvaluea 0.01 0.01 0.01 Pairwise comparison P. cubensis clade II (Cucumis spp.) vs P. cubensis clade I (Cucurbita spp.) P. cubensis clade II (Cucumis spp.) vs P. humuli (Humulus lupulus) P. cubensis clade I (Cucurbita spp.) vs P. humuli (Humulus lupulus) aPvalues were calculated using an analysis of molecular variance (AMOVA) 167 Table 4-4. Analysis of molecular variance (AMOVA) for Pseudoperonospora spp. grouped by host. The significance of variance was tested from 999 permutations of the dataa. MSD d Variance (%)e Source Between host Between samples within host Within samples Total Sigma df b SSD c 4033.0104 806.60208 68.370171 78.720821 5 2.353212 301.9755 2.709468 30.19755 10 16.128063 18.569711 85 1370.8853 16.12806 100 5705.8712 57.05871 86.851446 100 Pvalue 0.01 aSample sizes used in this analysis included 59, 34 and 8 samples collected from Cucumis spp., Humulus lupulus and Cucurbita spp., respectively. bDegrees of freedom. cSum of squared differences. dMean of squared differences. eVariance (%) was adjusted to zero for negative sigma values. 168 Table 4-5. Genotypic diversity estimates for Pseudoperonospora spp. samples grouped by host (clade). Clade/Host species P. cubensis clade I Cucurbita spp. and Citrullus lanatus P. cubensis clade II Cucucmis spp. P. humuli Humulus lupulus Total 1.67 N a MLG b eMLG c SE d H e 8 59 34 101 77 0 0.634 3.72 0.851 3.17 0.533 4.24 6 9.54 9.14 9.7 6 45 27 λ g 0.781 G f 4.57 36.64 0.973 19.27 0.948 59.65 0.983 Hexp i E5 h 0.831 0.215 0.889 0.192 0.803 0.217 0.86 0.206 aN is the number of individuals sampled of each region. bMLG is the number of multilocus genotypes observed. ceMLG is the number of expected MLGs at a sample size based on rarefaction. dStandard error eShannon-Wiener Index (H) fStoddart and Taylor’s Index (G) gSimpson’s index (lambda) hEvenness (E5) iWithin population gene diversity (Hexp). 169 Table 4-6. Analysis of molecular variance (AMOVA) of Midwest subpopulationsa of Pseudoperonospora cubensis collected from Cucumis sativus. The significance of variance was tested from 999 permutations of the dataa. Source Between subpopulations Between samples within subpopulations Within samples Total dfb SSDc 3 35.30 12 125.89 38 351.62 53 512.82 MSDd Sigma 11.76 0.012 10.49 0.449 9.25 9.253 9.714 9.67 % variancee Pvalue 0.125 0.21 4.625 95.248 100 aSample sizes used in this analysis included 15, 19, and 6 samples collected from Cucumis sativus in the east, west, and central Michigan, respectively. The Midwest-w/o-MI subpopulation (17 samples) was formed by all the samples collected in the Midwest states of the U.S. and Ontario, Ca, not including Michigan bDegrees of freedom. cSum of squared differences. dMean of squared differences. eVariance (%) was adjusted to zero for negative sigma values. 170 Table 4-7. Pairwise FST comparisons among subpopulations of Pseudoperonospora cubensis collected from C. sativus in the Midwesta. Subpopulation East Michigan Central Michigan West Michigan Midwest-w/o-Michigan b East Michigan -- 0.0083* 0.0034 0.0025 Central Michigan -- 0.0035 0.0039 West Michigan -- 1E-04 Midwest-w/o- Michigan -- aSample sizes used in this analysis included 15, 19, and 6 samples collected from Cucumis sativus in the east, west, and central Michigan, respectively. bThe Midwest-w/o-MI subpopulation (17 samples) was formed by all the samples collected in the Midwest states of the U.S. and Ontario, Ca, not including Michigan *Indicates a significant Fst value based on AMOVA 171 Table 4-8. Genotypic diversity estimates and index of association of Pseudoperonospora cubensis subpopulations collected from C. sativus in the Midwest. Subpopulation East Michigan Central Michigan West Michigan Midwest-w/o-Michiganj Total N a 15 6 19 17 57 MLG b eMLG c SE d 14 0.495 0 6 0.772 16 0.795 14 44 0.638 9.57 6 9.07 8.85 9.54 H e 2.62 1.79 2.7 2.56 3.69 G f 13.2 6 13.4 11.6 35.7 λ g 0.924 0.833 0.925 0.913 0.972 E5 h 0.965 1 0.893 0.887 0.886 Hexp i 0.197 0.178 0.189 0.19 0.192 aN is the number of individuals sampled of each region. bMLG is the number of multilocus genotypes observed. ceMLG is the number of expected MLGs at a sample size based on rarefaction. dStandard error eShannon-Wiener Index (H) fStoddart and Taylor’s Index (G) gSimpson’s index (lambda) hEvenness (E5) iWithin population gene diversity (Hexp). jThe Midwest-w/o-MI subpopulation was formed by all the samples collected in the Midwest states of the U.S. and Ontario, Ca, not including Michigan 172 Table 4-9. Analysis of molecular variance (AMOVA) of Michigan subpopulations of Pseudoperonospora humuli. The significance of variance was tested from 999 permutations of the dataa Source Between subpopulations Between samples within subpopulations Within samples Total Df b SSD c MSD d 1 40.187 40.18681 1.986518 11.249556 Sigma % variance e Pvalue 0.01 3 23 27 53.042 17.68075 0.449694 2.546595 350.12 15.22243 15.22243 86.203849 443.34 16.42018 17.65864 100 aSample sizes used in this analysis included 8 and 20 samples from two and three hop yards in west and north Michigan, respectively. bDegrees of freedom. cSum of squared differences. dMean of squared differences. eVariance (%) was adjusted to zero for negative sigma values. 173 Table 4-10. Pairwise FST comparisons among subpopulations of Pseudoperonospora humuli collected from hop yards in Michigana. Subpopulation/Yard West North D E B A C West D -- 0.00237 0.02269* 0.03052* 0.02879* North B -- E -- 0.00953 0.021636* 0.002614 0.019606* 0.001831 A -- 0.002613 C -- aSample sizes used in this analysis included 4, 4, 6, 4 and 10 samples collected from hop yards indicated as D, E, C, B and A, respectively. The hop yards in the central region were excluded from analysis due to a low number of samples. *Indicates a significant Fst value based on AMOVA 174 Table 4-11.Genotypic diversity estimates and index of association of Michigan subpopulations of Pseudoperonospora humuli. E5 h λ g 0.875 1 2.08 8 0.83 2.11 5.88 1.61 5 0.8 2.86 12.52 0.92 0.676 1 0.701 Hexp i 0.206 0.224 0.220 0.217 8 11 5 22 8 6.79 5 8.33 Subpopulation West North East-central Total N a MLG b eMLG c SE d H e G f 8 20 5 33 0 1.05 0 1.08 aN is the number of individuals sampled of each region. bMLG is the number of multilocus genotypes observed. ceMLG is the number of expected MLGs at a sample size based on rarefaction. dStandard error eShannon-Wiener Index (H) fStoddart and Taylor’s Index (G) gSimpson’s index (lambda) hEvenness (E5) iWithin population gene diversity (Hexp). 