INFECTION OF BLUEBERRIES BY COLLETOTRICHUM ACUTATUM: HOST DEFENSES, INHERITANCE OF RESISTANCE AND ENVIRONMENTAL EFFECTS By Timothy David Miles A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Plant Pathology 2011 i ABSTRACT INFECTION OF BLUEBERRIES BY COLLETOTRICHUM ACUTATUM: HOST DEFENSES, INHERITANCE OF RESISTANCE AND ENVIRONMENTAL EFFECTS By Timothy David Miles Anthracnose fruit caused by Colletotrichum acutatum is the most important postharvest disease of blueberries (Vaccinium corymbosum). In order to develop a disease forecasting model for anthracnose fruit rot in blueberries, the effects of temperature, wetness duration, relative humidity and wetness interruptions were studied on appressorium formation and infection of green and blue fruit by C. acutatum. Three-dimensional response surfaces were fitted to the data. A model combining all variables was developed. The resistance response in blueberry cultivar ‘Elliott’ was investigated. Using suppression subtraction hybridization, several defense-related genes as well as abiotic stress-related genes were found to be upregulated in ‘Elliott’ but not in ‘Jersey’ fruit within 24 hours after inoculation. Some genes were related to oxidative stress in plant tissues. Higher levels of hydrogen peroxide were also found in 'Elliott' than 'Jersey' fruit after inoculation. To investigate the role of antifungal compounds in resistance, a series of fruit extractions was performed with water, methanol, and ethyl acetate, and fractions were tested for antifungal activity in agar plate bioassays. The methanolic extract was the most biologically active; disease incidence was reduced by 88% when ‘Jersey’ fruit was pretreated with a 4% solution prior to inoculation with C. acutatum. Spectrophotometry showed that this fraction mainly contained flavonoids. Anthocyanins and flavonols were then quantified and identified in both cultivars using HPLC-MS. ‘Elliott’ fruit overall contained more anthocyanins than ‘Jersey’ fruit but the same compounds were found in both cultivars. However, two unique flavonols were present in ‘Elliott’. ii A total of 26 blueberry cultivars were screened for anthracnose resistance using several different techniques. A cut-fruit technique was promising for use as a rapid screening method. A positive linear correlation was observed between resistance and fruit sugar content but not fruit pH. The growth of C. acutatum in culture was restricted at higher sugar and lower pH levels. The inheritance of resistance was also investigated to facilitate future breeding work. Using inoculated fruit from F-1 populations of specific crosses between resistant and susceptible cultivars, resistance ratings based on AUDPC values were compared to resistance ratings predicted from previous studies. Significant correlations were observed, providing strong evidence that anthracnose resistance is highly heritable in highbush blueberries. The results from these studies provide a significant contribution to the understanding of the biology and management of anthracnose fruit rot in blueberry. iii For my parents Paula and David who have supported me since the beginning and my wife Laura who has been a great source of motivation and inspiration. iv ACKNOWLEDGMENTS I would like to express my genuine gratitude to Dr. Annemiek Schilder, who has provided me with guidance and support that was vital for the success of my Ph.D. I would like to thank the members of my guidance committee; Drs. Muraleedharan Nair, Brad Day, and Ray Hammerschmidt. Their direction was invaluable throughout my program. I would also like to thank members of my laboratory for helping me throughout my program. Special thanks go to Roger Sysak; Jerri Gillett, my second advisor; Stephen Jordan, my surrogate big brother; and undergraduate students that have assisted me: Christine Bates, Daniel Svoboda and Christopher Woelk. I would like to especially thank Michigan State University for the opportunity to meet my wife, throughout my Ph.D. Laura has provided me with love and support and I don’t think I could have made it without her. I would like to thank Dr. Andy Jarosz for help with infection modeling, Dr. James Hancock for his help in supporting my research ideas, and Dr. Daniel Jones and Christine Vandervoort for assistance with my HPLC-MS and MS-MS analyses. All of these faculty members have provided invaluable advice. A special thanks to goes to my mother Paula, my father David, and my grandmother Mary. v TABLE OF CONTENTS LIST OF TABLES .................................................................................................................... ix LIST OF FIGURES .................................................................................................................. xii CHAPTER I: LITERATURE REVIEW .................................................................................... 1 Introduction ................................................................................................................................. 1 Blueberry production in Michigan .............................................................................................. 9 Anthracnose fruit rot of blueberry ............................................................................................ 11 Environmental requirements for infection ................................................................................ 14 The infection process ................................................................................................................ 16 Passive host defenses ................................................................................................................ 18 Active host defenses ................................................................................................................. 21 Genetics of resistance ............................................................................................................... 23 Conclusions ............................................................................................................................... 25 CHAPTER II: THE EFFECT OF ENVIRONMENTAL FACTORS ON INFECTION OF BLUEBERRY FRUIT BY COLLETOTRICHUM ACUTATUM ............................................. 26 Abstract ..................................................................................................................................... 26 Introduction ............................................................................................................................... 27 Materials and Methods .............................................................................................................. 29 Results ....................................................................................................................................... 37 Discussion ................................................................................................................................. 50 Acknowledgments..................................................................................................................... 53 CHAPTER III: IDENTIFICATION OF DIFFERENTIALLY EXPRESSED GENES IN A RESISTANT VERSUS SUSCEPTIBLE CULTIVAR AFTER INFECTION BY COLLETOTRICHUM ACUTATUM......................................................................................... 54 Abstract ..................................................................................................................................... 54 Introduction ............................................................................................................................... 55 Materials and Methods .............................................................................................................. 58 Results ....................................................................................................................................... 67 Discussion ................................................................................................................................. 82 Acknowledgments..................................................................................................................... 87 CHAPTER IV: CHARACTERIZATION AND BIOLOGICAL ACTIVITY OF FLAVONOIDS FROM RIPE FRUIT OF AN ANTHRACNOSE-RESISTANT BLUEBERRY CULTIVAR 88 Abstract ..................................................................................................................................... 88 Introduction ............................................................................................................................... 89 vi Materials and Methods .............................................................................................................. 91 Results and Discussion ............................................................................................................. 