<<

Studies in the Management of Seed and Root Rot of : Efficacy of Seed Treatments, Screening Germplasm for Resistance, and Comparison of Quantitative Disease Resistance Loci to Three Species of Pythium and sojae

Thesis

Presented in Partial Fulfillment of the Requirements for the Degree Master of Science in

the Graduate School of The Ohio State University

By

Kelsey Lynn Scott

Graduate Program in Pathology

The Ohio State University

2018

Thesis Committee:

Dr. Anne Dorrance, Advisor

Dr. Jason Slot

Dr. Leah McHale

Dr. Melanie Lewis Ivey

1

Copyrighted by

Kelsey Lynn Scott

2018

2

Abstract

In Ohio, soybean seedling damping-off and seed rot are problems routinely encountered soon after planting. Reduced tillage systems that lead to inoculum build-up combined with saturated soil conditions are ideal environments for seedling diseases, which cause large losses of soybean stand and thus yield. Prior Ohio field surveys identified multiple species of Pythium and Phytophthora that contribute to soybean seedling damping-off. Among the most common and aggressive species are

Phytophthora sojae, , var. ultimum, and Pythium ultimum var. sporangiiferum. Fungicide seed treatment and host resistance are two management strategies that are used to minimize yield loss caused by these pathogens.

Thus, the objectives of these studies were to: i) evaluate new active ingredients for efficacy in the lab and field, and ii) identify and characterize new sources of resistance towards the most common seedling pathogens. These are key strategies for the development of effective strategies for the management of soybean seedling disease.

During 2014-2015, at two environments, ethaboxam seed treatments combined with metalaxyl on a susceptible cultivar significantly increased yield compared to other fungicide treatments containing metalaxyl or mefenoxam alone. treated with ethaboxam plus metalaxyl had significantly higher plant populations when compared to the nontreated control at all four 2016 field locations, while one environment had

iii significantly higher yield. In laboratory seed plate and greenhouse cup assays, ethaboxam plus metalaxyl in a commercial formulation provided equal or better protection against multiple species of Pythium when compared with other seed treatments that contained metalaxyl or mefenoxam only. These results indicate that ethaboxam with metalaxyl is effective at managing seed and rot root caused by the diverse species of Pythium and Phytophthora and provides another seed treatment fungicide available to producers which can be used in an integrated disease management program.

The parents that were used to develop six nested association mapping (NAM) populations were previously identified as segregating for resistance towards Phytophthora sojae, Pythium irregulare, Pythium ultimum var. ultimum, and Pythium ultimum var. sporangiiferum. Following inoculation in a cup assay, the resistance was quantitatively inherited in each of the NAM populations towards the four seedling pathogens. In total, 33 QDRL from the six populations surpassed the genome- wide logarithm of odds (LOD) threshold and there was a large number of suggestive

QDRL that surpassed the chromosomal LOD threshold. Of these 33 significant QDRL,

10 explained more than 15% of the phenotypic variation. Only four QDRL conferred resistance to more than one of the pathogens; one on chromosome 3, one on chromosome 17, and two located at separate locations on chromosome 13. This indicates that there may be multiple mechanisms for resistance to these root pathogens. Further analyses are needed to precisely map these QDRL so they may be selectively bred into highly resistant germplasm in order to manage seed and seedling damping-off. These

iv

NAM populations will serve as a rich resource for breeders to incorporate resistance into adapted soybean cultivars.

v

I would like to dedicate this thesis to my family, my friends, my Andy, and also Diego.

vi

Acknowledgements

I would like to thank and acknowledge Anne Dorrance for her constant support and direction through this entire process, from planting beans in the field to finishing this thesis. Thank you, Jason, Leah, and Melanie, for your knowledge, guidance, and time. A big thank-you to Valent, the Center for Applied Plant Sciences, the Soybean Center, The

Ohio Soybean Council and the United Soybean Board for making this research possible. I would also like to thank The Ohio State University Department for providing such an amazing space to learn and grow. Thank you, Christine Balk, who performed the research that is the basis of my own QTL study. Thanks Clifton, who organized the prior field research and was a fantastic mentor when I was an intern here at

OARDC. Many thanks to all members of the Dorrance lab who inspired me daily with their drive and dedication. I am so grateful to those that helped with this project, especially Deloris for the lab work, Jonell for the field and greenhouse work, and the small army of interns for everything in between. Thank you Jonell, Amilcar, Linda, and

Meredith for our fun times together and your friendship- my memories of you will always hold a special place in my heart. Most of all, I thank Andy for everything.

vii

Vita

November 18, 1993……………………………………………………..Born- Elba, NY

2011………………………………………………………………….Elba Central School

2015……………………………………………………….B.S. Biology, SUNY Geneseo

2018…………………………..………..M.S. Plant Pathology, The Ohio State University

Publications

Scott, K., Eyre, M. and Dorrance, A. 2016. Screening soybean germplasm for resistance towards Pythium species. Phytopathology. 106:S12:90.

Scott, K., Vargas, A., Eyre, M. and Dorrance, A. 2016. Efficacy of three soybean fungicide seed treatments against Pythium species in seed plate and growth chamber assays. Phytopathology. 106:S12:68.

Fields of Study

Major Field: Plant Pathology

viii

Table of Contents

Abstract ...... iii Acknowledgements ...... vii Vita ...... viii List of Tables ...... xi List of Figures ...... xix Chapter 1. Literature Review ...... 1 Significance of soybean: ...... 1 Pathogens: Pythium, Phytophthora, and Phytopythium ...... 3 Pythium seed and root rot of soybean ...... 8 Disease management strategies ...... 10 Chapter 2: The efficacy of ethaboxam as a component in soybean seed treatments towards in Ohio ...... 19 Introduction ...... 19 Materials and Methods ...... 24 Field experiments ...... 24 Pythium, Phytopythium, and Phytophthora isolates ...... 26 Seed plate assay to evaluate seed treatments ...... 27 Cup assay to evaluate seed treatments ...... 29 Results ...... 31 Field trials ...... 31 Seed plate assay ...... 34 Cup assay ...... 35 Discussion ...... 37

ix

Chapter 3: Identifying the Quantitative Disease Resistance Loci in Nested Association Mapping populations towards Phytophthora sojae and three different species of Pythium ...... 76 Introduction ...... 76 Materials and methods ...... 80 NAM background and seed increase ...... 80 Assays and inoculum production to determine resistance phenotype ...... 80 QDRL Identification ...... 86 Results ...... 87 QDRL to Ph. sojae in four NAM RIL populations: ...... 87 QDRL to Py. irregulare in three RIL populations ...... 89 QDRL to Py. ultimum var. ultimum in three RIL populations: ...... 95 QDRL to Py. ultimum var. sporangiiferum in two RIL populations: ...... 101 Summary of QDRL for Ph. sojae ...... 105 Summary of QDRL for Py. irregulare ...... 106 Summary of QDRL for Py. ultimum var. ultimum...... 107 Summary of QDRL for Py. ultimum var. sporangiiferum ...... 108 Discussion ...... 109 References ...... 148

x

List of Tables

Table 2.1. Planting and harvest dates, precipitation, seeding rate, total row length counted for stand counts, and soil type for the field trials used to evaluate fungicide seed treatments in Ohio during 2014-2016. Boxes with a “-“ indicate there was no irrigation at that field site...... 43

Table 2.2. Fungicide seed treatments evaluated in each field environment. Select were also screened in the laboratory seed plate assay and the growth chamber cup assay...... 44

Table 2.3. List of cultivars used in each field site and year in order to determine the effect of cultivar choice on management of seed and root pathogens in Ohio fields...... 46

Table 2.4. List of Pythium (Py.), Phytophthora (Ph.), and Phytopythium (Pp.) isolates included in the plate assay used to evaluate the efficacy of three fungicide seed treatments and determine the seed-rotting ability of each isolate...... 47

Table 2.5. Isolates used in the growth chamber cup assay to evaluate the efficacy of fungicide seed treatments and determine the root-rotting ability of each isolate...... 49

Table 2.6. P-values for cultivar (cult.), fungicide seed treatment (fung.), and cultivar by fungicide interactions (CxF) for stand at stage V1/V2 and stand at stage V3/V4 data

(/Ha) for the 2014-2015 locations. Degrees of freedom (df) for cultivar (cult.), fungicide seed treatment (fung.), and cultivar by fungicide interactions (CxF) are shown.

xi

Inches of precipitation (natural and irrigated combined) in the 14 days after planting

(dap) are shown for each location (Loc.)...... 51

Table 2.7. Significance of cultivar (cult.), fungicide seed treatment (fung.), and cultivar by fungicide interactions (CxF) for plant stand at stage R8 (plants/Ha) and yield data

(Bu/A) for the 2014-2015 field locations...... 52

Table 2.8. P-values for cultivar (cult.), fungicide seed treatment (fung.), and cultivar by fungicide interactions (CxF) for stand at stage V1/V2 and stand at stage V3/V4 data

(plants/Ha) for the 2016 locations. Degrees of freedom (df) for cultivar (cult.), fungicide seed treatment (fung.), and cultivar by fungicide interactions (CxF) are shown. Inches of precipitation (natural and irrigated combined) in the 14 days after planting (dap) are shown for each location (Loc.)...... 53

Table 2.9. Significance of cultivar (cult.), fungicide seed treatment (fung.), and cultivar by fungicide interactions (CxF) for plant stand at stage R8 (plants/Ha) and yield data

(Bu/A) for the 2016 field locations...... 54

Table 2.10. Significance of cultivar, fungicide seed treatment, location, and interactions for stand at growth stages V1/V2, V3/V4, R8, and yield (Bu/A) for the 2016 locations. 55

Table 2.11. Comparison of different combinations on soybean population and yield in

Northwest Branch 2014 (Cv. Sloan). Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at each growth stage (plants/Ha), as well as yield (Kg/H and Bu/A). Treatments are compared to each other separately for each growth stage as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05)...... 56

xii

Table 2.12. Comparison of different fungicide combinations on soybean population and yield in Northwest Branch 2014 (Cv. Conrad). Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth stages as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05)...... 57

Table 2.13. Comparison of different fungicide combinations on soybean population and yield in Van Wert 2014 (Cv. Sloan). Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth stages as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05)...... 58

Table 2.14. Comparison of different fungicide active ingredients on soybean population and yield in Van Wert 2014 (Cv. Conrad). Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth stages as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05)...... 59

Table 2.15. Comparison of different fungicide active ingredients on soybean population and yield in Defiance 2015 (Cv. Conrad). Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to

xiii each other separately for growth stages as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05)...... 60

Table 2.16. Comparison of different fungicide active ingredients on soybean population and yield in Defiance 2015 (Cv. Sloan). Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth stages as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05)...... 61

Table 2.17. Comparison of different fungicide active ingredients on soybean population and yield in Defiance 2015 (Cv. Kottman). Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth stages as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05)...... 62

Table 2.18. Comparison of different fungicide active ingredients on soybean population and yield in Defiance 2015 (Cv. Lorain). Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth stages as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05)...... 63

Table 2.19. Comparison of different fungicide active ingredients on soybean population and yield in Northwest Branch 2015 (Cv. Conrad). Active ingredient rates are estimates

xiv based on label information. Overall mean across all treatments are shown for plant population at each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth stages as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05).

...... 64

Table 2.20. Comparison of different fungicide active ingredients on soybean population and yield in Northwest Branch 2015 (Cv. Kottman). Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth stages as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05).

...... 65

Table 2.21. Comparison of different fungicide active ingredients on soybean population and yield in Northwest Branch 2015 (Cv. Lorain). Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth stages as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05).

...... 66

Table 3.1. The six Nested Association Mapping (NAM) RIL populations evaluated for resistance towards Phytophthora sojae in a tray test and Pythium irregulare, Py. ultimum var. ultimum, and Py. ultimum var. sporangiiferum. Cells marked with an “X” indicate

xv that phenotypic screens and QTL analysis were carried out for the NAM RIL population x pathogen species combination. Data was collected by Balk (2014)...... 116

Table 3.2. Complete list of all QDRL, arranged by chromosome and position, for Ph. sojae (Phs), Py. irregulare (Pirr), Py. ultimum var. ultimum (Puu), and Py. ultimum var. sporangiiferum (Pus) identified by composite interval mapping (CIM) using F5 RILs in separate SoyNAM populations. Traits shown are total dry root weight (DRW), adjusted root weight (ARW), percent (%G), root rot score (RRS), and mean lesion length (MLL). Major QDRL (>15.0 PVE) are indicated in bold. Suggestive QTL are indicated in italics. QDRL that did not have a marker at the identified peak are marked with a dash. Marker range is the position in bp of the neighboring markers; some QDRL did not have both a right and a left marker...... 117

Table 3.3. Comparison of QTL for resistance towards oomycete pathogens Phytophthora sojae, Pythium irregulare, Py. ultimum var. ultimum, and Py. ultimum var. sporangiiferm in the six NAM populations generated by crossing IA3023 with 4J105-3-4, HS6-976,

LD02-9050, S06-13640, LG05-4832, and LG00-3372. Suggestive QDRL (significant at the chromosome LOD threshold) are indicated in bold. QDRL that overlap in position are noted by having a shared letter. Major QDRL (explanation of >15.0% phenotypic variance) are indicated in bold. Cells containing a “X” indicate that no major or minor

QDRL detected for the indicated NAM RIL population x pathogen combination. Cells containing “--” indicate that the NAM RIL population x pathogen combination was not screened with phenotype assays...... 131

xvi

Table 3.4. Comparison of major and minor QDRL significant at the genome-wide threshold mapped in the NAM RIL population generated from the cross of IA3023 x

LD02-9050. Traits shown are adjusted root weight (ARW), percent germination (%G), root rot score (RRS), and mean lesion length (MLL). Major QTL (>15.0 PVE) are indicated in bold...... 133

Table 3.5. Comparison of major and minor QDRL for resistance towards Ph. sojae (Phs),

Py. ultimum var. ultimum (Puu), Py. irregulare (Pirr), and Py. ultimum var. sporangiferum (Pus) significant at the genome-wide threshold mapped in the NAM F5

RIL population derived from the cross of HS6-3976 x IA3023. Traits shown are total dry root weight (DRW), adjusted root weight (ARW), root rot score (RRS), and mean lesion length (MLL). Major QTL (>15.0 PVE) are indicated in bold...... 135

Table 3.6. Comparison of QDRL to Py. irregulare significant at the genome-wide threshold mapped in the NAM RIL population generated from the cross of IA3023 x

LG00-3372. Traits shown are total dry root weight (DRW), adjusted root weight (ARW), and percent germination (%G)...... 139

Table 3.7. Comparison of major and minor QDRL for Py. ultimum var. ultimum which were significant at the genome-wide threshold mapped in the NAM F5 RIL population derived from the cross of IA3023 x 4J105-3-4. Traits shown are total dry root weight

(DRW), percent germination (%G), and root rot score (RRS). Major QDRL (>15.0 PVE) are indicated in bold...... 141

Table 3.8. Comparison of major QDRL to Py. ultimum var. ultimum significant at the genome-wide threshold mapped in the NAM RIL population generated from the cross of

xvii

IA3023 x S06-13640. Traits shown are percent germination (%G) and root rot score

(RRS). Major QDRL (>15.0 PVE) are indicated in bold...... 144

Table 3.9. Comparison of major and minor QDRL to Py. ultimum var. sporangiiferum significant at the genome-wide threshold mapped in the NAM RIL population generated from the cross of IA3023 x LG05-4832. Traits shown are total dry root weight (DRW), percent germination (%G), and root rot score (RRS). Major QTL (>15.0 PVE) are indicated in bold...... 147

xviii

List of Figures

Figure 2.1. Root rot scoring system used in the growth chamber cup to evaluate the pathogenicity and aggressiveness of isolates of Pythium recovered from soybean in Ohio.

For this score 1 = healthy roots with no infection, 2 = lesions covering 1-25% of roots, 3

= 26-75% of roots with lesions, 4 = 76-100% of roots with lesions, and 5 = total colonization of the seed with no germination...... 50

Figure 2.2. Early stand at stage V1/V2 (plants/Ha) of the seed planted at Northwest

Branch in 2016, with standard deviation shown. Bars sharing a letter are not significantly different from each other (p<0.05)...... 67

Figure 2.3. Yields (Bu/A) of the seed planted at Northwest Branch in 2016, with standard deviation shown. Bars sharing a letter are not significantly different from each other (p<0.05)...... 68

Figure 2.4. The relative pathogenicity of each Pythium (Py.), Phytophthora (Ph.), and

Phytopythium (Pp.) species evaluated in the seed plate assay, with standard error shown.

Results were calculated from the seed rot scores on the inoculated plates with nontreated seed (cv. Kottman). Bars sharing a letter are not significantly different from each other

(p<0.05). The higher the relative marginal effect, the greater the degree of seed rot...... 69

Figure 2.5. The response of the 27 isolates representing 14 oomycete species (n = number of isolates per the species) of Pythium and Phytopythium to Intego Suite,

xix

Acceleron, and CruiserMaxx Advanced fungicide seed treatments applied at the commercial rate. All the oomycete species were analyzed together and no within species differences were found within this group; data is separated into four separate graphs for visual clarity. Bars sharing a letter are not significantly different from one another

(p<0.05). Standard error is shown. The larger the relative marginal effect, the greater the degree of seed rot...... 70

Figure 2.6. The response of the six isolates representing two Phytopthora species (n = number of isolates per the species) to Intego Suite, Acceleron, and CruiserMaxx

Advanced fungicide seed treatments applied at the commercial rate. All oomycete spp. were analyzed together and no within species differences were found within this group;

Phytophthora data is shown separately for visual clarity. Bars sharing a letter are not significantly different from one another (p<0.05). Standard error is shown ...... 71

Figure 2.7. Oomycete species that exhibited a differential response when screened with the fungicide seed treatments. Seed treatments were compared to the nontreated control seed within each species. Bars with the same letter are not significantly different

(p=0.05). The larger the relative marginal effect, the greater the degree of seed rot...... 72

Figure 2.8. Response of each of the six Py. ultimum var. ultimum isolates to the three fungicide seed treatments used in the plate assay, with standard error shown. Within each isolate, the level of seed rot on the seed treatments was compared with the seed rot on the nontreated control seeds. Bars with the same letter are not significantly different

(p=0.05). The larger the relative marginal effect, the greater the degree of seed rot...... 73

xx

Figure 2.9. Relative pathogenicity of isolates of Pythium (Py.), Phytopythium (Pp.), and

Phytophthora (Ph.) tested in a growth chamber cup assay with surface sterilized soybean seed of the cultivar Kottman, with standard deviation shown. Results are based on the average adjusted root weights of the plants. Bars sharing a letter are not significantly different from each other. There was a significantly difference in the pathogenicity of the oomycete species (P=<0.0001). The lower the adjusted root weight, the higher the pathogenicity of the isolate...... 74

Figure 2.10. The response of the screened oomycete species to fungicide seed treatments in the cup assay, with standard deviation shown. The fungicide treated seed for each species was compared to the nontreated control within each species. Bars sharing a letter are not significantly different from one another (p=0.05). The lower the adjusted root weight, the more the roots were rotted by the oomycete pathogen...... 75

Figure 3.1. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks mean lesion length for RIL populations following inoculation with Ph. sojae in a tray assay. Shown are (A)

HS6-3976 x IA3023 and (B) IA3023 x LD02-9050 (isolate 371a92_Windfall_Ind) and populations (C) IA3023 x LG05-4832 and (D) IA3023 x LG00-3372 (isolate

(2)2_Dayton_739_LA_02). BLUE values not indicated on the graphs are (B) Conrad

(5.9), (C) Conrad (7.0), and (D) Resnik (3.6) and Conrad (7.0). BLUP values were inverted; lower values indicate more susceptible (S) lines and higher BLUP values indicate lines with greater resistance (R) to the pathogen ...... 132

xxi

Figure 3.2. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross HS6-3976 x IA3023 following inoculation with Py. irregulare in a greenhouse cup assay. BLUE values not indicated on the graphs are (A) Clermont (-12.5) and Sloan (-14.6), (B) Clermont (-0.58), Sloan (-

0.42), and Lorain (-0.25), (C) Clermont (-0.10) and Sloan (-0.06), and (D) Clermont (-

35.0). Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher BLUP values indicate more resistant (R) lines. .. 136

Figure 3.3. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross IA3023 x LD02-9050 following inoculation with Py. irregulare in a greenhouse cup assay. Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher

BLUP values indicate more resistant (R) lines...... 137

Figure 3.4. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross IA3023 × LG00-3372 following inoculation with Py. irregulare in a greenhouse cup assay. BLUE values not indicated on the graphs are (A) Lorain (-22.5), (B) Clermont (0.20), Sloan (-0.40), and Lorain (-0.8),

xxii

and (D) Lorain (-104.1). Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher BLUP values indicate more resistant

(R) lines...... 138

Figure 3.5. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross IA3023 x 4J105-3-4 following inoculation with Py. ultimum var. ultimum in a greenhouse cup assay. BLUE values not indicated on the graphs are (A) Lorain (13.75), (B) Lorain (0.30), and (C) Lorain (0.08).

Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher BLUP values indicate more resistant (R) lines...... 140

Figure 3.6. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross HS6-3976 x IA3023 following inoculation with Py. ultimum var. ultimum in a greenhouse cup assay. BLUE values not indicated on the graphs are (A) Clermont (10.4) and Sloan (6.3), (B) Clermont (0.33),

Sloan (0.17), and Lorain (0.08), and (D) Sloan (15.8) and Lorain (63.1). Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher BLUP values indicate more resistant (R) lines...... 142

Figure 3.7. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent

xxiii

germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross IA3023 x S06-13640 following inoculation with Py. ultimum var. ultimum in a greenhouse cup assay. BLUE values not indicated on the graphs are (B) Clermont (0.38) and Sloan (0.38), and (D) Clermont

(82.9). Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher BLUP values indicate more resistant (R) lines. .. 143

Figure 3.8. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross HS6-3976 x IA3023 following inoculation with Py. ultimum var. sporangiiferum in a greenhouse cup assay. BLUE values not indicated on the graphs are (A) Sloan (-18.8) and Clermont (15.6), (B)

Clermont (-0.17), Sloan (0.08), and Lorain (-0.5), (C) Sloan (-0.08), and (D) Sloan (-

126.3) and Lorain (-61.2). Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher BLUP values indicate more resistant

(R) lines...... 145

Figure 3.9. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross IA3023 x LG05-4832 following inoculation with Py. ultimum var. sporangiiferum in a greenhouse cup assay. BLUE values not indicated on the graphs are (B) Lorain (-0.50), (C) Clermont (-0.15) and

xxiv

Lorain (-0.18), and (D) Sloan (-118.6). Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher BLUP values indicate more resistant (R) lines...... 146

xxv

Chapter 1. Literature Review

Significance of soybean:

Soybean [Glycine max (L.) Merr.] is historically and currently the number one crop produced in Ohio. During 2014 and 2015, there were over 4.6 million harvested acres each year totaling over 2.5 billion dollars in value (National Agricultural Statistics

Service, USDA, 2015). The economic importance of soybean is not limited to Ohio; it is also a leading crop both in terms of acres planted and production value in many other states in the north central region. The United States (US) is the world’s top producer of soybean, followed closely by Argentina and Brazil (FAOSTAT, 2015). In 2014 the US exported 48.7 million metric tons, totaling over 30 trillion US dollar (USD) in value

(Global Agricultural Trade System Online, 2015).

The majority of soybeans harvested in Ohio are made into meal or oil for both feed and human consumption, respectively. Soybean has a relatively high amount of protein with low amounts of saturated fat, making it a valuable source of nutrition for both humans and livestock (Dwevedi and Kayastha 2011). Soybean is prized for its oil- producing capability; soybean-derived oils make up most of the vegetable oil consumed in the US (United States Census Bureau, 2010). Soybeans can also be processed into foods such as dairy or meat alternatives, flour, or tofu. Within the past

1

three decades there was an increase in demand for the “healthy-alternative” high oleic soybean oil; there is an active effort in the development of high oleic soybean and some cultivars are currently available (Pham et al., 2012; La et al., 2014). The livestock industry is highly dependent on high quality seed for soybean meal as feed for poultry, swine, and cattle due to its high protein content (Yin et al., 2007).

High quality soybeans and high yields are crucial to soybean growers’ profits as well as those of the many connected industries. The profitability of soybean in Ohio is at risk from a variety of factors, including weeds, nutrient deficiencies, insect pests, and pre- and post-harvest diseases (Beuerlein and Dorrance, 2004; Koenning and Wrather,

2010). Soybean seed and seedling diseases are caused by several different pathogens that target either the seed itself or the very young developing seedling. These diseases are regularly caused by a number of fungal and watermold species including: Fusarium,

Macrophomina, Pythium, Phytophthora, and Rhizoctonia (Ajayi-Oyetunde and Bradley,

2017; Broders et al., 2007a,b; Dorrance et al., 2003a,b; Dorrance et al., 2004; Navi and

Yang, 2016; Pearson et al., 1984; Rizvi and Yang 1996; Rojas et al., 2017a, b; Yang and

Feng, 2001).

In the past twenty years, based on annual soybean disease surveys, seed and seedling diseases have been consistently the third most important cause of yield loss behind soybean cyst nematode and sudden death syndrome in the US (Allen et al., 2017;

University of Illinois Extension and Outreach, 2016; Wrather et al., 2001; Wrather and

Koenning, 2006; Wrather et al., 2009). Based on reports from 2014 alone, the total yield loss due to disease was estimated to be nearly 500 million bushels, of which the loss of

2

approximately 60 million bushels were attributed to seed and seedling diseases (Allen et al., 2017). Pythium contributes to reduced yield by infecting and killing seedlings and seeds, thereby decreasing stand (Beuerlein and Dorrance, 2004). The presence of seedling disease in a field may also necessitate replanting and negatively impact the vigor of any surviving plants due to do damaged root systems (Yang, 1999).

In order to prevent yield loss, the most efficient and effective disease management strategies for seed and seedling diseases must be identified for Ohio soybean farmers to implement into their production practices.

Pathogens: Pythium, Phytophthora, and Phytopythium

The Pythium (Py.) has a high proportion of soilborne and generalist plant pathogens (Middleton, 1943; van der Plaats-Niterink, 1981). Pythium belongs in the class

Oomycota and family . It is characterized by the production of a thick-walled oospore, an elongated antheridium, cell walls made of cellulose, and irregularly branched, coenocytic hyphae (Schroeder et al., 2013).

Most species of Pythium are able to reproduce both sexually and asexually.

During asexual reproduction, a specialized structure called a will hold a number of mobile propagules called zoospores, which are formed inside a vesicle; however some species of Pythium do not produce zoospores (Middleton, 1943). The morphological shape of the sporangium can range from spherical to filamentous and can be used to help identify some species (Middleton, 1943).

3

Once mature, biflagellate motile zoospores are released from the vesicle and swim freely through water under saturated soil conditions. These motile zoospores act as the primary infective agent on susceptible seeds and seedlings. Zoospores exhibit a positive chemotropism response to sugars and amino acids in the surrounding environment

(Donaldson and Deacon, 1993; Jones et al. 1991). Root tips produce exudate, which is composed of sugars and amino acids, and attracts zoospores to this vulnerable growing region. After sensing and moving towards the growing soybean root, the zoospore will encyst on the root tissue, lose mobility, form an appressorium, and produce enzymes that will facilitate the germ tube’s direct penetration through the root barrier (Raftoyannis and

Dick 2006; van West et al., 2003). Pythium species produce pectinases and some species produce cellulases to partly dissolve the , to feed on the less complex sugars present within the host’s cells (Zerillo et al., 2013). Once inside the root, Pythium colonizes the surrounding healthy tissue and ultimately the root system, or in some cases the entire plant, to acquire nutrients.

Due to the zoospores’ ability to move through water the distribution of Pythium and subsequently, disease incidence is much higher in soils with a high moisture content.

Pythium species in particular can devastate greenhouse production operations if there is not proper prevention and care (Watanabe et al., 2008). The ability of zoospores to move through water and very accurately sense the apical portions of plant roots favors Pythium infection in hydroponic growth conditions (Jenkins and Averre, 1983; Sutton et al.,

2006).

4

Species of Pythium can either be heterothallic (mating of two different genotypes) or homothallic (self-fertilization) (Middleton, 1943). Sexual reproduction involves the transfer of nuclei from the antheridium to the via a fertilization tube, which results in the formation of the zygote (Marchant, 1968). To protect the newly formed zygote, the oogonium develops into the oospore by thickening the outer cell wall. These oospores are especially hardy and act as the survival structure through winter and over long periods of drought, either on infected plant material or in the soil (Schroeder et al.,

2013).

Chlamydospores are vegetative resting structures produced by Pythium that are also thick-walled and allow the oomycete to withstand harsh conditions. The mycelium of Pythium does not fare well under long-term stressful conditions, but oospores and chlamydospores allow Pythium to persist even in harsh conditions or over long periods of time in the soil. When the environmental conditions are suitable for growth or a host plant is sensed nearby, the oospores/chlamydospores germinate and hyphae grow toward sources of nutrients or directly produce sporangia to begin the disease cycle anew

(Middleton, 1943; van West et al., 2003). The pathogen has a rapid response to the detection of a suitable host; the oospores of Py. aphanidermatum and the sporangia of Py. ultimum germinate between 1.5 and 3 hours after being exposed to bean or pea exudates

(Lifshitz et al., 1986; Stanghellini and Burr, 1973; Stanghellini and Hancock, 1971a,b).

Most Pythium species are generalist and necrotrophic and/or saprophytic

(Middleton, 1943; Schroeder et al., 2013). Pythium can survive as mycelium on infected tissue in or on the soil surface. Reduced tillage or an excess of plant debris on the soil

5

surface is a conducive environment for Pythium, resulting in increased disease severity

(Larkin et al., 1995; Pankhurst et al., 1995). Low-tillage and no-till further compounds the problem due to reduced drainage, greater soil retention of water, and the buildup of inoculum in the top layers of soil (Bescansa et al., 2006; Workneh et al., 1999). Though no-till farming has benefits such as the reduction of erosion and decreased labor costs, the potential increase in disease incidence and disease severity from debris- and water-loving pathogens must be considered. The use of cover crops (such as grass, clover, or oilseed radish) can decrease the severity of many root rot pathogens, including Pythium, in some host systems (Abawi and Widmer, 2000). However, planting cover crops can also increase disease incidence or severity. Recently, planting a grass cover crop leads to increased infection rates in corn seedlings during the following growing season (Bakker et al., 2016). The type of cover crop must also be taken into consideration; cover crops have been associated with the increase of Pythium populations in soil (Rothrock et al., 1995). Since Pythium is generalist pathogen, is not an effective strategy in reducing inoculum (Pankhurst et al., 1995). For example, Py. irregulare commonly infects both corn and soybean, so the corn-soybean rotation commonly used in Ohio may not be an effective means of managing this pathogen (Broders et al., 2007). The impacts of the previously mentioned production practices on disease incidence are important to keep in mind, especially because no-tillage farming is firmly established in Ohio and the use of cover crops is increasing in the US (USDA, National Agricultural Statistics

Service).

6

Within the Pythium genus, there are many plant-pathogenic species that are economically important in agriculture. The pathogenicity of a species of Pythium depends on both the species in question as well as the environmental conditions (Dorrance et al.,

2004; Rojas et al., 2017a, b; Zhang and Yang, 2000). Soil type and nutrient availability also impact which Pythium species can survive in a location (Broders et al., 2009; Zitnik-

Anderson et al., 2017). Most Pythium species typically have a higher proliferation rate and pathogenicity at temperatures around 20ºC, with some exceptions (Gold and

Stanghellini, 1985; Littrell and McCarter, 1969; Thomson et al., 1971). For instance, Py. irregulare is more pathogenic at 20ºC than at 28ºC, whereas Py. aphanidermatum and

Py. myriotylum exhibit the reverse pattern and exhibits greater pathogenicity at higher temperatures (Ben-Yephet and Nelson, 1999; McCarter and Littrell, 1970; Littrell and

McCarter, 1969; Thomson et al., 1971). The general trend of Pythium species’ preference for wet conditions makes Pythium seed and root rot a key disease to scout for in the spring, soon after planting regardless of soil temperatures.

Oomycetes in the genus Phytophthora (Ph.) are closely related to the genus

Pythium and also contain a large number of economically important plant-parasitic pathogens (Cooke et al., 2000). Phytophthora sojae and Ph. sansomeana are commonly found on soybean in Ohio (Dorrance, 2003b; Schmitthenner, 1985; Zelaya-Molina et al.

2010). Pythium species tend to be more generalist, whereas Phytophthora sojae is limited to soybean. Phytophthora and Pythium are similar in appearance under the microscope and both can cause extensive root-rot on soybean seedlings, but there are several key differences. There are microscopic morphological differences between Pythium and

7

Phytophthora: Zoospores are produced directly in the sporangia in Phytophthora, and in

Pythium the zoospores form in a vesicle produced by the sporangia (de Cock and

Levesque, 2004). The sporangia of Phytophthora are typically ovoid to lemon-shaped, as opposed to the globose, filamentous, or oval-shaped sporangia of Pythium species. The sporangia of most species of Phytophthora are terminal, whereas the sporangia of

Pythium are either terminal or intercalary.

Phytopythium (Pp.) is a recently described genus that encompasses oomycete plant pathogens closely related to both Pythium and Phytophthora (de Cock et al., 2015).