175 Figure 4-1. Cucumis sativus planted acreage and number Humulus lupulus planted in Michigan by county (Adapted from Neufeld, 2017). A) C. sativus planted acreage, location and number of samples collected in the Michigan (N=40). A total of 26 samples were collected outside of Michigan in Ontario Canada (N=6) and the U.S. states of Florida (N=2), Alabama (N=1), North Carolina (N=2), South Carolina (N=2), California (N=1), New York (N=1), Wisconsin (N=1), Ohio (N=3), Iowa (N=1), Indiana (N=5), Georgia (N=1). (B) H. lupulus planted, location and number of samples collected from hop yards in Michigan (N=33) and Oregon (N=1). Hop yards with more than 800 plants (H. lupulus) are represented by stars, those with fewer are represented by circles. Hop yard sampled (A-F) are colored in red. (C) Population assignment by geographic location of Pseudoperonospora spp. samples in the Midwest (Wisconsin, Indiana, Ohio, Michigan (Mi) and Ontario, CA). 176 A t ) h p e d e g a r e v o C ( 0 1 g o l 3 2 1 R = -0.053 , p = 0.78 R = 0.077 , p = 0.59 Tec TES2 TES3 WGS B 3.5 ) h t p e d e g a r e v o C ( 0 1 g o l 3.0 2.5 2.0 1.5 1.0 set a a a s2 s3 wgs 2 S E T 3 S E T Sequencing approach S G W 0 25 50 ng of DNA 75 100 Figure 4-2. Depth coverage of high-quality SNPs across libraries. (A). Depth coverage of high-quality SNPs (log10) across libraries of Pseudoperonospora cubensis and Pseudoperonospora humuli sequenced using target enrichment with 24h (TES2) and 48h (TES3) of hybridization and whole genome sequencing (WGS). (B). Depth coverage of high-quality SNPs (log10) across libraries sequenced using target enrichment with different amounts of input DNA for library preparation. Only samples that were retained after quality filtering were included in the analysis. 177 A 5 3 8 1 . B 5 3 8 1 . A 2 3 B 2 3 Sample A 3 7 6 0.00 B 3 7 6 A 2 8 6 B 2 8 6 0.02 0.00 0.04 0.02 0.06 18−35A S37 0.04 0.08 0.06 Genetic distance (proportion of loci that are different) 2683 SNPs Genetic distance (proportion of loci that are different) 2683 SNPs A 2 8 6 B 2 8 6 A B Genetic distance 3 2 1 ) h ) t h p t p e e d d e e g g a a r e r v e o v C o ( c 0 ( 1 0 g 1 o g o L l Tec TES2 TES3 18-35B_S39 L32_S62 L682_S63 100 18-35A_S37 L32_S26 L682_S28 0.169 0.022 0.0104 0.176 0.011 0.174 0.011 0.176 0.173 100 100 100 100 L32 S62 L32 S26 L682 S63 L682 S28 L673A S4 100 18−35B S39 18−35A S37 7 3 S _ A 5 3 − 8 1 9 3 S _ B 5 3 − 8 1 6 2 S _ 2 3 L 2 6 S _ 2 3 L Sample Sample 4 S _ A 3 7 6 L 8 2 S _ 2 8 6 L 3 6 S _ 2 8 6 L 0.00 Genetic distance (proportion of loci that are different (2978 SNPs)) Genetic distance. Proportion of loci that are different (2978 SNPs) 0.04 0.02 0.06 0.08 Figure 4-3. Characterization of sequencing error among technical replicates. (A) Coverage distribution of the technical replicates sequenced. Violin plots are filled with 2,978 high-quality SNPs. Each sample is colored according to the sequencing batch. (B) Genetic differentiation among technical replicates. A UPGMA tree was reconstructed using 2,978 SNPs. The genetic distance represents the number of SNPs that are different among samples. 178 Cucurbita.spp Cucurbita.spp Cucurbita.spp C.sativus C.sativus C.sativus H.lupulus H.lupulus H.lupulus Cucurbita.spp Cucurbita.spp Cucurbita.spp Cucurbita moschata Cucumis sativus Humulus lupulus C.sativus C.sativus C.sativus Citrullus lanatus Cucurbita maxima Cucumis melo H.lupulus H.lupulus H.lupulus 86 100 100 100 93 100 100 86 100 84 78 Pseudoperonospora cubensis clade I 98 77 100 99 413 S36 *** ** 39 24 37 41 38 10D S8 L682 S63 L682 S28 L1335 S16 44 898 S15 156 S51 1484 S9 1755 S4 CDM207 S55 257 S50 CDM19 S22 L33 S3 1752 S32 L119 S27 L1621 S13 L1072 S21 CDM209 S11 CDM201 S9 CDM232 S59 238 S11 CDM154 S60 936 S35 1589 S7 L32 S62 L32 S26 L1755 S64 1649 S2 L573 S20 204 S10 CDM228 S6 CDM202 S12 253 S49 1335 S30 K2A S38 CDM191 S14 186 S3 CDM250 S17 200 S5 180 S14 CDM242 S23 14J S6 OH1−2 S20 229 S8 203 S1 L1024 S15 481 S13 CDM155 S31 L673A S4 CDM177 S25 413 S36 413 S36 CDM152 S30 CDM123 S7 671 S34 CDM110 S48 S17−1 S17 CPF7 S24 296 S12 CDM153 S29 42 40 40 44 44 43 18−25 S36 17−15 S33 18−23 S37 17−101 S74 18−32 S64 17 108B S41 18−26 S55 18−17 S52 17−8 S50 17−20 S76 17−112 S18 17−100 S73 18−31 S65 17−113 S72 A19 S21 18−7 S66 18−20 S54 18−6 S67 kk10−4 S75 17−103 S70 18−44 S59 18−42 S42 18−5 S56 18−15 S60 18−1 S69 18−43 S44 17−114 S71 58 S61 18−30 S53 18−41 S45 18−35B S39 18−35A S37 18−13 S27 25 *** ***** 100 93 100 100 100 39 24 37 41 38 10D S8 L682 S63 L682 S28 L1335 S16 898 S15 156 S51 1484 S9 1755 S4 CDM207 S55 257 S50 CDM19 S22 L33 S3 84 1752 S32 L119 S27 L1621 S13 L1072 S21 CDM209 S11 78 CDM201 S9 CDM232 S59 238 S11 CDM154 S60 936 S35 1589 S7 L32 S62 L32 S26 L1755 S64 1649 S2 L573 S20 204 S10 CDM228 S6 90 39 CDM202 S12 24 253 S49 37 1335 S30 41 K2A S38 38 CDM191 S14 10D S8 186 S3 L682 S63 CDM250 S17 L682 S28 200 S5 L1335 S16 180 S14 898 S15 CDM242 S23 98 156 S51 14J S6 1484 S9 OH1−2 S20 1755 S4 229 S8 CDM207 S55 203 S1 257 S50 L1024 S15 73 CDM19 S22 481 S13 L33 S3 CDM155 S31 1752 S32 L673A S4 L119 S27 CDM177 S25 L1621 S13 413 S36 L1072 S21 CDM152 S30 CDM209 S11 CDM123 S7 CDM201 S9 671 S34 100 CDM232 S59 CDM110 S48 238 S11 S17−1 S17 CDM154 S60 CPF7 S24 936 S35 296 S12 1589 S7 CDM153 S29 L32 S62 42 100 L32 S26 40 L1755 S64 44 100 1649 S2 43 L573 S20 18−25 S36 204 S10 17−15 S33 CDM228 S6 18−23 S37 CDM202 S12 17−101 S74 253 S49 18−32 S64 1335 S30 17 108B S41 K2A S38 18−26 S55 CDM191 S14 18−17 S52 186 S3 17−8 S50 82 CDM250 S17 17−20 S76 200 S5 17−112 S18 180 S14 17−100 S73 CDM242 S23 18−31 S65 97 14J S6 17−113 S72 OH1−2 S20 A19 S21 229 S8 18−7 S66 71 203 S1 18−20 S54 L1024 S15 