96 Conclusions ............................................................................................................................. 108 Acknowledgments................................................................................................................... 109 CHAPTER V: EVALUATION OF SCREENING METHODS AND FRUIT COMPOSITION IN RELATION TO ANTHRACNOSE FRUIT ROT RESISTANCE IN BLUEBERRIES .. 110 Abstract ................................................................................................................................... 110 Introduction ............................................................................................................................. 111 Materials and Methods ............................................................................................................ 114 Results ..................................................................................................................................... 119 Discussion ............................................................................................................................... 135 Acknowledgments................................................................................................................... 139 CHAPTER VI: INHERITANCE OF RESISTANCE TO COLLETOTRICHUM ACUTATUM IN HIGHBUSH BLUEBERRY (VACCINIUM CORYMBOSUM L.) FRUIT ............................ 141 Abstract ................................................................................................................................... 141 Introduction ............................................................................................................................. 141 Materials and Methods ............................................................................................................ 143 Results ..................................................................................................................................... 147 Discussion ............................................................................................................................... 151 Acknowledgments................................................................................................................... 155 CONCLUSIONS AND FUTURE DIRECTIONS ................................................................. 156 APPENDICES ........................................................................................................................ 161 Appendix A – Raw data figures from chapter II..................................................................... 162 Appendix B – Supplemental tables and figures from chapter III ........................................... 167 Appendix C – Supplemental figures from chapter IV ............................................................ 172 Appendix D – Pathologicity of various isolates of Colletotrichum acutatum on ‘Jersey’ blueberries ............................................................................................................................... 176 Appendix E – Screening daughter blueberry plants from a ‘Draper’ x ‘Jewel’ cross for resistance to Colletotrichum acutatum .................................................................................................... 178 Appendix F – The role of ethylene in blueberry anthracnose fruit rot resistance................... 184 Appendix G – First report of grape root rot caused by Roesleria subterranea in Michigan .. 187 Appendix H – First report of Pythium sterilum causing root rot of blueberry in the United States ................................................................................................................................................. 189 REFERENCES ....................................................................................................................... 191 vii LIST OF TABLES Table 1.1. Colletotrichum species that cause diseases on fleshy fruits. ........................................ 2 Table 2.1. Fitted-models for the development of melanized appressoria and infection level of immature and mature blueberry fruits by Colletotrichum acutatum using temperature (°C), wetness duration (h), wetness interruption (h), and relative humidity (%). Nine equations were developed based on curve fitting the raw data from a mixture of experiments and models were examined for 2 goodness of fit (R ), standard error of the estimate (SEE), the P value of the regression and the mean squared error (MSE) of the residuals (* denotes a chosen equation) .................................. 35 Table 2.2. Proposed model for the development of melanized appressoria and the infection level of immature and mature blueberry fruits by Colletotrichum acutatum. Nine equations were developed based on curve fitting the raw data from a mixture of experiments and these equations were used to make three combined models........................................................................................................ 39 Table 3.1. Characteristics and predicted physiological function of differentially expressed sequence tags (ESTs) from ripe fruit of highbush blueberry cultivars Elliott versus Jersey after inoculation with Colletotrichum acutatum.The hypothetical function is based on homology to sequences in translated nucleotide databases (DDBJ/EMBL/GenBank) using TBLASTX. This -5 table only displays putative plant sequences with homologous E values lower than 1 x 10 . .... 72 Table 3.2. Primer DNA sequences, guanine-cytosine (GC) percentages, calculated primer melting temperatures (Tm), expected product size, product melting temperature, Q-RT-PCR standard curve plot slope, and calculated primer efficiency used in semi-quantitative and quantitative RT-PCR expression analysis of differentially expressed sequence tags in blueberry fruit in response to C. acutatum inoculation.. ................................................................................................................... 74 Table 4.1. Anthocyanins from the resistant blueberry cultivar Elliott and the susceptible blueberry cultivar Jersey identified in HPLC/MS analysis and quantified using HPLC/PDA. Compounds were putatively identified based on previous research, available spectra, standards, retention times and m/z ratios. Anthocyanins were quantified at a wavelength of 520 nm using a standard curve plot of cyanidin 3-O-glucoside.. ................................................................................................. 100 Table 4.2. Non-anthocyanin flavonoids from the resistant blueberry cultivar Elliott and the susceptible blueberry cultivar Jersey identified in HPLC/MS analysis and quantified using HPLC/PDA. Compounds were putatively identified based on previous research, available spectra, standards, retention times and m/z ratios. Non-anthocyanin flavonoids were quantified at a wavelength of 255 nm using a standard curve plot of quercetin 3-O-rhamnoside. .................... 103 Table 5.1. Anthracnose fruit rot resistance profiles of different blueberry cultivars after artificial inoculation with Colletotrichum acutatum as measured by infection incidence (proportion of fruit infected), infection severity (the percentage of the fruit surface supporting sporulation), and sporulation capacity (number of conidia produced on the cut surface of a half berry). Ripe fruit was viii collected from a field planting in Grand Junction, MI, USA in July and August of 2008 and 2009 (n = 10 unless otherwise noted and SE = standard error of the mean). ...................................... 122 Table 5.2. Pearson correlation coefficients (r) and statistical significance (P) for regressions between different measures of anthracnose fruit rot resistance after artificial inoculation of a range of blueberry cultivars with Colletotrichum acutatum. Infection incidence (proportion of fruit infected), infection severity (the percentage of the fruit surface supporting sporulation), sporulation capacity (number of conidia produced on the cut surface of a half berry) and previously published resistance ratings (Polashock et al., 2005) were subjected to linear regression. All values were log transformed prior to regression. R-values in boldface are statistically significant at α < 0.05... 126 Table 5.3. Characteristics of fruit of different blueberry cultivars collected from a field planting in Grand Junction, MI, USA from 2005 to 2008. Values shown are averages and standard errors over 4 years. Five berries were used per replicate. ............................................................................. 128 Table 5.4. Pearson correlation coefficients (r) for regressions of various measures of anthracnose fruit rot resistance after artificial inoculation of blueberry fruit with Colletotrichum acutatum (Table 5.1) against fruit characteristics (Table 5.1). Infection incidence (proportion of fruit infected), infection severity (percentage of the fruit surface supporting sporulation), sporulation capacity (number of conidia produced on the cut surface of a half berry) and previously published resistance ratings (Polashock et al., 2005) were subjected to linear regression against fruit variables. Statistically significant r-values are indicated in boldface. ....................................... 