Previously, Pythium was grouped into distinct clades mainly based on molecular systematic analyses and further supported by morphological traits (Levesque and de

Cock, 2004). The previously described Clade K of Pythium had morphological traits of both Pythium and Phytophthora; this group was recently described as the separate genus

Phytopythium (de Cock et al., 2015). Similar to Pythium and Phytophthora,

Phytopythium can be pathogenic on soybeans (Broders et al., 2009; Dorrance et al., unpublished data; Radmer et al., 2017; Rojas et al., 2017a,b; Rojas-Flechas, 2016).

Examples of Phytopythium species found in Ohio on soybean are Pp. vexans and Pp. helicoides (Broders et al., 2009; Dorrance et al., unpublished data).

Pythium seed and root rot of soybean

Pythium is a major cause of seedling root rot in soybean and can cause both pre- and post-emergence damping off. Pythium can cause reduction in seedling vigor and early death, which ultimately leads to a reduced stand, yield, and seed quality (Broders et

8

al., 2007a; Hwang et al., 2001; Yang, 1999). Seedlings in more favorable growing conditions may withstand the initial attack and continue to grow, albeit with reduced yield, growth, and performance (Dodd and White, 1999). Lesions typically form on the crown of the seedling, weakening the tissue and eventually causing the seedling to collapse near the soil line. Root rot can be easily seen; the pathogen will infect and destroy the taproot or the more delicate lateral roots and cause brown-colored lesions.

Due to their general preference for cooler temperatures and high soil moisture conditions, most Pythium species cause severe early-season stand losses. Slower growth of soybean seedlings at lower temperatures leaves them more prone to infection and allows for a longer infection period, which then increases the risk of Pythium disease and yield loss (Sanei et al., 1978). Low quality seeds or seeds with low vigor are especially at risk of seedling disease (Hamman et al., 2002; Poag at al., 2005; Rupe et al., 2011). There are more than 35 different species of Pythium associated with seed and seedling rot on soybean in Ohio and pathogenicity is highly variable both within and among the species identified in a given field (Broders et al., 2007a; Broders et al., 2009; Ellis et al., 2013;

Dorrance et al., 2004; Dorrance et al., unpublished data). Some species of Pythium are more likely to cause seed decay (pre-emergence damping off), while others are more pathogenic on seedlings (post-emergence damping off) (Rojas-Flechas, 2016). A wide variety of soil and environmental conditions such as location, precipitation, temperature, soil electrical conductivity, and clay content dictate the community composition and distribution of oomycete species in Ohio and North America (Broders et al. 2009; Rojas et al., 2017b; Zitnik-Anderson et al. 2017).

9

Disease management strategies

Management of Pythium seed and root rot requires the integration of cultural practices, seed treatments, and host resistance. One management strategy against

Pythium-caused diseases is crop rotation; since Pythium species have different levels of pathogenicity to different species of plants, rotating in a crop that is more resistant can potentially reduce the inoculum level in the field (McCarter and Littrell, 1969). However, due to the generalist nature of Pythium and the diversity of Pythium species within a field, crop rotation may not be a reliable method of reducing inoculum (Njoroge et al,.

2009; Pankhurst et al., 1995). There are many species of Pythium that are highly pathogenic to both corn and soybean; since one of the most common crop rotations in both Ohio and Iowa is soybean-corn, this rotation may not actually result in a decrease in

Pythium inoculum (Broders et al., 2007a; Zhang and Yang 2000).

Seed treatments are a useful tool for managing Pythium seed and root rot. Several biological seed treatments are available, but present additional challenges and considerations. Inoculants of less aggressive or Penicillium oxalicum were effective at reducing the incidence of seed rot and pre-emergence damping off of chickpea caused by more virulent Pythium species such as Py. irregulare or Py. ultimum var. ultimum (Casas et al., 1990). Priming sweet corn seed with an isolate of

Pseudomonas fluorescens protected against Py. ultimum with effects similar to a metalaxyl treatment (Callan et al., 1990). Biocontrol seed treatments are an attractive option when chemical control is not available or not desired, but there are special

10

considerations: shelf life, efficacy of the biocontrol agent in different soil environments, longevity of the biocontrol agent in the field (Sharma et al., 2015; Walker et al., 2004), in addition to efficacy towards the diversity of the species found in these fields as well as the ability colonize and proliferate in the diversity of environmental conditions

(Dorrance, personal communication).

Fungicide seed treatments are another commonly deployed disease management strategy against seed and seedling diseases in Ohio. Seed treatments can provide protection to the seed prior to germination and during germination when systemic fungicides are used (Cohen et al., 1979; Munkvold and O’Mara, 2002). Fungicide seed treatments on high-quality seed remain one of the most effective ways of protecting plants from seedling diseases (Poag et al., 2005).

The effectiveness of the active ingredient of the fungicide is dependent on the targeted pathogen, as different fungicides have varying levels of efficacy against certain pathogens (Beuerlein and Dorrance, 2004; Bradley, 2008; Broders et al., 2007a,b;

Broders et al., 2009; Radmer et al., 2017; Rojas et al., 2017b; Weiland et al., 2014).

Metalaxyl is a common oomycete fungicide that has been shown to reduce seed rot caused by Pythium (Hwang et al., 2001); however the efficacy of this treatment varies among Pythium species (Brantner et al., 1998; Broders et al., 2007a,b; Dorrance et al.,

2004). There are multiple reports across the US of oomycete pathogen populations developing insensitivity to fungicides, including insensitivity to mefenoxam and metalaxyl (Broders et al. 2007a; Dorrance et al. 2004; Moorman and Kim, 2004; Taylor et al., 2002).

11

The interaction of fungicide and target pathogen can be complex. In addition to the variable susceptibility of fungicide resistance within a pathogen population, the effectiveness of a fungicide is also impacted by the active ingredient concentration and environmental conditions. Low concentrations of mefenoxam stimulate growth and increase the pathogenicity of several oomycete species (Garzón et al., 2011; Pradhan et al., 2017). Temperature also has an impact on the fungicide sensitivity of multiple

Pythium species; Py. torulosum has a high resistance to multiple fungicides at 13ºC, while Py. sylvaticum has higher resistance to fungicides at temperatures greater than 13ºC

(Matthiesen et al., 2016). To more effectively manage Pythium seed and root rot, as well as oomycete and fungal diseases in general, fungicide seed treatments require routine screening to evaluate efficacy and monitor for the development of insensitivity.

Mefenoxam, ethaboxam, and strobilurin fungicides have a range of effectiveness towards different species of Pythium and other root rot pathogens (Radmer et al., 2017).

Recently two new fungicide active ingredients, ethaboxam and oxathiapiprolin, have been introduced into the market. The availability of new active ingredients and the varying susceptibility of root pathogens to the different available fungicide active ingredients requires further research to determine the fungicide seed treatment best suited for any particular field environment.

In addition to cultural practices and seed treatments, another disease management strategy towards seed and seedling diseases is planting resistant cultivars. Resistance can be qualitative or quantitative. Each type of resistance has benefits over the other and both types are very useful, time-efficient, and economic means of controlling seed and root rot

12

of soybean. Some cultivars that express more resistance to certain Pythium spp. have been identified (Anderson et al., 2017; Balk, 2014; Balk et al., 2014; Bates et al., 2008;

Kirkpatrick et al., 2006; Rosso et al., 2008; Rupe et al. 2011).

Qualitative resistance, also called race-specific or R-gene mediated resistance, is conferred by a single resistance gene (R-gene) in the host plant and provides a strong level of protection against certain races of a particular pathogen (Chisholm et al., 2006).

When a pathogen infects a host with qualitative resistance to that particular isolate, the hypersensitive response is activated at points of infection (Morel and Dangl, 1997). This reaction results in the death of the few infected cells and prevents any further development of the disease. Identification and deployment of qualitative resistance genes in soybean is common for biotrophic or hemibiotrophic pathogens, such as the hemibiotrophic watermold Phytophthora sojae and Cercospora sojina

(Schmitthenner, 1985; Mian et al., 1999). Though qualitative resistance is very effective in protecting plants against disease, the intense selection pressure placed on a the invading pathogen population can result in the pathogen population overcoming these resistance genes over time (Abney et al. 1997; Dorrance et al., 2003b,c; Dorrence et al.,

2016; Schmitthenner, 1985). R-gene mediated resistance is thought to best control biotrophic pathogens; the resulting programmed cell death from the recognition of a biotrophic pathogen would prevent the cell being used as a food source (Glazebrook,

2005). On the other hand, the programmed death of a targeted cell by a necrotrophic pathogen would simply provide a nutrient source and be ineffective in preventing pathogen growth (Glazebrook, 2005). Although Pythium species are generally thought to

13

be necrotrophs, one R-gene has been proposed for in soybean

(Rosso et al., 2008; Rupe et al., 2011).

Quantitative resistance, also called partial resistance, is most often conferred by a multitude of genes at several loci in a host plant and is the most commonly deployed resistance against necrotrophic plant pathogens (St. Clair, 2010). This type of resistance is commonly effective against multiple races of a pathogen. Quantitative resistance results in an overall decrease, but not a total elimination, of disease symptoms in an infected plant. Regions of the genome that contribute to the overall resistance of the plant are called quantitative disease resistance loci (QDRL) (St. Clair, 2010). These individual genes may have minor separate effects, but their combined overall effects contribute to the plant having resistance towards the pathogen (Kou and Wang, 2010). Unlike qualitative resistance, partial resistance does not involve the hypersensitive response and typically allows for some symptoms to develop (Kou and Wang, 2010). This type of resistance is also thought to put less selection pressure on pathogen populations, so partial resistance is more durable long-term in preventing the development of pathogens that can overcome host crop resistance (Kou and Wang, 2010; St Clair, 2010; Michelmore, 2013).

To incorporate resistance genes into new cultivars, first the genetic source of resistance must be identified. The best sources of resistance are soybean cultivars that are already adapted to a region and have high yield. The second step is to identify the genomic regions where resistance is located. Genetic mapping has been used extensively to identify the regions in genomes that contribute to the expression of partial resistance to numerous pathogens (Collard et al., 2005; Lee et al., 2013a,b; Lindhout, 2002; Poland et

14

al., 2009; Young, 1996). Quantitative resistance to a few species of Pythium has been mapped in soybean. Urrea et al. (2017) found two QRL to Py. aphanidermatum on chromosome 4 and 7. Stasko et al. (2016) identified two QTL conferring resistance to Py. irregulare on chromosome 14 and chromosome 19-2. The region on chromosome 14 is especially interesting because it is near (approximately 12kb from) another previously identified Py. irregulare resistance QTL (Ellis et al., 2013). Ellis et al. (2013) found resistance to Py. irregulare QTL on chromosomes 1, 6, 8, 11, and 13 using composite interval mapping associated with partial resistance to Py. irregulare recombinant inbred line (RILs) populations (Ellis et al., 2013; Stasko et al., 2016). A comparison of regions that were associated with resistance to Py. irregulare were different from those that were associated with resistance to Fusarium graminearum, another necrotrophic pathogen and

Ph. sojae, that causes seedling disease (Stasko et al., 2016). The difference in location of resistance QTL in this one soybean RIL population indicates that there could be different mechanisms for resistance to Py. irregulare, F. graminearum and Ph. sojae in soybean.

Discovery and identification of the QTL responsible for soybean partial resistance to oomycete pathogens is necessary for the development of cultivars with durable resistance, which contribute to maintaining soybean production. There are many approaches to mapping and identification of QTL. One such technique is linkage association, which can identify QTL with relatively sparse marker coverage across the genome. However, this technique also requires a recent cross of inbred plant lines and produces genetic maps with low resolution (Beckmann and Soller, 1988; Darvasi and

Weller, 1992; Doerge, 2002). Another commonly used technique is association mapping

15

(aka linkage disequilibrium mapping), which does not require a recent cross because it uses the ‘historic recombination’ in the plant, and can produce maps with high resolution

(Hall et al., 2010; Zhu et al., 2008). This technique is reliant on either prior knowledge of a gene of interest, or scanning the genome with many markers (Hirschhorn and Daly,

2005). Nested association mapping (NAM) is the combination of linkage analysis and association mapping, and offers the advantages of both techniques and avoids the disadvantages (Yu et al., 2008). Once the genomic SNPs of a population are identified,

NAM can provide high resolution maps even with low marker density. The NAM technique was originally developed to be used in to identify the genetic sources of agronomic traits of interest, such as flowering time (Buckler et al, 2009; McMullen et al.,

2009). This technique has been implemented to identify QTL in maize associated with disease resistance, leaf architecture, and kernel composition (Cook et al., 2012; Poland et al., 2011; Tian et al., 2011). Nested association mapping populations are being developed for other plants species such as Arabidopsis thaliana, rapeseed, , and soybean

(Bajgain et al., 2016; Li et al., 2014; Song et al., 2017; Stich, 2009).

A soybean NAM population was developed by crossing 40 diverse lines with the common parent IA3023 and producing 40 different Recombinant Inbred Line (RIL) populations, each containing 140 F5-dervied RILs (Song et al., 2017). Each of the 40 donor parents were selected due to their diverse ancestry, high yield, or drought tolerance. Each of the NAM parents and resulting populations were genotyped with the

SoySNP6K Illumina Infinium BeadChip Genotyping Array and had a linkage map generated based on this genotype data (Song et al., 2017; SoyNAM, soybase.org). The

16

soybean NAM project’s goals were to map QTL controlling desirable traits across diverse soybean germplasm, and then use this information to develop improved germplasm

(Diers, 2015). The genotype information of the NAM parents and RILs is publicly available at http://soybase.org, making the NAM population a great resource for QTL research (SoyNAM, SoyBase.org).

The species Phytophthora sojae, Py. irregulare, Py. ultimum var. ultimum, and

Py. ultimum var. sporangiiferum were chosen for this study because they have been recovered from symptomatic seedlings from many fields across Ohio and can cause severe root rot on susceptible soybean seeds and seedlings (Broders et al., 2009; Balk,

2014; Eyre et al., 2016; Dorrance et al., unpublished data). Balk (2014) identified six

NAM parents that had lower levels of root rot compared to the hub parent to these oomycete species, making the resulting NAM populations from crosses of these parents with IA3023 ideal for identification of QDRL.

In summary, the objectives of this research were to investigate two different avenues for managing Pythium seed and root rot of soybean. The first objective was to evaluate the efficacy of commercially available fungicides applied as a seed treatment in preventing seed and root rot of soybean caused by multiple oomycete species. Soybean growers need to be able to make informed decisions on how to most economically manage seedling diseases while considering yearly and local field conditions. Evaluating ethaboxam as a new active ingredient is important so soybean yields can be protected from seedling disease, while also preventing the development of further fungicide- resistant oomycete pathogen populations. This research also provides the level of

17

pathogenicity of different oomycete species on soybean seeds and seedling roots. Testing the efficacy of several recently available fungicide seed treatments, screening commonly used cultivars, and characterizing the level of pathogenicity of common Ohio oomycete species would be of great use to both soybean producers and the seed treatment industry, as well as public and private soybean breeders. The second objective was to map QDRL to Ph. sojae, Py. irregulare, Py. ultimum var. ultimum, and Py. ultimum var. sporangiiferum in order to identify sources of resistance. This information is important because the results from QDRL mapping can used by soybean breeders to develop lines with more resistance to infection by different oomycete species.

18

Chapter 2: The efficacy of ethaboxam as a component in soybean seed treatments towards oomycetes in Ohio

Introduction

Soybean (Glycine max L. Merrill) seedling diseases caused by oomycetes are a major problem in the United States (US) (Allen et al., 2017; Bradley, 2008; Broders et al.,

2007a; Koenning and Wrather, 2010; Rojas et al., 2017a,b). The oomycetes comprise a diverse number of genera, some of which cause extensive soybean pre- and post- emergence damping-off. This results in reduced yields due to delayed planting and reduced vigor of the surviving plants due to root damage. Depending on the severity of the yield loss, Pythium seed and root rot may require that the soybean fields be replanted.

In Ohio, field surveys and subsequent pathogenicity assays have identified primarily

Pythium (Py.) and Phytophthora (Ph.) species that pose a serious risk to soybean stand and yield. (Broders et al., 2007a; Broders et al., 2009; Dorrance et al., 2003b; Dorrance et al., 2004; Eyre, 2016)

Members of the oomycete genus Pythium can cause Pythium seed and root rot on soybean. Pathogenic Pythium spp. injure and kill seedlings by wounding the roots of plants and leeching nutrients, thereby reducing stand and decreasing the vigor of any surviving seedlings (Broders et al., 2007a; Hong and Moorman, 2005; Martin and Loper,

19

1999). Saturated soil creates favorable environmental conditions for infection, development, and reproduction of the pathogen (Kirkpatrick et al., 2006). Soybeans at the seedling stage are the most vulnerable to this disease. Crop rotation is not an effective means to manage Pythium seed and root rot, as Pythium spp. are generalists and will also infect commonly rotated crops such as corn or wheat (Broders et al., 2007b; Ingram and

Cook, 1990). The combination of poorly drained soils and reduced tillage, which promotes inoculum build-up (Workneh et al., 1998), provides the perfect conducive conditions for large scale oomycete disease epidemics.

More than one species of Pythium has been isolated from a single infected seedling, indicating a probable Pythium disease complex (Broders et al., 2007a, 2009;

Wei et al., 2010). Pythium can also form a seed and seedling complex with other soilborne pathogens such as Phytophthora sojae and Rhizoctonia spp. (Rizvi and Yang,

1996).

Integrated disease management that combines resistant cultivars and fungicide seed treatments is recommended for the management of oomycete seedling diseases in soybean (Dorrance et al., 2009; Urrea et al., 2013). Albeit, few sources of host resistance to Pythium spp. have been identified in germplasm (Bates et al., 2008; Ellis et al. 2013;

Rosso et al., 2008; Stasko et al., 2016; Urrea et al., 2017), but overall very little is known about the level of resistance towards Pythium among cultivars currently used by farmers.

Due to these current limitations on the knowledge of host resistance towards Pythium,

Pythium seed and root rot is typically managed with fungicide seed treatments (Bradley,

2008; Esker and Conley, 2012; Gaspar et al., 2017). The management of Phytophthora

20

sojae involves similar challenges; the quantitative resistance towards Ph. sojae is not effective until the plant is at least in the V1 growth stage (when the seedling has one set of trifoliate leaves) (Dorrance and McClure, 2001). Additionally, the population diversity of Phytophthora on soybean in the north central region of the US has a greater population of isolates that have adapted to commonly used resistance genes (Dorrance et al., 2016;

Guy et al., 1989; Stewart et al., 2016). These challenges have led to fungicides and fungicide seed treatments becoming a popular method of management for Ph. sojae. and other oomycetes (Anderson and Buzzell, 1982; Bradley, 2008; Dorrance and McClure,

2001).

Fungicide seed treatments are a valuable means of managing early season diseases caused by oomycetes. The phenylamide active ingredients, metalaxyl and its active isomer mefenoxam, have been reliable active ingredients in controlling oomycete infections by reducing the losses from damping-off, thereby increasing yield (Bradley,

2008; Bradley et al., 2001; Dorrance and McClure, 2001, Gaspar et al., 2015; Guy et al.,

1989; Lueschen et al., 1991; Radmer et al., 2017), even when reduced tillage is being practiced (Guy and Oplinger, 1989). These active ingredients work by inhibiting the ribosomal ribonucleic acid (RNA) synthesis in oomycetes (Davidse et al., 1983). In Ohio and across the US, Pythium isolates have been recovered from corn and soybean fields that are insensitive to metalaxyl and mefenoxam (Broders et al., 2007a; Dorrance et al.,

2004; Taylor et al., 2002; Radmer et al., 2017; Rojas et al., 2017a,b). Quinone outside respiration inhibitor (QoIs) fungicides have also been used to manage some diseases caused by oomycetes and fungi (Bradley, 2008; Broders et al., 2007a,b; Ellis et al., 2011;

21

Matthiesen et al., 2016; Radmer et al., 2017), but as with phenylamide fungicides, they are at high risk for the development of resistance in the pathogen population (Broders et al., 2007a; Gisi and Sieotzki, 2002; Radmer et al., 2017). To maintain management of seedling diseases and prevent poor stands, replanting, and potential yield loss, new chemistries need to be evaluated for efficacy against soilborne seed and seedling pathogens such as Pythium and Phytophthora.

Ethaboxam is a fungicide that is reported to be effective in managing several diseases caused by different oomycete species (Kim et al., 2004; Matthiesen et al., 2016;

Radmer et al., 2017). The mode-of-action for ethaboxam is the disruption of microtubulin assembly, as modeled in Phytophthora infestans, the causal agent of late blight of

(Uchida et al., 2005). Ethaboxam is considered a low to medium-risk active ingredient for the development of pathogen resistance (Kim et al., 2004). Ethaboxam is thus a good candidate for an effective fungicide active ingredient for the management of seed and root disease caused by Pythium, Phytophthora, and Phytopythium (Pp.) species.

We hypothesize that ethaboxam in commercial formulations will be effective in controlling a wide range of oomycete species when compared to other commercial formulations without this fungicide. Our objective was to evaluate and compare the efficacy of fungicides with different active ingredients in field, greenhouse, and laboratory assays to determine the ability of ethaboxam to manage oomycete seed and seedling diseases. Our study will mainly focus on Phytophthora spp., Pythium spp., and

Phytopythium spp. that have previously been found to be prevalent and pathogenic on

22

soybean in Ohio. Preliminary reports of these studies were published previously

(Dorrance et al., 2012; Scott et al., 2016).

23

Materials and Methods

Field experiments

Fields to study the effects of seed treatments on soybean plant populations and yield were selected due to a prior history of seedling disease, high incidence of damping off, or poor drainage. The dates for planting and harvest, rainfall data, soil type, total row length counted for stand counts, and seeding rate for each location are listed in Table 2.1.

In this study, an “environment” refers to a single field site within a growing season (e.g.

VW-16 is the Van Wert field in 2016). Each year, there was one field planted at the Ohio

Agricultural Research and Development Center (OARDC) Northwest Agricultural

Research Station (NWB-14, -15, -16). These fields were irrigated to ensure saturated soil conditions for optimal disease development. Other field locations in Ohio were Defiance

(DEF-15, -16), Van Wert (VW-14, -16), and Snyder Farm at OARDC, Wooster (SNY-

16); these fields did not have irrigation and were dependent on natural rainfall to produce an environment conducive for oomycete seedling disease.

Two field environments were planted in 2014 to evaluate the effects of seven fungicide seed treatments and a nontreated control (Table 2.2) with the cultivars Conrad

(Fehr et al., 1989) and Sloan (Bahrenfus and Fehr, 1980) (Table 2.3). During 2015, three seed treatments were evaluated at Northwest Branch (NWB-15) and Defiance (DEF-15)

(Table 2.2). The 2015 field sites compared the effect on four cultivars with different levels and combinations of resistance towards Pythium and Phytophthora sojae; these cultivars were Conrad, Sloan, Kottman (St Martin et al., 2001), and Lorain (OSU-

OARDC) (Table 2.3).

24

Four field locations were planted in 2016 to evaluate the effect of one seed treatment formulation, ethaboxam with clothianidin plus ipconazole plus metalaxyl

(Intego Solo in combination with Inovate Pro, Table 2.2), on seven cultivars (Table 2.3) with different levels and combinations of resistance for early and late plant populations and yield. The seven cultivars Clermont (OSU-OARDC), Dennison (St Martin et al.,

2008), Kottman, Streeter (OSU-OARDC), Conrad, Lorain, and Sloan were planted at the following locations: Northwest Branch (NWB-16), Defiance (DEF-16), Van Wert (VW-

16), and Snyder Farm at OARDC, Wooster (SNY-16).

Each field study was planted as randomized split plot design with the cultivar as the whole plot and the seed treatments as the sub-plots with 4-6 replicates, dependent on how much space was available in each field (Table 2.1). From each plot at each location, data for stand on a subsection of each of the experimental plots was collected three times in each growing season: i) developmental stage when there were one or two trifoliate leaves (V1-V2), ii) at the three or four trifoliate leaf stage (V3-V4), and iii) the maturity stage when most of the pods are brown or tan (R7-R8). Stand counts were converted into the number of plants per hectare (plants/Ha).

Fields at Snyder were harvested with an Almaco Specialized Plot Combine

(Almaco SPC20) (Almaco, Nevada, IA) equipped with a Harvestmaster GrainGage system. Plot data were collected and stored with an Allegro MX (Juniper Systems,

Logan, UT) with Field Research Software (FRS) (Juniper Systems, Logan, UT). Fields at

NWB were harvested with a Kincaid 8-XP (Kincaid Equipment Manufacturing, Haven,

KS) paired with a High Capacity GrainGage weigh system (Juniper Systems, Logan,

25

UT). The Defiance and Van Wert fields were harvested by Tech Services, Inc. (Bluffton,

IN) with a Massey Ferguson 8-XP combine paired with an Almaco LRX electronic system (ALMACO, Nevada, Iowa). The yield data for all locations was standardized to

13.5% moisture and converted to bushels/acre and kilograms/hectare. The data collected from the field experiments were individually analyzed using ANOVA with PROC GLM in SAS (SAS version 9.4; SAS Institute, Cary, NC) and mean comparisons were performed with Fisher’s protected LSD (p<0.05). The 2016 field data was analyzed across all four of the locations in order to test for significance of the location, cultivar, fungicide, and each of the corresponding interactions on stand and yield; this analysis was done using ANOVA with PROC GLM in SAS (SAS version 9.4; SAS Institute,

Cary, NC) and mean comparisons performed with Fisher’s protected LSD (p<0.05).

Pythium, Phytopythium, and Phytophthora isolates

All isolates used in these experiments were recovered from Ohio fields from infected seedlings, many from the field experiments in 2015 or through soil baiting from fields with a history of losses due to seedling diseases or isolates from the soybean pathology lab collection. Each isolate was identified to the species level based on morphological characteristics from cultures grown on potato carrot agar plates or from grass blade cultures using a monographic key for the genus Pythium (van der Plaats-

Niterink, 1981; Waterhouse, 1968), as well as sequence analysis of the entire internal transcribed spacer (ITS) region, as previously described by Broders et al. (2007a).

Isolates were grown on V8 broth at ambient temperature for 2-6 days, and the mycelia

26

were collected on filter paper placed in a Buchner funnel, transferred to a mortar, frozen with liquid nitrogen, and ground into a powder. DNA was extracted using the method described by Zelaya-Molina et al. (2011). Cleaned amplicons were submitted for sequencing at the Ohio Agricultural Research and Development Center (OARDC)

Molecular and Cellular Imaging Center (MCIC) in Wooster, OH. Sequence was trimmed with codon-code aligner and then blasted against the sequences in the NCBI database.

Sequence identities of 99% were considered a match for species identification. The isolates used in the seed plate assay and the cup assay are listed in Table 2.4 and Table

2.5, respectively.

Seed plate assay to evaluate seed treatments

Three commercial formulations of seed treatments were compared to nontreated seed in two laboratory based assays (Table 2.2). The seed treatments were: i) ethaboxam combined with an insecticide, clothianidin plus metalaxyl and ipconazole, ii) thiamethoxam plus mefenoxam plus fludioxonil, and iii) pyraclostrobin plus metalaxyl plus fluxapyroxad plus an insecticide imidacloprid (Table 2.2). Seed of the cultivar

Kottman, which is susceptible to both Pythium spp. and some isolates of Phytophthora spp., was treated by Valent at Leland, MS. Seed was treated at the commercial rate for each fungicide listed in Table 2.2. Nontreated seeds were surface sterilized in a 10% bleach solution for 1.5 minutes, rinsed under running deionized water for one minute, and allowed to dry thoroughly before use.

27

The seed plate assay was used to compare the efficacy of the fungicide seed treatment and the relative pathogenicity of the isolates as previously described by Broders et al. (2007a). Briefly, Pythium and Phytopythium isolates were grown on 2% potato carrot agar (PCA) for 3 days at 20ºC in darkness. Phytophthora isolates were grown on

PCA for 4 days at 25ºC in the dark on 2% PCA. Mycelial plugs, 5 mm in diameter, were aseptically removed from the growing margins of the culture and transferred to the center of a 100 mm 1.2% PCA plate. Pythium and Phytopythium isolates were allowed to colonize the plates for 3 days at 20ºC and Phytophthora isolates were grown for 7 days at

25°C. The seeds (n=10) were pushed into the agar, evenly spaced, approximately 1 cm from the edge of the plate. Plates were incubated in the dark for 7 days. Data were collected on the percent germination and disease incidence (DI) for each plate. Seeds were only considered germinated if the radicle was >1 cm. A score was assigned to the ten seeds on each plate based on the following ordinal scale: 0 = 100% germination with no infection; 1 = 70-99% germination with lesions on the roots; 2 = 30-69% germination with lesions on the roots; 3 = 0-29% germination with lesions on the roots.

There were three replicate plates of each seed treatment by isolate combination and the experiment was replicated twice, for a total of six replicate plates for each treatment by isolate combination. Plates within an experiment were arranged in a randomized complete block design within the incubator. Isolates recovered from the field from the years 2014-2016 were used in this experiment. Isolates that did not cause seed rot were not included in the pathogenicity or fungicide efficacy analysis. A total of 44 oomycete isolates were pathogenic on nontreated soybean seed were used in this

28

experiment (Table 2.4). Subsets of the isolates were evaluated against the fungicide seed treatments over time in an incomplete block design for a total of 14 subsets, with a range of four to ten isolates evaluated in each experiment. An isolate of Py. ultimum var. ultimum (Miami 137) was used in each experiment to verify the responses of the isolates were the same across all experiments. All data were combined for final analysis. The data collected from this experiment were analyzed using a nonparametric relative marginal effects analysis as described by Shah and Madden (2004). Data were analyzed with

PROC MIXED in SAS (SAS Institute, Cary, NC).

The seed plate assay data for the six Py. ultimum var. ultimum isolates (Table 2.4) were also analyzed in a separate nonparametric relative marginal effects analysis as described by Shah and Madden (2004) with PROC MIXED in SAS (SAS Institute, Cary,

NC), in order to identify variation within a species.

Cup assay to evaluate seed treatments

The same treatments were evaluated against 11 isolates of Pythium spp., 3 isolates of Phytopythium spp., and 1 isolate of Ph. sansomeana (Table 2.5) using a cup assay modified from Ellis et al. (2013). Briefly, isolates were grown on potato carrot agar

(PCA) were grown at 20°C for three to four days. Eight 10-mm plugs of each isolate were taken from the growing margin and transferred to a sterilized Spawn bag (Myco

Supply, Pittsburgh, PA) containing 950 ml play sand (Quikrete, Ravenna, Ohio), 50 ml corn meal (Quaker Company, Chicago, IL), and 250 ml deionized water. The Spawn bags were closed with an electrical-impulse sealer (Harbor Freight Tools, Camarillo, CA)

29

to prevent contamination. These bags were grown at room temperature for ten days and shaken every other day to ensure the pathogen grew uniformly. The inoculated sand- cornmeal mixture was then mixed with fine vermiculite in a 1:4 ratio and placed into 600 ml polystyrene cups in a growth chamber at 20ºC, 16 hr light, and 60% RH. The cups were regularly watered with deionized water to provide a conducive environment for disease development. The cups were arranged in a randomized complete block design with 3 replications. Each cup had eight seeds placed directly on the inoculum, then covered with 100 ml of coarse vermiculate. The plants were grown for 14 days, then each cup had the plants removed and the roots washed. Data collected included: a visual root rot score, percent seed germination, total plant weight, and adjusted root weight. The root rot score ranged from 1-5, where 1 = no lesions on roots, 2 = 1-20% root have lesions, 3

= 20-70% roots have lesions, 4 = 71-100% root have lesions, and 5 = no germination of seed (Figure 2.1). Adjusted root weight was calculated by dividing the total root weight of the plants in a single cup by the number of plants.

Data analysis of the cup assay root rot scores used a nonparametric relative marginal effects analysis as described by Shah and Madden (2004) with PROC MIXED in SAS. All other data was analyzed with PROC GLM in SAS.

30

Results

Field trials

Fields were considered to have disease pressure if there were significant differences in plant population at the V1/V2 or V3/V4 stage between treated seed and nontreated control seed as well as observations of seedlings with damping-off (data not shown). Fungicide, cultivar, and fungicide by cultivar effects for each separately analyzed environment are listed in Tables 2.6-2.9. The location, cultivar, fungicide, and each corresponding interaction effects for the analysis across all 2016 locations are listed in Table 2.10.

There were significant differences for early plant population (V1/V2 stage) at the

NWB-14, DEF-15, NWB-16, VW-16 and SNY-16 environments (Tables 2.6 and 2.8), so these fields were considered to have high levels of disease pressure. The variation in disease pressure across the different environments was due in part to differences in total precipitation during the 14 days after planting (Table 2.1). Rainfall alone contributed to conducive environments for the development of seedling blight. However, the environment NWB-15 had flooding injury due to 5.7 in. of combined natural and irrigated precipitation and lost most of the plants in the study for the one cultivar Sloan.

The 2014 field studies indicated that the plant population and yield (kg/ha, Bu/A) were significantly impacted by cultivar and fungicide seed treatment for the NWB-14 environment (Tables 2.6 and 2.7). At NWB-14 for both Conrad and Sloan, the combination of metalaxyl plus ethaboxam resulted in significantly higher yield when compared to metalaxyl alone (6.2 µg active ingredient/seed) or mefenoxam alone (23.0

31

µg a.i./seed), as well as other seed treatment combinations based on mean separation with

Fishers Protected LSD (P<0.05) (Tables 2.11 and 2.12). For the cultivar Conrad, the higher rate of mefenoxam alone (23.0 µg a.i./seed) was not significantly different from both combinations of ethaboxam plus metalaxyl (Table 2.12). Interestingly, the difference or loss of yield due to seed and seedling rot at NWB-14 was 57.9% and 90.8% for Sloan and Conrad, respectively. For the second environment in 2014, VW-14, there was a significant effect of fungicide seed treatment for yield (P=0.047) (Table 2.7).