18−6 S67 481 S13 kk10−4 S75 CDM155 S31 17−103 S70 L673A S4 18−44 S59 93 CDM177 S25 18−42 S42 413 S36 18−5 S56 83 CDM152 S30 18−15 S60 83 CDM123 S7 18−1 S69 73 671 S34 18−43 S44 CDM110 S48 17−114 S71 S17−1 S17 58 S61 CPF7 S24 74 18−30 S53 296 S12 18−41 S45 CDM153 S29 18−35B S39 42 18−35A S37 40 18−13 S27 44 25 43 18−25 S36 17−15 S33 18−23 S37 17−101 S74 18−32 S64 17 108B S41 18−26 S55 18−17 S52 17−8 S50 17−20 S76 17−112 S18 17−100 S73 18−31 S65 17−113 S72 A19 S21 18−7 S66 18−20 S54 18−6 S67 kk10−4 S75 17−103 S70 18−44 S59 18−42 S42 18−5 S56 18−15 S60 18−1 S69 18−43 S44 17−114 S71 58 S61 18−30 S53 18−41 S45 18−35B S39 18−35A S37 18−13 S27 25 0.00 100 Cucurbita.spp Cucurbita.spp Cucurbita.spp C.sativus C.sativus C.sativus H.lupulus H.lupulus H.lupulus C.lanatus C.lanatus C.lanatus C. maxima C. maxima C. maxima C. melo C. melo C. melo 100 100 100 100 100 100 73 84 86 100 93 100 100 100 100 100 90 100 100 90 100 84 78 78 100 86 100 100 100 90 93 100 100 100 100 100 C. melo C. melo C. melo C. melo C. melo C. melo C.sativus C.sativus C.sativus H.lupulus H.lupulus H.lupulus C.lanatus C.lanatus C.lanatus C.lanatus C.lanatus C.lanatus C.lanatus C.lanatus C.lanatus C. maxima C. maxima C. maxima C. maxima C. maxima C. maxima C. maxima C. maxima C. maxima Cucurbita.spp Cucurbita.spp Cucurbita.spp cubensis clade II Pseudoperonospora 39 24 37 41 38 10D S8 L682 S63 L682 S28 L1335 S16 898 S15 156 S51 1484 S9 1755 S4 CDM207 S55 257 S50 CDM19 S22 L33 S3 1752 S32 L119 S27 L1621 S13 98 L1072 S21 CDM209 S11 CDM201 S9 CDM232 S59 238 S11 CDM154 S60 936 S35 1589 S7 L32 S62 L32 S26 L1755 S64 1649 S2 L573 S20 204 S10 CDM228 S6 CDM202 S12 253 S49 1335 S30 K2A S38 CDM191 S14 186 S3 CDM250 S17 200 S5 180 S14 CDM242 S23 14J S6 OH1−2 S20 229 S8 203 S1 L1024 S15 481 S13 82 CDM155 S31 L673A S4 80 CDM177 S25 413 S36 97 CDM152 S30 CDM123 S7 671 S34 CDM110 S48 S17−1 S17 CPF7 S24 296 S12 CDM153 S29 42 40 44 43 18−25 S36 17−15 S33 18−23 S37 17−101 S74 18−32 S64 17 108B S41 18−26 S55 18−17 S52 100 17−8 S50 17−20 S76 17−112 S18 17−100 S73 Figure 4-4. Genetic differentiation of Pseudoperonospora cubensis and 18−31 S65 17−113 S72 A19 S21 Pseudoperonospora humuli samples. 18−7 S66 18−20 S54 0.02 18−6 S67 kk10−4 S75 17−103 S70 A UPGMA tree was reconstructed using 2,978 single nucleotide polymorphism (SNP) 18−44 S59 80 18−42 S42 variants. Bootstrap support values are indicated above the branches. The genetic 18−5 S56 18−15 S60 18−1 S69 distance represents the proportion of loci that are different between samples. * Samples 18−43 S44 71 17−114 S71 58 S61 previously classified as clade I members. ** Samples previously classified as clade II 18−30 S53 18−41 S45 members. Technical replicates are enclosed in red squares. 18−35B S39 18−35A S37 18−13 S27 25 Genetic distance (proportion of loci that are different) Pseudoperonospora Genetic distance (proportion of loci that are different) Genetic distance (proportion of loci that are different) Genetic distance (proportion of loci that are different) Genetic distance (proportion of loci that are different) Genetic distance (proportion of loci that are different) Genetic distance (proportion of loci that are different) humuli 0.02 100 100 100 82 97 82 97 100 100 100 83 83 74 100 100 100 C. melo C. melo C. melo 0.00 98 93 83 71 73 71 73 98 73 0.04 80 73 100 74 100 80 100 93 83 100 100 100 93 83 73 83 179 74 100 Genetic distance (proportion of loci that are different) Genetic distance (proportion of loci that are different) Genetic distance (proportion of loci that are different) 66 60 74 Genetic distance (proportion of loci that are different) 2683 SNPs Genetic distance (proportion of loci that are different) 2683 SNPs 0.06 Subpopulation Subpopulation Figure 4-5. Ordination plots of Pseudoperonospora spp. based on 2,978 SNPs. ((A) Ordination plot of Pseudoperonospora cubensis and P. humuli samples according to host species. All points represent samples collected in Michigan unless indicated otherwise. (B) Ordination plot of P. cubensis samples from Cucumis sativus from 2007 to 2017 in the U.S. (n=57). (C) Ordination plot of P. humuli samples from hop yards of Michigan in 2017 and 2018 (n=34) 180 A C D North North West Central B East N = 57 MLG = 40 West Central Midwest-w/o-Mi Figure 4-6. Frequency and geographic distribution of Pseudoperonospora cubensis and Pseudoperonospora humuli genotypes (MLG). (A) Geographic distribution of P. cubensis genotypes collected from Cucumis sativus in the Midwest. Unique genotypes are colored in gray. (B) Frequency of P. cubensis genotypes collected from C. sativus in the Midwest. Unique genotypes are colored in gray. (C) Geographic distribution of P. humuli genotypes collected from Humulus lupulus in Michigan. Unique genotypes are color coded in gray. The circles and stars colored in red represent 181 different hop yards. (D) Frequency of P. humuli genotypes collected from H. lupulus in Michigan. Unique genotypes are colored in gray. Figure 4-7. The relationship between genetic differences among samples and geographic distances of the locations from which samples originated. (A) P. cubensis samples from Cucumis sativus and (B) P. humuli samples from Humulus lupulus in Michigan. 