130 Table 6.1. Previous anthracnose fruit rot resistance profiles of blueberry cultivars used in this study using proportion decayed values (24) and sporulation capacity (17). Both studies used artificial inoculation, however, Polashock et al. (2005) inoculated green fruit and rated on disease incidence (proportion decayed) and Miles et al. (2011b) inoculated the cut surface of ripe fruit and rated on the quantity of conidia produced on a cut fruit surface (sporulation capacity).. .......... 145 Table 6.2. Area under the disease progress (AUDPC) values of the incidence of Colletotrichum acutatum on various blueberry cultivars and cross families screened by artificial inoculation in 2010 and 2011 at the Southwest Michigan Research and Extension Center (Benton Harbor, MI). Fruit of 4 to 7-year old bushes were inoculated when immature, harvested when ripe, incubated, and rated at 5, 8 and 12 days. A statistically significant effect of year (P = 0.008) so data from the two years was analyzed separately. ............................................................................................ 149 Table B.1. Characteristics and predicted physiological function of differentially expressed sequence tags (ESTs) from ripe fruit of highbush blueberry cultivar Elliott versus cultivar Jersey after inoculation with Colletotrichum acutatum. The hypothetical function is based on homology to sequences in translated nucleotide databases (DDBJ/EMBL/GenBank) using TBLASTX. Some ESTs are likely of fungal origin based on homology with fungal genes. ................................... 167 Table C.1. Bioactivity of extracts and fractions from the anthracnose-resistant blueberry cultivar Elliott and susceptible cultivar Jersey as measure by their ability to inhibit microconidiation of Colletotrichum acutatum on solid media (+ = activity, - = no activity) .................................... 172 ix Table D.1. Pathogenicity of various isolates of Colletotrichum acutatum on detached fruit of the susceptible blueberry cultivar Jersey. Isolates were obtained from various hosts including highbush blueberry (Vaccinium corymbosum), currant (Ribes spp.), cranberry (Vaccinium macrocarpon), raspberry (Rubus spp.), blackberry (Rubus spp.), strawberry (Fragaria x ananassa), and grape (Vitis spp.). Ripe ‘Jersey’ fruits were harvested from a field planting in Harrietta, MI in August 6 2006, spray-inoculated with 10 conidia per ml, incubated at 100% relative humidity, and rated for incidence of disease 10 days post inoculation (n = 5 and SE = standard error of the mean). .... 176 Table E.1. Anthracnose resistance ratings of detached blueberryfruits from several daughters of a Draper x Jewel cross. Fruits were harvested in Interlachen, FL between 26 April and 12 May in 2011, stored, and shipped overnight to Michigan State University and inoculated with 106 spores/ml by spraying (whole fruit assays) or by a 50 µl droplet on a cut fruit surface (cut fruit assays). Fruits incubated at 100% relative humidity at 22-24°C and rated at 3 days (for cut fruit assays) or 10 days ( for whole fruit assays). For each daughter 5 fruits were screened per replicate (25 in total, and n = 5) per type of inoculation. .......................................................................... 178 Table E.2. Anthracnose resistance ratings of detached blueberryfruits from several daughters of a Draper x Jewel cross. Fruits were harvested in Corvallis, OR between 20 July and 11 August in 2011, stored, and shipped overnight to Michigan State University and inoculated with 106 spores/ml by a 50 µl droplet on a cut fruit surface. Fruits incubated at 100% relative humidity at 22-24°C and rated at 3 days. For each daughter 5 fruits were screened per replicate (25 in total, and n = 5) per type of inoculation. .............................................................................................. 181 x LIST OF FIGURES Figure 1.1. Representative symptoms on ripe fruit of infection by Colletotrichum spp. A) apple (Malus domestica Borkh. - image 1233007) B) blueberry (Vaccinium corymbosum L.), courtesy A. Schilder, C) lychee (Litchi chinensis Sonn. – image 5403527), D) cantaloupe (Cucumis melo L. – image 5405331), E) peach (Prunus persica L. Batsch – image 5407867), F) sweet pepper (Capsicum annuum L. – image 5435606), G) tomato (Lycopersicon esculentum Mill. – image 1236163), and H) strawberry (Fragaria × ananassa Duchesne), courtesy A. Schilder. Numbered images courtesy of IPM images (http://www.ipmimages.org/); photographer information included, Clemson University - USDA Cooperative Extension Slide Series (A and G), Yuan-Min Shen, Taichung District Agricultural Research and Extension Station (C and F), and Paul Bachi, University of Kentucky Research and Education Center (D and E). For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this thesis......................................................................................................................................... 4 Figure 1.2. Conidium germination of Colletotrichum acutatum on parafilm-covered glass slides. A) Ungerminated conidia B) Germinated conidium with germ tube, C) Unmelanized appressorium and D) Melanized appressoria with internal light spots (ILS) (Bars = 10 µm).. ............................ 7 Figure 1.3. A) A close-up of the skin of an infected blueberry reveals small open blisters (acervuli) in which the spores are produced.B) Acervulus bursting through the fruit epidermis (electron micrograph). Conidia can be seen in the matrix (Bar = 50 µm)... ................................... 7 Figure 1.4. Chemical structures of two common flavonoids present in fruit of highbush blueberries. A) Anthocyanin backbone which is typically glycosylated at the R3 position. B) Flavonol backbone. ....................................................................................................................... 10 Figure 1.5. Disease cycle of anthracnose fruit rot caused by Colletotrichum acutatum on blueberries. (Illustration by Jennifer Pagan.). ............................................................................... 13 Figure 2.1. The effect of temperature on mycelial growth of Colletotrichum acutatum (isolate #0001) on quarter-strength potato dextrose agar. Colony diameter was measured in two perpendicular directions and averaged after a 10-day incubation at each temperature. Values are the average of three replications and error bars denote the standard error of the mean. .............. 38 Figure 2.2. The effect of temperature and wetness duration on the predicted development of melanized appressoria of Colletotrichum acutatum on parafilm-covered glass slides. A Gaussian equation (Table 2.2, Equation A) was used to fit the data. Shading intensity of the surface indicates changes in the percentage of melanized appressoria at 20% intervals starting at zero................. 40 Figure 2.3. The effect of temperature and wetness duration on the predicted infection level of immature (A) and mature (B) blueberry fruits by Colletotrichum acutatum. Separate Gaussian equations were fit to the data for immature fruit (Table 2.2, Equation D) and mature fruit (Table 2.2, Equation G). Shading intensity of the surface indicates changes in the percentage of infected fruit at 20% intervals starting at zero... ......................................................................................... 41 xi Figure 2.4. The effect of interrupted wetness periods (0-, 1-, 4-, and 16-h of dry time in the middle of 12-, 24-, and 48-h wet periods) on the development of melanized appressoria (%) of Colletotrichum acutatum (A-C) and the percentage immature (D-F) and mature (G-I) blueberry fruit infected by C. acutatum. Values are the average of four replications and error bars denote the standard error of the mean... ......................................................................................................... 42 Figure 2.5. The effect of wetness duration and interrupted wetness periods on the predicted development of melanized appressoria of Colletotrichum acutatum on parafilm-coated glass slides. Values are calculated as the percent melanized appressoria relative to the continuous wetness treatment. A Planer equation (Table 2.2, Equation B) was fit to the data. Shading intensity of the surface indicates changes with the percentage of melanized appressoria at 20% intervals starting at zero................................................................................................................................................ 44 Figure 2.6. The effect of wetness duration and interrupted wetness periods on the predicted infection level of immature (A) and mature (B) blueberry fruit by Colletotrichum acutatum. Values are calculated as the percent infected fruit relative to the continuous wetness control. Planer equations (Table 2.2, Equation E) were fit to the data for immature fruit and mature fruit (Table 2.2, Equation H). Shading intensity of the surface indicates changes in the percentage of infected fruit at 20% intervals starting at zero. ........................................................................................... 