However, only mefenoxam alone at the highest rate (23.0 µg a.i./seed) resulted in significantly higher yield when compared with the nontreated seed for both cultivars

(Tables 2.13-2.14).

Since most seed treatments are sold and marketed as mixtures, the 2015 evaluations focused on the comparison of three formulations on four cultivars. There was a cultivar by fungicide interaction in DEF-15 for stand at stages V1/V2 and V3/V4, and for yield (Tables 2.6 and 2.7). For seed of cultivar Conrad, only the seed treated with ethaboxam plus metalaxyl (Table 2.2) had the mean early plant populations significantly differ from the nontreated seed (Table 2.15). For the cultivar Sloan, there was a similar trend for mean early plant populations, but both the ethaboxam plus metalaxyl and the pyraclostrobin plus metalaxyl seed treatment was also significantly higher than the nontreated control (Table 2.16). The cultivars with resistance to Ph. sojae, Kottman and

Lorain, had significantly higher early plant populations with ethaboxam plus metalaxyl and mefenoxam plus thiamethoxam plus fludioxonil when compared to the nontreated

(Tables 2.17 and 2.18). However, the cultivars Kottman and Lorain did not have any

32

significant differences between any of the seed treatments for yield in DEF-15 (Tables

2.17 and 2.18). At NWB-15 there was not a significant fungicide effect on yield (Table

2.7); the analysis broken down by each of the cultivars is recorded in Tables 2.19-2.21.

There was a significant interaction for fungicide seed treatment and cultivar by location, thus the effect of cultivar and the fungicide seed treatment was analyzed separately for all four locations in 2016 (Table 2.10). The fungicide seed treatment, metalaxyl plus ethaboxam (Table 2.2), had a significant effect on mean early plant population and yield at 3 of the 4 environments (Table 2.10). For NWB-16, VW-16, and

SNY-16, seed treated with ethaboxam plus metalaxyl fungicide seed treatment resulted in significantly greater stands when compared with the nontreated seed taken at stages

V1/V2, V3/V4, and R8 (Tables 2.8 and 2.9).

Of all the 2016 field sites, only NWB-16 had a significant fungicide effect on yield (P=0.0002) (Table 2.9). For early stand (V1/V2 stage), the seed treated with the ethaboxam plus metalaxyl fungicide resulted in significantly greater stand when compared with the nontreated control seed (P=0.0002) (Figure 2.2). Overall, the treated seed also had a significantly increased yield when compared with the nontreated control seed (Figure 2.3). The yield at the DEF-16 environment was very low; this location received heavy rainfall 3 to 4 week after planting which may have impacted the overall study.

33

Seed plate assay

There was a wide range of mean seed rot severity on the nontreated seed, which indicates the relative pathogenicity for each oomycete species (Figure 2.4). Pythium ultimum var. ultimum, Py. diclinum, Py. cryptoirregulare, and Pp. helicoides had the greatest degree of seed rot on the nontreated Kottman seed. Isolates of Ph. sojae and Ph. sansomeana caused a relatively low amount of seed rot, which may be an indication of the level of resistance of the cultivar Kottman to both of these species. Kottman has resistance genes Rps1k and Rps3a (Dorrance et al., 2009). Only the isolates that had a relative marginal effect score of >0.30 on the nontreated seed were examined for the effects of seed treatments.

There was a significant effect of fungicide seed treatment (P<0.0001) and oomycete species (P<0.0001) on seed rot severity, as well as a species by fungicide seed treatment interaction (P<0.0001). The seed treated with any one of the three fungicide seed treatments had significantly less (P<0.05) seed rot when compared with nontreated seed for isolates of 19 species of oomycetes (Figures 2.5-2.7).

There was a significant treatment by species interaction (P<0.0001); for the species Py. ultimum var. ultimum, Py. oopapillum, and Pp. helicoides, the seed treatments comparatively differed in their efficacy (Figure 2.7). Pp. helicoides had a significant reduction in seed rot with all three of the seed treatments, but CruiserMaxx Advanced was less effective than Acceleron or Intego Suite (Figure 2.7). Py. oopapillum also had a significant reduction in seed root for each treatment, but Intego Suite and CruiserMaxx

Advanced were less effective in preventing seed rot than Acceleron (Figure 2.7).

34

Similarly, Py. ultimum var. ultimum had a reduction in seed rot with each of the seed treatments, but CruiserMaxx Advanced and Acceleron were less effective than Intego

Suite (Figure. 2.7).

Within the six screened Py. ultimum var. ultimum isolates, there was a differential response to the three fungicide seed treatments (Figure 2.8). Two of the included isolates did not have a significant reduction in seed rot when treated with CruiserMaxx

Advanced. In all six of the included isolates, Intego Suite resulted in a significantly reduced seed rot when compared with the nontreated control seed.

Cup assay

There was a wide range of adjusted root weight values for each of the oomycete species for the nontreated seed, which is the primary indicator of the relative pathogenicity for each oomycete isolate (Figure 2.9). There was a significant effect of isolate on the adjusted root weight (P=0.0009), though 10 of the isolates used in the cup assay did not have a significantly different adjusted root weight compared to the uninoculated control (Figure 2.9). The isolates of Pythium ultimum var. ultimum, Py. torulosum, Py. irregulare, and Pp. helicoides had root rot on the nontreated seed; the nontreated seed either did not germinate or the seedlings damped-off (Figure 2.9). One isolate of Ph. sansomeana resulted in a significantly higher adjusted root weight when compared with the uninoculated control (Figure 2.9). Only the isolates that had a significantly lower adjusted root weight than the noninoculated control were analyzed for fungicide effects.

35

There was a significant effect (P<0.0001) of fungicide seed treatment and an isolate by treatment interaction (P=0.0328) on adjusted root weight. Each of the fungicides significantly increased the adjusted root weight when compared with the nontreated control for each of the four highly pathogenic species, with the exception of

CruiserMaxx Advanced to Py. irregulare (Figure 2.10). The Acceleron seed treatment provided the most protection against root rot from the species Py. irregulare and Py. ultimum var. ultimum. For each of the four pathogenic species in the cup assay, Py. irregulare, Py. ultimum var. ultimum, Py. torulosum, and Pp. helicoides, the Intego Suite seed treatment resulted in a significantly increased adjusted root weight when compared with the nontreated control seed (Figure 2.10).

36

Discussion

In Ohio and other soybean production areas, seedling diseases are typically managed using a combination of fungicide seed treatment and cultivars with host resistance (Bradley, 2008; Esker and Conley, 2012; Dorrance et al., 2009; Gaspar et al.,

2017; Urrea et al., 2013). In recent years, some oomycetes species that cause seed rot and damping off have been found that are insensitive to metalaxyl and strobilurins (Broders et al., 2007a; Dorrance et al., 2004; Matthiesen et al., 2016; Radmer et al., 2017; Rojas et al., 2017a,b).

This study investigated the efficacy of ethaboxam in combination with metalaxyl in different concentrations as a seed treatment in order to protect against seed and seedling rot in field and laboratory assays. Ethaboxam in a wettable powder form was initially used in Korea to manage oomycete pathogens that cause downy mildew, potato late blight, and pepper Phytophthora blight starting in the late 1990s (Kim et al., 1999; Kim et al., 2003). In recent years, ethaboxam has been labeled for soybean seed treatment in the US.

Pathogen population and diversity are highly influenced by environmental conditions such as the level of soil saturation, soil type, nutrient availability, amount of crop debris, and degree of tillage (Bescansa et al., 2006; Broder et al., 2009; Kirkpatrick et al., 2006; Workneh et al., 1999; Zitnik-Anderson et al., 2017). This environmental variation plays a part in the differences observed in the fungicide effects across the field location. In particular, the amount of rainfall in the 14 days after planting has a significant impact on the amount of oomycete disease pressure, as noted in prior field

37

seed treatment studies (Bradley, 2008; Dorrance et al., 2009). Four out of the five field sites that received greater than 1.0 in. of water 14 dap (days after planting) had a significant fungicide effect on yield. Since the efficacy of fungicide seed treatments can only be determined with adequate disease pressure, efforts should be taken to properly irrigate research field sites to encourage pathogen development. However, overwatering or flooding may mask any effects by a fungicide or otherwise impact the study (Henshaw et al., 2007a,b; Kirkpatrick et al., 2006; VanToai et al., 1994). This can be seen in the field site NWB-15 with 5.7 in. of water 14 dap and in VW-16 and DEF-16 with nearly

19.0 in. water total in the growing season; despite these fields being having a history of

Pythium and Phytophthora infection, there were no significant fungicide effects on yield.

The field sites in 2014 compared ethaboxam and metalaxyl combinations to metalaxyl alone. The metalaxyl treatment (6.0 µg a.i./seed) did not protect against seed and root disease for both cultivars Conrad and Sloan. This finding is consistent with previous reports of low rates of metalaxyl having limited efficacy towards Ph. sojae

(Dorrance and McClure, 2001). Additionally, this supports previous research that there is the emergence of populations of Pythium with insensitivity to this fungicide (Broders et al., 2007a; Dorrance et al., 2004; Matthiesen et al., 2016; Radmer et al., 2017; Rojas et al., 2017a,b).

At the DEF-15 field site, there was a significant fungicide by cultivar interaction;

Conrad and Sloan exhibited yield differences among fungicide treatments whereas cultivars Lorain and Kottman did not. This difference in fungicide effect may be due to the host resistance of Lorain and Kottman performing better against the particular

38

pathogen population in that environment. Kottman has resistance genes Rps1k and Rps3a

(Dorrance et al., 2009) and Lorain has resistance gene Rps1c (OSU-OARDC), and both of these cultivars have moderately high levels of partial resistance to Pythium ultimum var. ultimum, Py. ultimum var. sporangiiferum, and Py. irregulare (Balk, 2014; Dorrance lab, unpublished data). Comparatively, Conrad and Sloan have relatively low amounts of partial resistance to Pythium spp. and no Rps genes for Ph. sojae.

The 2016 locations also evaluated cultivar by fungicide seed treatment interactions; the plant populations were higher for seed treatment at three of the four locations and there was a fungicide effect on yield at only one environment (NWB-16).

Soybean cultivars with different levels of resistance to Ph. sojae and Pythium spp. did not respond the same to the seed treatment at each location (Table 2.9). In a highly favorable environment for seedling disease, the more susceptible cultivars performed better with the ethaboxam seed treatment when compared to the nontreated seed and there were no significant differences between treatments for cultivars with more resistance.

Overall, the cultivars treated with Inovate Pro plus Intego Solo had significantly greater yield (P=0.0002) and early stand at V1/V2 (P<0.0001) when compared with the nontreated seed. There was also a significant cultivar effect for both yield (P<0.0001) and early stand (PP=0.0206). The NWB-16 field environment had the more susceptible cultivars perform better with the ethaboxam seed treatment when compared to nontreated seed; there was no significant differences between treatments for the more resistant cultivars (Figures 2.2 and 2.3). Since the effects of fungicides were more easily seen on the cultivars with greater susceptibility to oomycete pathogens, future field studies should

39

include cultivars with a range of level of host resistance, taking care to include highly susceptible cultivars. This study is in agreement with other studies that suggest that cultivar selection should be used in combination with fungicide use to protect against fungi and oomycete pathogen diseases (Dorrance et al., 2009; Kandel et al., 2016;

Matthiesen et al., 2016; Marburger et al., 2014).

Overall, our findings from the field assays support prior studies that found that fungicide seed treatment have stand- and yield-protecting benefits against a variety of fungal and oomycete pathogens (Bradley, 2008; Esker and Conley, 2012; Gaspar et al.,

2014; Gaspar et al., 2017; Guy and Oplinger, 1989; Guy et al., 1998; Poag et al., 2005;

Urrea et al., 2013).

A seed plate assay is used to determine the effect of fungicide treated seed on a single isolate, compared to nontreated seed with the same isolate. This method allows commercial fungicide combinations to be screened with any pathogenic isolate. The seed plate assay was used previously to evaluate the relative pathogenicity of each oomycete species on nontreated seed (Broders et al., 2007a), as well as to determine the effect of temperature on Pythium spp. aggressiveness (Matthiesen et al., 2016). Similar findings were reported in this study as there was a wide range of pathogenicity among the

Phytophthora, Pythium, and Phytopythium isolates recovered from these same fields, as well as a varied response to the seed treatments within the screened Py. ultimum var. ultimum isolates. All pathogenic isolates had a reduction in seed rotting ability for each of the fungicide seed treatments compared to the nontreated seed. For the three Pythium species that had a differential response to the different fungicides, the treatment

40

containing ethaboxam (Intego Suite) provided equal or greater protection against seed rot when compared with the Acceleron package or CruiserMaxx Advanced.

The majority of isolates screened in the cup assay were also screened in the plate assay, though these assays differed in their assessment of isolate pathogenicity. All of the included isolated in the seed plate assay had a significantly greater degree of seed root than the uninoculated control, indicating that each of the isolates were pathogenic on the soybeans. However, the cup assay only identified some isolates of Py. ultimum var. ultimum, Py. irregulare, Py. torulosum, and Pp. helicoides that resulted in significantly decreased adjusted root weight when compared with the uninoculated control. Both assays identified P. ultimum var. ultimum and Pp. helicoides as highly pathogenic species. Py. torulosum had opposite results across the laboratory assays; the cup and plate assay identified it as strong and weakly pathogenic, respectively. The cup assay may be more comparable to field studies; the plants are allowed to mature further and the greenhouse mimics field conditions more closely than a Petri dish. However, the seed plate assay still provides invaluable information on how a single isolate reacts to a fungicide seed treatment.

Research from both the field studies and the laboratory assays suggest that ethaboxam can be used in combination to manage a variety of oomycete species

(Phytophthora, Pythium, and Phytopythium) commonly isolated from Ohio fields that are able to cause seedling damping off on soybean. Results indicate that the active ingredient ethaboxam has equal or greater effects in controlling these species when compared to other commonly used active ingredients, such as mefenoxam, metalaxyl, and strobilurins

41

used in combination in commercially available fungicide seed treatments. Due to the low risk of pathogens developing resistance towards ethaboxam and efficacy in controlling multiple oomycete species, this active ingredient could be a valuable disease management strategy for soybean seedling disease in Ohio.

Overall, the results from these studies indicate that fungicide seed treatments containing ethaboxam have comparable efficacy to other commercially available fungicide seed treatments tested in this study. Cultivar selection was also shown to influence yield, with more susceptible cultivars benefiting more (more yield protection) from fungicide seed treatments under heavy disease pressure than more resistant cultivars. These field experiments highlight how both cultivar selection and fungicide treatment must be taken into consideration when managing seed and root disease. Ideally, both a resistant cultivar and an effective fungicide seed treatment should be used in order to protect yield from soybean seed and root rot pathogens.

42

Table 2.1. Planting and harvest dates, precipitation, seeding rate, total row length counted for stand counts, and soil type for the field

trials used to evaluate fungicide seed treatments in Ohio during 2014-2016. Boxes with a “-“ indicate there was no irrigation at that

field site.

Total Seeding Total row No. Planting Harvest Irrigation Rainfall Precip.c rate length Env.a reps date date (in.) 14-dapb (in.) (in.) (seeds/ft) counted (ft.) Soil Type VW-14 5 26-May 13-Oct - 1.90 17.2 8.0 30.0 Latty Silty clay NWB-14 5 3-Jun 24-Oct 2.92 4.10 12.5 4.1 12.0 Hoytville clay NWB-15 4 22-Jun 20-Oct 1.60 5.71 18.8 6.0 12.0 Hoytville clay DEF-15 4 26-May 27-Sep - 3.23 19.4 8.0 30.0 Silty clay

43

NWB-16 5 1-Jun 17-18-Oct 2.8 3.25 13.49 5.0 12.0 Hoytville Clay Wooster Riddles SNY-16 6 20-May 6-Oct - 0.67 13.95 4.1 12.0 Silt Loam VW-16 6 25-May 8-Oct - 0.97 18.86 4.1 12.0 Latty Silty Clay DEF-16 6 26-May 15-Oct - 0.97 18.86 4.1 12.0 Paulding Clay a Indicates the field location and year (environment). b Rainfall and irrigation combined in the 14 days after planting (dap). c Total precipitation indicates the total rainfall and irrigation combined from planting date to harvest date.

Table 2.2. Fungicide seed treatments evaluated in each field environment. Select fungicides were also screened in the laboratory seed plate assay and the growth chamber cup assay.

Active ingredient and rate Environment (µg a.i./seed) Trade name NWB-14 Inovate Pro (2.47 lb Clothianidin (75.1) + ipconazole VW-14 a.i./gal) + Sebring (2.65 lb (3.8) + metalaxyl (6.2) a.i./gal) NWB-14 VW-14 Clothianidin (75.1) + ipconazole Inovate Pro (2.47 lb NWB-16 (3.8) + metalaxyl (3.2) + a.i./gal) + Intego Solo (3.2 VW-16 ethaboxam (12.0) lb a.i./gal) SNY-16 DEF-16 Inovate Pro (2.47 lb Clothianidin (75.1) + ipconazole NWB-14 a.i./gal) + Sebring (2.65 lb (3.8) + metalaxyl (6.2) + VW-14 a.i./gal) + Intego Solo (3.2 ethaboxam (12.0) lb a.i./gal) Thiamethoxam (77.8) + NWB-14 CruiserMaxx Plus (2.47 lb mefenoxam (11.7) + fludioxonil VW-14 a.i./gal) (3.8) Thiamethoxam (77.8) + CruiserMaxx Plus (2.47 lb NWB-14 mefenoxam (23.0) + fludioxonil a.i./gal) + Apron XL (3.0 VW-14 (3.8) lb a.i./gal) Evergol Energy (1.47 lb Metalaxyl (24.9) Prothioconazole NWB-14 a.i./gal) + Allegiance-FL (8.1) + Penflufen (4.0) + VW-14 (2.65lb a.i./gal) + Gaucho imidacloprid (101.0) (5 lb a.i./gal) Acceleron package: NWB-14 Acceleron DX 109 (0.8 fl VW-14 Pyraclostrobin (16.9) + metalaxyl oz/ cwt) + Acceleron DX NWB-15 (13.2) + fluxapyroxad (8.3) + 309 (0.4 fl oz/cwt) + DEF-15 imidacloprid (127.0) Acceleron DX 612 (0.240 Seed plate assay fl oz/ cwt) + Acceleron IX Cup assay 409 (2.00 fl oz/ cwt) NWB-15 Clothianidin (81.0) + ipconazole DEF-15 Intego Suite (3.37 fl oz/ (4.0) + metalaxyl (3.2) + Seed plate assay cwt) ethaboxam (12.0) Cup assay Continued

44

Table 2.2 continued.

Active ingredient and rate Environment (µg a.i./seed) Trade name NWB-15 Thiamethoxam (77.9) + DEF-15 CruiserMaxx Advanced mefenoxam (11.7) + fludioxonil Seed plate assay (3.2 fl oz/ cwt) (3.8) Cup assay NWB-16 Clothianidin (75.1) + ipconazole Inovate Pro (2.47 lb VW-16 (3.8) + metalaxyl (3.2) + a.i./gal) + Intego Solo (3.2 SNY-16 ethaboxam (12.0) lb a.i./gal) DEF-16

45

Table 2.3. List of cultivars used in each field site and year in order to determine the effect of cultivar choice on management of seed and root pathogens in Ohio fields.

Environment Cultivars planted NWB-14 Conrad, Sloan VW-14 NWB-15 Conrad, Sloan, Kottman, Lorain DEF-15 NWB-16 DEF-16 Conrad, Sloan, Kottman, Lorain, Clermont, VW-16 Dennison, Streeter SNY-16

46

Table 2.4. List of Pythium (Py.), Phytophthora (Ph.), and Phytopythium (Pp.) isolates included in the plate assay used to evaluate the efficacy of three fungicide seed treatments and determine the seed-rotting ability of each isolate.

Oomycete Spp. Isolate Code Year Location Py. attrantheridium MDa4.2b 2015 Defiance MVa12.5b 2015 Van Wert 17SNY111.3- 2017 Snyder Py. cryptoirregulare Erie 2-6-4 2006/2007 Erie Co. Craw 1-1-10 2006/2007 Crawford Co. Py. diclinum MVa3.5b 2015 Van Wert Py. dissotocum MDa2.1a 2015 Defiance D411.1 2016 Defiance N405.1 2016 NWB Py. heterothallicum MDa6.5b 2015 Defiance Py. inflatum MDa11.2a 2015 Defiance MVa3.3a 2015 Van Wert MVa5.3a 2015 Van Wert Py. irregulare Mont 1-4-7 2006/2007 Montgomery Co. Sand 1-6-22 2006/2007 Sandusky Co. Wood 1-6-9 2006/2007 Wood Co. Py. lutarium MVa20.4b 2015 Van Wert MVa23.5b 2015 Van Wert Py. nodosum MWa9.1a1 2015 Waterman Py. oopapillum MDa2.4b 2015 Defiance MVa1.4a 2015 Van Wert N110.1 2016 NWB Py. pleroticum MVa13.3b 2015 Van Wert Py. torulosum MDa4.4b 2015 Defiance Co. MVa7.2a 2015 Van Wert Py. ultimum var. ultimum MDa5.2b 2015 Defiance Co. Miami137 2006/2007 Miami Co. MVa1.2a 2015 Van Wert MVa21.4b 2015 Van Wert N508.1 2016 NWB N201.2(2) 2016 NWB Continued

47

Table 2.4. continued.

Oomycete Spp. Isolate Code Year Location Ph. sansomeana MVa2.5b 2015 Van Wert MVa8.4b 2015 Van Wert 17SNY208.5- 2017 Snyder 17CLN111.2 2017 Clinton Co. Ph. sojae OH.12108.06.03 2012 Van Wert R25_1-97-01 Standard isolate Pp. helicoides MVa14.4b 2015 Van Wert D502.1 2016 Defiance Pp. litorale D206.3 2016 Defiance Pp. mercurial N109.1(2) 2016 NWB Pp. vexans MWa12.1a 2015 Wood Co. 17SNY101.1+ 2017 Snyder 17CLN111.1 2017 Clinton Co.

48

Table 2.5. Isolates used in the growth chamber cup assay to evaluate the efficacy of fungicide seed treatments and determine the root-rotting ability of each isolate.

Isolate Oomycete spp. code Year Location Py. heterothallicum MDa6.5b 2015 Defiance Py. irregulare Br_235 2006/2007 Brown Co. Gr_151 2006/2007 Greene Co. Py. lutarium MVa23.5b 2015 Van Wert Py. nodosum MWa9.1a1 2015 Wood Co. Py. oopapillum MVa1.4a 2015 Van Wert Py. torulosum MDa4.4b 2015 Defiance MVa12.3a 2015 Van Wert MVa5.4a 2015 Van Wert Py. ultimum var. ultimum MVa1.2a 2015 Van Wert MVa21.4b 2015 Van Wert Ph. sansomeana MVa2.5b 2015 Van Wert Pp. helicoides MVa14.4b 2015 Van Wert MVa5.1a 2015 Van Wert Pp. vexans MWa12.1a 2015 Wood Co.

49

50

Figure 2.1. Root rot scoring system used in the growth chamber cup to evaluate the pathogenicity and aggressiveness of isolates of

Pythium recovered from soybean in Ohio. For this score 1 = healthy roots with no infection, 2 = lesions covering 1-25% of roots, 3 =

26-75% of roots with lesions, 4 = 76-100% of roots with lesions, and 5 = total colonization of the seed with no germination.

Table 2.6. P-values for cultivar (cult.), fungicide seed treatment (fung.), and cultivar by fungicide interactions (CxF) for stand at stage

V1/V2 and stand at stage V3/V4 data (plants/Ha) for the 2014-2015 locations. Degrees of freedom (df) for cultivar (cult.), fungicide

seed treatment (fung.), and cultivar by fungicide interactions (CxF) are shown. Inches of precipitation (natural and irrigated combined)

in the 14 days after planting (dap) are shown for each location (Loc.).

Environment df V1/V2 V3/V4 rain 14 Loc. dap Cult Fung CxF Mean Cult. Fung. CxF Mean Cult. Fung. CxF

51 VW-14 1.9'' 1 7 7 268,471.8 0.5789 0.9486 0.5650 - - - - a a a a

NWB-14 4.1'' 1 7 7 181,995.8 0.0075 <0.0001 0.3721 182,301.2 0.0059 <0.0001 0.3008 a a a a a a DEF-15 3.2'' 3 3 9 225,238.0 <0.0001 <0.0001 0.0169 224,991.1 <0.0001 <0.0001 0.0120 b a NWB-15 5.7'' 2 3 6 262,004.0 0.6600 0.8213 0.0598 257,645.0 0.6540 0.7308 0.0462 - indicates no data collected. a indicates significance at 95% confidence using ANOVA. b indicates significance at 90% confidence using ANOVA.

Table 2.7. Significance of cultivar (cult.), fungicide seed treatment (fung.), and cultivar by fungicide interactions (CxF) for plant stand

at stage R8 (plants/Ha) and yield data (Bu/A) for the 2014-2015 field locations.

Environment R8 Bu/A rain Location 14 dap Mean Cult. Fung. CxF Mean Cult. Fung. CxF a a VW-14 1.9'' 282,277.7 0.0438 0.8109 0.9572 46.5 0.1873 0.0470 0.9877 a a a a NWB-14 4.1'' 178,255.6 0.0129 <0.0001 0.3800 23.2 0.0233 <0.0001 0.3592 a a a a DEF-15 3.2'' 242,769.5 0.5245 0.0141 0.1985 20.6 <0.0001 <0.0001 0.0179 b NWB-15 5.7'' 243,635.0 0.5196 0.5979 0.0600 31.9 0.3042 0.3686 0.2832 a indicates significance at 95% confidence using ANOVA. 52 b indicates significance at 90% confidence using ANOVA.

Table 2.8. P-values for cultivar (cult.), fungicide seed treatment (fung.), and cultivar by fungicide interactions (CxF) for stand at stage

V1/V2 and stand at stage V3/V4 data (plants/Ha) for the 2016 locations. Degrees of freedom (df) for cultivar (cult.), fungicide seed

treatment (fung.), and cultivar by fungicide interactions (CxF) are shown. Inches of precipitation (natural and irrigated combined) in

the 14 days after planting (dap) are shown for each location (Loc.).

Environment df V1/V2 V3/V4 rain 14 Location dap Cult Fung CxF Mean Cult. Fung. CxF Mean Cult. Fung. CxF a a b a a b NWB-16 3.23'' 6 1 6 359,316.8 0.0206 <0.0001 0.0822 392,452.1 0.0015 0.0001 0.0920 53 DEF-16 0.97'' 6 1 6 215,472.8 0.4308 0.1655 0.4591 239,990.1 0.0136a <0.0001a 0.1540

a VW-16 0.97'' 6 1 6 303,945.7 0.1745 0.0064 0.3991 302,408.0 0.1956 0.2909 0.6601 a a SNY-16 0.67'' 6 1 6 166,068.1 0.6334 0.0138 0.9092 165,384.7 0.5337 0.0428 0.8956 a indicates significance at 95% confidence using ANOVA. b indicates significance at 90% confidence using ANOVA.

Table 2.9. Significance of cultivar (cult.), fungicide seed treatment (fung.), and cultivar by fungicide interactions (CxF) for plant stand

at stage R8 (plants/Ha) and yield data (Bu/A) for the 2016 field locations.

Environment R8 Bu/A rain 14 Location dap Mean Cult. Fung. CxF Mean Cult. Fung. CxF a a b a a NWB-16 3.23'' 377,678.4 0.0012 0.0003 0.0975 47.3 <0.0001 0.0002 0.1315 a a a DEF-16 0.97'' 222,881.6 0.0002 0.0181 0.3708 21.0 <0.0001 0.2239 0.3601 a VW-16 0.97'' 241,755.6 0.4258 0.3718 0.8218 37.8 <0.0001 0.4787 0.1421 b b b SNY-16 0.67'' 162,565.7 0.4993 0.0813 0.9643 37.6 0.0790 0.0742 0.7110 a indicates significance at 95% confidence using ANOVA. b

54 indicates significance at 90% confidence using ANOVA.

Table 2.10. Significance of cultivar, fungicide seed treatment, location, and interactions for stand at growth stages V1/V2, V3/V4, R8, and yield (Bu/A) for the 2016 locations.

Stand at Stand at V1/V2 V3/V4 Stand at R8 Yield Cultivar 0.0771b 0.0002a <0.0001a <0.0001a Location <0.0001a <0.0001a <0.0001a <0.0001a Fungicide <0.0001a <0.0001a <0.0001a 0.1757 Fungicide x Location 0.0639b 0.0037a 0.0362a 0.0070a Cultivar x Location 0.0263a <.0001a 0.0003a <0.0001a Cultivar x Fungicide 0.2726 0.1125 0.5153 0.8226 a indicates significance at 95% confidence using ANOVA. b indicates significance at 90% confidence using ANOVA.

55

Table 2.11. Comparison of different combinations on soybean population and yield in Northwest Branch 2014 (Cv. Sloan). Active

ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at each

growth stage (plants/Ha), as well as yield (Kg/H and Bu/A). Treatments are compared to each other separately for each growth stage

as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05).

Plant population / Ha Yield Rate Treatment (µg a.i./ seed) V1/V2 V3/4 R7 Kg/H Bu/A Metalaxyl1 6.2 71,758 c 76,063 c 78,934 b 1,073.7 cd 16.0 cd 56 Metalaxyl + ethaboxam1 3.0 + 12.0 299,948 a 302,818 a 285,237 a 2,275.8 a 33.8 a

Metalaxyl + ethaboxam1 6.0 + 12.0 289,902 a 294,207 a 281,650 a 2,200.2 a 32.7 a Mefenoxam2 11.7 68,888 c 71,758 c 66,376 b 775.7 d 11.5 d Mefenoxam2 23.0 134,905 bc 137,775 bc 71,758 b 1,113.6 cd 16.6 cd Metalaxyl + Prothioconazole 24.9 + 8.1 + 126,294 bc 130,599 bc 123,782 b 1,143.5 c 17.0 c + Penflufen + imidacloprid 4.0 + 101.0 Pyraclostrobin + metalaxyl + 16.9 + 13.2 + 152,127 b 150,691 b 141,722 b 1,631.8 b 24.3 b fluxapyroxad + imidacloprid 8.3 + 127.0 Nontreated check 0 87,545 bc 83,239 c 68,170 b 872.4 cd 13.0 cd Overall mean - 153,920.9 155,893.9 139,703.6 1,385.8 20.6 LSD (P < 0.05) - 66,329 66,990 77,329 357.9 5.3 1Seed also treated with clothianidin (75.1 µg/seed) and ipconazole (3.8 µg/seed) 2Seed also treated with thiamethoxam (77.8 µg/seed) and fludioxonil (3.8 µg/seed)

Table 2.12. Comparison of different fungicide combinations on soybean population and yield in Northwest Branch 2014 (Cv.

Conrad). Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant

population at each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth

stages as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05).

Plant population / Ha Yield Rate Treatment (µg a.i./ seed) V1/V2 V3/4 R7 Kg/H Bu/A Metalaxyl1 6.2 169,349 cd 172,219 bc 166,837 bc 1,186.4 d 17.6 d Metalaxyl + ethaboxam1 3.0 + 12.0 322,910 a 325,781 a 317,528 a 2,409.7 a 35.8 a

57 Metalaxyl + ethaboxam1 6.0 + 12.0 295,642 ab 302,818 a 308,559 a 2,461.7 a 36.6 a Mefenoxam2 11.7 166,478 cd 159,302 c 179,395 bc 1,331.6 cd 19.8 cd Mefenoxam2 23.0 254,023 abc 258,328 ab 260,122 ab 1,994.7 ab 29.7 ab Metalaxyl + Prothioconazole 24.9 + 8.1 + 179,395 cd 182,265 bc 175,807 bc 1,609.7 bcd 23.9 bcd + Penflufen + imidacloprid 4.0 + 101.0 Pyraclostrobin + metalaxyl + 16.9 + 13.2 + 196,616 bcd 166,837 bc 200,922 abc 1,878.9 abc 27.9 abc fluxapyroxad + imidacloprid 8.3 + 127.0 Nontreated check 0 96,156 d 99,026 c 98,667 c 1,033.8 d 15.4 d Overall mean - 210,071.1 209,385.7 213,479.6 1,738.3 25.8 LSD (P < 0.05) - 100,863 94,507 124,455 632.8 9.4 1Seed also treated with clothianidin (75.1 µg/seed) and ipconazole (3.8 µg/seed) 2Seed also treated with thiamethoxam (77.8 µg/seed) and fludioxonil (3.8 µg/seed)

Table 2.13. Comparison of different fungicide combinations on soybean population and yield in Van Wert 2014 (Cv. Sloan). Active

ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at each

growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth stages as well as

yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05).