182 e B 0.12 c e d A 0.125 0.100 a 0.075 0.050 d a c ) A i I ( n o i t a c o s s a f o x e d n I 0.025 0.000 b e g a k n i l _ o N s i s n e b u c . P e g a k n e g a k n e g a k n i l _ % 5 2 i l _ % 0 5 i l _ % 5 7 i n o i t a c o s s a f o x e d n I e g a k n i l _ % 0 0 1 0.08 a 0.04 0.00 i l u m u h . P b e g a k n i l _ o N a e g a k n e g a k n e g a k n i l _ % 5 2 i l _ % 0 5 i l _ % 5 7 e g a k n i l _ % 0 0 1 Data set Figure 4-8. Estimation of the degree of linkage disequilibrium within Pseudoperonospora cubensis and Pseudoperonospora humuli of Michigan. (A) Observed index of association (IA) distribution of P. cubensis samples compared with the IA distribution values of 0, 25, 50, 75, and 100% linkage. (B) Observed IA distribution of P. humuli samples compared with the IA distribution values of 0, 25, 50, 75, and 100% linkage. Groupings based on the Kruskal-Wallis rank-sum test are noted by the letters over the boxplots, in which the P. cubensis and the P. humuli population datasets are grouped with the simulated 50% and 25% linkage data, respectively Data set 183 SUPPLEMENTARY TABLES Supplementary Table 4-1: Samples sequenced using target enrichment (TE), whole genome sequencing (WGS) or low coverage whole genome sequencing (Lc-WGS) Hybr. Timea Enri. setb Log10 Depth State County / Yard Region Year Host Species Collector Sample 10D_S8 Input DNA (ng) 2 48 h ES3 1.9 Michigan Newaygo Eastern 2015 1335_S30 83 24 h ES2 1.6 Michigan St. Clair Eastern 2009 C. sativus C. moschata J. Bello L, Quesada- Ocampo, L, Quesada- Ocampo, 2009 C. sativus 2015 C. sativus J, Bello 2017 C. sativus J, Bello 2009 C. sativus L, Quesada- Ocampo, L, Quesada- Ocampo, 2017 H. lupulus D, Higgins 2017 H. lupulus D, Higgins 2009 C. sativus Leelanau (B) (A) Northern 2017 H. lupulus D, Higgins 1484_S9 14J_S6 156_S51 1589_S7 8 4 14 17 48 h ES3 1.75 Michigan Newaygo Western 48 h ES3 48 h ES3 2.08 Michigan Newaygo Western 1.49 Michigan Muskegon Western 48 h ES3 2.69 Indiana 1649_S2 54 48 h ES3 2.46 Iowa 17-100_S73 17-101_S74 6 5 48 h ES3 48 h ES3 2.05 Michigan Barry (D) Western 1.34 Michigan Northern Leelanau - - Midwest Midwest 17-103_S70 54 48 h ES3 2.52 Michigan aHybridization time. bEnrichment set (ES2 = 24h, ES3=48h) 184 Sample 17_108B_S41 17-112_S18 17-113_S72 17-114_S71 17-15_S33 17-20_S76 1752_S32 1755_S4 17-8_S50 180_S14 Input DNA (ng) 54 85 54 34 38 5 72 23 25 4 Hybr. Timea Enri. setb 48 h ES3 48 h ES3 State County / Log10 Depth Yard 1.28 Michigan Genesee Leelanau 2.08 Michigan (F) 48 h ES3 2.33 Michigan 48 h ES3 2.49 Michigan 48 h ES3 1.45 Michigan Leelanau Leelanau Leelanau (A) (A) (A) (B) 24 h ES2 1.86 Michigan Berrien (E) Western 2017 48 h ES3 1.96 Michigan Ingham Central 2009 48 h ES3 24 h ES2 48 h ES3 1.61 Michigan Ingham 1.45 Michigan Genesee 2.84 Michigan Muskegon Western Central Central (F) 2009 2017 2018 Region Year Host Species Collector Central 2017 Northern 2017 Northern 2017 Northern 2017 Northern 2017 H. H. H. H. H. lupulus D, Higgins lupulus D, Higgins lupulus D, Higgins lupulus D, Higgins lupulus D, Higgins lupulus D, Higgins L, Quesada- Ocampo, L, Quesada- Ocampo, sativus lupulus D, Higgins sativus H. C. H. C. C. sativus J, Bello Supplementary Table 4-1. (cont’d) aHybridization time. bEnrichment set (ES2 = 24h, ES3=48h) 185 Sample 18-13_S27 Input DNA (ng) 32 18-15_S60 136 18-17_S52 18-1_S69 18-20_S54 18-23_S37 33 45 26 56 18-25_S36 145 18-26_S55 18-30_S53 30 42 18-31_S65 205 48 h 48 h 48 h 48 h 48 h 24 h 24 h 48 h 48 h 48 h ES3 ES3 ES3 ES3 ES3 ES2 ES2 ES3 ES3 ES3 Grand 2.84 Michigan 2.72 Michigan Traverse (C) Northern Traverse (C) Northern Traverse (C) Northern Traverse (C) Northern 1.56 Michigan Genesee (F) Central 1.63 Michigan Michigan Grand Grand 3 2018 2017 2017 2018 2017 1.08 Michigan Berrien (E) Western 2018 1.43 Michigan Berrien (E) Western 2018 1.51 Michigan Barry (D) Western 2018 2.95 Michigan Leelanau (A) Northern 2018 2.28 Michigan Leelanau (A) Northern 2018 H. H. H. H. H. H. H. H. H. H. lupulus lupulus lupulus lupulus lupulus lupulus lupulus lupulus lupulus lupulus J, Bello D, Higgins D, Higgins J, Bello D, Higgins J, Bello J, Bello J, Bello J, Bello J, Bello Supplementary Table 4-1. (cont’d) Hybr. Timea Enri. setb Log10 Depth State Region Year Host Species Collector County / Yard Grand aHybridization time. bEnrichment set (ES2 = 24h, ES3=48h) 186 Supplementary Table 4-1. (cont’d) Sample 18-32_S64 18-35A_S37 18-35B_S39 18-41_S45 18-42_S42 18-43_S44 18-44_S59 18-5_S56 186_S3 18-6_S67 Input DNA (ng) 342 54 54 37 62 26 20 41 26 37 Hybr. Timea Enri. setb Log10 Depth State County / Yard Region Year Host Species Collector 48 h 48 h 48 h 48 h 48 h 48 h 48 h 48 h 48 h 48 h ES3 ES3 ES3 ES3 ES3 ES3 ES3 ES3 ES3 ES3 1.3 Michigan Barry (D) Western 2018 3.02 Michigan 3.04 Michigan 2.79 Michigan 2.97 Michigan 2.94 Michigan 2.99 Michigan 2.98 Michigan Leelanau Leelanau (A) (A) (A) (B) (B) (B) (A) Leelanau Leelanau Leelanau Leelanau Leelanau Northern 2018 Northern 2018 Northern 2018 Northern 2018 Northern 2018 Northern 2018 Northern 2018 2.39 Michigan Muskegon Western 2018 2.07 Michigan Berrien (E) Western 2018 lupulus lupulus lupulus lupulus lupulus lupulus lupulus lupulus H. H. H. H. H. H. H. H. C. H. sativus lupulus J, Bello J, Bello J, Bello J, Bello J, Bello J, Bello J, Bello J, Bello J, Bello J, Bello aHybridization time. bEnrichment set (ES2 = 24h, ES3=48h) 187 Supplementary Table 4-1. (cont’d) Input DNA (ng) 392 Hybr. Timea Enri. setb Log10 Depth State 48 h ES3 1.76 Michigan County / Yard Barry (D) Region Year Host Species Collector Western 2018 Sample 18-7_S66 200_S5 203_S1 204_S10 229_S8 238_S11 lupulus sativus sativus sativus sativus sativus H. C. C. C. C. C. C. a C. C. moschat sativus J, Bello J, Bello J, Bello J, Bello J, Bello J, Bello Thomas, A J, Bello J, Bello sativus lupulus Thomas, A H. 20 74 8 14 7 48 h ES3 2.94 Michigan Saginaw Eastern 2018 48 h ES3 1.71 Michigan Saginaw Eastern 2018 48 h ES3 2.59 Michigan Saginaw Eastern 2018 48 h ES3 2.29 Michigan Berrien Western 2018 48 h ES3 1.87 Michigan Allegan Western 2018 24 NA wgs wgs 2.33 Alabama - Alabama 2013 253_S49 257_S50 7 3 48 h ES3 2.8 Michigan Saginaw Eastern 2018 48 h ES3 2.12 Michigan Saginaw Eastern 2018 25 NA wgs wgs 2.11 Oregon - Oregon 2012 aHybridization time. bEnrichment set (ES2 = 24h, ES3=48h) 188 Supplementary Table 4-1. (cont’d) Sample 296_S12 Input DNA (ng) 5 37 NA Enri. Hybr. Timea setb 48 h ES3 wgs wgs State Log10 Depth 1.56 Michigan 2.38 County / Yard Ingham - - - - Central South Carolina Florida South Carolina North Carolina South Carolina Florida South Carolina North Carolina 38 NA wgs wgs 2.43 39 NA wgs wgs 2.3 40 NA wgs wgs 1.97 Region Year Host Species Collector C. C. moschata 2018 C. sativus J, Bello Thomas, 2013 A 2013 C. lanatus Thomas, A Thomas, 2012 A 2013 C. maxima Thomas, moschata moschata 2007 C. melo A L, Quesada- Ocampo, Thomas, A L, 2008 C. sativus Quesada- Ocampo, 2012 C. sativus Thomas, A Thomas, 