45 Figure 2.7. The effect of temperature and relative humidity on the predicted development of melanized appressoria (%) of Colletotrichum acutatum (A), and the predicted infection level of mature blueberry fruits (B) by C. acutatum. A Gaussian equation (Table 2.2, Equation C) was fit to the data for melanized appressoria and a Lorentzian equation (Table 2.2, Equations F and I) was used to fit the data for mature fruit. Shading intensity of the surface indicates changes in the percentage of melanized appressoria and percentage infected fruit at 20% intervals, respectively, starting at zero.. ............................................................................................................................. 47 Figure 3.1. Symptoms and signs of Colletotrichum acutatum infection of highbush blueberry fruit A) Fruit appearance 8 d after inoculation in the susceptible cultivar Jersey, and B) the resistant cultivar Elliott. Scanning electron micrograph of acervuli on the fruit surface of C) ‘Jersey’ and D) ‘Elliott’. Bar = 0.5 mm.................................................................................................................. 68 Figure 3.2. A) Average diameter of acervuli of C. acutatum on fruit of the susceptible blueberry cultivar ‘Jersey’ (white bars) and resistant cultivar ‘Elliott’ (black bars) (n = 50). B) Quantity of C. acutatum conidia produced on ‘Jersey’ (open bars) and ‘Elliott’ (closed bars) after 8 d of incubation (n = 5). Error bars denote standard error of the mean. Means with the same letter are not significantly different from each other according to Student’s paired t-test (P < 0.05). .............. 69 Figure 3.3. The percentage of C. acutatum conidia that formed melanized appressoria at 24 h post inoculation on blueberry fruit at different stages of development in the susceptible cultivar Jersey (white bars) and resistant cultivar Elliott (black bars). Error bars denote the standard error of the mean (n = 5). Within each development stage, means with the same letter are not significantly different from each other according to Student’s paired t-test (P < 0.05) .................................... 70 xii Figure 3.4. Suppression subtractive hybridization to detect differential expression of genes in ‘Elliott’ (resistant) compared to ‘Jersey’ (susceptible) blueberry fruit after inoculation with C. acutatum. A) Colony PCR products obtained from forward-subtracted ‘Elliott’ cDNA library with T7 and Sp6 primers used for dot blot analysis and visualized through gel electrophoresis. Lanes 1 and 24, 1-kb+ DNA ladder; lane 2, negative control; lane 3, EV=empty vector; lanes 4 through 23 PCR products from ‘Elliott’ forward-subtracted cDNA library. B) PCR products from the forward ‘Elliott’ library hybridized with the reverse ‘Jersey’ library DIG probe. C) PCR products from the forward ‘Elliott’ library hybridized with the forward ‘Elliott’ DIG probe. Dotted lines surround the actin PCR product which served as an internal standard. ............................................................. 75 Figure 3.5. Temporal pattern of transcript accumulation and origin determination of putative genes in blueberry fruit of the resistant cultivar Elliott and the susceptible cultivar Jersey after inoculation with C. acutatum. A) Transcript accumulation of expressed sequence tags (Table 3.1) at 0, 24, 48, 96, and 144 hours post inoculation in semi-quantitative RT-PCR. B) The origin of the ESTs as determined by PCR using genomic DNA from ‘Elliott’, ‘Jersey’ and C. acutatum. In A and B, experiments were performed with specific primers constructed for each individual EST of interest (Table 3.2). Controls were performed with primers specific for actin (Table 3.2) and the fungal ITS region (ITS1F-ITS4).. ................................................................................................. 77 Figure 3.6. Ratio of transcript accumulation of expressed sequence tags (ESTs) relative to actin in quantitative RT-PCR at 0, 24, 48, 96, and 144 hours post inoculation of ripe blueberry fruit of a susceptible cultivar Jersey and a resistant cultivar Elliott with C. acutatum or water (control). Mean expression levels were normalized with mean endogenous actin levels, and an induced sample was used to generate the standard curve plots for each individual EST. ∆Rn thresholds varied between 0.15 and 0.75. Error bars denote standard error of the means of the replicates within a representative experiment (n = 3). .................................................................................. 78 Figure 3.7. H2O2 accumulation in epidermal peels of green and ripe fruit of the susceptible blueberry cultivar Jersey and resistant cultivar Elliott inoculated with C. acutatum or water (control) at 0, 12, 18, 24 and 48 hours post inoculation (n=5). H2O2 was detected by the Amplex® Red Hydrogen Peroxide/Peroxidase Kit. Results from a representative experiment out of three independent experiments are displayed ........................................................................................ 81 Figure 4.1. Anthracnose infection incidence in fruit of the susceptible cultivar Jersey pretreated with methanol-soluble fractions from ‘Jersey’ and ‘Elliott’ followed by inoculation with Colletotrichum acutatum. Bars denote the standard error of the mean (n = 5). From left to right: Negative control = uninoculated fruit with no extract applied, Positive control = inoculated fruit with no extract applied, J.M.S. = ‘Jersey’ methanol-soluble fraction, and E.M.S. = ‘Elliott’ methanol-soluble fraction. ............................................................................................................ 97 Figure 4.2. HPLC chromatograms of anthocyanins from the anthracnose-resistant blueberry cultivar Elliott (A) and susceptible cultivar Jersey (B) using 10-μl injections at λ = 520 nm... ... 99 Figure 4.3. HPLC chromatograms of non-anthocyanin flavonoids from the anthracnose-resistant blueberry cultivar Elliott (A) and susceptible cultivar Jersey (B) using 10-μl injections at λ 255 nm. Arrows denote compounds selected for further analysis. ........................................................... 102 xiii Figure 4.4. MS/MS and UV/Vis spectra of two flavonoid compounds unique to the anthracnoseresistant cultivar Elliott. (A) Daughter ions detected when selecting for the parent ion of quercetin 3-O-rhamnoside (compound 08F in Table 4.2). (B) UV/Vis spectrum of quercetin 3-O-rhamnoside (compound 08F in Table 4.2). (C) Daughter ions detected when selecting for the parent ion of dimethylmyricetin methyl pentoside (compound 09F in Table 4.2). (D) UV/Vis spectrum of dimethylmyricetin methyl pentoside (compound 09F in Table 4.2)........................................... 105 Figure 4.5. Change in the optical density (λ 590 nm) of liquid medium inoculated with conidia of Colletotrichum acutatum over time in the presence of various fractions (50 μg per 100 μl of culture) from ripe fruit of the anthracnose-resistant blueberry cultivar Elliott and susceptible cultivar Jersey. (A) Anthocyanin-containing extracts and fractions from ‘Elliott ‘and ‘Jersey’ as well as the standard cyanidin 3-O-glucoside. (B) Flavonol-containing extracts and fractions from ‘Elliott’ and ‘Jersey’ as well as the standards quercetin 3-O-glucoside and quercetin 3-Orhamnoside. The positive control denotes the fungus alone in the medium. Bars denote the standard error of the mean (n = 3). Note , Postive control = methanol only, Cy-gluc = cyanidin 3glucoside, Q-gluc = quercetin-3-O-glucoside, Q-rham = quercetin3-O-rhamnoside, J.M.E. = ‘Jersey’ methanol extract, E.M.E. = ‘Elliott’ methanol extract, E.A.F. = ‘Elliott’ anthocyanin fraction, J.A.F. = ‘Jersey’ anthocyanin fraction, E.N.A.F. = ‘Elliott’ non-anthocyanin flavonoid fraction, and J.N.A.F = ‘Jersey’ non-anthocyanin flavonoid fraction. ....................................... 107 Figure 5.1. Signs and symptoms of fruit infection by Colletotrichum acutatum on two different blueberry cultivars, Jersey (A to E), and Elliott (F to J) after inoculation with a conidial suspension using different techniques, such as applying a 10-µL droplet into the calyx cup (A and F), spraying the berries until runoff (B and G), injecting 50 µL into the interior of the fruit with a syringe (needle still visible) (C and H), and applying a 50-µL droplet to the open surface of a cut fruit (D and I). Fifty microliters of sterile deionized water served as a control in the cut fruit experiments (E and J). All treatments were incubated for 10 days after inoculation, except the cut-fruit treatments (3 days) ...................................................................................................................... 