Plant Population / Ha Yield Rate Treatment (µg a.i./ seed) V1/V2 R7 Kg/H Bu/A Metalaxyl1 6.0 272,961 a 261,468 a 2,937.2 ab 43.7 ab Metalaxyl + ethaboxam1 3.0 + 12.0 265,203 a 261,756 a 3,009.2 ab 44.8 ab

58 Metalaxyl + ethaboxam1 6.0 + 12.0 277,846 a 267,789 a 2,949.4 ab 43.8 ab Mefenoxam2 11.7 265,491 a 261,756 a 2,919.3 ab 43.4 ab Mefenoxam2 23.0 258,308 a 265,203 a 3,247.6 a 48.3 a Metalaxyl + Prothioconazole + 24.9 + 8.1 274,973 a 267,502 a 3,152.1 ab 46.9 ab Penflufen + imidacloprid +4.0 + 101.0 Pyraclostrobin + metalaxyl + 16.9 + 13.2 + 278,708 a 276,984 a 2,787.5 ab 41.4 ab fluxapyroxad + imidacloprid 8.3 + 127.0 Nontreated check 0 271,812 a 262,546 a 2,712.7 b 40.3 b Overall mean - 270,662.8 265,625.5 2,968.7 44.1 LSD (P < 0.05) - NS NS 524.19 7.8 1Seed also treated with clothianidin (75.1 µg/seed) and ipconazole (3.8 µg/seed) 2Seed also treated with thiamethoxam (77.8 µg/seed) and fludioxonil (3.8 µg/seed)

Table 2.14. Comparison of different fungicide active ingredients on soybean population and yield in Van Wert 2014 (Cv. Conrad).

Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at

each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth stages as well

as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05).

Plant Population / Ha Yield Rate Treatment (µg a.i./ seed) V1/V2 R7 Kg/H Bu/A Metalaxyl1 6.0 262,330 a 284,095 a 3,187.8 ab 47.4 ab Metalaxyl + ethaboxam1 3.0 + 12.0 268,364 a 277,271 a 3,256.5 ab 48.4 ab

59 Metalaxyl + ethaboxam1 6.0 + 12.0 268,077 a 273,680 a 3,037.9 ab 45.2 ab Mefenoxam2 11.7 284,167 a 294,511 a 3,324.6 a 49.4 a Mefenoxam2 23.0 273,249 a 280,863 a 3,346.5 a 49.8 a Metalaxyl + Prothioconazole + 24.9 + 8.1 + 266,640 a 284,814 a 3,241.1 ab 48.2 ab Penflufen + imidacloprid 4.0 + 101.0 Pyraclostrobin + metalaxyl + 16.9 + 13.2 + 267,789 a 293,074 a 2,910.7 b 43.3 b fluxapyroxad + imidacloprid 8.3 + 127.0 Nontreated check 0 257,158 a 271,166 a 2,863.3 b 42.6 b Overall mean - 268,471.8 282,434.3 3,127.5 46.5 LSD (P < 0.05) - NS NS 407.0 6.1 1Seed also treated with clothianidin (75.1 µg/seed) and ipconazole (3.8 µg/seed) 2Seed also treated with thiamethoxam (77.8 µg/seed) and fludioxonil (3.8 µg/seed)

Table 2.15. Comparison of different fungicide active ingredients on soybean population and yield in Defiance 2015 (Cv. Conrad).

Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at

each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth stages as well

as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05).

Plant Population / Ha Yield Rate Treatment (µg a.i./ seed) V1/V2 V3/V4 R7 Kg/H Bu/A Metalaxyl + ethaboxam + 3.0 + 12.0 + 253,926 a 260,750 a 267,574 a 18.1 a clothianidin + ipconazole 81.0 + 4.0 1,213.8 a

60 Mefenoxam + thiamethoxam 11.7 + 77.9 + 183,890 b 183,531 b 247,820 a 1,021.4 ab 15.2 ab + fludioxonil 3.8 Pyraclostrobin + metalaxyl + 16.9 + 13.2 + 194,665 ab 192,150 b 217,292 a 1,076.4 ab 16.0 ab fluxapyroxad + imidacloprid 8.3 + 127.0 Nontreated 0 151,925 b 151,925 b 195,742 a 952.7 b 14.1 b Overall mean - 196,101.5 197,088.9 232,107.0 1,066.1 15.9 LSD (P < 0.05) - 67,518 65,077 NS 213.9 3.2

Table 2.16. Comparison of different fungicide active ingredients on soybean population and yield in Defiance 2015 (Cv. Sloan).

Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at

each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth stages as well

as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05).

Plant population / Ha Yield Rate Treatment (µg a.i./ seed) V1/V2 V3/V4 R7 Kg/H Bu/A Metalaxyl + ethaboxam + 3.0 + 12.0 + 221,961 a 219,447 a 247,820 ab 1,381.9 a 20.5 a clothianidin + ipconazole 81.0 + 4.0

61 Mefenoxam + thiamethoxam 11.7 + 77.9 + 141,150 b 142,586 b 184,967 ab 901.3 cb 13.4 cb + fludioxonil 3.8 Pyraclostrobin + metalaxyl + 16.9 + 13.2 + 211,904 a 208,313 a 280,145 a 1,065.5 b 15.9 b fluxapyroxad + imidacloprid 8.3 + 127.0 Nontreated 0 107,389 b 102,361 b 159,826 b 740.3 c 11.0 c Overall mean - 170,601.0 168,176.5 218,190.0 1,022.3 15.2 LSD (P < 0.05) - 44,047 50,351 106,950 202.0 3.0

Table 2.17. Comparison of different fungicide active ingredients on soybean population and yield in Defiance 2015 (Cv. Kottman).

Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at

each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth stages as well

as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05).

Plant population / Ha Yield Rate Treatment (µg a.i./ seed) V1/V2 V3/V4 R7 Kg/H Bu/A Metalaxyl + ethaboxam + 3.0 + 12.0 + 281,581 a 272,961 a 271,166 a 1,559.3 a 23.2 a clothianidin + ipconazole 81.0 + 4.0

62 Mefenoxam + thiamethoxam 11.7 + 77.9 + 249,616 a 252,130 a 255,003 ab 1,346.2 a 20.0 a + fludioxonil 3.8 Pyraclostrobin + metalaxyl + 16.9 + 13.2 + 258,595 a 263,264 a 247,820 a 1,635.0 a 24.3 a fluxapyroxad + imidacloprid 8.3 + 127.0 Nontreated 0 211,545 b 210,827 b 199,334 b 1,523.8 a 22.7 a Overall mean - 250,334.3 249,795.5 243,331.0 1,516.1 22.5 LSD (P < 0.05) - 36,504 43,872 35,155 NS NS

Table 2.18. Comparison of different fungicide active ingredients on soybean population and yield in Defiance 2015 (Cv. Lorain).

Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant population at

each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth stages as well

as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05).

Plant population / Ha Yield Rate Treatment (µg a.i./ seed) V1/V2 V3/V4 R7 Kg/H Bu/A Metalaxyl + ethaboxam + 3.0 + 12.0 + 293,433 a 294,511 a 253,208 a 2,030.1 a 30.2 a clothianidin + ipconazole 81.0 + 4.0

63 Mefenoxam + thiamethoxam 11.7 + 77.9 + 286,250 a 284,454 a 276,553 a 1,931.5 a 28.7 a + fludioxonil 3.8 Pyraclostrobin + metalaxyl + 16.9 + 13.2 + 282,659 ab 280,863 a 307,082 a 1,952.2 a 29.0 a fluxapyroxad + imidacloprid 8.3 + 127.0 Nontreated 0 273,321 b 279,785 a 272,961 a 1,881.8 a 28.0 a 277,451. Overall mean - 283,915.8 284,903.4 1,948.9 29.0 0 LSD (P < 0.05) - 11,041 NS NS NS NS

Table 2.19. Comparison of different fungicide active ingredients on soybean population and yield in Northwest Branch 2015 (Cv.

Conrad). Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant

population at each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth

stages as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05).

Plant population / Ha Yield Rate Treatment (µg a.i./ seed) V1/V2 V3/V4 R7 Kg/H Bu/A Metalaxyl + ethaboxam + 3.0 + 12.0 + 267,896 a 258,328 a 251,152 a 2,141.0 a 31.8 a clothianidin + ipconazole 81.0 + 4.0

64 Mefenoxam + thiamethoxam 11.7 + 77.9 + 261,916 a 261,916 a 247,565 a 1,888.2 a 28.1 a + fludioxonil 3.8 Pyraclostrobin + metalaxyl + 16.9 + 13.2 + 203,314 a 198,530 a 181,787 a 1,595.3 a 23.7 a fluxapyroxad + imidacloprid 8.3 + 127.0 Nontreated 0 200,922 a 204,510 a 175,807 a 1,778.6 a 26.5 a Overall mean - 233,512.0 230,342.7 214,077.8 1,854.2 27.6 LSD (P < 0.05) - NS NS NS NS NS

Table 2.20. Comparison of different fungicide active ingredients on soybean population and yield in Northwest Branch 2015 (Cv.

Kottman). Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant

population at each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth

stages as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05).

Plant population / Ha Yield Rate Treatment (µg a.i./ seed) V1/V2 V3/V4 R7 Kg/H Bu/A Metalaxyl + ethaboxam + 3.0 + 12.0 + 310,353 a 310,353 a 258,328 a 2,502.7 ab 37.2 ab clothianidin + ipconazole 81.0 + 4.0

65 Mefenoxam + thiamethoxam 11.7 + 77.9 + 224,243 ab 224,243 ab 224,243 a 2,768.0 a 41.1 a + fludioxonil 3.8 Pyraclostrobin + metalaxyl + 16.9 + 13.2 + 251,152 ab 263,710 ab 245,771 a 2,227.9 ab 33.1 ab fluxapyroxad + imidacloprid 8.3 + 127.0 Nontreated 0 174,611 b 179,395 b 177,601 a 1,950.8 b 29.0 b Overall mean - 240,089.8 244,425.1 226,485.8 2,362.4 35.1 LSD (P < 0.05) - 86,426 88,213 NS 633.9 9.4

Table 2.21. Comparison of different fungicide active ingredients on soybean population and yield in Northwest Branch 2015 (Cv.

Lorain). Active ingredient rates are estimates based on label information. Overall mean across all treatments are shown for plant

population at each growth stage (plants/Ha) and yield (Kg/H and Bu/A). Treatments are compared to each other separately for growth

stages as well as yield data. Treatments sharing a letter indicate that they are not significantly different from each other (p<0.05).

Plant population / Ha Yield Rate Treatment (µg a.i./ seed) V1/V2 V3/V4 R7 Kg/H Bu/A Metalaxyl + ethaboxam + 3.0 + 12.0 + 218,861 a 213,480 a 206,304 a 1,931.4 a 28.7 a clothianidin + ipconazole 81.0 + 4.0

66 Mefenoxam + thiamethoxam 11.7 + 77.9 + 324,704 a 313,941 a 313,941 a 2,199.4 a 32.7 a + fludioxonil 3.8 Pyraclostrobin + metalaxyl + 16.9 + 13.2 + 306,765 a 299,589 a 281,650 a 1,994.7 a 29.7 a fluxapyroxad + imidacloprid 8.3 + 127.0 Nontreated 0 333,674 a 324,704 a 313,941 a 2,346.6 a 34.9 a Overall mean - 283,443.3 287,928.3 267,298.3 2,118.0 31.5 LSD (P < 0.05) - NS NS NS NS NS

600000

500000 y M Y g M Y A G B 400000 G a a h z Innovate Pro 300000 + Intego Solo Nontreated 200000

Plant population at V1/V2 (plants/Ha) V1/V2 at population Plant 100000

0 Clermont Conrad Dennison Kottman Lorain Streeter Sloan Cultivar

Figure 2.2. Early stand at stage V1/V2 (plants/Ha) of the seed planted at Northwest

Branch in 2016, with standard deviation shown. Bars sharing a letter are not significantly different from each other (p<0.05).

67

70 M Y g 60 A M A Y G a h 50 H a y Innovate Pro 40 y + Intego Solo

Nontreated

30 Yield (Bu/A) Yield

20

10

0 Clermont Conrad Dennison Kottman Lorain Streeter Sloan Cultivar

Figure 2.3. Yields (Bu/A) of the seed planted at Northwest Branch in 2016, with standard deviation shown. Bars sharing a letter are not significantly different from each other (p<0.05).

68

0.8 A A A AB BC BC BC BC 0.7 C CD 0.6 CD CDE DE DE DE 0.5 DE 0.4 DE E E 0.3 0.2 F

Relative marginal effect marginal Relative 0.1 0

Oomycete spp.

Figure 2.4. The relative pathogenicity of each Pythium (Py.), Phytophthora (Ph.), and

Phytopythium (Pp.) species evaluated in the seed plate assay, with standard error shown.

Results were calculated from the seed rot scores on the inoculated plates with nontreated seed (cv. Kottman). Bars sharing a letter are not significantly different from each other

(p<0.05). The higher the relative marginal effect, the greater the degree of seed rot.

69

1 A 0.9 A Pp. mercuriale (n=1) Py. attranthericium (n=3) 0.8 Pp. litorale (n=1 Py. cryptoirregulare (n=2) Pp. vexans (n=3) Py. diclinium (n=1) 0.7 0.6 B B 0.5 0.4 0.3 0.2 0.1 0 Relative marginal effectNontreated IntegoSuite Acceleron CruiserMaxx Nontreated IntegoSuite Acceleron CruiserMaxx seed Adv. seed Adv. 1 A 0.9 Py. dissotocum (n=3) A Py. lutarium (n=2) 0.8 Py. heterothallium (n=3) Py. nodosum (n=1) Py. inflatum (n=3) Py. pleroticum (n=1) 0.7 Py. irregulare (n=3) Py. torulosum (n=2) 0.6 B B 0.5 0.4 0.3 0.2 0.1 0 Nontreated IntegoSuite Acceleron CruiserMaxx Nontreated IntegoSuite Acceleron CruiserMaxx seed Adv. seed Adv. Seed Treatment

Figure 2.5. The response of the 27 isolates representing 14 oomycete species (n =

number of isolates per the species) of Pythium and Phytopythium to Intego Suite,

Acceleron, and CruiserMaxx Advanced fungicide seed treatments applied at the

commercial rate. All the oomycete species were analyzed together and no within species

differences were found within this group; data is separated into four separate graphs for

visual clarity. Bars sharing a letter are not significantly different from one another

(p<0.05). Standard error is shown. The larger the relative marginal effect, the greater the

degree of seed rot.

70

1 0.9 Ph. sonsomeana (n = 4) 0.8 Ph. sojae (n = 2) 0.7 0.6 0.5 0.4 0.3

Relative marginal effect marginal Relative 0.2 0.1 0 Nontreated seed IntegoSuite Acceleron CruiserMaxx Adv. Seed treatment

Figure 2.6. The response of the six isolates representing two Phytophthora species (n = number of isolates per the species) to Intego Suite, Acceleron, and CruiserMaxx

Advanced fungicide seed treatments applied at the commercial rate. All oomycete spp. were analyzed together and no within species differences were found within this group;

Phytophthora data is shown separately for visual clarity. Bars sharing a letter are not significantly different from one another (p<0.05). Standard error is shown

71

1 Nontreated seed IntegoSuite A a X Acceleron 0.8 CruiserMaxx Adv. B b 0.6 c bc C C YZ Y

0.4 Z Relative marginal effect marginal Relative 0.2

0 Pp. helicoides Py. oopapillum Py. ultimum var. ultimum Oomycete spp.

Figure 2.7. Oomycete species that exhibited a differential response when screened with the fungicide seed treatments. Seed treatments were compared to the nontreated control seed within each species. Bars with the same letter are not significantly different

(p=0.05). The larger the relative marginal effect, the greater the degree of seed rot.

72

1 A Y G a x 0.9 g AB y 0.8 B Z Z Z H 0.7 g B H yz 0.6 H z Nontreated 0.5 b h h IntegoSuite b b 0.4 Acceleron

0.3 CruiserMaxx Relative marginal effect marginal Relative

0.2

0.1

0 MDa5.2b MVa1.2a MVa21.4b Miami137 N201.2(2) N508.1 P. ultimum var. ultimum isolate

Figure 2.8. Response of each of the six Py. ultimum var. ultimum isolates to the three fungicide seed treatments used in the plate assay, with standard error shown. Within each isolate, the level of seed rot on the seed treatments was compared with the seed rot on the nontreated control seeds. Bars with the same letter are not significantly different

(p=0.05). The larger the relative marginal effect, the greater the degree of seed rot.

73

0.8 A 0.7 0.6 AB 0.5 CD BC 0.4 BC BC BC 0.3 CD CD CD 0.2 CD 0.1 Asjusted Asjusted root weight (g) D D D D D 0

Isolate

Figure 2.9. Relative pathogenicity of isolates of Pythium (Py.), Phytopythium (Pp.), and

Phytophthora (Ph.) tested in a growth chamber cup assay with surface sterilized soybean seed of the cultivar Kottman, with standard deviation shown. Results are based on the average adjusted root weights of the plants. Bars sharing a letter are not significantly different from each other. There was a significant difference in the pathogenicity of the oomycete species (P=<0.0001). The lower the adjusted root weight, the higher the pathogenicity of the isolate.

74

0.5 0.45 X a 0.4 0.35 XY Nontreated 0.3 0.25 IntegoSuite b b A A A x 0.2 y y Acceleron

0.15 Cruisermaxx YZ

0.1 Advanced Adjusted root weight (g) weight root Adjusted 0.05 Z c z B 0 Py. irregulare Py. ultimum Py. torulosum Pp.helicoides var. ulimum Oomycete spp.

Figure 2.10. The response of the screened oomycete species to fungicide seed treatments in the cup assay, with standard deviation shown. The fungicide treated seed for each species was compared to the nontreated control within each species. Bars sharing a letter are not significantly different from one another (p=0.05). The lower the adjusted root weight, the more the roots were rotted by the oomycete pathogen.

75

Chapter 3: Identifying the Quantitative Disease Resistance Loci in Nested Association Mapping populations towards Phytophthora sojae and three different species of Pythium

Introduction

Selection and planting of soybean [Glycine max (L.) Merr.] cultivars that are resistant to Pythium (Py.) (syn. Globisporangium) may be the most sustainable long-term approach to managing seed, seedling and root rot diseases. A few cultivars of soybean have been identified with resistance to some species of Pythium (Balk, 2014; Bates et al.,

2008; Ellis et al., 2013; Keeling, 1974; Kirkpatrick et al., 2006; Rosso et al., 2008).

Resistance to oomycetes has been studied most intensively for Phytophthora (Ph.) sojae

(Kaufmann and Gerdemann) -soybean interaction, including the investigation of possible mechanisms of defense (Chawla et al., 2013; Wang et al., 2010; Wang et al., 2012a,b;

Wilcox, 2001). Quantitative resistance is especially useful for disease management because this type of resistance is more durable than major gene resistance and is also effective against more than one race of a pathogen (Kou and Wang, 2010; Krattinger et al., 2009; Mundt, 2014). Only a few studies have identified, characterized, and mapped quantitative disease resistance loci (QDRL) associated with resistance to different species of Pythium (Ellis et al., 2013, 2014; Rosso et al., 2008; Stasko et al., 2016).

Past field surveys in Ohio have identified more than 35 different species of

Pythium associated with seed and seedling rot of soybean (Broders et al., 2007a; 2009;

76

Ellis et al., 2013; Dorrance et al., 2004; Dorrance Lab, unpublished data). Pythium irregulare (Buisman), Pythium ultimum Trow var. ultimum, and Pythium ultimum Trow var. sporangiiferum, which are commonly found in production fields in Ohio were identified as being aggressive towards soybean (Balk, 2014; Broders et al., 2007a, 2009;

Dorrance lab, unpublished data; Eyre, 2016). Thus, it is important to identify resistance to these pathogens, in addition to Ph. sojae to improve management guidelines. Previous studies identified Resistance-gene (R-gene) mediated Py. aphanidermatum and QDRL to

Py. irregulare and Py. aphanidermatum (Ellis et al., 2013; Rosso et al., 2008; Stasko et al., 2016; Urrea et al., 2017). Soybean resistance towards Pythium has been understudied and with the number of species that can affect this crop, characterization of the newly identified sources of resistance (Balk, 2014; Lerch et al., in progress) as well as a comparison of the QDRL for these soilborne root pathogens should be performed to provide better direction for breeding for resistance.

Before a region of the genome associated with disease resistance can be incorporated into a cultivar for management purposes, a source of resistance needs to be identified and characterized. This can be achieved in part with large mapping populations that are advanced in generation (St. Clair, 2010). Quantitative trait locus (QTL) mapping is essential to identify the regions in the host genome that contribute to the quantitative resistance towards plant pathogens and pests, including root pathogens of soybean

(Collard et al., 2005; Lindhout, 2002; Poland et al., 2009; Young, 1996).

Quantitative disease resistance loci towards Py. irregulare resistance were mapped previously in two separate studies (Ellis et al., 2013; Stasko et al., 2016). Two

77

QDRL to Py. irregulare were identified on chromosomes 14 and 19-2 in a recombinant inbred line (RIL) population derived from a cross of two moderately susceptible cultivars, Conrad and Sloan (Stasko et al., 2016). The QDRL on chromosome 14 is especially interesting because it is very near (approximately 12 kb) another previously identified Py. irregulare suggestive of a resistance QTL from recombinant inbred line

(RIL) populations derived from two three-way crosses OHS 303 x (Williams x PI

424354) and Dennison x (Williams x PI 424354) (Ellis et al., 2013). In these same populations resistance to Py. irregulare was also identified on chromosomes 1, 6, 8, 11, and 13 (Ellis et al., 2013). Stasko et al. (2016) compared regions of the genome that were associated with resistance to Py. irregulare to those that were associated with resistance to the necrotrophic pathogen Fusarium graminearum and the hemibiotroph Ph. sojae, in a Conrad x Sloan RIL population. QDRL were located to different regions of the genome for each, indicating that the mechanism for resistance to Py. irregulare may be different than the mechanisms for F. graminearum and Ph. sojae resistance in this population.

The nested association mapping (NAM) approach is currently being utilized in several plant systems to map several traits including resistance (Bajgain et al., 2016;

Diers, personal communication; Li et al., 2014; Stich, 2009; Xavier et al., 2017). The

Soybean Nested Association Mapping Project developed a large NAM population derived from the crosses of exotic germplasm, plant introductions, and high-yielding lines to one common, high-yielding hub parent (IA3023) with the goal of identifying

QTL associated with yield and other quality traits (Diers, 2014; Grant et al., 2009). Lines created from these crosses were advanced by single seed descent until the F5 generation

78

and ultimately produced 40 related RIL populations. Every F5 RIL within the NAM population was genotyped with the SoySNP6K Illumina Infinium BeadChip Genotyping

Array (Song et al., 2017) and this data is publicly available online at SoyBase

(http://soybase.org, accessed April 2017). A composite linkage map and a separate linkage map for each NAM RIL population, was generated and is available for QTL mapping of resistance to pathogens or other desirable traits (Song et al., 2017).

The soyNAM population has been previously used to successfully identify loci associated with yield and environmental interactions (Xavier et al., 2018). NAM populations have been used in maize to successfully identify QDRL for multiple pathogens (Benson et al., 2015; Kump et al., 2011; Poland et al., 2011). Previous studies have shown that several of the parents used to develop the soyNAM populations were resistant to one or more species of Pythium and Ph. sojae (Balk, 2014; Lerch, 2017;

Wickramasinghe et al., 2012). The objective of this research was to identify and compare

QDRL for resistance to Ph. sojae, Py. irregulare, Py. ultimum var. ultimum, and Py. ultimum var. sporangiiferum in these individual NAM RIL populations, and compare these to other QDRL identified in other studies.

79

Materials and methods

NAM background and seed increase

Of the 40 soybean NAM diverse donor parents, several were previously identified as having resistance based on greenhouse and laboratory experiments to one or more of the following pathogens: Py. ultimum var. ultimum, Py. ultimum var. sporangiiferum, Py. irregulare, and Ph. sojae (Balk, 2014; Wickramasinghe, 2012; Lerch, 2017). Seed from six separate F5 populations comprised of 140 RILs each derived from the crosses of

IA3023 (Fehr, Iowa State University) with 4J105-3-4 (Leroy, Purdue University), HS6-

3976 (St. Martin, Ohio State University), LD02-9050 (Diers, University of Illinois), S06-

13640 (Shannon, University of Missouri), LG05-4832 (Nelson, USDA-ARD, University of Illinois), and LG00-3372 (Nelson and Johnson, 2011) were provided by Brian Diers

(Univ. of Illinois) (soybase.com) were last increased in 2011. Information regarding the background, parentage, and resistance of the NAM populations is in Table 3.1.

For each of the six populations, 100 seeds from each RIL were planted in separate field plots for seed increase at the Ohio Agricultural Research and Development Center

Wooster on 25-26 May 2016. Each RIL from each plot was bulk hand-harvested and individually threshed to obtain fresh seed to be used in greenhouse phenotyping experiments.

Assays and inoculum production to determine resistance phenotype

Resistance towards Pythium Each NAM population was independently evaluated for resistance separately to either one or multiple Pythium species (Table 3.1) using a greenhouse cup assay that was 80

previously described (Ellis et al., 2013; Stasko et al., 2016). Briefly, isolates of Py. ultimum var. ultimum, Py. ultimum var. sporangiiferum, and Py. irregulare were grown on potato carrot agar (PCA) at 20ºC for 3 days. Eight 10-mm plugs of the isolate were placed into a sterilized Spawn bag (Myco Supply, Pittsburgh, PA) containing 950 ml play sand (Quikrete, Ravenna, Ohio), 50 ml corn meal (Quaker Oats Company, Chicago, IL), and 250 ml deionized water. These bags were closed with an electrical-impulse sealer

(Harbor Freight Tools, Camarillo, CA) and placed in a 20ºC incubator for ten days while mixed every other day to ensure even growth of the pathogen. The sand-cornmeal mixture was then mixed with fine vermiculite in a 1:4 ratio and placed into 600 ml polystyrene cups in either a growth chamber or greenhouse. The cups were watered three times over 24 hours with deionized water to ensure a suitable environment for the pathogen. Eight seeds of each RIL were placed on the inoculum and the cups were arranged in a randomized complete block design with three replications. The cups were watered twice a day to maintain a conducive environment for root disease.

Experiments to evaluate resistance towards Py. ultimum var. ultimum and Py. irregulare were done in a greenhouse at temperatures ranging between 18-23ºC; while those for Py. ultimum var. sporangiiferum were placed in a growth chamber set at 20ºC.

All assays had a 16 hr light: 8 hr dark cycle with 60% relative humidity (RH). For each

NAM population there was a minimum of two separate experiments with four to six replicates. Each experiment used the same isolates of Py. irregulare (Br.2-3-5), Py. ultimum var. ultimum (Miami 1-3-7), and Py. ultimum var. sporangiiferum (Will 1-6-7), except for populations HS6-3976 x IA3023 and IA3023 x S06-13640. These populations

81

had three replicates of one pathogenic isolate of Py. ultimum var. ultimum (Miami 1-3-7) followed by three replicates of Py. ultimum var. ultimum (N201.2(2)).

Within each replication the following cultivars with different levels of resistance towards Pythium spp. were included as controls: Clermont (OSU-OARDC), Dennison (St

Martin et al., 2008), Kottman (St Martin et al., 2001), Lorain (OSU-OARDC), Sloan

(Bahrenfus and Fehr, 1980), and Williams 82 (Bernard and Cremeens, 1988). Clermont was previously found to be susceptible while Dennison, Kottman, and Sloan were moderately susceptible to Py. ultimum var. ultimum (Balk, 2014). Clermont, Dennison,

Kottman, and Sloan were previously found to be moderately resistant, susceptible, moderately susceptible, and moderately susceptible to Py. ultimum var. sporangiiferum, respectively (Balk, 2014; Dorrance lab, unpublished data). Dennison, Kottman, Sloan, were previously found to be moderately susceptible to Py. irregulare, and Lorain was moderately resistant (Balk, 2014; Dorrance lab, unpublished data). If the cultivars that were used as controls did not develop the anticipated level of disease towards the pathogen, the experiment was repeated.

For each experiment, plants were grown for 14 days, then the roots were washed and data for disease severity for root rot, percent seed germination, total plant weight, and root weight were collected. Disease severity was based on a root rot score from 1 to 5: 1

= healthy roots with no infection, 2 = lesions covering 1-25% of roots, 3 = 26-75% of roots with lesions, 4 = 76-100% of roots with lesions, and 5 = total colonization of the seed with no germination. Roots were dried in an oven for 7 days and then weighed to

82

obtain the dry root weight. Adjusted root weight was calculated based on the root weight of all plants in a single cup divided by the number of plants in that cup.

The data for percent germination, root rot score, average root weight, and total dry root weight were analyzed for each population separately. The best linear unbiased predictor (BLUP) (Stroup, 1989) for the response of each RIL population to each pathogen was separately calculated by using PROC MIXED in SAS (SAS 9.1, SAS

Institute, Cary, NC, USA).

The assays for Py. ultimum var. ultimum and Py. irregulare were set up in a randomized complete block design (RBCD). Three replicate cups for each RIL were included in a single experiment; two experiments were performed for each NAM population x pathogen spp. combination for a total of six replicate cups for each RIL. The model for the RCBD experiments was Yijkl = µ + Ei + R(E)ij + Ck + G(C)kl + εijkl, where µ is the overall mean, Ei is the effect of the ith experiment, R(E)ij is the effect of the jth replication in the ith experiment, Ck is the effect of the kth class of entry (checks:

Clermont, Dennison, Kottman, Lorain, Sloan, Williams 82, parents, and RILs), G(C)kl is the effect of lth genotype within class (genotypic variance, σG2), εijkl is the experimental error (σ2). Class of entry was assumed to be a fixed effect and all other terms were considered random effects. The model of this analysis allowed checks and parents to be considered fixed effects and the RILs considered as random effects.

The assays for Py. ultimum var. sporangiiferum were set up in a randomized incomplete block design (RIBD). Due to limited growth chamber space, two NAM populations were evaluated simultaneously with a complete set of replicate cups for each

83

RIL for one NAM population, plus one third of another NAM population in randomized incomplete block design. This design was repeated over time in the same growth chamber until population HS6-3976 x IA3023 had a total of four replicate cups and population

IA3023 x LG05-4832 had a total of five replicate cups. The model for the RIBD experiments was Yijkl = µ + Ri + R(E)ij + Ck + G(C)kl + εijkl, where µ is the overall mean,

Ri is the effect of the ith growth chamber experiment, R(E)ij is the effect of the jth growth chamber experiment in the ith replicate, Ck is the effect of the kth class of entry (checks:

Clermont, Dennison, Kottman, Lorain, Sloan, Williams 82, parents, and RILs), G(C)kl is the effect of lth genotype within class (genotypic variance, σG2), εijkl is the experimental error (σ2). Class of entry was assumed to be a fixed effect and all other terms were considered random effects. The model of this analysis allowed checks and parents to be considered fixed effects and the RILs considered as random effects.

Resistance towards Phytophthora A tray test was used to evaluate the level of partial resistance towards Ph. sojae in the NAM populations, as previously described (Wang et al., 2010, Wang et al., 2012a,

Tucker et al., 2010; Stasko et al., 2016). Briefly, ten 7-day-old seedlings from each RIL were placed on a cotton wicking pad on a plastic tray and a 10-mm scratch was made on the tap root with a sterile syringe 20-mm below the crown. An agar-mycelial slurry made from a 7-day old culture of a Ph. sojae isolate was placed over the wound. Populations

HS6-3976 x IA3023 and IA3023 x LD02-9050 were inoculated with Ph. sojae isolate

371a92_Windfall_Ind; populations IA3023 x LG05-4832 and IA3023 x LG00-3372 were inoculated with isolate (2)2_Dayton_739_LA_02. The trays were placed in a 26.5 L 84

bucket and moved to a growth chamber at 25ºC with 14 hr light at 60% RH. For each plant, the length of the lesion was measured from the inoculation site to the edge of the lesion margin 7 dai (days after inoculation). This experiment was conducted twice in total; each RIL evaluated with this tray assay had a total of two replicate trays over time

(20 plants from each RIL total). Each of the buckets contained 24 trays, and within each replication there were three sets of check lines, and the parents of the RIL population.

Within each replication the following cultivars with different levels of resistance to Ph. sojae were included as controls: Conrad (Fehr et al., 1989), Resnik (McBlain et al.,

1990), OX-20, and Williams (Bernard and Lindahl, 1972), which were previously found to be moderately resistant, moderately resistant, highly susceptible, and susceptible towards Ph. sojae, respectively (Mideros et al., 2007).

Data from each replicate were combined and the mean lesion length of the ten seedlings from each RIL was analyzed. The best linear unbiased predictor (BLUP) of each RIL was calculated by using PROC MIXED in SAS (SAS 9.1, SAS Institute, Cary,

NC, USA) (Stroup, 1989). The model was Yijkl = µ + Ri + B(R)ij + Ck + G(C)kl + εijkl, where µ is the overall mean, Ri is the effect of the ith replication, B(R)ij is the effect of the jth bucket in the ith replication, Ck is the effect of the kth class of entry (checks: Conrad,

OX20, Resnik, Sloan, Williams, parents, and RILs), G(C)kl is the effect of lth genotype

2 within class for recombinant inbred lines only (genotypic variance, σG ), εijkl is the experimental error (σ2). Class of entry was assumed to be a fixed effect and all other terms were considered random effects. The model of this analysis allowed checks and parents to be considered fixed effects and the RILs considered as random effects.