2013 C. melo A 2008 C. 413_S36 26 48 h ES3 1.32 Michigan Clinton Eastern 41 NA wgs wgs 1.93 Georgia 42 NA wgs wgs 1.97 California 43 NA wgs wgs 44 NA wgs wgs 2.33 North Carolina 2.42 New York - - - - Georgia California Ncarolina New York aHybridization time. bEnrichment set (ES2 = 24h, ES3=48h) 189 Hybr. Timea Enri. setb Log10 Depth State County / Yard Region Year Host Species Collector 481_S13 5 48 h ES3 0.9 Indiana - Midwest 2007 C. sativus 58_S61 197 48 h ES3 3.07 Michigan Ingham (G) Central L, Quesada- Ocampo, 2017 H. lupulus D, Higgins L, Quesada- Ocampo, L, Quesada- Ocampo, L, Quesada- Ocampo, Midwest 2008 C. sativus Midwest 2008 C. sativus Midwest 2008 C. sativus 48 h 24 h 24 h 24 h 24 h ES3 ES2 ES2 ES2 ES2 2019 H. lupulus J, Bello 2.41 Michigan 2017 C. sativus J, Bello 1.2 Michigan 1.11 Michigan 2017 C. sativus J, Bello 1.18 Michigan Muskegon Western 2017 C. sativus J, Bello 1.52 Michigan Muskegon Western 2017 C. sativus J, Bello Ingham Central Ingham Central Ingham Central Supplementary Table 4-1. (cont’d) Sample Input DNA (ng) 48 h ES3 1.3 Ontario 48 h ES3 1.92 Ohio 48 h ES3 2.63 Ontario - - - 671_S34 898_S15 936_S35 A19_S21 CDM110_S48 CDM123_S7 CDM152_S30 CDM153_S29 7 2 7 67 16 10 8 12 aHybridization time. bEnrichment set (ES2 = 24h, ES3=48h) 190 Supplementary Table 4-1. (cont’d) Sample Input DNA (ng) Hybr. Timea Enri. setb Log10 Depth State County / Yard Region Year Host Species Collector L32_S26 10 48 h ES3 2.43 Michigan Monroe Eastern 2007 C. sativus L32_S62 10 24 h ES2 2.61 Michigan Monroe Eastern 2008 C. sativus L33_S3 31 24 h ES2 1.8 Michigan Monroe Eastern 2007 C. sativus L573_S20 5 24 h ES2 2.47 Florida - Florida 2008 C. sativus L673A_S4 20 24 h ES2 1.08 Ontario L682_S28 22 24 h ES2 NA Ontario L682_S63 11 24 h ES2 2.1 Ontario OH1-2_S20 S17-1_S17 55 100 48 h 48 h ES3 ES3 1.9 1.11 Ohio Indiana aHybridization time. bEnrichment set (ES2 = 24h, ES3=48h) - - - - - 191 L, Quesada- Ocampo, L, Quesada- Ocampo, L, Quesada- Ocampo, L, Quesada- Ocampo, L, Quesada- Ocampo, L, Quesada- Ocampo, L, Quesada- Ocampo, Midwest 2008 C. sativus Midwest 2008 C. sativus Midwest 2008 C. sativus Midwest Western 2017 C. sativus J, Bello moschata J, Bello 2017 C. Supplementary Table 4-1. (cont’d) Sample CDM232_S59 Input DNA (ng) 26 CDM242_S23 CDM250_S17 18 10 Hybr. Timea Enri. setb Log10 Depth State County / Yard Region Year Host Species Collector 24 h ES2 2.35 Michigan Allegan Western 2018 24 h ES2 3.4 Michigan Berrien Western 2018 24 h ES2 2.95 Michigan Berrien Western 2018 sativus sativus sativus C. C. C. C. C. H. sativus sativus lupulus C. sativus C. sativus C. sativus C. sativus J, Bello J, Bello J, Bello J, Bello J, Bello J, Bello L, Quesada- Ocampo, L, Quesada- Ocampo, L, Quesada- Ocampo, L, Quesada- Ocampo, CPF7_S24 162 48 h ES3 1.11 Indiana - Midwest 2017 K2A_S38 kk10-4_S75 L1024_S15 L1072_S21 32 15 6 8 48 h ES3 2.41 Michigan Bay Eastern 2016 48 h ES3 2.25 Michigan Leelanau (B) Northern 2018 24 h ES2 1 Indiana - Midwest 2008 24 h ES2 2.02 Michigan Monroe Eastern 2008 L119_S27 22 24 h ES2 1.74 Ohio - Midwest 2007 L1335_S16 2 48 h ES3 1.86 Michigan St. Clair Eastern 2009 aHybridization time. bEnrichment set (ES2 = 24h, ES3=48h) 192 Supplementary Table 4-1. (cont’d) Hybr. Timea Enri. setb Log10 Depth State County / Yard Region Year Host Species Collector sativus sativus sativus C. C. C. C. C. C. C. C. C. sativus sativus sativus sativus sativus sativus C. sativus C. sativus J, Bello J, Bello J, Bello J, Bello J, Bello J, Bello J, Bello J, Bello J, Bello L, Quesada- Ocampo, L, Quesada- Ocampo, Sample CDM154_S60 CDM155_S31 CDM191_S14 CDM19_S22 CDM201_S9 CDM202_S12 CDM207_S55 CDM209_S11 CDM228_S6 L1621_S13 Input DNA (ng) 44 32 24 83 52 56 92 28 25 40 24 h 24 h 24 h 24 h 24 h 24 h 24 h 24 h 24 h ES2 ES2 ES2 ES2 ES2 ES2 ES2 ES2 ES2 1.74 Michigan Muskegon Western 2017 1 Michigan Muskegon Western 2017 2.75 Michigan Muskegon Western 2018 1.85 Ontario - Midwest 2016 2.28 Michigan Saginaw Eastern 2018 2.67 Michigan Saginaw Eastern 2018 1.93 Michigan Berrien Western 2018 2.35 Michigan Berrien Western 2018 2.58 Michigan Berrien Western 2018 24 h ES2 2.16 Wisconsin - Midwest 2009 L1755_S64 22 24 h ES2 2.57 Michigan Ingham Central 2009 aHybridization time. bEnrichment set (ES2 = 24h, ES3=48h) 193 LITERATURE CITED 194 LITERATURE CITED Ali, S., Gautier, A., Leconte, M., Enjalbert, J., de Vallavieille-Pope, C. (2011) A rapid genotyping method for an obligate fungal pathogen, Puccinia striiformis f.sp. tritici, based on DNA extraction from infected leaf and multiplex PCR genotyping. BMC Res Notes 4: 240 Baxter, L., Tripathy, S., Ishaque, N., Boot, N., Cabral, A., Kemen, E., Thines, M., Ah- Fong, A., Anderson, R., Badejoko, W., et al (2010) Signatures of adaptation to obligate biotrophy in the Hyaloperonospora arabidopsidis genome. Science (80) 330: 1549 LP – 1551 Bello, J.C., Sakalidis, M.L., Perla, D.E., Hausbeck, M.K. (2020) Detection of airborne sporangia of Pseudoperonospora cubensis in Michigan using Burkard spore traps coupled to qPCR. Plant Dis. - Manuscript in review Blum, M., Waldner, M., Olaya, G., Cohen, Y., Gisi, U., Sierotzki, H. (2011) Resistance mechanism to carboxylic acid amide fungicides in the cucurbit downy mildew pathogen Pseudoperonospora cubensis. Pest Manag Sci 67: 1211–1214 Bolger, A.M., Lohse, M., Usadel, B. (2014) Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30: 2114–2120 Bourret, T.B., Choudhury, R.A., Mehl, H.K., Blomquist, C.L., McRoberts, N., Rizzo, D.M. (2018) Multiple origins of downy mildews and mito-nuclear discordance within the paraphyletic genus Phytophthora. PLoS One 13: e0192502 Brown, A.H.D., Feldman, M.W., Nevo, E. (1980) Multilocus structure of natural populations of Hordeum spontaneum. Genetics 96: 523-536. Burkhardt, A., Buchanan, A., Cumbie, J..S, Savory, E.A., Chang, J.H., Day, B. (2014) Alternative splicing in the obligate biotrophic oomycete pathogen Pseudoperonospora cubensis. Mol Plant-Microbe Interact 28: 298–309 Burkhardt, A., Day, B. (2016) Transcriptome and small RNAome dynamics during a resistant and susceptible