120 Figure 5.2. Conidium production on fruit of blueberry cultivars Jersey and Elliott after inoculation with a Colletotrichum acutatum conidial suspension using different inoculation methods (applying a 10-µL droplet into the calyx cup, spraying the berries until runoff, injecting 50 µL into the interior of the fruit with a syringe, and applying a 50-µL droplet to the open surface of a cut fruit). All treatments were incubated for 10 days after inoculation, except the cut-fruit treatment (3 days). Bars denote the standard error of the mean (n = 5 with 10 fruits per replicate).. ....................... 121 Figure 5.3. Relationship between the concentration of an aqueous suspension of Colletotrichum acutatum conidia and optical density at 590 nm. A) Conidia produced on PDA. B) Conidia from cut fruit surfaces from a Draper x Jewel F1 population. C) Conidia produced on cut fruit surfaces of selected blueberry cultivars .................................................................................................... 125 Figure 5.4. Mycelial growth of Colletotrichum acutatum on minimal medium with different sugar (50/50 D-glucose/D-fructose) concentrations: 8% (w/v) (A), 12% (w/v) (B), and 16% (w/v) (C) after 7 days. D) The effect of sugar content in minimal medium on mycelial growth of C. acutatum after 7 days using D-glucose, D-fructose and a 50/50 D-glucose/D-fructose mixture. E) The effect xiv of sugar content in minimal medium on mycelial growth of C. acutatum after 7 days using a 50/50 D-glucose/D-fructose mixture in the physiological range for blueberries. In D and E, error bars denote standard error of the mean ............................................................................................... 132 Figure 5.5. A) The effect of initial pH in potato dextrose broth and minimal medium on the amount of mycelial growth (dry weight) of Colletotrichum acutatum after 7 days in 3ml cultures. Initial pH is indicated here since the pH changed as cultures grew. Error bars denote standard error of the mean. B) The effect of initial pH and sugar content (using a 50/50 D-glucose/D-fructose mixture) on mycelial growth (dry weight) of C. acutatum after 7 days. Different degrees of shading indicate different mycelial weights, with darker shades indicating less growth......................... 133 Figure 6.1. The proportion of all blueberry fruit infected over time by Colletotrichum acutatum in 2010 and 2011 at 5, 8 and 12 days post incubation averaged over all of the cross families and cultivars... .................................................................................................................................... 148 Figure 6.2. Distribution of anthracnose fruit rot resistance (expressed as area under the disease progress curves (AUDPC)) in a susceptible x moderate cross (‘Nelson’ x ‘Ozarkblue’) and a resistant x resistant cross (‘Bluegold’ x ‘Elliott’) in 2010 and 2011. A) ‘Nelson’ x ‘Ozarkblue’ in 2010, B) ‘Nelson’ x ‘Ozarkblue’ in 2011, C) ‘Bluegold’ x ‘Elliott’ in 2010, and D) ‘Bluegold’ x ‘Elliott’ in 2011. Discrete levels were defined every 50 AUDPC values.. ................................. 150 Figure 6.3. Correlation between the average anthracnose fruit rot incidence on various blueberry cultivars and cross families expressed as area under the disease progress curves (AUDPC) against actual and predicted proportion decayed values from Polashock et al. (2005) in (A) 2010 and (B) 2011. Abbreviations for parent cultivars: BG = Bluegold; BR = Brigitta; DU = Duke; E = Elliott; LE = Legacy; N = Nelson;O = Ozarkblue. ................................................................................. 152 Figure 6.4. Correlation between the average anthracnose fruit rot incidence on various blueberry cultivars and cross families expressed as area under the disease progress curves (AUDPC) against actual and predicted sporulation capacity values from Miles et al. (2011b) in (A) 2010 and (B) 2011. Abbreviations for parent cultivars: BR = Brigitta; DU = Duke; DR = Draper; E = Elliott; LI = Liberty; N = Nelson.. ............................................................................................................... 153 Figure A.1. The effect of temperature and wetness duration on the development of melanized appressoria of Colletotrichum acutatum on parafilm slides. Discrete transitions indicate levels of melanized appressoria every 20% starting at zero ...................................................................... 162 Figure A.2. The effect of temperature and wetness duration on the infection level of immature (A) and mature (B) blueberry fruits by Colletotrichum acutatum. Discrete transitions indicate levels of infected fruit every 20% starting at zero.. ................................................................................... 163 Figure A.3. The effect of wetness duration and interrupted wetness periods on the development of melanized appressoria of Colletotrichum acutatum on parafilm slides. Values are calculated as the melanized appressoria (%) relative to the no interruption control. Discrete transitions indicate levels of melanized appressoria every 20% starting at zero. ...................................................... 164 xv Figure A.4. The effect of wetness duration and interrupted wetness periods on the infection level of immature (A) and mature (B) blueberry fruits by Colletotrichum acutatum. Values are calculated as the infected fruit (%) relative to the no interruption control. Discrete transitions indicate levels of infected fruit every 20% starting at zero.. ...................................................... 165 Figure A.5. The effect of temperature and relative humidity on the development of melanized appressoria (%) of Colletotrichum acutatum (A), and the infection level of mature blueberry fruits (B) by C. acutatum. Discrete transitions indicate levels of melanized appressoria every 20% starting at zero and in mature fruit levels are indicated every 20% starting at zero. .................. 166 Figure B.1. Optimization of cDNA amplification using various PCR amplification cycles (15-30), for C. acutatum-inoculated fruit of blueberry cultivars Jersey (lanes 2-7) and Elliott (lanes 9-14) (5 µL loaded per well). Lanes 1 and 8, 1-kb+ DNA ladder; lanes 2 and 9, 15 cycles; lanes 3 and 10, 18 cycles; lanes 4 and 11, 21 cycles; lanes 5 and 12, 24 cycles; lanes 6 and 13, 27 cycles; lanes 7 and 14, 30 cycles. Twenty one cycles was found to be optimal for both libraries and those products were used for digestion and adapter ligation... ........................................................................... 169 Figure B.2. Restriction enzyme digestion of pooled cDNA libraries of the blueberry cultivars Jersey and Elliott after inoculation with C. acutatum for the generation of shorter, blunt-end fragments (necessary for adaptor ligation and subtraction). Lanes 1 and 4, 1-kb+ DNA ladder; lanes 2 and 5, samples from ‘Jersey’ and ‘Elliott’ before digestion, respectively; lanes 3 and 6, samples from ‘Jersey’ and ‘Elliott’ after digestion, respectively.. ............................................. 170 Figure B.3. Primary (lanes 2-5) and secondary (lanes 7-10) PCR products of pooled cDNA libraries in subtracted and nonsubtracted samples of blueberry cultivars Elliott and Jersey after inoculation with C. acutatum. These products were used for the construction of DIG probes and for cloning of the subtractive libraries for differential screening. Lanes 1 and 6, 1-kb+ DNA ladder; lanes 2 and 3, ‘Elliott’ and ‘Jersey’ subtracted primary PCR products, respectively; lanes 4 and 5, ‘Elliott’ and ‘Jersey’ nonsubtracted primary PCR products, respectively; lanes 7 and 8, ‘Elliott’ and ‘Jersey’ subtracted secondary PCR products, respectively; lanes 9 and 10, ‘Elliott’ and ‘Jersey’ nonsubtracted secondary PCR products, respectively.. .............................................................. 171 Figure C.1. TLC plate assay showing inhibition of microconidiation of Colletotrichum acutatum and UV/Vis spectra of methanolic extracts from ripe fruit of the anthracnose-resistant blueberry cultivar Elliott and susceptible cultivar Jersey. (A) Cellulose TLC plate after inoculation with conidia of C. acutatum and stained with iodine crystals. (B) UV/Vis spectra of the boxed area in the TLC plate. (Note: 1 = methanolic extract of uninoculated ‘Elliott’ fruit, 2 = methanolic extract of ‘Elliott’ fruit 4 days after inoculation, 3 methanolic extract of uninoculated ‘Jersey’ fruit, and 4 = methanolic extract of ‘Jersey’ fruit 4 days after inoculation.. ................................................. 