85

QDRL Identification

Using the linkage maps previously generated for each NAM population (Song et al., 2017), the program MapQTL® 5 (Van Ooijen, 2004) was used to identify QDRL in each NAM population for each pathogen separately. Interval mapping (IM) was used to identify preliminary regions of interest, followed by multiple-QTL method and automatic cofactor selection with the walking speed set to 1 cM. The LOD thresholds for chromosome- and genome-wide analysis were calculated through permutation tests with

1000 permutations at α = 0.05; the genome- and chromosome-wide LOD threshold values for each analysis are indicated in Table 3.2. The LOD thresholds calculated at the individual chromosome level are lower than the genome-wide LOD threshold and are a useful way of identifying suggestive QTL or comparing QTL between experiments (Van

Ooijen, 1999). Quantitative disease resistance loci that had a LOD value that surpassed the calculated genome-wide LOD threshold were significant at α = 0.05; QDRL that did not surpass the genome-wide LOD threshold but did surpass the calculated chromosome- wide threshold were considered to be “suggestive QDRL” and are useful for comparative purposes only. Quantitative disease resistance loci have the percent of phenotypic variation they control calculated; if this value is 15.0% or above, the QDRL in questions was considered a “major QDRL”.

86

Results

Each of the NAM populations that were evaluated to the four pathogens had the resistance phenotypes segregate as a quantitative trait. A summary of the minor and major QDRL identified in this study can be found in Table 3.3 and these are summarized by pathogen below.

QDRL to Ph. sojae in four NAM RIL populations:

Four RIL populations derived from crosses of IA3023 with HS6-3976 (122 RILs),

LD02-9050 (91 RILs), LG05-4832 (123 RILs), and LG00-3372 (116 RILs) all responded with a quantitative resistance to Ph. sojae (Figure 3.1). The check cultivars Conrad,

Resnik, Williams, and OX20-8 also responded as expected. Overall, the difference in mean lesion lengths among the RILs in each cross were significantly different for the populations generated from the crosses HS6-3976 x IA3023 (P<0.0001), IA3023 x

LD05-9050 (P=0.0421), and IA3023 x LG05-4832 (P=0.0008). The BLUP values calculated from the mean lesions length data were inverted for presentation of data; a lower BLUP value corresponds to a higher level of resistance (Figure 3.1).

There were four QDRL for Ph. sojae identified in two populations (Table 3.3).

Two of these were major QDRL (> 15.0 % phenotypic variation, PVE) on chromosomes

6 and 18 from population IA3023 x LD02-9050 (Table 3.4). Identified from population

HS6-3976 x IA3023 was a major QDRL on chromosome 13, as well as a minor QDRL at a different location on chromosome 13 from (Table 3.5). There were several suggestive

QDRL for Ph. sojae identified in each population that were significant at the chromosome-wide LOD threshold (Table 3.2). Interestingly, all identified QDRL for Ph. 87

sojae were identified on different chromosomes between the two screened populations

(Table 3.3). This indicates that there could be multiple types or multiple sources of quantitative resistance to Ph. sojae in these NAM parents.

RIL HS6-3976 x IA3023: The mean lesion lengths for the 122 RILs ranged from

6.0 to 64.0 mm and the mean lesion length among the RILs was significantly different

(P<0.0001). The mean lesion lengths for the checks Williams, Conrad, Resnik, Sloan, and OX20, were 28.6, 30.8, 31.7, 46.1, and 54.6 mm, respectively. The mean lesion length of the hub parent IA3023 (35.8 mm) was significantly larger than the mean lesion length of the donor parent HS6-3976 (12.0 mm) (P<0.0001). The mean lesion length of all RILs in the population was 21.6 mm. The BLUP values of the RILs ranged from -12.9 to 7.0 (Figure 3.1A). One major QDRL and one minor QDRL were identified on different locations on chromosome 13 (Table 3.5), with additional suggestive QDRL identified on chromosomes 5 and 17 (Table 3.2).

RIL IA3023 x LD02-9050: The mean lesion lengths for the 91 RILs ranged from

6.2 to 58.5 mm and the mean lesion lengths among the RILs was significantly different

(P=0.0421). The mean lesion lengths for the checks Conrad, Williams, Resnik, Sloan, and OX20 were 20.1, 25.4, 28.9, 45.8, and 64.3 mm, respectively. The hub parent

IA3023 mean lesion length (27.1 mm) was significantly higher than the mean length of parent LD02-9050 (15.2 mm) (P<0.0001). The mean lesion length of the RILs were 24.2 mm. The BLUP values of the RILs ranged from -5.8 to 3.4 (Figure 3.1B). Two major

QDRL were identified on chromosome 6 and 18 (Table 3.4), with additional suggestive

QDRL identified on chromosomes 6, 10, and 14 (Table 3.2).

88

RIL IA3023 x LG05-4832: The mean lesion lengths for the 123 RILs ranged from

9.6 to 43.5 mm and the mean lesion lengths of the RILs were significantly different

(P=0.0008). The mean lesion lengths for the checks Conrad, Resnik, Williams, Sloan, and OX20 were 25.7, 30.2, 32.7, 36.6, and 43.5 mm respectively. The hub parent IA3023 did not have a significantly higher mean lesion length (30.55 mm) than the donor parent

LG05-4832 (24.1 mm) (P=0.1687). The mean lesion length of the RILs was 27.6 mm.

The BLUP values of the RILs ranged from -3.2 to 5.0 (Figure 3.1C). No QDRL were identified that surpassed the genome-wide LOD threshold, but two suggestive QDRL were identified on chromosomes 10 and 15 (Table 3.2).

RIL IA3023 x LG00-3372: The mean lesion length for the 116 RILs ranged from

6.9 to 59.6 mm and the mean lesion length of the RILs were not significantly different

(P=0.3715). The mean lesion lengths for the checks Conrad, Resnik, Williams, Sloan, and OX20 were 27.9, 31.5, 34.8, 44.5, and 46.2 mm, respectively. The hub parent

IA3023 did not have a significantly higher mean lesion length (29.3 mm) than the donor parent LG00-3372 (24.3 mm) (P=0.7086). The mean lesion length of the RILs was 26.7 mm. The BLUP values of the RILs ranged from -1.6 to 1.4 (Figure 3.1D). No QDRL were identified that surpassed the genome-wide LOD threshold, but three suggestive

QDRL were identified on chromosomes 5, 14, and 17 (Table 3.2).

QDRL to Py. irregulare in three RIL populations

The RIL populations derived from the crosses of IA3023 with HS6-3976 (122

RILs), LD02-9050 (91 RILs), and LG00-3372 (116 RILs) were evaluated for their

89

quantitative resistance towards Py. irregulare in a greenhouse cup assay. For all traits, a higher BLUP value corresponds to higher resistance to Py. irregulare. Check lines

(Clermont, Dennison, Kottman, Lorain, Sloan, and Williams 82) as well as parents

LD02-9050 and LG00-3372 were also screened for resistance to Py. irregulare in these assays. Root rot score data for population HS6-3976 x IA3023 was not used in the analysis due to large amounts of missing data, but was included in Figure 3.2. Dry root weight for population IA3023 x LD02-9050 was not used in the analyses due to a large number of missing data points, but was included in Figure 3.3.

QDRL for Py. irregulare were identified in populations IA3023 x LD02-9050 and

IA3023 x LG00-3372. One major QDRL on chromosome 4 was identified from population IA3023 x LD02-9050 (Table 3.4). There were multiple minor QDRL identified from populations IA3023 x LD02-9050 and IA3023 x LG00-3372 (Tables 3.4 and 3.6). There were additional suggestive QDRL identified in all three populations that were significant at the chromosome-wide LOD threshold (Table 3.2). The same minor

QDRL was identified in IA3023 x LD02-9050 and IA3023 x LG00-3372 on chromosome

17 with the resistance allele contributed by hub parent IA3023 (Table 3.3). There were two QDRL identified at different positions on chromosome 16 in populations IA3023 x

LD02-9050 and IA3023 x LG00-3372 (Table 3.2). All other identified QDRL for Py. irregulare were on different chromosomes, indicating that there could be multiple types or mechanisms for resistance to Py. irregulare in these RIL parents.

RIL HS6-3976 x IA3023: The percent germination for the 122 RILs ranged from

0.0 to 100.0% and the percent germination among the RILs was significantly different

90

(P=0.0002). The mean percent germination for the checks Dennison, Williams 82,

Kottman, Lorain, Clermont, and Sloan were 85.4%, 83.3%, 58.3%, 55.2%, 45.3%, and

43.8% respectively. The hub parent IA3023 had an average percent germination of 37.5% and the average percent germination for all RILs was 64.9%. The BLUP values calculated from the percent germination of the RILs ranged from -9.5 to 9.2 (Figure

3.2A). No QTL were identified in HS6-3976 population when using the percent germination data from Py. irregulare assays.

The adjusted root weight for the 122 RILs ranged from 0.00 to 0.68 g and the adjusted root weight among the RILs was significantly different (P<0.0001). The average adjusted root weight of the checks Dennison, Kottman, Lorain, Williams 82, Sloan, and

Clermont were, 0.43, 0.36, 0.31, 0.27, 0.31, 0.18 g, respectively. The hub parent IA3023 had an average adjusted root weight of 0.26 g and the average adjusted root weight of all

122 RILs were 0.31 g. The BLUP values calculated from the adjusted root weight of the

RILs ranged from -0.04 to 0.04 (Figure 3.2C). No QTL were identified in HS6-3976 population when using the average root weight data from Py. irregulare assays.

The total dry root weight for the 122 RILs ranged from 0.0 to 390.1 mg and the dry root weight among the RILs was significantly different (P<0.0001). The average total dry root weight of the checks Dennison, Kottman, Lorain, Williams 82, Clermont, and

Sloan were 187.1, 162.2, 112.7, 99.4, 69.9, 69.9 mg, respectively. The hub parent IA3023 had an average total dry root weight of 63.8 mg and the average of all 122 RILs dry root weight was 134.5 mg. The BLUP values calculated from the total dry root weights ranged from -31.0 to 42.7 (Figure 3.2D). No QTL surpassing the genome-wide LOD threshold

91

were identified in HS6-3976 population when using the total dry root weight data from

Py. irregulare assays.

RIL IA3023 x LD02-9050: The percent germination of the 91 RILs ranged from

0.0% to 100.0% and the percent germination among the RILs were significantly different

(P<0.0001). The average percent germination of the checks Kottman, Williams 82,

Clermont, Dennison, Lorain, and Sloan were 96.3%, 96.3%, 93.8%, 92.5%, 87.5%,

86.3%, respectively. The average percent germination of parents IA3023 (54.2%) and

LD02-9050 (100.0%) was significantly different (P<0.05). The average percent germination of all 91 RILs was 80.8%. The BLUP values calculated from the percent germination data ranged -33.8 to 15.0 (Figure 3.3A). Using the percent germination data from population IA3023 x LD02-9050, two minor QDRL were identified on chromosome

2 and 17, and a major QTL on chromosome 4 (Table 3.4). Three additional suggestive

QDRL were identified on chromosome 5, 7, and 15 (Table 3.2).

The root rot score of the 91 RILs ranged from 1 to 5 and the scores among the

RILs were significantly different (P<0.0001). The average root rot scores of the checks

Clermont, Dennison, Kottman, Lorain, Williams 82, and Sloan were 2.3, 2.3, 2.3, 2.4,

2.4, and 2.5, respectively. The root rot scores of parents IA3023 (3.2) and LD02-9050

(2.0) were significantly different (P<0.05). The average root rot score of all 91 RILs was

2.6. The BLUP values calculated from the root rot score data ranged from -0.53 to 0.29

(Figure 3.3B). Using the root rot score data from population IA3023 x LD02-9050, three minor QDRL were identified on chromosome 4, 13, and 16 (Table 3.4). Three additional suggestive QDRL were identified on chromosome 6, 17, and 19 (Table 3.2).

92

The adjusted root weights for the 91 RILs ranged from 0.0 to 2.0 g and the adjusted root weights among the RILs was significantly different (P<0.0001). The average adjusted root weight of the checks Dennison, Clermont, Kottman, Lorain,

Williams 82, and Sloan were 0.94, 0.80, 0.77, 0.76, 0.70, and 0.69 g, respectively. The adjusted root weight of parent LD02-9050 (1.20 g) was significantly greater than hub parent IA3023 (0.44 g) (P<0.05). The average adjusted root weight of all 91 RILs was

0.74 g. The BLUP values calculated from the adjusted root weight data ranged from -0.13 to 0.19 (Figure 3.3C). Three minor QDRL were identified on chromosome 2, 5, and 11

(Table 3.4); an additional five suggestive QDRL were identified on chromosome 3, 9, 16,

18, 20 (Table 3.2).

RIL IA3023 x LG00-3372: The percent germination for the 116 RILs ranged from

12.5% to 100.0% and the percent germination among the RILs were significantly different (P<0.0001). The average percent germination of the checks Dennison, Kottman,

Sloan, Clermont, Williams 82, and Lorain were 96.2%, 86.2%, 77.5%, 78.8%, 67.5%, and 56.2%, respectively. The percent germination of the parents IA3023 and LG00-3372 were 87.5% and 83.3%, respectively and were not significantly different (P=0.7376). The average of all 116 RILs percent germination was 85.2%. The BLUP values calculated from the percent germination data ranged -18.6 to 8.8 (Figure 3.4A). Two minor QDRL were identified on chromosome 3 and 18 (Table 3.6); three additional suggestive QDRL were identified on chromosome 3, 13, and 15 (Table 3.2).

The root rot score for the 116 RILs ranged from 1 to 5 and the root rot scores among the RILs was significantly different (P<0.0001). The average root rot scores of the

93

checks Clermont, Williams 82, Dennison, Kottman, Sloan, and Lorain were 2.1, 2.3, 2.4,

2.7, 2.7, and 3.1, respectively. The average root rot score of the parents IA3023 and

LG00-3372 were 2.3 and 2.0, respectively and were not significantly different

(P=0.1161). The average root rot score of all 116 RILs was 2.2. The BLUP values calculated from the average root rot data ranged from -0.34 to 0.11 (Figure 3.4B). Two suggestive QTL were identified on chromosomes 6 and 12 (Table 3.2). No major or minor QDRL were identified when using the root rot score data from this population.

The adjusted root weights for the 116 RILs ranged from 0.09 to 1.60 g and the adjusted root weights among the RILs was significantly different (P<0.0001). The average adjusted root weights of the checks Dennison, Clermont, Kottman, Williams 82,

Sloan, and Lorain were, 0.73, 0.56, 0.56, 0.54, 0.48, and 0.41 g, respectively. The adjusted root weight the parent LG00-3372 (0.59 g) was significantly greater than hub parent IA3023 (0.41 g) (P=0.0014). The average adjusted root weight of all 116 RILs was 0.63 g. The BLUP values calculated from the adjusted root weight data ranged from

-0.09 to 0.10 (Figure 3.4C). Three minor QDRL were identified on chromosome 1, 3, and

10 (Table 3.6); an additional four suggestive QDRL were identified on chromosome 4,

16, 19, and 20 (Table 3.2).

The total dry root weight for the 116 RILs ranged from 29.8 to 747.2 mg and the total dry root weight was significantly different among the RILs (P<0.0001). The dry root weight of the checks Dennison, Kottman, Williams 82, Clermont, Sloan, and Lorain were

379.6, 290.4, 250.4, 227.9, 192.7, and 146.3 mg, respectively. The average total dry root weight of the parents IA3023 (242.5 mg) and LG00-3372 (273.7 mg) were significantly

94

different (P<0.0001). The average total dry root weight of the RILs was 302.7 mg. The

BLUP values calculated from the dry root weight data ranged from -87.4 to 55.2 (Figure

3.4D). Three minor QDRL were identified on chromosome 10, 16, and 17 (Table 3.6); an additional suggestive QDRL was identified on chromosome 16 (Table 3.2).

QDRL to Py. ultimum var. ultimum in three RIL populations:

The RIL populations derived from the crosses of IA3023 with 4J105-3-4 (94

RILs), HS6-3976 (122 RILs), and S06-13640 (75 RILs) were evaluated for their quantitative resistance to Py. ultimum var. ultimum infection. For all traits a higher BLUP value corresponds to higher resistance to Py. ultimum var. ultimum. Checks (Clermont,

Dennison, Kottman, Lorain, Sloan, and Williams 82) as well as the parents of each population were also screened for resistance to Py. ultimum var. ultimum in these assays.

QDRL for Py. ultimum var. ultimum were identified in populations IA3023 x

4J105-3-4, HS6-3976 x IA3023, and IA3023 x S06-13640. There were four major QDRL identified on chromosomes 3, 5, 13, and 17, with four minor QTL identified on chromosomes 1, 7, 13, and 17 (Table 3.3). Also identified were additional suggestive

QDRL from all three populations that were significant at the chromosome-wide LOD threshold (Table 3.2). Populations IA3023 x S06-13640 and HS6-3976 x IA3023 each had one QDRL identified on chromosome 17, each QDRL were at different locations on the chromosome (Table 3.3).

RIL IA3023 x 4J105-3-4: The percent germination for the 94 RILs ranged from

0.0 to 100.0% and the percent germination among the RILs was significantly different

95

(P<0.0001). The mean percent germination for the checks Dennison, Kottman, Lorain,

Clermont, Williams 82, and Sloan were 85.0%, 76.3%, 73.8%, 68.8%, 60.0%, and

50.0%, respectively. The percent germination of the hub parent IA3023 (52.1%) and the donor parent 4J105-3-4 (16.7%) were significantly different (P=0.0267). The mean percent germination for all RILs was 69.1%. The BLUP values calculated from the percent germination of the RILs ranged from -14.4 to 11.3 (Figure 3.5A). One major

QDRL was identified on chromosome 3 (Table 3.7); an additional suggestive QDRL was identified on chromosome 9 (Table 3.2).

The root rot scores for the 94 RILs ranged from 2 to 5 and the root rot scores among the RILs were significantly different (P=0.0024). The mean root rot score for the checks Dennison, Kottman, Sloan, Lorain, Williams 82, and Clermont were 2.8, 3.0, 3.0,

3.3, 3.4, and 3.5, respectively. The average root rot score of hub parent IA3023 (4.0) and donor parent 4J105-3-4 (3.7) were not significantly different (P=0.4226). The mean score for all RILs was 3.2. The BLUP values calculated from the root rot scores of the RILs ranged from -0.27 to 0.28 (Figure 3.5B). One major QDRL was identified on chromosome 13 and one minor QDRL was identified on chromosome 3 (Table 3.7). Two

QDRL were identified on chromosome 13 (Table 3.3), and one additional suggestive

QDRL was identified on chromosome 19 (Table 3.2).

The adjusted root weights for the 94 RILs ranged from 0.00 to 0.80 g and the adjusted root weights among the RILs were significantly different (P<0.0001). The average adjusted root weight for the checks Dennison, Kottman, Lorain, Williams 82,

Sloan, and Clermont were 0.54, 0.45, 0.41, 0.34, 0.31, and 0.30 g, respectively. The

96

adjusted root weight of the hub parent IA3023 (0.20 g) and the donor parent 4J105-3-4

(0.18 g) were not significantly different (P=0.0577). The average adjusted root weight for all RILs was 0.32 g. The BLUP values calculated from the adjusted root weight of the

RILs ranged from -0.05 to 0.06 (Figure 3.5C). No major or minor QDRL were identified using the adjusted root weight data. Two suggestive QDRL were identified on chromosome 15 and 17 (Table 3.2).

The average total dry root weight for the 94 RILs ranged from 0.0 mg to 782.4 mg and the average root weights among the RILs were significantly different (P=0.0034).

The average total dry root weight for the checks Dennison, Kottman, Lorain, Clermont,

Williams 82, and Sloan were 254.6, 216.7, 192.7, 153.1, 150.4, and 121.2 mg, respectively. The average total dry root weight of the hub parent IA3023 (370.8) and the donor parent 4J105-3-4 (304.4) were not significantly different (P=0.1837). The mean average total dry root weight for all RILs was 304.7 mg. The BLUP values calculated from the total dry root weight of the RILs ranged from -33.2 to 27.4 (Figure 3.5D). One minor QDRL was identified on chromosome 1 (Table 3.7); three suggestive QDRL were identified on chromosome 15, 17, and 17 (Table 3.2).

RIL HS6-3976 x IA3023: The percent germination for the 122 RILs ranged from

0.0 to 100.0% and the percent germination among the RILs was significantly different

(P<0.0001). The mean percent germination for the checks Dennison, Kottman, Lorain,

Clermont, Sloan, and Williams 82 were 87.5%, 75.0%, 63.5%, 58.3%, 54.1%, and

47.9%, respectively. The average percent germination of the parents IA3023 (70.8%) and

HS6-3976 (58.3%) were not significantly different (P=0.1835). The mean percent

97

germination for the RILs was 70.2%. The BLUP values calculated from the percent germination of the RILs ranged from -18.5 to 12.9 (Figure 3.6A). No major or minor

QDRL were identified using the percent germination data from this population. Three suggestive QDRL were identified on chromosome 2, 5, and 10 (Table 3.2).

The scores among the 122 RILs ranged from 1 to 5 and the root rot scores among the RILs was significantly different (P<0.0001). The mean root rot scores for the checks

Dennison, Kottman, Clermont, Lorain, Sloan, and Williams 82 were 2.6, 2.9, 3.1, 3.3,

3.3, and 3.4, respectively. The root rot scores of the parents IA3023 (3.0) and HS6-3976

(2.7) were not significantly different (P>0.05). The mean root rot for the RILs was 3.0.

The BLUP values calculated from the root rot scores of the RILs ranged from -0.46 to

0.31 (Figure 3.6B). One minor QDRL was identified on chromosome 17 (Table 3.5); an additional four suggestive QDRL were identified on chromosome 2, 10, 11, and 14

(Table 3.2).

The adjusted root weights for the 122 RILs ranged from 0.00 to 1.33 g and the adjusted root weights among the RILs was significantly different (P<0.0001). The average adjusted root weights for the checks Dennison, Kottman, Lorain, Sloan,

Clermont, and Williams 82 were 0.64, 0.54, 0.49, 0.46, 0.41, and 0.38 g, respectively.

The adjusted root weights of the parents IA3023 (0.43 g) and HS6-3976 (0.42) were not significantly different (P=0.9389). The average adjusted root weight for the RILs was

0.53 g. The BLUP values calculated from the adjusted root weights of the RILs ranged from -0.11 to 0.12 (Figure 3.6C). One minor QDRL was identified on chromosome 2

98

(Table 3.5); an additional 3 suggestive QDRL were identified on chromosome 11, 18, and

19 (Table 3.2).

The total dry root weight for the 122 RILs ranged from 0.0 to 622.1 mg and the average total dry root weight among the RILs was significantly different (P<0.0001).

The mean total dry root weight for the checks Kottman, Lorain, Clermont, Williams 82,

Dennison, and Sloan were 123.6, 123.3, 114.0, 90.1, 80.8, and 69.1 mg, respectively. The average total dry root weight of the parents IA3023 (193.3 mg) and HS6-3976 (146.9) were not significantly different (P=0.3906). The average total dry root weight for the

RILs was 227.2 mg. The BLUP values calculated from the dry root weights of the RILs ranged from -82.8 to 66.1 (Figure 3.6D). One minor QDRL was identified on chromosome 2 (Table 3.5); two additional suggestive QDRL were identified on chromosomes 10 and 18 (Table 3.2).

RIL IA3023 x S06-13640: The percent germination for the 75 RILs ranged from

0.0 to 100.0% and the percent germination among the RILs was significantly different

(P<0.0001). The mean percent germination for the checks Dennison, Kottman, Clermont,

Williams 82, Sloan, and Lorain were 92.2%, 78.1%, 70.3%, 62.5%, 54.7%, and 51.6%, respectively. The hub parent IA3023 had an average percent germination of 81.3% and the mean percent germination for the RILs was 61.7%. The BLUP values calculated from the percent germination of the RILs ranged from -24.0 to 15.5 (Figure 3.7A). Two major

QDRL were identified on chromosome 5 and 17 (Table 3.8); and additional suggestive

QDRL was identified on chromosome 3 (Table 3.2).

99

The scores for the 75 RILs ranged from 2 to 5 and the average scores among the

RILs was significantly different (P<0.0001). The mean root rot scores for the checks

Dennison, Clermont, Kottman, Sloan, Lorain, and Williams 82 were 2.5, 2.9, 2.9, 2.9,

3.0, and 3.3, respectively. The hub parent IA3023 had an average root rot score of 2.7 and the mean root rot for the RILs was 2.9. The BLUP values calculated from the root rot scores of the RILs ranged from -0.22 to 0.25 (Figure 3.7B). One major QDRL was identified on chromosome 7 (Table 3.8); and additional two suggestive QDRL were identified on chromosome 2 and 8 (Table 3.2).

The adjusted root weight for the 75 RILs ranged from 0.00 to 1.03 g and the adjusted root weights among the RILs was significantly different (P=0.0002). The average adjusted root weight for the checks Dennison, Kottman, Lorain, Clermont, Sloan, and Williams 82 were 0.68, 0.60, 0.49, 0.48, 0.48, and 0.34 g, respectively. The hub parent IA3023 had an average adjusted root weight of 0.47 g and the average adjusted root weight for the RILs was 0.49 g. The BLUP values calculated from the adjusted root weights of the RILs ranged from -0.08 to 0.12 (Figure 3.7C). No QDRL were identified using the average root weight data from this population (data not shown).

The total dry root weight for the 75 RILs ranged from 0.0 to 477.6 mg and the average total dry root weight among the RILs was significantly different (P<0.0001).

The mean total dry root weight for the checks Dennison, Kottman, Clermont, Sloan,

Lorain, and Williams 82 were 323.6, 245.5, 211.6, 158.7, 142.3, and 128.8 mg, respectively. The hub parent IA3023 had an average total dry root weight of 216.0 mg and the average total dry root weight for the RILs was 169.2 mg. The BLUP values

100

calculated from the dry root weights of the RILs ranged from -59.5 to 74.7 (Figure 3.7D).

One suggestive QDRL was identified on chromosome 17 (Table 3.2). No major or minor

QDRL were identified using the total dry root data from this population (data not shown).

QDRL to Py. ultimum var. sporangiiferum in two RIL populations:

Two RIL populations derived from the crosses of IA3023 with HS6-3976 (122

RILs) and LG05-4832 (123 RILs) were evaluated for their quantitative resistance to Py. ultimum var. sporangiiferum. For all traits, a higher BLUP value corresponds with higher resistance to Py. ultimum var. sporangiiferum. Checks (Clermont, Dennison, Kottman,

Lorain, Sloan, and Williams 82) as well as the parents of each population were also screened for resistance to Py. ultimum var. sporangiiferum in these assays.

QDRL for Py. ultimum var. sporangiiferum were identified in populations HS6-

3976 x IA3023 and IA3023 x LG05-4832. There was one major QDRL identified on chromosome 3 and four minor QTL identified on chromosome 3, 5, 11, and 17 (Table

3.3). Also identified were additional suggestive QDRL from these two populations that were significant at the chromosome-wide LOD threshold (Table 3.2).

RIL HS6-3976 x IA3023: The percent germination for the 122 RILs ranged from

25.0 to 100.0% and the percent germination among the RILs was significantly different

(P=0.0038). The mean percent germination for the checks Dennison, Kottman, Clermont,

Lorain, Williams 82, and Sloan were 96.9%, 94.8%, 83.3%, 69.8%, 67.7%, and 49.0%, respectively. The average percent germination of the parents IA3023 (91.7%) and HS6-

3976 (75.0%) were not significantly different (P=0.2697). The mean percent germination

101

for the RILs was 75.2%. The BLUP values calculated from the percent germination of the

RILs ranged from -12.0 to 5.5 (Figure 3.8A). Two suggestive QDRL were identified on chromosome 5 and 12 (Table 3.2). No major or minor QDRL were identified using the percent germination data for this population (data not shown).

The scores for the 122 RILs ranged from 2 to 4 and the average scores among the

RILs was not significantly different (P=0.2000). The mean root rot scores for the checks

Dennison, Kottman, Sloan, Williams 82, Clermont, and Lorain were 2.1, 2.3, 2.3, 2.3,

2.5, and 2.8, respectively. The root rot scores of the parents IA3023 (3.0) and HS6-3976

(2.7) were not significantly different (P=0.4226). The mean root rot score for the RILs was 2.6. The BLUP values calculated from the root rot scores of the RILs ranged from -

0.10 to 0.06 (Figure 3.8B). Two minor QDRL were identified on chromosome 5 and 17

(Table 3.5); an additional two suggestive QDRL were identified on 12 and 18 (Table

3.2).

The adjusted root weights for the 122 RILs ranged from 0.18 to 2.73 g and the adjusted root weight among the RILs was significantly different (P<0.0001). The average adjusted root weight for the checks Dennison, Kottman, Williams 82, Sloan,

Clermont, and Lorain were 0.96, 0.82, 0.71, 0.63, 0.60, and 0.56 g, respectively. The adjusted root weights of the parents IA3023 (0.59 g) and HS6-3976 (0.70) were not significantly different (P=0.2881). The average adjusted root weight for the RILs was

0.65 g. The BLUP values calculated from the adjusted root weights of the RILs ranged from -0.07 to 0.14 (Figure 3.8C). One minor QDRL was identified on chromosome 11

102

(Table 3.5); an additional three suggestive QDRL were identified on chromosome 1, 3, and 16 (Table 3.2).

The total dry root weight for the 122 RILs ranged from 62.6 to 995.3 mg and the average total dry root weight among the RILs was not significantly different (P=0.1777).

The mean total dry root weight for the checks Dennison, Kottman, Clermont, Williams

82, Lorain, and Sloan were 573.6, 537.4, 400.7, 377.7, 316.5, and 251.4 mg, respectively.

The average total dry root weight of the parents IA3023 (482.2 mg) and HS6-3976 (463.6 mg) were not significantly different (P=0.8016). The average total dry root weight for the

RILs was 413.2 mg. The BLUP values calculated from the dry root weight of the RILs ranged from -33.9 to 27.7 (Figure 3.8D). One suggestive QDRL was identified on chromosome 15 (Table 3.2). No major or minor QDRL were identified using the total dry root data from this population (data not shown).

RIL IA3023 x LG05-4832: The percent germination for the 123 RILs ranged from

0.0% to 100.0% and the percent germination among the RILs was significantly different

(P<0.0001). The mean percent germination for the checks Dennison, Kottman, Clermont,

Lorain, Sloan, and Williams 82 were 96.4%, 92.9%, 79.5%, 72.3%, 67.0%, and 54.5%, respectively. The average percent germination of the parents IA3023 (93.8%) and LG05-

4832 (50.0%) were significantly different (P=0.0273). The mean percent germination for the RILs was 74.7%. The BLUP values calculated from the percent germination of the

123 RILs ranged from -24.4 to 11.7 (Figure 3.9A). One major QDRL was identified on chromosome 3 (Table 3.9), and an additional suggestive QDRL was identified on chromosome 2 (Table 3.2).

103

The root rot scores for the 123 RILs ranged from 2 to 5 and the average scores among the RILs were significantly different (P=0.0056). The mean root rot scores for the checks Dennison, Kottman, Clermont, Sloan, Williams 82, and Lorain were 2.2, 2.4, 2.5,

2.4, 2.4, and 2.9, respectively. The root rot scores of the parents IA3023 (2.5) and LG05-

4832 (2.8) were not significantly different (P=0.3910). The mean root rot score for the

RILs was 2.6. The BLUP values calculated from the root rot scores of the RILs ranged from to -0.30 to 0.20 9 (Figure 3.9B). One minor QDRL was identified on chromosome 3

(Table 3.9); and additional suggestive QDRL was identified on chromosome 13 (Table

3.2).

The adjusted root weights for the 123 RILs ranged from 0.00 to 1.43 g and the adjusted root weight among the RILs was significantly different (P=0.0267). The average adjusted root weight for the checks Dennison, Kottman, Williams 82, Sloan,

Clermont, and Lorain were 0.97, 0.78, 0.73, 0.61, 0.59, and 0.56 g, respectively. The adjusted root weight of the parents IA3023 (0.57 g) and LG05-4832 (0.55 g) were not significantly different (P=0.8082). The average adjusted root weight for the RILs was

0.64 g. The BLUP values calculated from the adjusted root weights of the RILs ranged from -0.13 to 0.12 9 (Figure 3.9C). No major or minor QDRL were identified using the adjusted root weight data from this population. Three suggestive QDRL were identified on chromosome 1, 5, and 19 (Table 3.2).

The total dry root weight for the 123 RILs ranged from 0.0 to 1043.7 mg and the average total dry root weight among the RILs was significantly different (P=0.0257).

The average total dry root weight for the checks Dennison, Kottman, Clermont, Williams

104

82, Lorain, and Sloan, were 625.8, 564.3, 433.4, 432.0, 370.5, and 313.5 mg, respectively. The average dry root weight of the parents IA3023 (433.9 mg) and LG05-

4832 (178.6 mg) were not significantly different (P=0.2206). The average total dry root weight for the RILs was 365.3 mg. The BLUP values calculated from the total dry root weights of the RILs ranged from -83.6 to 77.8 (Figure 3.9D). One minor QDRL was identified on chromosome 3 (Table 3.9); an additional suggestive QDRL was identified on chromosome 12 (Table 3.2).

Summary of QDRL for Ph. sojae

There was one minor QDRL and one major QDRL for Ph. sojae identified on chromosome 13 from the population HS6-3976 x IA3023, explaining 7.2% and 42.2% of the phenotypic variation, respectively (Table 3.5). The minor QDRL on chromosome 13 for Ph. sojae overlaps in position for a minor Py. ultimum var. ultimum QDRL identified on chromosome 13 from population IA3023 x 4J105-3-4 (Table 3.3).