interaction between cucumber and downy mildew. Plant Genome. doi: 10.3835/plantgenome2015.08.0069 Carleson, N.C., Fieland, V.J., Scagel, C.F., Weiland, J.E., Grünwald, N.J. (2019) Population structure of Phytophthora plurivora on rhododendron in Oregon nurseries. Plant Dis 103: 1923–1930 Chee, H.Y., Nelson, M.E., Grov,e G.G., Eastwell, K.C., Kenny, S.T., Klein, R.E. (2006) Population biology of Pseudoperonospora humuli in Oregon and Washington. Plant Dis 90: 1283–1286 195 Choi, Y-J., Hong, S-B., Shin, H-D. (2005) A re-consideration of Pseudoperonospora cubensis and P. humuli based on molecular and morphological data. Mycol Res 109: 841–848 Choi, Y-J., Thines, M. (2015) Host jumps and radiation, not co-divergence drives diversification of obligate pathogens. A case study in Downy mildews and Asteraceae. PLoS One 10: e0133655 Clark, M.J., Chen, R, Lam, H.Y.K., Karczewski, K.J., Chen, R., Euskirchen, G., Butte, A.J., Snyder, M. (2011) Performance comparison of exome DNA sequencing technologies. Nat Biotechnol 29: 908–914 Cruz-Dávalos, D.I., Llamas, B., Gaunitz, C., Fages, A., Gamba, C., Soubrier, J., Librado, P., Seguin-Orlando, A., Pruvost, M., Alfarhan, A.H., et al (2017) Experimental conditions improving in-solution target enrichment for ancient DNA. Mol Ecol Resour 17: 508–522 Cui, C., Herlihy, J.H., Bombarely, A., McDowell, J.M., Haak, D.C. (2019) Draft assembly of Phytophthora capsici from long-read sequencing uncovers complexity. Mol Plant- Microbe Interact 32: 1559–1563 Dray, S., Dufour, A. (2007) The ade4 package: implementing the duality diagram for ecologists. J Stat Softw 22(4): 1–20 Faircloth, B.C., Branstetter, M.G., White, N.D., Brady, S.G. (2015) Target enrichment of ultraconserved elements from arthropods provides a genomic perspective on relationships among Hymenoptera. Mol Ecol Resour 15: 489–501 Fletcher, K., Gil, J., Bertier, L.D., Kenefick, A., Wood, K.J., Zhang, L., Reyes-Chin-Wo, S., Cavanaugh, K., Tsuchida, C., Wong, J., et al (2019) Genomic signatures of heterokaryosis in the oomycete pathogen Bremia lactucae. Nat Commun 10: 2645 Gent, D.H., Adair, N., Knaus, B.J., Grünwald, N.J. (2019) Genotyping-by-sequencing reveals fine-scale population differentiation in populations of Pseudoperonospora humuli. Phytopathology. doi: 10.1094/PHYTO-12-18-0485-R Gent, D.H., Nelson, M.E., Farnsworth, J.L., Grove, G.G. (2009) PCR detection of Pseudoperonospora humuli in air samples from hop yards. Plant Pathol 58: 1081– 1091 Gent, D.H., Ocamb, C.M., Farnsworth, J.L. (2010) Forecasting and management of hop downy mildew. Plant Dis 94: 425–431 Gisi, U., Sierotzki, H. (2008) Fungicide modes of action and resistance in downy mildews. Eur J Plant Pathol 122: 157–167 196 Gnirke, A., Melnikov, A., Maguire, J., Rogov, P., LeProust., E. M., Brockman, W., Fennell, T., Giannoukos, G., Fisher, S., Russ, C., et al (2009) Solution hybrid selection with ultra-long oligonucleotides for massively parallel targeted sequencing. Nat Biotechnol 27: 182–189 Goldenhar, K.E., Hausbeck, M.K. (2019) Fungicides for control of downy mildew on pickling cucumber in Michigan. Plant Heal Prog 20: 165–169 Goudet, J. (2005) Hierfstat, a package for r to compute and test hierarchical F-statistics. Mol Ecol Notes 5: 184–186 Grünwald, N.J., Everhart, S.E., Knaus, B.J., Kamvar, Z.N. (2017) Best practices for population genetic analyses. Phytopathology 107: 1000–1010 Grünwald, N.J., McDonald, B.A., Milgroom, M.G. (2016) Population genomics of fungal and oomycete pathogens. Annu Rev Phytopathol 54: 323–346 Holdsworth, W.L., Summers, C.F., Glos, M., Smart, C.D., Mazourek, M. (2014) Development of downy mildew-resistant cucumbers for late-season production in the Northeastern United States. HortScience horts 49: 10–17 Johnson, D.A, Alldredge, J.R, Allen, J.R, Allwine, R. (1991) Spatial pattern of downy mildew in hop yards during severe and mild disease epidemics. Phytopathology 81: 1369–1374 Jombart, T., and Ahmed, I. (2011) adegenet 1.3-1: new tools for the analysis of genome-wide SNP data. Bioinformatics 27: 3070–3071 Jones, L., Riaz, S., Morales-Cruz, A., Amrine, K.C.H., McGuire, B., Gubler, W.D., Walker, M.A., Cantu, D. (2014) Adaptive genomic structural variation in the grape powdery mildew pathogen, Erysiphe necator. BMC Genomics 15: 1081 Jouet, A., Saunders, D.G.O., McMullan, M., Ward, B., Furzer, O., Jupe, F., Cevik, V., Hein, I., Thilliez, G.J.A., Holub, E., van Oosterhout, C. and Jones, J.D.G. (2019), Albugo candida race diversity, ploidy, and host-associated microbes revealed using DNA sequence capture on diseased plants in the field. New Phytol, 221: 1529-1543 Kamvar, Z.N., Brooks, J.C., Grünwald, N.J. (2015) Novel R tools for analysis of genome-wide population genetic data with emphasis on clonality. Front Genet 6: 208 Kamvar, Z.N., Tabima, J.F., Grünwald, N.J. (2014) Poppr: an R package for genetic analysis of populations with clonal, partially clonal, and/or sexual reproduction. PeerJ 2: e281 197 Kanetis, L., Vasiliou, A., Neophytou, G., Samouel, S., Tsaltas, D. (2013) First report of downy mildew caused by Peronospora belbahrii on sweet basil (Ocimum basilicum) in Cyprus. Plant Dis 98: 283 Kitner, M., Lebeda, A., Sharma, R., Runge, F., Dvořák, P., Tahir, A., Choi, Y-J., Sedláková, B., Thines, M. (2015) Coincidence of virulence shifts and population genetic changes of Pseudoperonospora cubensis in the Czech Republic. Plant Pathol 64: 1461–1470 Knaus, B.J., Grünwald, N.J. (2017) vcfr: a package to manipulate and visualize variant call format data in R. Mol Ecol Resour 17: 44–53 Kozarewa, I., Armisen, J., Gardner, A.F., Slatko, B.E., Hendrickson, C.L. (2015) Overview of target enrichment strategies. Curr Protoc Mol Biol 112: 7.21.1-7.21.23 Kunjeti, S.G., Anchieta, A., Martin, F.N., Choi, Y-J., Thines, M., Michelmore, R.W., Koike, S.T., Tsuchida, C., Mahaffee, W., Subbarao, K.V., et al (2016) Detection and quantification of Bremia lactucae by spore trapping and quantitative PCR. Phytopathology 106: 1426–1437 Lane, C.R., Beales, P.A., O’Neill, T.M., McPherson, G.M., Finlay, A.R., David, J., Constantinescu, O., Henricot, B. (2005) First