173 Figure C.2. Schematic presentation of the extraction procedures for biochemical analysis of blueberry fruit using fresh fruit material and the solvents water, methanol, and ethyl acetate.. 174 Figure C.3. Schematic presentation of the extraction procedures for biochemical analysis of blueberry fruit using lyophilized fruit material and the solvents methanol, and ethyl acetate. .. 175 xvi Figure E.1. Relationship between the concentration of an aqueous suspension of Colletotrichum acutatum from anthracnose resistance cut fruit screenings (590 nm) (Table E.1 and Table E.2 values) of detached blueberry fruits from several daughters of a Draper x Jewel cross ............ 183 Figure F.1. External differences in ethylene concentration in ‘Elliott’ and ‘Jersey’ fruits after inoculation of Colletotrichum acutatum or sterile deionized water. Fruits were spray-inoculated 6 with 10 conidia/ml, incubated at 100% relative humidity for 48 hours at 22-24°C, placed in quart sized mason jars, ethylene concentration was measured using then was measured 2 hours later using previously described methods (1). Error bars denote standard error of the mean. ........... 184 Figure F.2. Internal differences in ethylene concentration within ‘Elliott’ and ‘Jersey’ fruits. Fruits were immersed in a saturated solution of KCl, a vaccum was applied and headspace analysis was used to measure the concentration of ethylene in accordance with previous methods (1, 2). Error bars denote standard error of the mean... .................................................................................... 185 Figure F.3. Effect of ethylene on anthracnose symptom development in ‘Elliott’ and ‘Jersey’ fruits. A) Amount of conidia produced per berry in ‘Elliott’ and ‘Jersey’ fruits in each chamber (ethylene or air). B) Levels of ethylene in both chambers throughout a 14 day period. Two 30gallon tanks were used to simulate an environment with 1ppm ethylene or air. Fruits were sprayinoculated with 106 conidia/ml, incubated at 100% relative humidity for 14 days at 22-24°C , ethylene concentration was measured in accordance with previous methods (1). Closed circles represent the ethylene chamber, open circles represent the air chamber. Error bars denote standard error of the mean.. ....................................................................................................................... 186 xvii CHAPTER I: LITERATURE REVIEW INTRODUCTION Colletotrichum species are ubiquitous fungi and have been recovered from almost every plant species. Several members of the genus commonly cause anthracnose fruit rot of fleshy fruits, including C. acutatum J. H. Simmonds, C. coccodes (Wallr.) S. Hughes , and C. gloeosporioides (Penz.) Penz. & Sacc. (Table 1.1, Figure 1.1). In this case, tomatoes and peppers are included here also as they are fruits in a botanical sense. The disease cycle for Colletotrichum spp. that infect fruit varies by host and climate. In temperate regions, the fungus overwinters in infected host tissues. During rainy periods in the spring, conidia are released that infect green fruit. The infections remain latent until the fruit ripens (200). The fungus can also survive and reproduce without causing symptoms during this stage (112). The latent phase of the infection often causes problems in estimating whether the disease is present and when the infection actually takes place. When fruits ripen, the initial symptoms are softening and shriveling on the fruit, followed by the appearance of acervuli bursting through the fruit surface. Secondary infections will also occur on nearby ripe fruit via splash dispersal of conidia produced on the surface of infected fruit. The infection process typically starts when a conidium lands on the host, attaches itself, and begins to germinate. Germination gives rise to a germ tube (Figure 1.2B), followed by the formation of an appressorium (Figure 1.2C), which will eventually become melanized (Figure 1.2D) and develops an internal light spot that corresponds to the penetration pore (54). The melanin inside the appressorium then generates a hypotonic environment and increased turgor pressure allows the fungus to directly penetrate the host epidermis. 1 Table 1.1. Colletotrichum species that cause diseases on fleshy fruits. Colletotrichum spp. C. acutatum J. H. Simmonds Host common name (scientific name) Apple (Malus × domestica Borkh.) Blueberry (Vaccinium corymbosum L.) Cherry (Prunus spp.) Citrus (Citrus spp.) Cranberry (Vaccinium macrocarpon Ait.) Grape (Vitis spp.) Guava (Psidium guajava L.) Mango (Mangifera indica L.) Olive (Olea europaea L.) Peach and nectarine (Prunus persica L.) Pear (Pyrus communis L.) Pepper (Capsicum spp.) Strawberry (Fragaria × ananassa Duch.) Tomato (Lycopersicon esculentum Mill.) C. ananas Garud Pineapple (Ananas comosus L. Merr.) C. capsici (Syd.) E. J. Butler & Bisby Pepper (Capsicum spp.) C. coccodes (Wallr.) S. J. Hughes Pepper (Capsicum spp.) Tomato (Lycopersicon esculentum Mill.) C. fragariae A. N. Brooks Strawberry (Fragaria × ananassa Duch.) C. gloeosporioides (Penz.) Penz. & Sacc. Apple (Malus × domestica Borkh.) Avocado (Persea Americana Mill.) Blueberry (Vaccinium corymbosum L.) Caneberry (Rubus spp.) Cherry (Prunus spp.) Citrus (Citrus spp.) Coffee (Coffea spp.) Cranberry (Vaccinium macrocarpon Ait.) Guava (Psidium guajava L.) Grape (Vitis spp.) Mango (Mangifera indica L.) Olive (Olea europaea L.) Papaya (Carica papaya L.) Passion fruit (Passiflora edulis Sims.) Peach and nectarine (Prunus persica L.) Pear (Pyrus communis L.) Pepper (Capsicum spp.) Pineapple (Ananas comosus L. Merr.) Strawberry (Fragaria × ananassa Duch.) Tomato (Lycopersicon esculentum Mill.) 2 Table 1.1. (cont’d.) C. kahawae Waller & Bridge Coffee (Coffea spp.) C. musae (Berk. & M. A. Curtis) Arx Banana and plantain (Musa spp.) C. orbiculare (Berk. & Mont.) Arx Cucurbits (Citrullus spp., Cucumis spp., Cucurbita spp., and others) Information from the USDA Systematic Mycology and Microbiology Laboratory Fungus-Host Distribution Database (http://nt.ars-grin.gov/fungaldatabases/fungushost/fungushost.cfm) 3 A C B E D F G H Figure 1.1. 4 Figure 1.1 (cont’d.). Representative symptoms on ripe fruit of infection by Colletotrichum spp. A) apple (Malus domestica Borkh. - image 1233007) B) blueberry (Vaccinium corymbosum L.), courtesy A. Schilder, C) lychee (Litchi chinensis Sonn. – image 5403527), D) cantaloupe (Cucumis melo L. – image 5405331), E) peach (Prunus persica L. Batsch – image 5407867), F) sweet pepper (Capsicum annuum L. – image 5435606), G) tomato (Lycopersicon esculentum Mill. – image 1236163), and H) strawberry (Fragaria × ananassa Duchesne), courtesy A. Schilder. Numbered images courtesy of IPM images (http://www.ipmimages.org/); photographer information included, Clemson University - USDA Cooperative Extension Slide Series (A and G), Yuan-Min Shen, Taichung District Agricultural Research and Extension Station (C and F), and Paul Bachi, University of Kentucky Research and Education Center (D and E). For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this thesis. 5 Following direct penetration of host tissues, Colletotrichum species generally have one of two different host colonization strategies, depending on the tissue or host being colonized: intracellular hemibiotrophy and subcuticular intramural necrotrophy (139). Intracellular hemibiotrophy is the direct invasion of the initial host cell by a primary infection vesicle, followed by the proliferation of thick primary hyphae and thin secondary hyphae; this strategy is common in many non-fruit hosts of Colletotrichum spp., including bean (137), pea (138), sorghum (199), and tobacco (173). In subcuticular intramural necrotrophy, Colletotrichum spp. will invade superficially under the cuticle, generally producing thinner necrotrophic hyphae sooner, which will not invade the host cell intracellularly. This type of strategy is known to exclusively occur in infection by C. capsici (Schwein.) Andrus & W.D. Moore on the leaves of cowpea (151) and cotton (162). Following colonization, acervuli will begin to develop and eventually burst through the host epidermis (Figure 1.3). In a number of pathosystems, correlations have been found between the presence of pectinolytic enzymes, disease symptoms, and pathogen virulence (61). In the C. lindemuthianum (Sacc. & Magn.) Scribn.