There were two major QDRL for Ph. sojae identified on chromosome 6 and 18 from population IA3023 x LD02-9050 that explained 17.0% and 24.6% phenotypic variation, respectively (Table 3.4). The major QDRL on chromosome 6 overlaps with a suggestive QDRL for Py. irregulare from population IA3023 x LD02-9050 and these

QDRL had different parental sources of the resistance allele (Table 3.2).

Populations IA3023 x LG05-4832 and IA3023 x LG00-3372 did not have any

QDRL that surpassed the genome-wide LOD threshold, though there were some suggestive QDRL identified (Table 3.2).

105

Summary of QDRL for Py. irregulare

From population IA3023 x LD02-9050, eight QDRL were identified for Py. irregulare on chromosomes 2, 4, 5, 11, 13, 16, and 17 (Table 3.3). Only the QDRL identified on chromosome 4 explained more than 15.0% of the phenotypic variation

(Table 3.4). The minor QDRL identified on chromosome 2 overlaps in position with a suggestive QDRL for Py. ultimum var. ultimum from population HS6-3976 x IA3023; these QDRL had different parental sources of the resistance allele (Table 3.2).

From population IA3023 x LG00-3372, seven minor QDRL on chromosome 1, 3,

10, 16, 17, and 18 were identified. The Py. irregulare QDRL identified on chromosome

17 from the populations IA3023 x LD02-9050 and IA3023 x LG00-3372 are responsible for 13.6% and 11.0% of phenotypic variation, respectively. In both populations the resistance allele is contributed by IA3023 and overlaps in position. The two minor QDRL identified on chromosome 16 from populations IA3023 x LD02-9050 and IA3023 x

LG00-3372 do not overlap in position and the resistance alleles are contributed by each population’s respective non-hub parent. The minor QDRL on chromosome 18 identified from population IA3023 x LG00-3372 overlaps with a suggestive Py. ultimum var. ultimum QDRL; the resistance allele is sourced from different parents (Table 3.2).

There were no QDRL that surpassed the genome-wide LOD threshold identified towards Py. irregulare from population HS6-3976, though there were several suggestive

QDRL that surpassed the chromosome LOD thresholds (Table 3.2).

106

Summary of QDRL for Py. ultimum var. ultimum

There were four QDRL towards Py. ultimum var. ultimum that surpassed the genome-wide LOD threshold for the population IA3023 x 4J105-3-4 on chromosomes 1,

3, and 13; the two QDRL on chromosome 13 did not overlap in position (Table 3.7). The major QDRL identified on chromosome 13 overlaps in position with the minor QDRL for

Py. irregulare identified on chromosome 13 from population IA3023 x LD02-9050. The minor QDRL identified on chromosome 13 overlaps in position with the minor Ph. sojae

QDRL on chromosome 13 from population HS6-3976 x IA3023. The minor QDRL on chromosome 3 overlaps in position with a suggestive QDRL for Py. ultimum var. sporangiiferum identified from population IA3023 x LG05-4832 and both QDRL have the resistant allele from hub parent IA3023 (Table 3.2).

In population HS6-3976 x IA3023, one minor QDRL was identified on chromosome 17 that explained 12.2% of the phenotypic variation (Table 3.5). This minor

QDRL overlaps in position with a suggestive QDRL for Ph. sojae, also from population

HS6-3976 x IA3023, and both resistant alleles come from HS6-3976. In addition, it also overlaps with a Py. ultimum var. ultimum suggestive QDRL from IA3023 x 4J105-3-4; however, the resistant allele of the suggestive QDRL is from the parent 4J105-3-4.

There were three major QDRL identified from IA3023 x S06-13640 on chromosomes 5, 7, and 17 (Table 3.8). The QDRL identified from this population on chromosome 17 explained 24.2% of the phenotypic variation and overlaps in position with the minor QDRL for Py. irregulare identified from populations IA3023 x LD02-

9050 and IA3023 x LG00-3372. The major QDRL identified on chromosome 5 overlaps

107

in position with a suggestive Py. irregulare QDRL from population HS6-3976 x IA3023; the QDRL have different parental sources of the resistant allele. The major QDRL on chromosome 17 overlaps with two different minor Py. irregulare QDRL from populations IA3023 x LG00-3372 and IA3023 LD02-9050. These QDRL also overlap with a suggestive QDRL for Py. ultimum var. ultimum from population IA3023 x S06-

13640. For all QDRL at this location on chromosome 17, the resistant allele was contributed by the hub parent IA3023 (Table 3.2).

Summary of QDRL for Py. ultimum var. sporangiiferum

There were three Py. ultimum var. sporangiiferum QDRL that surpassed the genome-wide LOD threshold from population HS6-3976 x IA3023 on chromosome 5,

11, and 17. The QDRL on chromosome 5 overlaps in position with the major Py. ultimum var. ultimum QDRL identified on chromosome 5 from population IA3023 x S06-13640; both sources of the resistant allele are from the hub parent IA3023. Also overlapping in this position is a suggestive QDRL for Py. irregulare identified on population HS6-3976 x IA3023, however the resistant allele for this suggestive QDRL was contributed by HS6-

3976.

There were two QDRL identified on chromosome 3 from population IA3023 x

LG05-4832; the major QDRL on this chromosome overlaps in position with the major

Py. ultimum var. ultimum QDRL on chromosome 3 identified on population IA3023 x

4J105-3-4; both resistant alleles were contributed by hub parent IA3023 (Table 3.2).

108

Discussion

Prior resistance screenings of the NAM parents with Ph. sojae, Py. irregulare, Py. ultimum var. ultimum, and Py. ultimum var. sporangiiferum suggested that partial resistance to these pathogens could be identified in select NAM populations (Balk, 2014).

As such, there was a diversity of loci identified and thus potential mechanisms for quantitative disease resistance as measured in these tray assays for Ph. sojae and in cup assays towards Py. irregulare, Py. ultimum var. ultimum, and Py. ultimum var. sporangiiferum. BLUP value distributions for the NAM populations indicated that there was significant genetic variation for root traits and lesion sizes in most of these experiments. In populations where we were not able to identify clear and significant differences among the RILs or for LOD values, population size may have been a factor.

RIL populations are considered a valuable and powerful means of identifying

QTL of many different phenotypic traits. However, the power of RIL populations to identify QTL with precision decreases with the fewer the number of lines included in the population (Ferreira et al., 2006; St. Clair, 2010). Due to heavy losses in the field, NAM populations 4J105-3-4 x IA3023, IA3023 x LD02-9050, and IA3023 x S06-13640 each had less than 100 RILs. The QTL analysis may have been more precise or been able to identify more QDRL if each of the populations had more RILs. The large number of suggestive QDRL identified in this study may also be a reflection of the small population sizes used in this experiment. If these experiments were repeated with the entire 140 RILs per NAM population, it is possible that some of the previously identified suggestive

QDRL may surpass the genome-wide LOD threshold.

109

An advantage of analyzing the BLUPs from several different root-related traits in a single experiment is the possibility of identifying and potentially verifying the QDRL at the same genomic location. For instance, a QDRL for Py. ultimum var. ultimum from

HS6-3976 x IA3023 was identified at the same location on chromosome 2 for both adjusted root weight and total dry root weight (Table 3.5).

An advantage of screening several populations for multiple pathogens is the chance of identifying QDRL in the same genomic regions across populations. This was a key goal of this study to ascertain if these pathogens, all with similar lifestyles as watermolds, would have similar QDRL and also would support the importance of that resistance loci in managing root disease. Co-localization of disease resistance QTL to different pathogens has been noted before in crop species, such as in rice (Kou and

Wang, 2012). There was co-localization for major QDRL on chromosome 3 for resistance to Py. ultimum var. ultimum identified from population IA3023 x 4J015-3-4 and another major QDRL for Py. ultimum var. sporangiiferum identified from population

IA3023 x LG05-4832 (Table 3.3). There were two examples of co-localization on chromosome 13. At one location on chromosome 13 there was one major QDRL identified for Py. ultimum var. ultimum and a minor QDRL for Py. irregulare identified from populations IA3023 x 4J105-3-4 and IA3023 x LD02-9050, respectively (Table

3.3). The other example of co-localization on chromosome 13 involves the overlapping minor QDRL for Py. ultimum var. ultimum and Ph. sojae from populations IA3023 x

4J105-3-4 and HS6-3976 x IA3023, respectively (Table 3.3). The last example of co- localization of QDRL found in this study was three overlapping QDRL on chromosome

110

17; two minor QDRL for Py. irregulare were identified from populations IA3023 x

LD05-9050 and LG00-3372 and one major QDRL was identified for Py. ultimum var. ultimum from population IA3023 x S06-13640 (Table 3.3).

Both populations IA3023 x LG05-4832 and IA3023 x LG00-3372 did not have any identified QDRL that were significant at the genome-wide LOD threshold, though they did have some suggestive QDRL (Table 3.2). The lack of significant QDRL not due to a lack of RILs in the populations, as population IA3023 x LG05-4832 and IA3023 x

LG00-3372 have 123 and 111 RILs, respectively. There was also adequate marker coverage in these populations; population IA3023 x LG05-4832 had 3422 polymorphic markers and population IA3023 x LG00-3372 had 3022 polymorphic markers. In both population IA3023 x LG05-4832 and IA3023 x LG00-3372, there was no significant difference between the mean lesion lengths of the two parents after inoculation with Ph. sojae isolate (2)2_Dayton_739_LA_02. Additionally, the mean lesion lengths of the RILs from population IA3023 x LG00-3372 were not significantly different. Without significant differences in the resistance phenotype of the RILs, it is not possible to map or identify QDRL. It is possible that a tray assay using a different Ph. sojae isolate could yield significant results by producing significantly different mean lesion lengths between the RILs. Additionally, there was variability between the two replicate tray assays for populations HS6-3976 x IA3023 and IA3023 x LD02-9050; potentially significant QDRL could be found using these populations if the assays were repeated and care taken to ensure more consistent conditions.

111

Though there were a few exceptions, most of the QDRL identified in this study did not overlap between pathogen species (Table 3.3). These results were unexpected as one would expect the soybean seedlings to have the same QDRL, corresponding to the same or similar mechanisms of defense, between the closely related Pythium species. The wide diversity of QDRL identified to the four oomycete species indicates that there may be a diversity of defense mechanisms or genes involved in host resistance against the different pathogenic oomycete species. Many of the QDRL identified in this study are novel, though there are a few that overlap with previously identified resistance QTL to other pathogens or qualitative resistance genes.

In a previous study by Lee et al. (2014), QDRL for Ph. sojae were identified on chromosomes 1, 3, 12, 13, 16, 18, and 19 in a joint linkage analysis. The major QDRL identified on chromosome 13 in this study is a short distance away from the multiple Ph. sojae QDRL identified on chromosome 13 from Lee et al. (2014). Lee et al. (2013a) and

Lee et al. (2013b) also identified multiple QDRL towards Ph. sojae that surpassed the genome-wide LOD threshold that overlaps in this same region; these QDRL were identified in other studies as being located near R-gene rich regions (Tucker et al., 2010;

Wang et al., 2010, 2012a,b). The minor QDRL identified on chromosome 13 is similar in location to a Ph. sojae QDRL previously identified in three RIL populations; in each of the three populations, the QDRL accounted for >20% of the phenotypic variation

(Burnham et al. 2003).

The minor Ph. sojae QDRL on chromosome 13 is near another minor Ph. sojae

QDRL that was identified using both laboratory and field experiments (Han et al., 2008;

112

Li et al., 2010). This same minor QDRL is near the approximate region of a suggestive

QDRL for Ph. sojae previously identified using interval mapping on a Conrad x Sloan

F4:6 population (Wang et al., 2012a). The minor QDRL on chromosome 13 found in this study also maps near regions of the genome previously reported to have resistance to F. graminearum (Ellis et al., 2012), Sclerotinia sclerotiorum (Arahana et al., 2001; Guo et al., 2008), Py. aphanidermatum (Rosso et al., 2008), and multiple other pest and pathogen QDRL (Gordon et al., 2006).

Interestingly, the QDRL identified on chromosome 18 in this study was not located near Ph. sojae QDRL identified in previous studies (Stasko et al., 2016; Lee et al., 2013a,b; Lee et al, 2014; Wang et al., 2012a,b), which are located near the cluster of

Rps4, Rps5, and Rps6, and QDRL conferring resistance to soybean cyst nematode (SCN)

(Kabelka et al., 2005).

Many QDRL for Py. irregulare were identified in this study that surpassed the genome-wide LOD threshold (Table 3.3). Interestingly, none of the QDRL identified from the three populations overlaps with QDRL for Py. irregulare in prior studies (Ellis et al., 2013; Stasko et al., 2016). The minor Py. irregulare QDRL identified on chromosome 1 overlaps with a QDRL identified for SDS resistance (F. virguliforme)

(Abdelmajid et al., 2012) and adjacent to a previously identified QDRL for resistance to

S. sclerotiorum (Kim and Diers, 2000). The minor QDRL identified on chromosome 13 overlaps with a previously identified QDRL for Phytophthora root rot (Han et al., 2008).

One of the minor QDRL on chromosome 17 that was identified in two populations

(IA3023 x LD02-9050 and IA3023 x LG00-3372) not only overlaps with a major QDRL

113

identified for Py. ultimum var. ultimum from population IA3023 x S06-13640, but also overlaps with a QDRL previously identified for S. sclerotiorum resistance (Arahana et al., 2001). The QDRL on chromosome 18 is located near several QDRL previously identified in other studies for resistance to SDS (Iqbal et al., 2001; Kazi et al., 2008) and

Ph. sojae (Tucket et al., 2010).

Many QDRL for Py. ultimum var. ultimum were identified in this study that surpassed the genome-wide LOD threshold (Table 3.3). The QDRL identified on chromosome 5 overlaps in position with a QDRL identified for resistance to SCN (Jiao et al., 2015). The major QDRL on chromosome 13 not only overlaps in position with the

QDRL identified for Py. irregulare in this study, it is close in location to previously identified resistance to F. graminearum (Ellis et al., 2012), S. sclerotiorum (Arahana et al., 2001; Guo et al., 2008), Py. aphanidermatum (Rosso et al., 2008), and multiple other pest and pathogen QDRL (Gordon et al., 2006). The QDRL identified on chromosome 17 is close to previously identified resistance to S. sclerotiorum (Arahana et al., 2001).

Ultimately, the goal of QTL analysis for is to assist in the breeding of high-yielding cultivars that also have the desired disease resistance. This study has provided the basis of identifying genetic markers that are linked to multiple

QDRL to four different soybean root oomycete pathogens. Once a marker that is tightly linked to a desirable QDRL is identified, plant breeders can use that information to expedite selection for that QDRL. Individual plants that contain the marker of interest and therefore the desired QDRL can be screened for the presence of a genetic marker early in the seedling stage rather than waiting for the phenotype to be revealed in time-

114

consuming greenhouse or field trials (Tanksley et al., 1989). The use of genetic markers can also assist in avoiding “linkage drag” when trying to introgress a QTL into a cultivar;

“linkage drag” is the inclusion of undesirable genes that are physically located close to the desired QTL (Concibido et al., 2003). The QDRL identification in this study is the first step in one day being able to use marker selection for the breeding of oomycete root rot resistance cultivars of soybean.

In this study we identified many QDRL to Ph. sojae, Py. irregulare, Py, ultimum var. ultimum, and Py. ultimum var. sporangiiferum in six soybean NAM populations. We expected to be able to identify a large number of QDRL with precision due to the scale of the soyNAM population and the dense genetic map generated from the results of the

SoySNP6K BeadChip. Overall, we identified a wide variety of putative QDRL that suggest there is a range of many genes that are involved with partial resistance, which supports previous hypotheses (Poland et al., 2009). Some of the QDRL identified in this study overlap or are close in location to QDRL to the same pathogen, or to a variety of other soybean oomycete and fungal pathogens as well as those at novel locations.

Potentially, the QDRL identified in this study can be combined in a single cultivar to give resistance to multiple types of root rot pathogens.

115

Table 3.1. The six Nested Association Mapping (NAM) RIL populations evaluated for resistance towards Phytophthora sojae in a

tray test and Pythium irregulare, Py. ultimum var. ultimum, and Py. ultimum var. sporangiiferum. Cells marked with an “X” indicate

that phenotypic screens and QTL analysis were carried out for the NAM RIL population x pathogen species combination. The data

from the individual parents was reported by Balk (2014).

Segregates for resistance to: IA3023 x No. of RILs in Donor each population Parent Py. ultimum Py. ultimum var. Parent used in this study Characteristic Ph. sojae Py. irregulare var. ultimum sporangiiferum 4J105-3-4 94 high yield X 116 HS6-3976 122 high yield X X X X LD02-9050 91 high yield X X S06-13640 75 high yield X LG05-4832 123 diverse ancestry X X LG00-3372 116 diverse ancestry X X

Table 3.2. Complete list of all QDRL, arranged by chromosome and position, for Ph. sojae (Phs), Py. irregulare (Pirr), Py. ultimum

var. ultimum (Puu), and Py. ultimum var. sporangiiferum (Pus) identified by composite interval mapping (CIM) using F5 RILs in

separate SoyNAM populations. Traits shown are total dry root weight (DRW), adjusted root weight (ARW), percent germination

(%G), root rot score (RRS), and mean lesion length (MLL). Major QDRL (>15.0 PVE) are indicated in bold. Suggestive QTL are

indicated in italics. QDRL that did not have a marker at the identified peak are marked with a dash. Marker range is the position in bp

of the neighboring markers; some QDRL did not have both a right and a left marker.

Parent Peak

117 NAM LOD GW Chr contrib. Marker Marker a

Population Path Trait Chr PVE Score LOD LOD Res. Allele Marker at Peak pos (bp) range (bp) 8388480- 4J105-3-4 Puu DRW 1 12.1 4.2 3.2 1.5 IA3023 - - 10556016 50295199- LG00-3372 Pirr ARW 1 6.9 3.3 3.3 1.5 LG00-3372 Gm01_49641478_A_G 50525642 50583510 51320677- HS6-3976 Pus ARW 1 6.8 2.32 2.9 1.8 HS6-3976 - - 51781178 52400089- LG05-4832 Pus ARW 1 6.9 2.26 3.1 2.0 IA3023 Gm01_52577638_G_A 53462529 53523589 a Phenotypic variation explained by the individual QDRL, calculated in MapQTL ® 5 (van Ooijen, 2004).

Continued

Table 3.2 continued.

Parent Peak NAM LOD GW Chr contrib. Res. Marker Marker Population Path Trait Chr PVEa Score LOD LOD Allele Marker at Peak pos (bp) range (bp) 5054610- LD02-9050 Pirr %G 2 9.6 3.67 3.0 2.0 LD02-9050 Gm02_5035934_C_A 5096021 5321601 5555227- LG05-4832 Pus %G 2 8.4 2.78 3.3 2.1 LG05-4832 Gm02_5529675_T_C 5588023 5893075 6572325- LD02-9050 Pirr ARW 2 7.8 3.52 2.9 2.0 LD02-90501 Gm02_6529620_G_A 6605936 7031201 6640715-

118 HS6-3976 Puu %G 2 7.4 2.53 3.1 2.0 IA3023 Gm02_6952838_T_C 7031201 7104809

13877996- HS6-3976 Puu ARW 2 9.2 3.34 3.2 2.2 HS6-3976 - - 14102777 13877996- HS6-3976 Puu DRW 2 12.8 4.12 3.0 2.0 HS6-3976 Gm02_13904897_A_G 14102777 14206854 13877996- HS6-3976 Puu RRS 2 6.0 2.07 3.2 2.0 HS6-3976 Gm02_13904897_A_G 14102777 14206854 46948328- S06-13640 Puu RRS 2 12.4 2.97 3.2 1.5 S06-13640 Gm02_50204121_A_G 47124597 47168791 8046- LG05-4832 Pus DRW 3 9.7 3.25 3.2 1.8 IA3023 Gm03_140242_G_A 139472 172048 Continued

Table 3.2 continued.

Parent Peak NAM LOD GW Chr contrib. Marker Marker Population Path Trait Chr PVEa Score LOD LOD Res. Allele Marker at Peak pos (bp) range (bp) 215829- S06-13640 Puu %G 3 8.0 2.2 3.1 1.8 IA3023 Gm03_172818_A_G 172048 247749 425209- LG05-4832 Pus %G 3 15.9 5.13 3.3 1.9 IA3023 Gm03_511376_C_T 510431 587640 425209- LG05-4832 Pus RRS 3 10.1 3.38 3.2 2.0 IA3023 - - 510431 510431-

119 4J105-3-4 Puu %G 3 16.8 5.6 3.1 1.5 IA3023 Gm03_588585_C_T 587640 853885

2078116- LG00-3372 Pirr %G 3 7.7 3.02 3.1 1.9 IA3023 Gm03_2153294_A_C 2162466 2428982 7688949- HS6-3976 Pus ARW 3 5.1 1.88 2.9 1.7 HS6-3976 Gm03_10546906_C_T 10067608 16852562 31912038- LG00-3372 Pirr %G 3 12.3 4.66 3.1 1.9 LG00-3372 - - 33553037

LD02-9050 Pirr ARW 3 3.8 1.98 2.9 1.4 IA3023 Gm03_43348646_G_T 41335531 42438340

43485660- LG00-3372 Pirr ARW 3 6.9 3.37 3.3 1.4 IA3023 Gm03_45516951_G_A 43512609 43849572 Continued

Table 3.2 continued.

Parent Peak Marker NAM LOD GW Chr contrib. Marker range Population Path Trait Chr PVEa Score LOD LOD Res. Allele Marker at Peak pos (bp) (bp) LD02-9050 Pirr %G 4 20.0 6.58 3.0 1.6 IA3023 - 5314249- - 5903949 LD02-9050 Pirr RRS 4 12.2 3.93 3.1 1.6 IA3023 Gm04_5837752_G_A 5314249- 5903949 6972200 LD02-9050 Pirr %G 4 7.5 2.0 3.0 1.6 LD02-9050 - 5903949- - 6972200 LG00-3372 Pirr ARW 4 4.9 2.39 3.3 1.6 IA3023 Gm04_8184443_T_C 6980768-

120 8256473 8418530

HS6-3976 Phs MLL 5 4.4 2.52 3.0 1.4 HS6-3976 Gm05_8965870_G_T 86517- 152028 234355 LG00-3372 Phs MLL 5 4.2 1.28 3.1 1.1 LG00-3372 Gm05_7682777_T_C 1441870 410455

LD02-9050 Pirr ARW 5 12.2 5.23 2.9 1.4 IA3023 Gm05_389226_T_C 2064410 2220637

LD02-9050 Pirr %G 5 5.2 2.04 3.0 1.4 IA3023 Gm05_1429125_A_G 3040462- 3145925 3175240 Continued

Table 3.2 continued.

Parent Peak NAM LOD GW Chr contrib. Marker Marker Population Path Trait Chr PVEa Score LOD LOD Res. Allele Marker at Peak pos (bp) range (bp) 4938108- HS6-3976 Pus %G 5 4.5 1.42 3.0 1.3 IA3023 Gm05_3493714_C_A 5210013 5905686 34454262- LG05-4832 Pus ARW 5 6.8 2.21 3.1 1.8 LG05-4832 - - 35163430 35147835- HS6-3976 Pirr ARW 5 8.2 2.78 3.2 1.7 IA3023 Gm05_34863934_A_C 35134643 35268299

1 HS6-3976 Puu %G 5 6.9 2.38 3.1 1.7 IA3023 Gm05_34877126_C_T 35147835 35134643

21

HS6-3976 Pirr DRW 5 5.8 2.06 3.1 1.8 IA3023 Gm05_34877126_C_T 35147835 35134643

38388163- S06-13640 Puu %G 5 18.5 4.67 3.1 1.6 IA3023 Gm05_41903142_C_T 38435220 38446748 38639960- HS6-3976 Pus RRS 5 10.2 4.02 3.2 1.6 IA3023 Gm05_41540078_C_A 38797782 38838119 39791394- HS6-3976 Pirr ARW 5 7.4 2.64 3.2 1.7 HS6-3976 Gm05_40523972_A_G 39814902 39885811

HS6-3976 Pirr %G 6 4.3 1.45 3.2 1.1 HS6-3976 Gm06_1976435_T_G 1994260 2569042 Continued

Table 3.2 continued.

Parent Peak NAM LOD GW Chr contrib. Marker Marker Population Path Trait Chr PVEa Score LOD LOD Res. Allele Marker at Peak pos (bp) range (bp) 11721439- LD02-9050 Pirr RRS 6 3.9 1.46 3.1 1.3 LD02-9050 Gm06_11701693_C_A 11728261 11803252 11725151- LD02-9050 Phs MLL 6 17.0 5.5 3.1 1.3 IA3023 Gm06_11776489_C_A 11803252 11728261 13924125- LG00-3372 Pirr RRS 6 8.6 2.44 3.2 1.8 LG00-3372 - - 16001372 16710120-

122 LD02-9050 Phs MLL 6 4.5 1.6 3.1 1.6 LD02-9050 Gm06_16710123_A_G 16755458 17024674

17442842- HS6-3976 Pirr ARW 6 4.9 1.71 3.2 1.1 IA3023 Gm06_17407746_T_C 17462416 17781823 17559069- LD02-9050 Pirr %G 7 5.3 2.08 3.0 1.3 IA3023 - - 17969328 30099305- S06-13640 Puu RRS 7 16.3 3.8 3.2 2.0 S06-13640 Gm07_36955973_T_C 36852524 36907555

S06-13640 Puu RRS 8 8.2 2.03 3.2 1.5 S06-13640 Gm08_10821603_A_C 10732481 10524776

8966209- 4J105-3-4 Puu %G 9 4.7 1.73 3.1 1.3 4J105-3-4 Gm09_31130602_G_A 33755797 35489229 Continued

Table 3.2 continued.

Parent Peak NAM LOD GW Chr contrib. Res. Marker Marker Population Path Trait Chr PVEa Score LOD LOD Allele Marker at Peak pos (bp) range (bp) LD02-9050 Pirr ARW 9 3.8 1.99 2.9 1.4 LD02-9050 Gm09_39615772_A_G 42293052 42424537

HS6-3976 Puu DRW 10 8.7 2.86 3.0 1.3 HS6-3976 Gm10_150503_T_C 146001 105491

453608- HS6-3976 Puu %G 10 4.5 1.58 3.1 1.4 HS6-3976 Gm10_474540_A_C 469309 638673 694881-

123 HS6-3976 Puu RRS 10 3.6 1.32 3.2 1.3 HS6-3976 Gm10_702278_C_T 697074 849820

14651686- LG05-4832 Phs MLL 10 4.4 1.35 3.3 1.0 LG05-4832 Gm10_11679545_C_T 11871077 28285220

LD02-9050 Phs MLL 10 8.3 2.9 3.1 1.5 LD02-9050 Gm10_43638272_A_G 44218338 44180065

HS6-3976 Pirr ARW 10 5.0 1.74 3.2 1.5 HS6-3976 Gm10_43608539_A_G 44188605 44180065

LG00-3372 Pirr ARW 10 10.9 5.07 3.3 1.6 LG00-3372 Gm10_43821942_T_C 44402011 44218338

LG00-3372 Pirr DRW 10 10.0 3.94 3.1 1.6 LG00-3372 Gm10_43821942_T_C 44402011 44218338 Continued

Table 3.2 continued.

Parent Peak NAM LOD GW Chr contrib. Marker Marker Population Path Trait Chr PVEa Score LOD LOD Res. Allele Marker at Peak pos (bp) range (bp) HS6-3976 Pirr %G 10 5.9 1.6 3.2 1.6 HS6-3976 - 45247863- - 48551669 HS6-3976 Pirr DRW 10 7.5 2.14 3.1 1.6 HS6-3976 - 45247863- - 48551669 HS6-3976 Puu ARW 11 6.3 2.35 3.2 1.7 HS6-3976 Gm11_4827840_A_C 3397697- 4837437 4953034 HS6-3976 Puu RRS 11 6.9 2.38 3.2 1.7 HS6-3976 - 7465799-

124 - 8050212

HS6-3976 Pus ARW 11 9.2 3.12 2.9 1.5 IA3023 Gm11_36517294_T_C 32042169 32006970

LD02-9050 Pirr ARW 11 10.2 4.49 2.9 1.7 LD02-9050 Gm11_38289103_C_T 34177149- 34244773 34296488 LG05-4832 Pus DRW 12 5.0 1.71 3.2 1.1 IA3023 Gm12_3152390_G_A 2843851- 3157788 3153069 HS6-3976 Pus %G 12 6.4 2.0 3.0 1.7 IA3023 Gm12_6888204_T_C 6716834- 6926701 7623357 Continued

Table 3.2 continued.

Parent Peak NAM LOD GW Chr contrib. Res. Marker Marker Population Path Trait Chr PVEa Score LOD LOD Allele Marker at Peak pos (bp) range (bp) 7843948- LG00-3372 Pirr RRS 12 6.2 1.84 3.2 1.5 IA3023 Gm12_7855367_G_A 7893592 8476059 7893592- HS6-3976 Pus RRS 12 4.6 1.84 3.2 1.7 IA3023 Gm12_7879643_G_T 7917868 8100860 22901190- LD02-9050 Pirr RRS 13 12.4 3.81 3.1 1.2 IA3023 - -

125 25230180

25230180- 4J105-3-4 Puu RRS 13 17.2 5.23 3.2 1.8 IA3023 - - 26955004 29273755- HS6-3976 Pirr %G 13 9.2 2.8 3.2 1.7 IA3023 - - 29287182 29273755- HS6-3976 Pirr DRW 13 8.8 2.98 3.1 1.8 IA3023 - - 29287182 29547527- 4J105-3-4 Puu RRS 13 8.6 3.1 3.2 1.8 4J105-3-4 Gm13_28561728_C_A 29759427 29784074 30125163- HS6-3976 Phs MLL 13 42.2 17.58 3.0 1.8 HS6-3976 Gm13_29043806_T_C 30243463 30154255 Continued

Table 3.2 continued.

Parent Peak NAM LOD GW Chr contrib. Res. Marker Marker Population Path Trait Chr PVEa Score LOD LOD Allele Marker at Peak pos (bp) range (bp) LG00-3372 Pirr %G 13 5.0 2.07 3.1 1.5 LG00-3372 Gm13_34762391_T_C 35971086 36038013

36512047- 4J105-3-4 Puu RRS 13 8.3 2.89 3.2 1.7 4J105-3-4 - - 36968979 39586198- LG05-4832 Pus RRS 13 6.7 2.36 3.2 2.2 LG05-4832 Gm13_38179911_C_T 39357388 39701102 40233656-

126 HS6-3976 Phs MLL 13 7.2 4.65 3.0 1.5 HS6-3976 Gm13_39560450_G_A 40739398 42919730

40935278- 4J105-3-4 Puu RRS 13 12.7 4.44 3.2 1.7 IA3023 Gm13_40441579_G_T 41887172 41953362 405270- HS6-3976 Puu RRS 14 4.1 1.45 3.2 1.4 HS6-3976 Gm14_743883_T_G 750283 919420 4964419- LD02-9050 Phs MLL 14 5.5 1.95 3.1 1.1 LD02-9050 Gm14_4887898_T_C 4972000 5084093 33034668- LD02-9050 Phs MLL 14 6.3 2.51 3.1 1.1 LD02-9050 Gm14_36271043_C_A 19436633 41321004 43825978- LG00-3372 Phs MLL 14 6.8 1.93 3.1 1.5 LG00-3372 Gm14_46331412_T_C 45609958 46642456 Continued

Table 3.2 continued.

Parent Peak NAM LOD GW Chr contributing Marker Marker Population Path Trait Chr PVEa Score LOD LOD Res. Allele Marker at Peak pos (bp) range (bp) 6791744- 4J105-3-4 Puu ARW 15 9.6 3.04 3.2 2.1 4J105-3-4 Gm15_6881423_T_C 6899908 6910313 7238417- 4J105-3-4 Puu DRW 15 6.5 2.39 3.2 1.9 IA3023 Gm15_8355608_T_C 8396570 8405044 12377604- LG05-4832 Phs MLL 15 10.6 3.04 3.3 2.1 LG05-4832 Gm15_12365188_G_A 12386061 12418026 12524905-

127 LG00-3372 Pirr %G 15 6.0 2.46 3.1 1.9 IA3023 Gm15_12471408_C_T 12492550 12561402

15322381- LD02-9050 Pirr %G 15 4.3 1.72 3.0 1.6 IA3023 Gm15_15664314_G_A 15689811 15893057 15398976- HS6-3976 Pus DRW 15 5.4 1.53 3.1 1.4 IA3023 Gm15_15310205_C_T 15335780 15893057 1034335- LD02-9050 Pirr ARW 16 5.2 2.33 2.9 1.2 LD02-9050 - - 1367524 1034335- LG00-3372 Pirr DRW 16 8.3 3.33 3.1 1.9 LG00-3372 Gm16_2780183_T_C 2805691 3225680 3282223- LG00-3372 Pirr ARW 16 5.8 2.79 3.3 2.1 LG00-3372 Gm16_3315591_C_T 3340686 3333343 Continued

Table 3.2 continued.