report of Impatiens downy mildew (Plasmopara obducens) in the UK. Plant Pathol 54: 243 Li, H., Durbin, R. (2009) Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics. doi: 10.1093/bioinformatics/btp324 Li, H., Handsaker, B., Wysoker, A., Fennell, T., Ruan J., Homer, N., Marth, G., Abecasis, G., Durbin, R., Subgroup 1000 genome project data processing (2009) The sequence alignment/map format and SAMtools. Bioinformatics 25: 2078–2079 Lim, H.C., Braun, M.J. (2016) High-throughput SNP genotyping of historical and modern samples of five bird species via sequence capture of ultraconserved elements. Mol Ecol Resour 16: 1204–1223 Ma, X., Shao, Y., Tian, L., Flasch, D.A., Mulder, H.L., Edmonson, M.N., Liu, Y., Chen, X., Newman, S., Nakitandwe, J., et al (2019) Analysis of error profiles in deep next- generation sequencing data. Genome Biol 20: 50 Mantel, N. (1967) The detection of disease clustering and a generalized regression approach. Cancer Res 27: 209–220 McCormack, J.E., Tsai, W.L.E., Faircloth, B.C. (2016) Sequence capture of ultraconserved elements from bird museum specimens. Mol Ecol Resour 16: 1189– 1203 198 Milgroom, M.G. (2015) Population biology of plant pathogens : genetics, ecology, and evolution. The American Phytopathological Society, St. Paul, Minnesota Milgroom, M.G. (1996) Recombination and the multilocus structure of fungal populations. Annu Rev Phytopathol 34: 457–477 Mitchell, M.N., Ocamb, C.M., Grünwald, N.J., Mancino, L.E., Gent, D.H. (2011) Genetic and pathogenic relatedness of Pseudoperonospora cubensis and P. humuli. Phytopathology 101: 805–818 Naegele, R.P., Quesada-Ocampo, L.M., Kurjan, J.D., Saude, C., Hausbeck, M.K. (2016) Regional and temporal population structure of Pseudoperonospora cubensis in Michigan and Ontario. Phytopathology 106: 372–379 Nei, M (1972) Genetic distance between populations. Am Nat 106: 283–292 Neufeld, K. N. (2017) Epidemiology of cucurbit downy mildew: predicting risk of disease infection and validation of the CDM ipmPIPE forecasting system. (Doctoral dissertation thesis, North Carolina State University, Raleigh, USA). Retrieved from https://repository.lib.ncsu.edu/bitstream/handle/1840.20/33592/etd.pdf?sequence=1&is Allowed=y Perea, C., De La Hoz, J.F., Cruz, D.F., Lobaton, J.D., Izquierdo, P., Quintero, J.C., Raatz, B., Duitama, J. (2016) Bioinformatic analysis of genotype by sequencing (GBS) data with NGSEP. BMC Genomics 17: 498 Potapov, V., Ong, J.L. (2017) Examining sources of error in PCR by single-molecule sequencing. PLoS One 12: e0169774 Purayannur, S., Cano, L.M., Bowman, M.J., Childs, K.L., Gent, D.H., Quesada-Ocampo, L.M. (2020) The effector repertoire of the hop downy mildew pathogen Pseudoperonospora humuli. Front Genet 11: 910 Quesada-Ocampo, L.M., Granke, L.L., Olsen, J., Gutting, H.C., Runge, F., Thines, M., Lebeda, A., Hausbeck, M.K. (2012) The genetic structure of Pseudoperonospora cubensis populations. Plant Dis 96: 1459–1470 Quinlan, AR, Hall, IM (2010) BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26: 841–842 Rahman, A., Góngora-Castillo, E., Bowman, M.J., Childs, K.L., Gent, D.H., Martin, F.N., Quesada-Ocampo, L.M. (2019) Genome sequencing and transcriptome analysis of the hop downy mildew pathogen Pseudoperonospora humuli reveal species-specific genes for molecular detection. Phytopathology 109: 1354–1366 199 Rivera, Y., Salgado-Salazar, C., Gulya, T.J., Crouch, J.A. (2016) Newly emerged populations of Plasmopara halstedii infecting rudbeckia exhibit unique genotypic profiles and are distinct from sunflower-infecting strains. Phytopathology 106: 752–761 Savory, E.A., Adhikari, B.N., Hamilton, J.P., Vaillancourt, B., Buell, C.R., Day, B. (2012a) mRNA-Seq Analysis of the Pseudoperonospora cubensis transcriptome during cucumber (Cucumis sativus L.) infection. PLoS One 7: e35796 Savory, E.A., Zou, C., Adhikari, B.N., Hamilton, J.P., Buell, C.R., Shiu, S-H., Day, B. (2012b) 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 Simpson, E. H. (1949) Measurement of diversity. Nature 163: 688 Sirrine, R., Lizotte, E., Brown, D., O’Brien, T., and Leach, A. 2014. Estimated costs of producing hops in Michigan. Michigan State University Extension E3236. Online publication. Shannon, C. E. (1948). A mathematical theory of communication. Bell System Technical Journal 27:379-423 Sharma, R., Xia, X., Cano, L.M., Evangelisti, E., Kemen, E., Judelson, H., Oome, S., Sambles, C., van den Hoogen, D.J., Kitner, M., et al (2015) Genome analyses of the sunflower pathogen Plasmopara halstedii provide insights into effector evolution in downy mildews and Phytophthora. BMC Genomics 16: 741 Starrett, J., Derkarabetian, S., Hedin, M., Bryson, Jr. R.W., McCormack, J.E., Faircloth, B.C. (2017) High phylogenetic utility of an ultraconserved element probe set designed for Arachnida. Mol Ecol Resour 17: 812–823 Stassen, J., Seidl, M., Vergeer, P., Nijman, I., Snel, B., Cuppen, E., Van den ackerveken, G. (2012) Effector identification in the lettuce downy mildew Bremia lactucae by massively parallel transcriptome sequencing. Mol Plant Pathol 13: 719–731 Stoddart, J.A., Taylor, J.F. (1988) Genotypic diversity: estimation and prediction in samples. Genetics 118: 705 LP – 711 Summers C.F., Gulliford C.M., Carlson C.H., Lillis J.A., Carlson M.O., Cadle-Davidson L., Gent, D.H., Smart, C.D. (2015) Identification of genetic variation between obligate plant pathogens Pseudoperonospora cubensis and P. humuli using RNA sequencing and genotyping-by-sequencing. PLoS One 10: e0143665 Tabima, J.F., Coffey, M.D., Zazada, I.A., Grünwald, N.J. (2018) Populations of Phytophthora rubi show little differentiation and high rates of migration among states in the western United States. Mol Plant-Microbe Interact 31: 614–622 200 Thines, M. (2014) Phylogeny and evolution of plant pathogenic oomycetes—a global overview. Eur J Plant Pathol 138: 431–447 Thilliez, G.J.A., Armstrong, M.R., Lim, T.