-bean pathosystem, the pathogen uses a number of pectin-degrading enzymes such as pectin lyase (a.k.a. pectin methyltranseliminase). It is the only known pectinase capable of degrading highly esterified pectins into small molecules via β-elimination without producing toxic methanol by enzymatically cleaving 1-4-alpha-D-galacturonan methyl esters (205). Pectin lyases are usually characterized as being alkaline or acidic. In most Colletotrichumplant interactions following pectin lyase secretion, a shift to a more alkaline pH is usually detected in host tissues (204). Other pectin-degrading enzymes that have been investigated in the C. lindemuthianum-bean pathosystem include endopolygalacturonase (endoPG) which 6 A B C D Figure 1.2. Conidium germination of Colletotrichum acutatum on parafilm-covered glass slides. A) Ungerminated conidia B) Germinated conidium with germ tube, C) Unmelanized appressorium and D) Melanized appressoria with internal light spots (ILS) (Bars = 10 µm). A B Figure 1.3. A) A close-up of the skin of an infected blueberry reveals small open blisters (acervuli) in which the spores are produced.B) Acervulus bursting through the fruit epidermis (electron micrograph). Conidia can be seen in the matrix (Bar = 50 µm). 7 enzymatically hydrolyzes 1,4-alpha-D-galactosiduronic linkages, and the resulting oligogalacturonides have been shown to elicit the production of pathogensis-related (PR) related proteins (32, 110). A variety of defense responses against infection by Colletotrichum species are employed by plant hosts during the infection process, including host cell wall modifications, a hypersensitive response and the production of phytoalexins, pathogenesis-related proteins, and reactive oxygen species. Cell wall modifications and the hypersensitive response constitute the initial physical line of defense against Colletotrichum spp. and make it difficult for the pathogen to colonize plant cells. The fortification of the plant cell walls, reduces the available nutrients and their production can be utilized as a signal to neighboring plant cells (81, 135, 186). A hypersensitive response in tangerines has been associated with a reduced risk of anthracnose infection (24). These researchers observed host cell death and a collapse of the cytoplasm against the cell wall distal from the area of fungal penetration. In some cases, host antimicrobial compounds (phytoalexins) against Colletotrichum spp. are induced upon inoculation and are antimicrobial by preventing pathogen growth. Their chemical structure varies significantly such as 3-deoxyanthocyanidin in sorghum (Sorghum bicolor (L.) Moench) (117), methylhipariochromene A in mistflower (Eupatorium riparium Regel) (77) and lα- and lβ-hydroxydihydrocomin aglycones in Alibertia macrophylla K. Schum (77) . Peroxidases (82), β-1, 3 glucanases (92, 110), hydroxyproline-rich glycoproteins (186) and chitinases (45) are also thought to play a role in the host resistance response because a number of Colletotrichum-resistant cultivars boast increased levels of these compounds. The goal of this review is to discuss blueberry production in Michigan, information about anthracnose fruit rot of blueberries, 8 the environmental requirements for infection, different infection strategies, passive and active host defenses and the heritability of resistance in Colletotrichum fruit pathosystems. BLUEBERRY PRODUCTION IN MICHIGAN The United States is the largest blueberry producing country in the world, with a production of 203,663 metric tons per year. Michigan is the leading state producing blueberries in the US, with 44,906 metric tons of fruit representing 22% of US production and worth $102 million U.S. dollars in 2009 (106). Maine is the second largest producer (20% of total production), however, this consists predominantly of wild lowbush blueberries (V. angustifolium Aiton), unlike other states which raise cultivated highbush (V. corymbosum L.) or rabbiteye (V. ashei J. M. Reade) blueberries. In Michigan, the majority of blueberry fields (75%) are irrigated and the most common methods are overhead sprinkler (57% of total irrigated hectarage) and drip (trickle) irrigation (21% of total irrigated hectarage) (105). In 2007, there were 7,810 harvested blueberry hectares in Michigan and planted primarily (78%) to three cultivars Bluecrop, Elliott and Jersey (105). Demand for blueberries and other fruit crops continues to rise as the health benefits are becoming increasingly apparent due to compounds with antioxidant (96, 116, 147) and antiinflammatory properties (94) that can reduce cardiovascular and neurodegenerative diseases in humans (174). In fact, anthocyanins, the predominant class of flavonoids in blueberries, have been shown to aid in obesity prevention (91), cardiovascular health (48), act as antiinflammatory compounds (196) and have anti-cancer effects (18, 172, 179) . Another common flavonoid present in highbush blueberries are flavonol glycosides such as quercetin and myricetin are also common in highbush blueberries (Figure 1.4.) 9 A B Figure 1.4. Chemical structures of two common flavonoids present in fruit of highbush blueberries. A) Anthocyanin backbone which is typically glycosylated at the R3 position. B) Flavonol backbone. 10 ANTHRACNOSE FRUIT ROT OF BLUEBERRY Anthracnose fruit rot or ripe rot is the most common and widespread fruit disease of blueberries in Michigan and the United States. In northern growing regions, this disease is typically caused by Colletotrichum acutatum, whereas, C. gloeosporioides may be more common in the southern regions of the United States (40). Anthracnose fruit rot can lead to substantial fruit losses, including reduced yield, shelf life and fruit quality. Pre-harvest crop losses may reach 10-20%, and in severe cases, post-harvest losses up to 100% (128). In addition, high levels of anthracnose fruit rot may contribute to unacceptable mold counts in frozen processed fruit. Warm, wet seasons are particularly conducive to disease development (169). Symptoms Infected immature green berries are symptomless. Fruit rot symptoms usually do not become apparent until the fruit ripens, hence the name “ripe rot” (128). Initial symptoms are softening and shriveling of the fruit, particularly on the “king” berry which is the first to ripen, followed by the appearance of small orange dots (spore masses) on the surface (Figure 1.1A). Spores (conidia) are produced in fruiting structures (acervuli), which break through the fruit epidermis (Figure 1.1B). Depending on relative humidity, spore masses may look wet or dry. Infected berries eventually shrivel up and fall off the bush. Even berries that look perfectly healthy on the bush can rot soon after harvest. However, refrigeration delays rot development. Colletotrichum acutatum may also cause blighting of blossoms and twigs but these symptoms are difficult to distinguish from Phomopsis twig blight or Botrytis blossom blight 11 without doing fungal isolations from the infected tissues. Cane cankers, and occasionally leaf spots caused by C. acutatum, have been observed during particularly rainy seasons (123). Disease cycle In spring and summer, conidia are produced on dead twigs, spent fruit trusses and dormant buds (49) which are dispersed by rain splash (85, 146). Although fruit infections are only obvious at maturity, fruit may be infected anytime between flowering and harvest. Reportedly, the fungus requires at least 12 hours of continuous wetness to penetrate the fruit surface and initiate an infection. The optimum temperature for fruit infection is 25ºC (74) but infections can take place at temperatures as low as 10ºC (193). After the fungus enters the developing fruit using a penetration peg produced from a melanized appressorium, it remains quiescent until the berry starts to ripen (Figure 1.2). At this point, the fungus begins to grow and produce enzymes (such as pectinases) that destroy host tissue, resulting in collapse and shriveling of the fruit (46). Within days, spore masses develop and eventually cover most of the berry. The time from infection of ripe fruit to the first appearance of symptoms is about 4 days (201). Secondary infections occur by direct contact or when spores are splashed from infected to nearby healthy berries. Contaminated surfaces of harvesting or sorting equipment can also be a source of infection. The fungus survives the winter in old fruit trusses and live buds that are infected in late summer or early fall (Figure 1.5) (49). Management Anthracnose is very common in blueberries, except in dry regions (146). Since fruit rot symptoms do not become apparent until close to harvest, preventive control strategies are 12 LATE SPRING Latent infections of developing fruit between flowering and harvest Rain splash dispersal of conidia EARLY SPRING EARLY SUMMER Sporulation on infected bud scales and twigs Rain splash dispersal of conidia Overwintering infections in buds and former fruit trusses Conidia produced in an acervulus Spore germination and appressorium formation MID-LATE SUMMER Ripe rot of infected berries WINTER Figure 1.5. Disease cycle of anthracnose fruit rot caused by Colletotrichum acutatum on blueberries. (Illustration by Jennifer Pagan.). 