Parent Peak NAM LOD GW Chr contributing Marker Marker Population Path Trait Chr PVEa Score LOD LOD Res. Allele Marker at Peak pos (bp) range (bp) 6100279- LG00-3372 Pirr DRW 16 7.6 3.05 3.1 1.9 LG00-3372 Gm16_18602792_A_G 21710419 27665746 7741637- HS6-3976 Pus ARW 16 5.3 1.97 2.9 1.7 IA3023 Gm16_7569316_C_T 7727875 7811960

LD02-9050 Pirr RRS 16 12.4 4.0 3.1 1.6 LD02-9050 Gm16_27322120_C_T 27665746 28348383

4610230-

128 LG00-3372 Pirr DRW 17 11.0 4.26 3.1 1.7 IA3023 - - 6517544

4949843- LD02-9050 Pirr RRS 17 9.7 3.08 3.1 1.5 IA3023 - - 6517544 4949843- S06-13640 Puu %G 17 24.4 5.67 3.1 1.8 IA3023 - - 6517544 4949843- S06-13640 Puu DRW 17 14.2 2.41 3.1 1.8 IA3023 - - 6517544 6517544- LD02-9050 Pirr %G 17 13.6 5.0 3.0 1.4 IA3023 - - 7100289 7523745- LG00-3372 Phs MLL 17 7.7 2.2 3.1 1.7 IA3023 Gm17_8054983_A_C 7785530 7816928 Continued

Table 3.2 continued.

Parent Peak Marker NAM LOD GW Chr Path Trait Chr PVEa contrib. Res. Marker at Peak Marker range Population Score LOD LOD Allele pos (bp) (bp) 7621659- 4J105-3-4 Puu DRW 17 6.4 2.35 3.2 1.3 4J105-3-4 Gm17_8023911_C_T 7754466 7900590 10172361- 4J105-3-4 Puu DRW 17 5.4 2.02 3.2 1.3 4J105-3-4 Gm17_10666981_A_G 10380761 10480175 37799552- HS6-3976 Pus RRS 17 9.8 3.58 3.2 1.5 HS6-3976 - - 38965056 40457644- HS6-3976 Phs MLL 17 3.0 1.72 3.0 1.6 HS6-3976 Gm17_41060022_G_A 40793218 40876232

129 4J105-3-4 Puu ARW 17 8.9 2.83 3.2 1.7 4J105-3-4 Gm17_41060022_G_A 40457644-

40793218 40868443 HS6-3976 Puu RRS 17 12.2 4.06 3.2 1.5 HS6-3976 Gm17_41060022_G_A 40457644- 40793218 40876232 1685460- LD02-9050 Phs MLL 18 24.6 7.54 3.1 1.1 LD02-9050 - - 2103034 LD02-9050 Pirr ARW 18 5.0 2.34 2.9 1.1 IA3023 Gm18_2908988_T_G 2919356 2405952 6000564- HS6-3976 Puu ARW 18 6.0 2.24 3.2 1.6 IA3023 Gm18_6684198_A_C 6709316 6903088 9205527- LG00-3372 Pirr %G 18 9.3 3.35 3.1 1.8 LG00-3372 - - 10045551 Continued

Table 3.2 continued.

Parent Peak NAM LOD GW Chr contributing Marker Marker Population Path Trait Chr PVEa Score LOD LOD Res. Allele Marker at Peak pos (bp) range (bp) 9205527- HS6-3976 Pus RRS 18 4.9 1.97 3.2 1.6 IA3023 Gm18_9892865_G_T 9925334 9968176

HS6-3976 Puu DRW 18 4.2 1.49 3.0 1.6 IA3023 Gm18_54230950_G_A 49956537 49731234

1818179- HS6-3976 Puu ARW 19 5.9 2.23 3.2 1.3 IA3023 Gm19_1813813_C_T 1858113 2036298 37844154-

130 LD02-9050 Pirr RRS 19 9.1 3.01 3.1 1.3 LD02-9050 Gm19_38457940_C_T 38670962 39194527

41539674- LG00-3372 Pirr ARW 19 3.1 1.64 3.3 1.5 IA3023 Gm19_41699897_C_A 41902571 41987315 41902571- LG05-4832 Pus ARW 19 8.1 2.61 3.1 1.6 IA3023 Gm19_41727752_A_G 41930426 41987315 42117954- HS6-3976 Pirr %G 19 6.3 2.01 3.2 1.6 IA3023 Gm19_42102616_T_C 42303692 42308104 2144167- LD02-9050 Pirr ARW 20 4.8 2.25 2.9 1.6 LD02-9050 Gm20_2336786_T_C 2330680 2319783 2308267- LG00-3372 Pirr ARW 20 3.8 1.88 3.3 1.7 LG00-3372 Gm20_2325889_A_C 2319783 2334285

Table 3.3. Comparison of QTL for resistance towards oomycete pathogens Phytophthora sojae, Pythium irregulare, Py. ultimum var.

ultimum, and Py. ultimum var. sporangiiferum in the six NAM populations generated by crossing IA3023 with 4J105-3-4, HS6-3976,

LD02-9050, S06-13640, LG05-4832, and LG00-3372. QDRL that overlap in position between different pathogen species or

populations are noted by having a shared letter. The overlapping QDRL identified using different trait data (percent germination, root

rot score, etc.) within a single pathogen x population combination are reported as a single QDRL. Major QDRL (explanation of

>15.0% phenotypic variance) are indicated in bold. Cells containing a “X” indicate that no major or minor QDRL detected for the

indicated NAM RIL population x pathogen combination. Cells containing “--” indicate that the NAM RIL population x pathogen

131 combination was not screened with phenotype assays.

Segregates for resistance to: IA3023 x Py. ultimum var. Py. ultimum var. Ph. sojae Py. irregulare Donor Parent #RILs ultimum sporangiiferum 4J105-3-4 94 -- -- 1a, 3b, 13a, 13c -- HS6-3976 122 13b, 13c X 2c,17c 5c,11a, 17b LD02-9050 91 6, 18a 2a, 2b, 4, 5a, 11b, 13a, 16b, 17a -- -- S06-13640 75 -- -- 5b, 7, 17a -- LG05-4832 123 X -- -- 3a, 3b LG00-3372 116 X 1b, 3c, 3d, 10, 16a, 17a, 18b -- --

Figure 3.1. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks mean lesion length for RIL populations following inoculation with Ph. sojae in a tray assay. Shown are (A)

HS6-3976 x IA3023 and (B) IA3023 x LD02-9050 (isolate 371a92_Windfall_Ind) and populations (C) IA3023 x LG05-4832 and (D) IA3023 x LG00-3372 (isolate

(2)2_Dayton_739_LA_02). BLUE values not indicated on the graphs are (B) Conrad

(5.9), (C) Conrad (7.0), and (D) Resnik (3.6) and Conrad (7.0). BLUP values were inverted; lower values indicate more susceptible (S) lines and higher BLUP values indicate lines with greater resistance (R) to the pathogen

132

Table 3.4. Comparison of major and minor QDRL towards Py. irregulare (Pirr) and Ph. sojae (Phs) significant at the genome-wide

threshold mapped in the NAM RIL population generated from the cross of IA3023 x LD02-9050. Traits shown are adjusted root

weight (ARW), percent germination (%G), root rot score (RRS), and mean lesion length (MLL). Major QTL (>15.0 PVE) are

indicated in bold.

LOD GW Parent contrib. Peak Marker Neighboring marker Path. Trait Chr. PVEa Score LOD resistance allele Marker at Peak Position (bp) range (bp) Pirr %G 2 9.6 3.67 3.0 LD02-9050 Gm02_5035934_C_A 5096021 5054610-5321601 Pirr ARW 2 7.8 3.52 2.9 LD02-9050 Gm02_6529620_G_A 6605936 6572325-7031201

133 Pirr %G 4 20.0 6.58 3.0 IA3023 - - 5314249-5903949

Pirr RRS 4 12.2 3.93 3.1 IA3023 Gm04_5837752_G_A 5903949 5314249-6972200 Pirr ARW 5 12.2 5.23 2.9 IA3023 Gm05_389226_T_C 2064410 2220637 Phs MLL 6 17.0 5.5 3.1 IA3023 Gm06_11776489_C_A 11803252 11725151-11728261 Pirr ARW 11 10.2 4.49 2.9 LD02-9050 Gm11_38289103_C_T 34244773 34177149-34296488 Pirr RRS 13 12.4 3.81 3.1 IA3023 - - 22901190-25230180 a Phenotypic variation explained by the individual QDRL, calculated in MapQTL ® 5 (van Ooijen, 2004).

Continued

Table 3.4 continued.

LOD GW Parent contrib. Peak Marker Neighboring marker Path. Trait Chr. PVEa Score LOD resistance allele Marker at Peak Position (bp) range (bp) Pirr RRS 16 12.4 4.0 3.1 LD02-9050 Gm16_27322120_C_T 27665746 28348383 Pirr %G 17 13.6 5.0 3.0 IA3023 - - 6517544-7100289 Phs MLL 18 24.6 7.54 3.1 LD02-9050 - - 1685460-2103034

134

Table 3.5. Comparison of major and minor QDRL for resistance towards Ph. sojae (Phs), Py. ultimum var. ultimum (Puu), Py.

irregulare (Pirr), and Py. ultimum var. sporangiiferum (Pus) significant at the genome-wide threshold mapped in the NAM F5 RIL

population derived from the cross of HS6-3976 x IA3023. Traits shown are total dry root weight (DRW), adjusted root weight

(ARW), root rot score (RRS), and mean lesion length (MLL). Major QTL (>15.0 PVE) are indicated in bold.

LOD GW Parent contrib. Peak Marker Neighboring marker Path. Trait Chr. PVEa Score LOD resistance allele Marker at Peak Position (bp) range (bp) Puu ARW 2 9.2 3.34 3.2 HS6-3976 - - 13877996-14102777 Puu DRW 2 12.8 4.12 3.0 HS6-3976 Gm02_13904897_A_G 14102777 13877996-14206854

135 Pus RRS 5 10.2 4.02 3.2 IA3023 Gm05_41540078_C_A 38797782 38639960-38838119

Pus ARW 11 9.2 3.12 2.9 IA3023 Gm11_36517294_T_C 32042169 32006970 Phs MLL 13 42.2 17.58 3.0 HS6-3976 Gm13_29043806_T_C 30243463 30125163-30154255 Phs MLL 13 7.2 4.65 3.0 HS6-3976 Gm13_39560450_G_A 40739398 40233656-42919730 Pus RRS 17 9.8 3.58 3.2 HS6-3976 - - 37799552-38965056 Puu RRS 17 12.2 4.06 3.2 HS6-3976 Gm17_41060022_G_A 40793218 40457644-40876232 a Phenotypic variation explained by the individual QDRL, calculated in MapQTL ® 5 (van Ooijen, 2004).

Figure 3.2. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross HS6-3976 x IA3023 following inoculation with Py. irregulare in a greenhouse cup assay. BLUE values not indicated on the graphs are (A) Clermont (-12.5) and Sloan (-14.6), (B) Clermont (-0.58), Sloan (-

0.42), and Lorain (-0.25), (C) Clermont (-0.10) and Sloan (-0.06), and (D) Clermont (-

35.0). Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher BLUP values indicate more resistant (R) lines.

136

Figure 3.3. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross IA3023 x LD02-9050 following inoculation with Py. irregulare in a greenhouse cup assay. Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher

BLUP values indicate more resistant (R) lines.

137

Figure 3.4. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross IA3023 × LG00-3372 following inoculation with Py. irregulare in a greenhouse cup assay. BLUE values not indicated on the graphs are (A) Lorain (-22.5), (B) Clermont (0.20), Sloan (-0.40), and Lorain (-0.8), and (D) Lorain (-104.1). Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher BLUP values indicate more resistant

(R) lines.

138

Table 3.6. Comparison of QDRL to Py. irregulare significant at the genome-wide threshold mapped in the NAM RIL population

generated from the cross of IA3023 x LG00-3372. Traits shown are total dry root weight (DRW), adjusted root weight (ARW), and

percent germination (%G).

LOD Parent contributing Marker at Peak Marker Path. Trait Chr. PVEa Score GW LOD resistance allele Peak Position (bp) ARW 1 6.9 3.3 3.3 LG00-3372 Gm01_49641478_A_G 50525642 50295199-50583510 %G 3 12.3 4.66 3.1 LG00-3372 - - 31912038-33553037 ARW 3 6.9 3.37 3.3 IA3023 Gm03_45516951_G_A 43512609 43485660-43849572

139 ARW 10 10.9 5.07 3.3 LG00-3372 Gm10_43821942_T_C 44402011 44218338

DRW 10 10.0 3.94 3.1 LG00-3372 Gm10_43821942_T_C 44402011 44218338 DRW 16 8.3 3.33 3.1 LG00-3372 Gm16_2780183_T_C 2805691 1034335-3225680 DRW 17 11.0 4.26 3.1 IA3023 - - 4610230-6517544 %G 18 9.3 3.35 3.1 LG00-3372 - - 9205527-10045551 a Phenotypic variation explained by the individual QDRL, calculated in MapQTL ® 5 (van Ooijen, 2004).

Figure 3.5. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross IA3023 x 4J105-3-4 following inoculation with Py. ultimum var. ultimum in a greenhouse cup assay. BLUE values not indicated on the graphs are (A) Lorain (13.75), (B) Lorain (0.30), and (C) Lorain (0.08).

Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher BLUP values indicate more resistant (R) lines.

140

Table 3.7. Comparison of major and minor QDRL for Py. ultimum var. ultimum which were significant at the genome-wide threshold

mapped in the NAM F5 RIL population derived from the cross of IA3023 x 4J105-3-4. Traits shown are total dry root weight (DRW),

percent germination (%G), and root rot score (RRS). Major QDRL (>15.0 PVE) are indicated in bold.

Parent LOD GW contributing Peak Marker Neighboring Trait Chr. PVEa Score LOD resistance allele Marker at Peak Position (bp) marker range (bp) DRW 1 12.1 4.2 3.2 IA3023 - - 8388480-10556016 %G 3 16.8 5.6 3.1 IA3023 Gm03_588585_C_T 587640 510431-853885

141 RRS 13 17.2 5.23 3.2 IA3023 - - 25230180-26955004

RRS 13 12.7 4.44 3.2 IA3023 Gm13_40441579_G_T 41887172 40935278-41953362 a Phenotypic variation explained by the individual QDRL, calculated in MapQTL ® 5 (van Ooijen, 2004).

Figure 3.6. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross HS6-3976 x IA3023 following inoculation with Py. ultimum var. ultimum in a greenhouse cup assay. BLUE values not indicated on the graphs are (A) Clermont (10.4) and Sloan (6.3), (B) Clermont (0.33),

Sloan (0.17), and Lorain (0.08), and (D) Sloan (15.8) and Lorain (63.1). Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher BLUP values indicate more resistant (R) lines.

142

Figure 3.7. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross IA3023 x S06-13640 following inoculation with Py. ultimum var. ultimum in a greenhouse cup assay. BLUE values not indicated on the graphs are (B) Clermont (0.38) and Sloan (0.38), and (D) Clermont

(82.9). Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher BLUP values indicate more resistant (R) lines.

143

Table 3.8. Comparison of major QDRL to Py. ultimum var. ultimum significant at the genome-wide threshold mapped in the NAM

RIL population generated from the cross of IA3023 x S06-13640. Traits shown are percent germination (%G) and root rot score

(RRS). Major QDRL (>15.0 PVE) are indicated in bold.

LOD Parent contributing Marker at Peak Marker Path. Trait Chr. PVEa Score GW LOD resistance allele Peak Position (bp) %G 5 18.5 4.67 3.1 IA3023 Gm05_41903142_C_T 38435220 38388163-38446748 RRS 7 16.3 3.8 3.2 S06-13640 Gm07_36955973_T_C 36852524 30099305-36907555 %G 17 24.4 5.67 3.1 IA3023 - - 4949843-6517544 a

144 Phenotypic variation explained by the individual QDRL, calculated in MapQTL ® 5 (van Ooijen, 2004).

Figure 3.8. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross HS6-3976 x IA3023 following inoculation with Py. ultimum var. sporangiiferum in a greenhouse cup assay. BLUE values not indicated on the graphs are (A) Sloan (-18.8) and Clermont (15.6), (B)

Clermont (-0.17), Sloan (0.08), and Lorain (-0.5), (C) Sloan (-0.08), and (D) Sloan (-

126.3) and Lorain (-61.2). Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher BLUP values indicate more resistant

(R) lines.

145

Figure 3.9. Frequency distributions of the best linear unbiased predictors (BLUP) values for the RILs and best linear unbiased estimates (BLUE) of the checks for percent germination (A), root rot score (B), average root weight (C), and total dry root weight (D) from the NAM population derived from the cross IA3023 x LG05-4832 following inoculation with Py. ultimum var. sporangiiferum in a greenhouse cup assay. BLUE values not indicated on the graphs are (B) Lorain (-0.50), (C) Clermont (-0.15) and

Lorain (-0.18), and (D) Sloan (-118.6). Root rot score values were inverted; for all traits, lower BLUP values indicate more susceptible (S) lines and higher BLUP values indicate more resistant (R) lines.

146

Table 3.9. Comparison of major and minor QDRL to Py. ultimum var. sporangiiferum significant at the genome-wide threshold

mapped in the NAM RIL population generated from the cross of IA3023 x LG05-4832. Traits shown are total dry root weight (DRW),

percent germination (%G), and root rot score (RRS). Major QTL (>15.0 PVE) are indicated in bold.

LOD Parent contributing Marker at Peak Marker Path. Trait Chr. PVEa Score GW LOD resistance allele Peak Position (bp) DRW 3 9.7 3.25 3.2 IA3023 Gm03_140242_G_A 139472 8046-172048 %G 3 15.9 5.13 3.3 IA3023 Gm03_511376_C_T 510431 425209-587640 RRS 3 10.1 3.38 3.2 IA3023 - - 425209-510431 a

147 Phenotypic variation explained by the individual QDRL, calculated in MapQTL ® 5 (van Ooijen, 2004).

References

Abawi, G. S., and Widmer, T. L. 2000. Impact of soil health management practices on

soilborne pathogens, nematodes and root diseases of vegetable crops. Appl. Soil

Ecol. 15:37-47.

Abdelmajid, K.M., Ramos, L., Leandro, L., Mbofung, G., Hyten, D.L., Kantartzi, S.K.,

Grier, I.V., Robert, L., Njiti, V.N. and Meksem, K. 2012. The ‘PI 438489B’by

‘Hamilton’ SNP-based genetic linkage map of soybean [Glycine max (L.) Merr.]

identified quantitative trait loci that underlie seedling SDS resistance. J. of Plant

Gen. Sci. 1:18-30.

Abney, T. S., Melgar, J. C., Richards, T. L., Scott, D. H., Grogan, J., and Young, J. 1997.

New races of Phytophthora sojae with Rps1-d virulence. Plant Dis. 81:653-655.

Ajayi-Oyetunde, O.O. and Bradley, C.A. 2017. Identification and characterization of

Rhizoctonia species associated with soybean seedling disease. Plant Dis. 101:520-

533.

Allen, T. W., Bradley, C. A., Sisson, A. J., Byamukama, E., Chilvers, M. I., Coker, C.

M., Collins, A. A., Damicone, J. P., Dorrance, A. E., Dufault, N. S., Esker, P. D.,

Fraske, T. R., Giesler, L. J., Grybauskas, A. P., Hershman, D. E., Hollier, C. A.,

Isakeit, T., Jardine, D. J., Kelly, H. M., Kemerait, R. C., Kleczewski, N. M.,

Koenning, S. R., Kurle, J. E., Malvick, D. K., Markell, S. G., Mehl, H. L.,

148

Mueller, D. S., Mueller, J. D., Mulrooney, R. P., Nelson, B. D., Newman, M. A.,

Osborne, L., Overstreet, C., Padgett, G. B., Phipps, P. M., Price, P. P., Sikora, E.

J., Smith, D. L., Spurlock, T. N., Tande, C. A., Tenuta, A. U., Wise, K. A., and

Wrather, J. A. 2017. Soybean yield loss estimates due to diseases in the United

States and Ontario, Canada, from 2010 to 2014. Plant Health Prog. 18:19-27.

Anderson, T.R. and Buzzell, R.I. 1982. Efficacy of metalaxyl in controlling Phytophthora

root and stalk rot of soybean cultivars differing in field tolerance. Plant Dis.

66:1144-1145.

Anderson, G., Campbell, B. and Kurle, J.E. 2017. Characteristics of resistance to seed

and root rots caused by Pythium ultimum and Pythium irregulare in early maturity

soybean lines. Phytopathology. 107:122.

Arahana, V.S., Graef, G.L., Specht, J.E., Steadman, J.R. and Eskridge, K.M. 2001.

Identification of QTLs for resistance to Sclerotinia sclerotiorum in soybean. Crop

Sci. 41:180-188.

Bahrenfus, J.B., and Fehr, W.R. 1980. Registration of ‘Sloan’ soybean. Crop Sci. 20:673.

Bajgain, P., Rouse, M.N., Tsilo, T.J., Macharia, G.K., Bhavani, S., Jin, Y. and Anderson,

J.A. 2016. Nested association mapping of stem rust resistance in wheat using

genotyping by sequencing. PloS One. 11. 0155760.

Bakker, M. G., Acharya, J., Moorman, T. B., Robertson, A. E., and Kaspar, T. C. 2016.

The potential for cereal rye cover crops to host corn seedling pathogens.

Phytopathology. 591-601.

149

Balk, C. S. 2014. Assessment of resistance in soybean to Pythium ultimum and sensitivity

of Ohio’s diverse Pythium species towards metalaxyl. M.S. Thesis, The Ohio

State University.

Balk, C., Fitzgibbon, T., Novakowiski, J.H., Eyre, M. and Dorrance, A.E. 2014.

Assessment of resistance in soybean towards Pythium ultimum var. ultimum and

Pythium ultimum var. sporangiiferum. Phytopathology. 104:S3:166.

Bates, G. D., Rothrock, C. S., and Rupe, J. C. 2008. Resistance of the soybean cultivar

‘Archer’ to Pythium damping-off and root rot caused by several Pythium spp.

Plant Dis. 92:763-766.

Beckmnn, J.S. and Soller, M. 1988. Detection of linkage between marker loci and loci

affecting quantitative traits in crosses between segregating populations. Theor.

Appl. Genet. 76:228-236.

Benson, J.M., Poland, J.A., Benson, B.M., Stromberg, E.L. and Nelson, R.J. 2015.

Resistance to gray leaf spot of maize: genetic architecture and mechanisms

elucidated through nested association mapping and near-isogenic line analysis.

PLoS Genet. 11. p.e1005045.

Ben-Yephet, Y., and Nelson, E. B. 1999. Differential suppression of damping-off caused

by Pythium aphanidermatum, P. irregulare, and P. myriotylum in composts at

different temperatures. Plant Dis. 83:356-360.

Bernard, R.L. and Cremeens, C.R. 1988. Registration of ‘Williams 82’soybean. Crop Sci.

28:1027-1028.

150

Bernard, R.L., and Lindahl, D.A. 1972. Registration of ‘Williams’ soybean. Crop Sci.

12:716.

Bescansa, P., Imaz, M. J., Virto, I., Enrique, A., and Hoogmoed, W. B. 2006. Soil water

retention as affected by tillage and residue management in semiarid Spain. Soil

Tillage Res. 87:19-27.

Beuerlein, J., and Dorrance, A. E. 2004. Ohio Agronomy Guide. Soybean Prod. 14:57-71.

Bradley, C. University of Illinois Extension and Outreach: Estimates of soybean yield

reduction caused by disease in the United States. University of Illinois. Retrieved

September 2016, from:

http://extension.cropsciences.illinois.edu/fieldcrops/diseases/yield_reductions.php

Bradley, C. A. 2008. Effect of fungicide seed treatments on stand establishment, seedling

disease, and yield of soybean in North Dakota. Plant Dis. 92:120-125.

Bradley, C. A., Wax, L. M., Ebelhar, S. A., Bollero, G. A. and Pedersen, W. L. 2001. The

effect of fungicide seed protectants, seeding rates, and reduced rates of herbicides

on no-till soybean. Crop Protect. 20:615-622.

Brantner, J. R., and Windels, C. E. 1998. Variability in sensitivity to metalaxyl in vitro,

pathogenicity, and control of Pythium spp. on sugar beet. Plant Dis. 82:896-899.

Broders, K.D., Lipps, P.E., Paul, P.A. and Dorrance, A.E. 2007. Characterization of

Pythium spp. associated with corn and soybean seed and seedling disease in Ohio.

Plant Dis. 91:727-735.

151

Broders, K.D., Lipps, P.E., Paul, P.A. and Dorrance, A.E. 2007. Evaluation of Fusarium

graminearum associated with corn and soybean seed and seedling disease in

Ohio. Plant Dis. 91:1155-1160.

Broders, K. D., Wallhead, M. W., Austin, G. D., Lipps, P. E., Paul, P. A., Mullen, R. W.

and Dorrance, A. E. 2009. Association of soil chemical and physical properties

with Pythium species diversity, community composition, and disease incidence.

Phytopathology. 99:957-967.

Buckler, E.S., Holland, J.B., Bradbury, P.J., Acharya, C.B., Brown, P.J., Browne, C.,

Ersoz, E., Flint-Garcia, S., Garcia, A., Glaubitz, J.C. and Goodman, M.M. 2009.

The genetic architecture of maize flowering time. Science. 325:714-718.

Burnham, K.D., Dorrance, A.E., VanToai, T.T. and St Martin, S.K. 2003. Quantitative

trait loci for partial resistance to Phytophthora sojae in soybean. Crop Sci.

43:1610-1617.

Callan, N. W., Mathre, D. and Miller, J. B. 1990. Bio-priming seed treatment for

biological control of Pythium ultimum preemergence damping-off in sh-2 sweet

corn. Plant Dis. 74:368-372.

Casas, A. T., Kaiser, W. J., and Ingramn, D. M. 1990. Control of Pythium seed rot and

preemergence damping-off of chickpea in the U.S. Pacific Northwest and Spain.

Plant Dis. 74:563-569.

Chawla, S., Bowen, C. R., Slaminko, T. L., Hobbs, H. A., and Hartman, G. L. 2013. A

public program to evaluate commercial soybean cultivars for pathogen and pest

resistance. Plant Dis. 97:568-578.

152

Chisholm, S. T., Coaker, G., Day, B., and Staskawicz, B. J. 2006. Review host-microbe

interactions: Shaping the evolution of the plant immune response. Cell. 124:803-

814.

Cohen, Y., Reuveni, M., and Eyal, H. 1979. The systemic antifungal activity of Ridomil

against Phytophthora infestans on plants. Phytopathology. 69:645-649.

Concibido, V., La Vallee, B., Mclaird, P., Pineda, N., Meyer, J., Hummel, L., Yang, J.,

Wu, K. and Delannay, X. 2003. Introgression of a quantitative trait locus for yield

from Glycine soja into commercial soybean cultivars. Theor. Appl. Genet.

106:575-582.

Cook, J. P., McMullen, M. D., Holland, J. B., Tian, F., Bradbury, P., Ross-Ibarra, J.,

Buckler, E. S. and Flint-Garcia, S. A. 2012. Genetic architecture of maize kernel

composition in the nested association mapping and inbred association panels.

Plant Phys. 158:824-834.

Cooke, D.E.L., Drenth, A., Duncan, J.M., Wagels, G. and Brasier, C.M. 2000. A

molecular phylogeny of Phytophthora and related oomycetes. Fun. Genet. Bio.

30:17-32.

Collard, B. C. Y., Jahufer, M. Z. Z., Brouwer, J. B. and Pang, E. C. K. 2005. An

introduction to markers, quantitative trait loci (QTL) mapping and marker-

assisted selection for crop improvement: the basic concepts. Euphytica. 142:169-

196.

153

Darvasi, A. and Weller, J. I. 1992. On the use of the moments method of estimation to

obtain approximate maximum likelihood estimates of linkage between a genetic

marker and a quantitative locus. Heredity. 68:43-46.

Davidse, L. C., Hofman, A. E. and Velthuis, G. C. 1983. Specific interference of

metalaxyl with endogenous RNA polymerase activity in isolated nuclei from

Phytophthora megasperma f. sp. medicaginis. Exp. Myco. 7:344-361. de Cock, A. W. A. M., and Levesque, C. A. 2004. New species of Pythium and

Phytophthora. Stud. Mycol. 50:481-487. de Cock, A. W. A. M., Lodhi, A. M., Rintoul, T. L., Bala, K., Robideau, G. P., Abad, Z.

G., Coffey, M. D., Shahzad, S. and Lévesque, C. A., 2015. Phytopythium:

molecular phylogeny and systematics. Persoonia. 34:25-39.

Diers, B.W. 2014. SoyNAM Project Update. Soybean Breeders Workshop, St. Louis

MO.

http://soybase.org/meeting_presentations/soybean_breeders_workshop/SBW_201

4/presentations/Diers_SBW2014.pdf

Diers, B. 2015. Nested association mapping of agronomic traits in soybean [Powerpoint

slides]. Available at: https://www.soybase.org/SoyNAM/soynamdetails.php.

Dodd, J.L. and White, D.G. 1999. Seed rot, seedling blight, and damping-off. Pages 10-

11 in: Compendium of Corn Diseases. 3rd ed. D. G. White, ed. The American

Phytopathological Society. Saint Paul, MN.

Doerge, R.W. 2002. Mapping and analysis of quantitative trait loci in experimental

populations. Nature Rev. Genetics. 3:43.

154

Donaldson, S. P., and Deacon, J. 1993. Effects of amino acids and sugars on zoospore

taxis, encystment and cyst germination in Pythium aphanidermatum (Edson)

Fitzp., P. catenulatum Matthews and P. dissotocum Drechs. New Phytol.

123:289-295.

Dorrance, A.E., Berry, S.A., Bowen, P. and Lipps, P.E. 2004. Characterization of

Pythium spp. from three Ohio fields for pathogenicity on corn and soybean and

metalaxyl sensitivity. Plant Health Prog. doi:10.1094/PHP-2004-0202-01-RS.

Dorrance, A. E., Ellis, M. L., McDuffee, D., and Arthur, K. 2012. Efficacy of ethaboxam

toward species of Phytophthora and Pythium. Phytopathology. 102:S4.31.

Dorrance, A.E., Kleinhenz, M.D., McClure, S.A. and Tuttle, N.T. 2003. Temperature,

moisture, and seed treatment effects on root rot of soybean.

Plant Dis. 87:533-538.

Dorrance, A. E., Kurle, J., Robertson, A. E., Bradley, C. A., Giesler, L., Wise, K. and

Concibido, V. C. 2016. Pathotype diversity of Phytophthora sojae in eleven states

in the United States. Plant Dis.100:1429-1437.

Dorrance, A. E. and McClure, S. A. 2001. Beneficial effects of fungicide seed treatments

for soybean cultivars with partial resistance to Phytophthora sojae. Plant Dis.

85:1063-1068.

Dorrance, A. E., McClure, S. A., and DeSilva, A. 2003. Pathogenic diversity of

Phytophthora sojae in Ohio soybean fields. Plant Dis. 87:139-146.

155

Dorrance, A. E., McClure, S. A. and St. Martin, S. K. 2003. Effect of partial resistance on

Phytophthora stem rot incidence and yield of soybean in Ohio. Plant Dis. 87:308-

312.

Dorrance, A. E., Robertson, A. E., Cianzo, S., Giesler, L. J., Grau, C. R., Draper, M. A.,

Tenuta, A. U. and Anderson, T. R. 2009. Integrated management strategies for

Phytophthora sojae combining host resistance and seed treatments. Plant Dis.

93:875-882.

Dwevedi, A., and Kayastha, A. M. 2011. Soybean: A multifaceted legume with enormous

economic capabilities in soybean. Biochem. Chem. and Physiol. 165-188.

Ellis, M.L., Broders, K.D., Paul, P.A. and Dorrance, A.E. 2011. Infection of soybean seed

by Fusarium graminearum and effect of seed treatments on disease under

controlled conditions. Plant Dis. 95:401-407.

Ellis, M. L., McHale, L. K., Paul, P. A., St. Martin, S. K., and Dorrance, A. E. 2013.

Soybean germplasm resistant to Pythium irregulare and molecular mapping of

resistance quantitative trait loci derived from the soybean accession PI 424354.

Crop Sci. 53:1008-1021.

Ellis, M.L., Wang, H., Paul, P.A., St Martin, S.K., McHale, L.K. and Dorrance, A.E.

2012. Identification of soybean genotypes resistant to Fusarium graminearum and

genetic mapping of resistance quantitative trait loci in the cultivar Conrad. Crop

Sci. 52:2224-2233.

Esker, P. D. and Conley, S. P. 2012. Probability of yield response and breaking even for

soybean seed treatments. Crop Sci. 52:351-359.

156

Eyre, M. M. 2016. Characterizing the Effect of Increased Phosphorus and Potassium on

Seedling Diseases and Sudden Death Syndrome of Soybean in Ohio. M.S. Thesis,

The Ohio State University.

Fehr, W. R., Cianzo, S. R., Voss, B. K., and. Schultz, S.P. 1989. Registration of ‘Conrad’

soybean. Crop Sci. 29:830-831.

Ferreira, A., Silva, M.F.D. and Cruz, C.D. 2006. Estimating the effects of population size

and type on the accuracy of genetic maps. Genet. Mol. Bio. 29:187-192.

Food and Agriculture Organization Corporate Statistical Database (FAOSTAT). Food

and Agriculture Organization (FAO). Retrieved from: www.fao.org/faostat/en/.