-Y., Baker, K., Jouet, A., Ward, B., van Oosterhout, C., Jones, J.D.G., Huitema, E., Birch, P.R.J. and Hein, I. (2019), Pathogen enrichment sequencing (PenSeq) enables population genomic studies in oomycetes. New Phytol, 221: 1634.-1648 Thomas, A., Carbone, I., Choe, K., Quesada-Ocampo, L.M., Ojiambo, P.S. (2017) Resurgence of cucurbit downy mildew in the United States: Insights from comparative genomic analysis of Pseudoperonospora cubensis. Ecol Evol 7: 6231–6246 USDA (2020) United States Department of Agriculture. 2019. Vegetables Annual Summary. Wallace, E.C., D’Arcangelo, K.N., Quesada-Ocampo, L.M. (2020) Population analyses reveal two host-adapted clades of Pseudoperonospora cubensis, the causal agent of cucurbit downy mildew, on commercial and wild cucurbits. Phytopathology. doi: 10.1094/PHYTO-01-20-0009-R Wallace, E.C., Quesada-Ocampo, L.M. (2017) Analysis of microsatellites from the transcriptome of downy mildew pathogens and their application for characterization of Pseudoperonospora populations. PeerJ 5: e3266 Withers, S., Gongora-Castillo, E., Gent, D., Thomas, A., Ojiambo, P.S., Quesada- Ocampo, L.M. (2016) Using next-generation sequencing to develop molecular diagnostics for Pseudoperonospora cubensis, the cucurbit downy mildew pathogen. Phytopathology 106: 1105–1116 Wong, F.P, and Wilcox, W.F (2001) Comparative physical modes of action of azoxystrobin, mancozeb, and metalaxyl against Plasmopara viticola (Grapevine Downy Mildew). Plant Dis 85: 649–656 201 CHAPTER V 202 CONCLUSIONS The research reported in this dissertation provides an assessment of the population structure of Pseudoperonospora cubensis, the causal agent of cucurbit downy mildew (CDM) in Michigan and evaluates the performance of spore traps coupled to qPCR for the monitoring of airborne Pseudoperonospora spp. sporangia. The population structure of P. cubensis was investigated using a target enrichment protocol that allowed the genotyping of environmental samples with low concentrations of a mix of plant and pathogen DNA. A significant effect of the host type on the population structure of P. cubensis was observed while no evidence of location-based differentiation was detected within the P. cubensis population of Michigan. In addition, this study identified an improved detection system for the monitoring of P. cubensis sporangia that allowed the differentiation between Pseudoperonospora spp. and the detection of P. cubensis DNA before symptoms were observed in commercial cucumber fields. This study provided evidence of significant genetic differentiation among the P. cubensis population from squash (clade I) and cucumber (clade II) but there was insufficient evidence to conclude that location (region) within Michigan has a significant effect on the distribution of the genetic variation of the P. cubensis population. Contrasting evidence was found for the population of Pseudoperonospora humuli (the causal agent of hop downy mildew), and a significant effect of location on the genetic variation of the population was detected in the state. The differences in the distribution of genetic variation between the population of each species could be explained by differences in the dispersion patterns between them. Although both pathogens 203 propagate via asexual sporangia that are aerially dispersed, empirical evidence indicates that only P. cubensis sporangia may spread unrestricted over long distances. The isolation and divergence of geographically distant populations of Michigan may be limited by the exchange of migrants between them. This was supported by the detection of the same multilocus genotypes of P. cubensis in multiple geographically distant populations. On the other hand, despite the potential for airborne dispersal of hop downy mildew, only one MLG was detected in two geographically distant populations which is consistent with the restricted pattern of dispersion suggested for this pathogen. This was also supported by a significant correlation between the genetic and geographic distance among P. humuli samples, that was not detected for the P. cubensis population. In Michigan, CDM occurs annually due to an influx of aerially dispersed P. cubensis sporangia from overwinter sources. Thus, timely alerts of an influx of the airborne inoculum of P. cubensis can assist Michigan growers in assessing the need to initiate fungicide sprays. In the current study, we reported the use of a highly specific and sensitive qPCR assay that allows the differentiation between P. humuli, and each host-adapted clade of P. cubensis (clade I and II) on spore trap samples. A distinction that was not possible using light microscopy only. After two years of monitoring using a Burkard and impaction spore traps coupled with qPCR in cucumber fields, P. humuli DNA was detected more frequently than P. cubensis early during the growing season from May to June. P. cubensis clade II DNA was detected in spore trap samples approximately 2-7 days before CDM symptoms were observed in commercial cucumber fields in July or August, while P. cubensis clade I DNA was never detected. The 204 differences in the airborne inoculum of each clade documented here are likely the result of differences in the total area planted of the susceptible hosts to each clade in Michigan. In addition, this study sheds light on the utilization of Burkard and impaction spore traps for the airborne monitoring or Pseudoperonospora spp. sporangia. In agreement with theoretical expectations, our results suggest that the Burkard spore traps are a more efficient instrument for the detection of airborne sporangia at low concentrations (<100 sporangia/day) than impaction spore traps. Adjustments can be made to increase the efficiency of the detection of P. cubensis using impaction spore traps. This includes the utilization of multiple impaction spore traps per location and increasing the sampling surface width and the exposition time of the impaction rods. The use of spore traps couple with qPCR could be used as part of a CDM risk advisory system to time fungicide applications that protect cucurbit crops in Michigan. 205