Garzón, C.D., Molineros, J.E., Yánez, J.M., Flores, F.J., del Mar Jiménez-Gasco, M. and

Moorman, G.W. 2011. Sublethal doses of mefenoxam enhance Pythium damping-

off of geranium. Plant Dis. 95:1233-1238.

Gaspar, A.P., Marburger, D.A., Mourtzinis, S. and Conley, S.P. 2014. Soybean seed yield

response to multiple seed treatment components across diverse environments.

Agronomy J. 106:1955-1962.

Gaspar, A. P., Mitchell, P. D. and Conley, S. P. 2015. Economic risk and profitability of

soybean fungicide and insecticide seed treatments at reduced seeding rates. Crop

Sci. 55:924-933.

Gaspar, A. P., Mueller, D. S., Wise, K. A., Chilvers, M. I., Tenuta, A. U. and Conley, S.

P. 2017. Response of broad-spectrum and target-specific seed treatments and

seeding rate on soybean seed yield, profitability, and economic risk. Crop Sci.

57:2251-2262.

157

Gisi, U., Sierotzki, H., Cook, A. and McCaffery, A. 2002. Mechanisms influencing the

evolution of resistance to Qo inhibitor fungicides. Pest Manag. Sci. 58:859-867.

Glazebrook, J. 2005. Contrasting mechanisms of defense against biotrophic and

necrotrophic pathogens. Ann. Rev. Phytopath. 43:205-227.

Global Agricultural Trade System Online. United States Department of Agriculture.

Retrieved 2015 from: https://apps.fas.usda.gov/gats/default.aspx.

Gold, S. E., and Stanghellini, M. E. 1985. Effects of temperature on Pythium root rot of

spinach grown under hydroponic conditions. Phytopathology. 75:333-337.

Gordon, S.G., S.K. St. Martin, and A.E. Dorrance. 2006. Rps8 maps to a resistance gene

rich region on soybean molecular linkage group F. Crop Sci. 46:168-173.

doi:10.2135/cropsci2004.04-0024

Grant, D., Nelson, R.T., Cannon, S.B. and Shoemaker, R.C. 2009. SoyBase, the USDA-

ARS soybean genetics and genomics database. Nucleic Acids Res.

38(suppl_1)pp.D843-D846.

Guo, X., Wang, D., Gordon, S.G., Helliwell, E., Smith, T., Berry, S.A., St Martin, S.K.

and Dorrance, A.E. 2008. Genetic mapping of QTLs underlying partial resistance

to Sclerotinia sclerotiorum in soybean PI 391589A and PI 391589B. Crop Sci.

48:1129-1139.

Guy, S. O. and Oplinger, E. S. 1989. Soybean cultivar performance as influenced by

tillage system and seed treatment. J. Prod. Ag. 2:57-62.

Guy, S. O., Oplinger, E. S. and Grau, C. R. 1989. Soybean cultivar response to metalaxyl

applied in furrow and as a seed treatment. Agro. J. 81:529-532.

158

Hall, D., Tegström, C. and Ingvarsson, P.K. 2010. Using association mapping to dissect

the genetic basis of complex traits in plants. Brief. Funct. Genom. 9:157-165.

Hamman, B., Egli, D.B. and Koning, G. 2002. Seed vigor, soilborne pathogens,

preemergent growth, and soybean seedling emergence. Crop Sci. 42:451-457.

Han, Y., Teng, W., Yu, K., Poysa, V., Anderson, T., Qiu, L., Lightfoot, D.A. and Li, W.

2008. Mapping QTL tolerance to Phytophthora root rot in soybean using

microsatellite and RAPD/SCAR derived markers. Euphytica. 162:231-239.

Henshaw, T.L., Gilbert, R.A., Scholberg, J.M.S. and Sinclair, T.R. 2007. Soya bean

(Glycine max L. Merr.) genotype response to early‐season flooding: I. Root and

nodule development. J. of Ag. Crop Sci. 193:177-188.

Henshaw, T.L., Gilbert, R.A., Scholberg, J.M.S. and Sinclair, T.R. 2007. Soya bean

(Glycine max L. Merr.) genotype response to early‐season flooding: II.

Aboveground growth and biomass. J. of Ag. Crop Sci. 193:189-197.

Hirschhorn, J.N. and Daly, M.J. 2005. Genome-wide association studies for common

diseases and complex traits. Nature Rev. Genet. 6:95.

Hong, C.X. and Moorman, G.W. 2005. Plant pathogens in irrigation water: challenges

and opportunities. Crit. Rev. Plant Sci. 24:189-208.

Hwang, S. F., Gossen, B. D., Chang, K. F., Turubull, G. D., and Howard, R. J. 2001.

Effect of seed damage and metalaxyl seed treatment on Pythium seedling blight

and seed yield of field pea. Can. J. Plant Sci. 81:509-517.

Ingram, D.M. and Cook, R.J. 1990. Pathogenicity of four Pythium species to wheat,

barley, peas and . Plant Path. 39:110-117.

159

Iqbal, M.J., Meksem, K., Njiti, V.N., Kassem, M.A. and Lightfoot, D.A. 2001.

Microsatellite markers identify three additional quantitative trait loci for

resistance to soybean sudden-death syndrome (SDS) in Essex × Forrest RILs.

Theor. Appl. Genet. 102:187-192.

Jenkins, S. F., and Averre, C. W. 1983. Root diseases of in hydroponic culture

systems in North Carolina Greenhouses. Plant Dis. 67:968-970.

Jiao, Y., Vuong, T.D., Liu, Y., Meinhardt, C., Liu, Y., Joshi, T., Cregan, P.B., Xu, D.,

Shannon, J.G. and Nguyen, H.T. 2015. Identification and evaluation of

quantitative trait loci underlying resistance to multiple HG types of soybean cyst

nematode in soybean PI 437655. Theor. Appl. Genet. 128:15-23.

Jones, S. W., Donaldson, S. P., and Deacon, J. W. 1991. Behavior of zoospores and

zoospore cysts in relation to root infection by Pythium aphanidermatum. New

Phytol. 117:289-301.

Kabelka, E.A., Carlson, S.R. and Diers, B.W. 2005. Localization of two loci that confer

resistance to soybean cyst nematode from Glycine soja PI 468916. Crop Sci.

45:2473-2481.

Kandel, Y.R., Wise, K.A., Bradley, C.A., Chilvers, M.I., Tenuta, A.U. and Mueller, D.S.

2016. Fungicide and cultivar effects on sudden death syndrome and yield of

soybean. Plant Dis. 100:1339-1350.

Kazi, S., Shultz, J., Afzal, J., Johnson, J., Njiti, V.N. and Lightfoot, D.A. 2008. Separate

loci underlie resistance to root infection and leaf scorch during soybean sudden

death syndrome. Theor. Appl. Genet. 116:967-977.

160

Keeling, B. L. 1974. Soybean seed rot and the relation of seed exudate to host

susceptibility. Phytopathology. 64:1445-1447.

Kim, H.S. and Diers, B.W. 2000. Inheritance of partial resistance to Sclerotinia stem rot

in soybean. Crop Sci. 40:55-61.

Kim, D. S., Chun, S. J., Jeon, J. J., Lee, S. W. and Joe, G. H. 2004. Synthesis and

fungicidal activity of ethaboxam against oomycetes. Pest Manag. Sci. 60:1007-

1012.

Kim, D.S., Prak, H.C., Chun, S.J., Yu, S.H., Park, K.J., Oh, J.H., Shin, K.H., Koh, Y.J.,

Kim, B.S., Hahm, Y.I. and Chung, B.K. 1999. Field performance of a new

fungicide ethaboxam against cucumber downy mildew, potato late blight and

pepper Phytophthora blight in Korea. Plant Path. J. 15:48-52.

Kim, B.S., Zhang, X.Z., Chung, E.K., Kim, D.S., Chun, S.J. and Park, W.B. 2003.

Sensitivity of Phytophthora infestans isolates to fungicides metalaxyl and

ethaboxam in Korea. Plant Path. J. 19: 143-147.

Kirkpatrick, M. T., Rothrock, C. S., Rupe, J. C., and Gbur, E. E. 2006. The effect of

Pythium ultimum and soil flooding on two soybean cultivars. Plant Dis. 90:597-

602.

Koenning, S. R. and Wrather, J. A. 2010. Suppression of soybean yield potential in the

continental United States by plant diseases from 2006 to 2009. Plant Health Prog.

10. doi:10.1094/PHP-2010-1122-01-RS.

Kou, Y., and Wang, S. 2010. Broad-spectrum and durability: understanding of

quantitative disease resistance. Curr. Opin. Plant Biol. 13:181-185.

161

Kou, Y. and Wang, S. 2012. Toward an understanding of the molecular basis of

quantitative disease resistance in rice. J. of Biotech. 159:283-290.

Krattinger, S.G., Lagudah, E.S., Spielmeyer, W., Singh, R.P., Huerta-Espino, J.,

McFadden, H., Bossolini, E., Selter, L.L. and Keller, B. 2009. A putative ABC

transporter confers durable resistance to multiple fungal pathogens in wheat.

Science. 323:1360-1363.

Kump, K.L., Bradbury, P.J., Wisser, R.J., Buckler, E.S., Belcher, A.R., Oropeza-Rosas,

M.A., Zwonitzer, J.C., Kresovich, S., McMullen, M.D., Ware, D. and Balint-

Kurti, P.J. 2011. Genome-wide association study of quantitative resistance to

southern leaf blight in the maize nested association mapping population. Nature

Genet. 43:163.

La, T.C., Pathan, S.M., Vuong, T., Lee, J.D., Scaboo, A.M., Smith, J.R., Gillen, A.M.,

Gillman, J., Ellersieck, M.R., Nguyen, H.T. and Shannon, J.G. 2014. Effect of

high-oleic acid soybean on seed oil, protein concentration, and yield. Crop Sci.

54:2054-2062.

Larkin, R. P., English, J. T., and Mihail, J. D. 1995. Identification, distribution and

comparative pathogenicity of Pythium spp associated with alfalfa seedlings. Soil

Biol. Biochem. 27:357-364.

Lee, S., Mian, R., McHale, L.K., Sneller, C.H. and Dorrance, A.E. 2013. Identification of

quantitative trait loci conditioning partial resistance to Phytophthora sojae in

soybean PI 407861A. Crop Sci. 53:1022-1031.

162

Lee, S., Mian, M.R., McHale, L.K., Wang, H., Wijeratne, A.J., Sneller, C.H. and

Dorrance, A.E. 2013. Novel quantitative trait loci for partial resistance to

Phytophthora sojae in soybean PI 398841. Theor. Appl. Genet. 126:1121-1132.

Lee, S., Mian, M.R., Sneller, C.H., Wang, H., Dorrance, A.E. and McHale, L.K. 2014.

Joint linkage QTL analyses for partial resistance to Phytophthora sojae in

soybean using six nested inbred populations with heterogeneous conditions.

Theor. Appl. Genet. 27:429-444.

Lerch, E.R. 2017. Soybean-Pythium Pathosystem: Search for Host Resistance and

Interaction with Fusarium Species. Doctoral Dissertation, Iowa State University.

Levesque, C.A. and De Cock, A.W. 2004. Molecular phylogeny and of the

genus Pythium. Myco. Res. 108:1363-1383.

Li, X., Han, Y., Teng, W., Zhang, S., Yu, K., Poysa, V., Anderson, T., Ding, J. and Li,

W. 2010. Pyramided QTL underlying tolerance to Phytophthora root rot in mega-

environments from soybean cultivars ‘Conrad’and ‘Hefeng 25’. Theor. Appl.

Genet. 121:651-658.

Li, N., Shi, J., Wang, X., Liu, G. and Wang, H. 2014. A combined linkage and regional

association mapping validation and fine mapping of two major pleiotropic QTLs

for seed weight and silique length in rapeseed (Brassica napus L.). BMC Plant

Biol. 14:114.

Lifshitz, R., Windham, M.T. and Baker, R. 1986. Mechanism of biological control of

preemergence damping-off of pea by seed treatment with spp.

Phytopathology. 76:720-725.

163

Lindhout, P. 2002. The perspectives of polygenic resistance in breeding for durable

disease resistance. Euphytica. 124:217-226.

Littrell, R. H., and McCarter, S. 1969. Effect of soil temperature on virulence of Pythium

aphanidermatum and to rye and tomato. Phytopathology.

60:704-707.

Lueschen, W. E., Evans, S. D., Ford, J. H., Hoverstad, T. R., Kanne, B. K., Orf, J. H.,

Staricka, J. A., Stienstra, W. C., Warnes, D. D. and Hicks, D. R. 1991. Soybean

production as affected by tillage in a corn and soybean management system: II.

Seed treatment response. J. Prod. Ag. 4:580-585.

Marchant, R. 1968. An ultrastructural study of sexual reproduction in Pythium ultimum.

New Phytol. 67:167-171.

Marburger, D.A., Conley, S.P., Esker, P.D., Lauer, J.G. and Ané, J.M. 2015. Yield

response to crop/genotype rotations and fungicide use to manage Fusarium-

related diseases. Crop Sci. 55:889-898.

Martin, F.N. and Loper, J.E. 1999. Soilborne plant diseases caused by Pythium spp.:

Ecology, epidemiology, and prospects for biological control. Critical Rev. Plant

Sci. 18:111-181.

Matthiesen, R. L., Ahmad, A. A., Robertson, A. E. 2016. Temperature affects

aggressiveness and fungicide sensitivity of four Pythium spp. that cause soybean

and corn damping off in Iowa. Plant Dis. 100:583-591.

164

McCarter, S. M., and Littrell, R. H. 1969. Comparative pathogenicity of Pythium

aphanidermatum and Pythium myriotylum to twelve plant species and

intraspecific variation in virulence. Phytopathology. 60:264-268.

McBlain, B.A., Fioritto, R.J., St Martin, S.K., Calip-DuBois, A., Schmitthenner, A.F.,

Cooper, R.L. and Martin, R.J. 1990. Registration of 'Resnik' soybean. Crop Sci.

30:424-425.

McMullen, M.D., Kresovich, S., Villeda, H.S., Bradbury, P., Li, H., Sun, Q., Flint-

Garcia, S., Thornsberry, J., Acharya, C., Bottoms, C. and Brown, P. 2009. Genetic

properties of the maize nested association mapping population. Science. 325:737-

740.

Mian, M.A., Wang, T., Phillips, D.V., Alvernaz, J. and Boerma, H.R. 1999. Molecular

mapping of the Rcs3 gene for resistance to frogeye leaf spot in soybean. Crop Sci.

39:1687-1691.

Michelmore, R.W., Christopoulou, M. and Caldwell, K.S. 2013. Impacts of resistance

gene genetics, function, and evolution on a durable future. Ann. Rev. Phytopath.

51:291-319.

Middleton, J. T. 1943. The taxonomy, host range and geographical distribution of the

genus Pythium. Mem. Torrey Bot. Club. 20:1-171.

Mideros, S., Nita, M. and Dorrance, A.E. 2007. Characterization of components of partial

resistance, Rps2, and root resistance to Phytophthora sojae in soybean.

Phytopathology. 97:655-662.

165

Moorman, G.W. and Kim, S.H. 2004. Species of Pythium from greenhouses in

Pennsylvania exhibit resistance to propamocarb and mefenoxam. Plant Dis.

88:630-632.

Mora, S. X. M. 2006. Study of incomplete resistance to Phytophthora sojae in soybean.

Master’s Thesis. Ohio State University.

Morel, J.B. and Dangl, J.L. 1997. The hypersensitive response and the induction of cell

death in plants. Cell Death Dif. 4:671-683.

Mundt, C.C. 2014. Durable resistance: a key to sustainable management of pathogens and

pests. Infect. Genet., and Evol. 27:446-455.

Munkvold, G. P., and O’Mara, J. K. 2002. Laboratory and growth chamber evaluation of

fungicidal seed treatments for maize seedling blight caused by Fusarium species.

Plant Dis. 86:143-150.

National Statistics for Soybeans. National Agricultural Statistics Service, USDA.

Retrieved from:

https://www.nass.usda.gov/Publications/Ag_Statistics/2015/index.php

Navi, S. S. and Yang, X. B. 2016. Sudden death syndrome-A growing threat of losses in

soybeans. CAB Rev. 11:1-13.

Nelson, R.L. and Johnson, E.O.C. 2011. Registration of soybean germplasm line LG00-

3372. J. Plant Reg. 5:403-405.

Njoroge, S. M. C., Kabir, Z., Martin, F. N., Koike, S. T., and Subbarao, K. V. 2009.

Comparison of crop rotation for Verticillium wilt management and effect on

166

Pythium species in conventional and organic strawberry production. Plant Dis.

93:519-527.

Pankhurst, C. E., McDonald, H. J., and Hawke, B. G. 1995. Influence of tillage and crop-

rotation on the epidemiology of Pythium infections of wheat in a red-brown earth

of South Australia. Soil Biol. Biochem. 27:1065-1073.

Pearson, C.A.S., Schwenk, F.W., Crowe, F.J. and Kelley, K. 1984. Colonization of

soybean roots by Macrophomina phaseolina. Plant Dis. 68:1086-1088.

Pham, A. T., Shannon, J. G., and Bilyeu, K. D. 2012. Combinations of mutant FAD2 and

FAD3 genes to produce high oleic acid and low linolenic acid soybean oil. Theor.

Appl. Genet. 125:503-515.

Poag, P.S., Popp, M., Rupe, J., Dixon, B., Rothrock, C. and Boger, C. 2005. Economic

evaluation of soybean fungicide seed treatments. Ag. J. 97:1647-1657.

Poland, J.A., Balint-Kurti, P.J., Wisser, R.J., Pratt, R.C. and Nelson, R.J. 2009. Shades of

gray: the world of quantitative disease resistance. Trends Plant Sci. 14:.21-29.

Poland, J.A., Bradbury, P.J., Buckler, E.S. and Nelson, R.J. 2011. Genome-wide nested

association mapping of quantitative resistance to northern leaf blight in maize.

Proc. of the Nat. Acad. of Sci. 108:6893-6898.

Pradhan, S., Flores, F.J., Molineros, J.E., Walker, N.R., Melouk, H. and Garzon, C.D.

2017. Improved assessment of mycelial growth stimulation by low doses of

mefenoxam in plant pathogenic Globisporangium species. European J. Plant Path.

147:477-487.

167

Radmer, L., Anderson, G., Malvick, D.M., Kurle, J.E., Rendahl, A. and Mallik, A. 2017.

Pythium, Phytophthora, and Phytopythium spp. associated with soybean in

Minnesota, their relative aggressiveness on soybean and corn, and their sensitivity

to seed treatment fungicides. Plant Dis. 101:62-72.

Raftoyannis, Y. and Dick, M.W. 2006. Zoospore encystment and pathogenicity of

Phytophthora and Pythium species on plant roots. Microbio. Res. 161:1-8.

Rizvi, S. S. A., and Yang, X. B. 1996. Fungi associated with soybean seedling disease in

Iowa. Plant Dis. 80:57-60.

Rojas-Flechas, J. A. 2016. Diversity of Oomycetes Associated with Soybean Seedling

Diseases. Doctoral Dissertation, Michigan State University.

Rojas, J. A., Jacobs, J. L., Napieralski, S., Karaj, B., Bradley, C. A., Chase, T., Esker, P.

D., Giesler, L. J., Jardine, D. J., Malvick, D. K. and Markell, S.G. 2017.

Oomycete species associated with soybean seedlings in North America-Part I:

Identification and pathogenicity characterization. Phytopathology. 107:280-292.

Rojas, J. A., Jacobs, J. L., Napieralski, S., Karaj, B., Bradley, C. A., Chase, T., Esker, P.

D., Giesler, L. J., Jardine, D. J., Malvick, D. K. and Markell, S. G. 2017.

Oomycete species associated with soybean seedlings in North America—part II:

diversity and ecology in relation to environmental and edaphic factors.

Phytopathology. 107:293-304.

Rosso, M.L., Rupe, J.C., Chen, P. and Mozzoni, L.A. 2008. Inheritance and genetic

mapping of resistance to damping-off caused by Pythium aphanidermatum in

‘Archer’ soybean. Crop Sci. 48:2215-2222.

168

Rothrock, C. S., Kirkpatrick, T. L., Frans, R. E., and Scott, H. D. 1995. The influence of

winter legume cover crops on soilborne plant-pathogens and cotton seedling

diseases. Plant Dis. 79:167-171.

Rupe, J.C., Rothrock, C.S., Bates, G., Rosso, M.L., Avanzato, M.V. and Chen, P. 2011.

Resistance to Pythium seedling disease in soybean. Soybean-Molecular Aspects

of Breeding. A. Sudaric, ed. InTech, Rjeka, Croatia. Doi:10.5772/15301.

Sanei, S. J., Mehri, Z., and Lotfalinezhad, E. 1978. Effects of soybean seed coat cracks

on seed exudation and seedling quality in soil infested with Pythium ultimum.

Phytopathology. 68:1186.

Schmitthenner, A.F. 1985. Problems and progress in control of Phytophthora root rot of

soybean. Plant Dis. 69:362-368.

Scott, K., Vargas, A., Eyre, M. and Dorrance, A. 2016. Efficacy of three soybean

fungicide seed treatments against Pythium species in seed plate and growth

chamber assays. Phytopathology. 106:S4:68.

Schroeder, K., Martin, F. N., de Cock, A., Levesque, C. A., Spies, C., Okubara, P., and

Paulitz, T.C. 2013. Molecular detection and quantification of Pythium species-

Evolving taxonomy, new tools and challenges. Plant Dis. 97:4-20.

Shah, D.A. and Madden, L.V. 2004. Nonparametric analysis of ordinal data in designed

factorial experiments. Phytopathology. 94:33-43.

Sharma, K.K., Singh, U.S., Sharma, P., Kumar, A. and Sharma, L. 2015. Seed treatments

for sustainable agriculture- A review. J. App. Nat. Sci. 7:521-539.

169

Song, Q., Yan, L., Quigley, C., Jordan, B.D., Fickus, E., Schroeder, S., Song, B.H.,

Charles An, Y.Q., Hyten, D., Nelson, R. and Rainey, K. 2017. Genetic

characterization of the soybean nested association mapping population. Plant

Genet. 10:1-14.

SoyNAM. Nested Associated Mapping (NAM) of Genes Controlling Soybean Yield and

Other Key Traits. Retrieved September 2016, from:

http://soybase.org/projects/SoyBase.C2014.02.php

St. Clair, D.A. 2010. Quantitative disease resistance and quantitative resistance loci in

breeding. Ann. Rev. Phytopath. 48:247-268.

St. Martin S. K., Feller M. K., McIntyre S. A., Fioritto R. J., Dorrance A.E., Berry S. A.,

Sneller C.H. 2008. Registration of ‘Dennison’ soybean. J. Plant Reg. 2:21.

St. Martin S. K., Mills G. R., Fioritto R. J., McIntyre S. A., Schmitthenner A., Dorrance

A. E., Cooper R. L. 2001. Registration of ‘Kottman’ soybean. Crop Sci. 41:590-

591.

Stanghellini, M.E., and Burr, T.J. 1973. Germination in vivo of Pythium aphanidermatum oospores and sporangia. Phytopathology. 63:1493-1496.

Stanghellini, M.E., and Hancock, J.G. 1971. The sporangium of Pythium ultimum as a survival structure in soil. Phytopathology. 61:157-164.

Stanghellini, M.E., and Hancock, J.G. 1971. Radial extent of the bean spermosphere and

its relation to the behavior of Pythium ultimum. Phytopathology. 61:165-168.

170

Stasko, A.K., Wickramasinghe, D., Nauth, B.J., Acharya, B., Ellis, M.L., Taylor, C.G.,

McHale, L.K. and Dorrance, A.E. 2016. High-density mapping of resistance QTL

toward Phytophthora sojae, Pythium irregulare, and Fusarium graminearum in

the same soybean population. Crop Sci. 56:2476-2492.

Stewart, S., Robertson, A. E., Wickramasinghe, D., Draper, M. A., Michel, A. and

Dorrance, A. E. 2016. Population structure among and within Iowa, Missouri,

Ohio, and South Dakota populations of Phytophthora sojae. Plant Dis. 100:367-

379.

Stich, B. 2009. Comparison of mating designs for establishing nested association

mapping populations in maize and Arabidopsis thaliana. Genetics. 183:1525-

1534.

Stroup, W.W. 1989. Why mixed models? Applications of mixed models in agriculture

and related disciplines. South Coop Ser Bull 343:1-8. Louisiana Agricultural

Experiment Station, Baton Rouge, Louisiana.

Sutton, J. C., Sopher, C. R., Owen-going, T. N., Liu, W., Grodzinski, B., Hall, J. C., and

Benchimol, R.L. 2006. Etiology and epidemiology of Pythium root rot in

hydroponic crops: Current knowledge and perspectives. Summa Phytopath.

32:307-321.

Tanksley, S.D., Young, N.D., Paterson, A.H., and Bonierbale, M.W. 1989. RFLP

mapping in plant breeding: new tools for an old science. Nat. Biotech. 7:257-264.

171

Taylor, R.J., Salas, B., Secor, G.A., Rivera, V. and Gudmestad, N.C. 2002. Sensitivity of

North American isolates of Phytophthora erythroseptica and Pythium ultimum to

mefenoxam (metalaxyl). Plant Dis. 86:797-802.

Thomson, T. B., Athow, K. L., and Laviolette, F. A. 1971. The effect of temperature on

the pathogenicity of Pythium aphanidermatum, P. debaryanum, and P. ultimum

on soybean. Phytopathology. 61:933-935.

Tian, F., Bradbury, P.J., Brown, P.J., Hung, H., Sun, Q., Flint-Garcia, S., Rocheford,

T.R., McMullen, M.D., Holland, J.B. and Buckler, E.S. 2011. Genome-wide

association study of leaf architecture in the maize nested association mapping

population. Nature Genet. 43:159-162.

Uchida, M., Roberson, R. W., Chun, S. J. and Kim, D. S. 2005. In vivo effects of the

fungicide ethaboxam on microtubule integrity in Phytophthora infestans. Pest

Manag. Sci. 61:787-792.

United States Census Bureau. United States Department of Agriculture. Retrieved 2010

from: http://www.fao.org/faostat/en/.

Urrea, K., Rupe, J. C. and Rothrock, C. S. 2013. Effect of fungicide seed treatments,

cultivars, and soils on soybean stand establishment. Plant Dis. 97:807-812.

Urrea, K., Rupe, J., Chen, P. and Rothrock, C.S. 2017. Characterization of seed rot

resistance to Pythium aphanidermatum in soybean. Crop Sci. 57:1394-1403.

Van Ooijen, J.W. 1999. LOD significance thresholds for QTL analysis in experimental

populations of diploid species. Heredity. 83:613-624.

172

Van Ooijen, J. W. 2004. MapQTL® 5. Software for the mapping of quantitative trait loci

in experimental populations. Kyazma BV, Wageningen. van der Plaats-Niterink, A.J. 1981. Monograph of the Genus Pythium. Studies in

Mycology No. No. 21. Baarn, the Netherlands: Centraalbureau voor

Schimmelcultures. van West, P., Appiah, A. A., and Gow, N. A. R. 2003. Advances in research on oomycete

root pathogens. Physiol. Mol. Plant Pathol. 62:99-113.

VanToai, T.T., Beuerlein, A.F., Schmitthenner, S.K. and St Martin, S.K. 1994. Genetic

variability for flooding tolerance in soybeans. Crop Sci. 34:1112-1115.

Walker, R., Rossall, S. and Asher, M.J.C. 2004. Comparison of application methods to

prolong the survival of potential biocontrol bacteria on stored sugar‐beet seed. J.

of App. Microbio. 97:293-305.

Wang, H., Waller, L., Tripathy, S., St Martin, S.K., Zhou, L., Krampis, K., Tucker, D.M.,

Mao, Y., Hoeschele, I., Maroof, S. and Tyler, B.M. 2010. Analysis of genes

underlying soybean quantitative trait loci conferring partial resistance to

Phytophthora sojae. The Plant Genet. 3:23-40.

Wang, H., St Martin, S.K. and Dorrance, A.E. 2012. Comparison of phenotypic methods

and yield contributions of quantitative trait loci for partial resistance to

Phytophthora sojae in soybean. Crop Sci. 52:609-622.

Wang, H., Wijeratne, A., Wijeratne, S., Lee, S., Taylor, C.G., St Martin, S.K., McHale,

L. and Dorrance, A.E. 2012. Dissection of two soybean QTL conferring partial

173

resistance to Phytophthora sojae through sequence and gene expression analysis.

BMC Genomics. 13:428.

Watanabe, H., Kageyama, K., Taguchi, Y., Horinouchi, H., and Hyakumachi, M. 2008.

Bait method to detect Pythium species that grow at high temperatures in

hydroponic solutions. J. Gen. Plant Pathol. 74:417-424.

Waterhouse GM, 1968. Monograph of the genus Pythium. Mycologie. 21:1-242.

Wei, L., Xue, A., Cober, E., Babcock, C., Zhang, J., Zhang, S., Li, W., Wu, J. and Liu, L.

2010. Pathogenicity of Pythium species causing seed rot and damping-off in

soybean under controlled conditions. Phytoprotection. 91:3-10.

Weiland, J. E., Santamaria, L., and Grünwald, N. J. 2014. Sensitivity of Pythium

irregulare, P. sylvaticum, and P. ultimum from forest nurseries to mefenoxam and

fosetyl-Al, and control of Pythium damping-off with fungicide. Plant Dis. 98:937-

942.

Wickramasinghe, D., Acharya, B., Plewa, D., and Dorrance, A.E. 2012. Screening for

resistance to Phytophthora sojae, Fusarium graminearum, and Pythium

irregulare in Nested Association Mapping (NAM) populations. Phytopathology.

102:S5.9.

Wilcox, J. R. 2001. Sixty years of improvement in publicly developed elite soybean lines.

Crop Sci. 41:1711-1716.

Workneh, F., Yang, X. B. and Tylka, G. L. 1998. Effect of tillage practices on vertical

distribution of Phytophthora sojae. Plant Dis. 82:1258-1263.

174

Workneh, F., Yang, X. B., and Tylka, G. L. 1999. Soybean brown stem rot, Phytophthora

sojae, and Heterodera glycines affected by soil texture and tillage relations.

Phytopathology. 89:844-850.

Wrather, J. A. and Koenning, S. R. 2006. Estimates of disease effects on soybean yields

in the United States 2003 to 2005. J. of Nema. 38:173.

Wrather, J. A., Stienstra, W. C. and Koenning, S. R. 2001. Soybean disease loss estimates

for the United States from 1996 to 1998. Can. J. Plant Path. 23:122-131.

Xavier, A., Hall, B., Casteel, S., Muir, W. and Rainey, K.M. 2017. Using unsupervised

learning techniques to assess interactions among complex traits in soybeans.

Euphytica. 213:200. DOI 10.1007/s10681-017-1975-4.

Xavier, A., Jarquin, D., Howard, R., Ramasubramanian, V., Specht, J.E., Graef, G.L.,

Beavis, W.D., Diers, B.W., Song, Q., Cregan, P.B. and Nelson, R., Mian, R.,

Grover Shannon, J., McHale, L., Wang, D., Schapaugh, W., Lorenz, A., Xu, S.,

Muir, W., and Rainey, K. 2018. Genome-wide analysis of grain yield stability and

environmental interactions in a multiparental soybean population. G3: Genes,

Genomes, Genet. 8:519-529.

Yang, X.B. 1999. Pythium damping-off and root rot. Pages 42-44 in: Compendium of

Soybean Diseases. G.L. Hartman, J.B. Sinclair, and J.C. Rupe, ed. 4th edition.

APS Press. St. Paul. MN 55121.

Yang, X. B. and Feng, F. 2001. Ranges and diversity of soybean fungal diseases in North

America. Phytopathology. 91:769-775.

175

Yin, Y., Fatufe, A. A., and Blachier, F. 2007. Soya bean meal and its extensive use in

livestock feeding and nutrition. Hany El-Shemy (Ed.). In Soybean and Nutrition.

269-384.

Young, N.D., 1996. QTL mapping and quantitative disease resistance in plants. Ann.

Rev. Phytopath. 34:479-501.

Yu, J., Holland, J.B., McMullen, M.D. and Buckler, E.S. 2008. Genetic design and

statistical power of nested association mapping in maize. Genetics. 178:539-551.

Zelaya-Molina, L.X., Ellis, M.L., Berry, S.A., and Dorrance, A.E. 2010. First report of

Phytophthora sansomeana causing wilting and stunting on corn in Ohio. Plant

Dis. 94: 125.

Zelaya-Molina, L. X., Ortega, M. A. and Dorrance, A. E. 2011. Easy and efficient

protocol for oomycete DNA extraction suitable for population genetic analysis.

Biotech. Letters. 33:715-720.

Zerillo, M. M., Adhikari, B. N., Hamilton, J. P., Buell, C. R., Lévesque, C. A. and

Tisserat, N. 2013. Carbohydrate-active enzymes in Pythium and their role in plant

cell wall and storage polysaccharide degradation. PLoS One. 8: p.e72572.

Zhang, B. Q., and Yang, X. B. 2000. Pathogenicity of Pythium Populations from corn-

soybean rotation fields. Plant Dis. 84:94-99.

Zitnick-Anderson, K.K., Norland, J.E., Luis, E., Fortuna, A.M. and Nelson, B.D. 2017.

Probability models based on soil properties for predicting presence-absence of

Pythium in soybean roots. Microb. Eco. 74:550-560.

176