13 important, especially in fields with a history of the disease. Timely harvesting, processing, and cooling reduce pre- and post-harvest losses (84). Pruning out dead twigs and old fruit trusses can reduce the incidence of fruit infection. However, the high cost associated with this type of detailed pruning may make it uneconomical. Another option is to plant resistant cultivars or less susceptible cultivars (150). A preventive fungicide spray program is most effective when focused on the period from pink bud to harvest (169). ENVIRONMENTAL REQUIREMENTS FOR INFECTION Anthracnose fruit rot in many crops is often difficult to control and disease prediction models have been developed to predict melanized appressorium development and infection. Many disease prediction models use regression equations, such as those based on polynomials (65, 170), logistic equations (26, 170) and complex three-dimensional response surfaces (21, 29, 207). Important environmental variables to consider include temperature, wetness duration, interrupted wetness periods and relative humidity. Temperature and wetness duration have been shown to have a direct effect on the development of melanized appressoria and fruit infection. In the C. coccodes-tomato pathosystem; melanized appressoria can form on foliage and fruit as early as 6 hours after incubation at optimal temperatures (16°C to 28°C) (27, 167). In strawberries, C. acutatum infection was investigated on both immature and mature fruit (206). In that study, disease incidence was generally higher in mature fruit and always increased with longer wetness durations (up to 48 hours) except at extreme temperatures (≤ 10°C and ≥ 30°C). Wilson et al. (1990) used a regression model with the logit of disease incidence of C. acutatum as the dependent variable which accurately described infection incidence as a function of wetness 14 duration and temperature. Temperature has also been identified as having a strong effect on the sporulation of C. acutatum on strawberry fruit (optimum temperatures of 22°C to 26°C) (104). Researchers have also looked at the role of interruptions in the wetness duration and the role of relative humidity on the development of melanized appressoria and infection in Colletotrichum pathosystems. In Stylosanthes scabra Vogel, infection by C. gloeosporioides was relatively unaffected by brief interruptions (2 to 4 hours) in wetness periods when they occurred after the initial 12 hours of wetness and were followed by another 24 hours of continuous wetness (33). However, they did not investigate longer periods of dryness or at different stages of the infection process. Relative humidity has also been investigated in relationship to the development of melanized appressoria of C. gloeosporioides on 1-month-old mango leaves (56). These researchers found that conidia were able to germinate and form appressoria at relative humidity values between 95 and 100% even though free water was only visible at 100%. In blueberries, a previous study showed that temperatures of ≤ 10°C and ≥ 30°C inhibited mycelial growth of C. gloeosporioides on potato dextrose agar. Based on in vitro spore germination studies, the authors suggested that 12 hours of continuous wetness at 15-27°C are required for conidial infection (85). In another study in British Columbia, the optimum temperature for anthracnose fruit rot infection was 20°C but infection was possible at temperatures between 7°C and 30°C (192). These researchers then conducted further studies on environmental requirements by placing potted plants in a commercial field at weekly intervals with weather monitoring. Fruit was allowed to ripen naturally and evaluated for anthracnose infection. When comparisons were made between peak infection periods and weather data, 15 regression values were not significant but infection occurred after a minimum of 10 hours of wetness at 11°C, 15°C, and 14°C in 2001, 2002, and 2003, respectively (193). THE INFECTION PROCESS Host tissue and its relationship to infection strategy Multiple studies have investigated the infection strategies of Colletotrichum species on a variety of fruits, including almond (53), avocado (41), blueberry (201), olive (75), and strawberry (43). As previously stated, Colletotrichum species generally have two types of colonization strategies, intercellular hemibiotrophy and subcuticular intramural necrotrophy. Intracellular hemibiotrophy requires a close cytoplasmic interaction between the host and the pathogen. Pathogens with this strategy are considered specialists, while subcuticular intramural necrotrophy does not require this closeness, therefore, these pathogens could be considered generalists (146). C. acutatum is known to use both infection strategies on almond (53), blueberries (201), olive (75) and strawberry (43). In blueberries, the type of infection strategy was associated with a susceptible or a resistant interaction (201). However, while both these strategies were observed in olives, it was not related to host plant resistance (75). Different infection strategies in strawberry appear to be associated with tissue type: intercellular hemibiotrophy occurs on leaves (9) and subcuticular intramural necrotrophy occurs on petioles, stolons and leaves (9, 43). In blueberries, a lower rate of conidium germination and appressorium formation was observed on a resistant cultivar compared to a susceptible cultivar (201). A possible reason for this phenomenon could be differences in the structure or composition of the waxy cuticle in susceptible fruit that stimulates conidium germination and appressorium formation. In avocados, 16 cuticular wax has been shown to trigger conidium germination and appressorium formation of C. gloeosporioides. However, waxes from non-host plants strongly inhibited appressorium formation (148). pH modulation by the pathogen Studies have shown that ammonia secretion by Colletotrichum spp. increases the pH of host tissues, and tissues that are originally unsuitable for growth become inhabitable. The modulation of fruit pH is important for pathogenicity on almond (55), apple (157), avocado (157) and tomato (3, 157). Additional research suggests that tissue alkalization and the presence of nitrogen are important factors for C. gloeosporioides attack and development in a fruit host and these factors appear to be independent (59). The process of pH modulation involves ammonia accumulation at the site of infection, which is influenced by the initial environmental pH (3). In controlled experiments using a lowpH buffer at the infection site, the authors have observed high levels of ammonia secretion and increased virulence of C. coccodes as compared to similar treatments using a neutral-pH buffer. Furthermore, researchers have identified a set of genes involved in ammonia accumulation and regulation by C. gloeosporioides. These genes include a NAD+-specific glutamate dehydrogenase (GDH2) and the ammonia exporter AMET. Genes involved in ammonia uptake (MEP) import glutamate (GLT) and glutamine synthase (GS1). The expression of these genes was dependent on ambient pH and involved in the metabolism and catabolism of ammonia (129). In almonds, alkalinization of host tissue has been visualized using pH-indicating fluorescent dyes and occurred between 24 and 72 hours post inoculation (55). 17 Cell wall-degrading enzymes In beans, necrotrophy was associated with enhanced excretion of plant cell walldegrading enzymes such as endo-polygalacturonases by C. lindemuthianum (204, 205). Pectin lyase activity correlates well with the onset of necrotrophy and the subsequent development of lesions. Pectin lyases have been identified as the main pathogenicity factor in a number of Colletotrichum species including C. acutatum (214), C. coccodes (15) and C. gloeosporioides (19, 198). As avocado fruits mature, the pectin lyase activity of C. gloeosporioides increases (159). This increase has been linked to the increase in ambient pH associated with fruit ripening (209) and a slow decrease in the flavonoid epicatechin which acts as an inhibitor of pectin lyase (159). Laccases (p-diphenol:dioxygen oxido-reductases) produced by C. gloeosporioides have also been shown to degrade this inhibitor and increase pathogenicity on avocado fruits (79). PASSIVE HOST DEFENSES Fruit physiological factors The infection process of Colletotrichum spp. on fruits has been studied in a number of plant pathosystems and in general, as fruits start to ripen they become increasingly susceptible to infection (36, 131, 159, 206). During fruit ripening, many physiological changes occur, such as a reduction in fruit firmness, changes in pH and cell wall composition and an increase in soluble sugars and secondary metabolites, such as anthocyanins (20, 164). In avocado, several factors have been associated with increased fruit susceptibility to infection by Colletotrichum gloeosporioides as fruit ripens, including an increase in fruit pH (159), a decrease in preformed antimicrobial compounds (155) and pathogenicity factor inhibitors, such as epicatechin (79). 18 Soluble sugars may also play a role in defense responses during ripening. Guava cultivars that contained high levels of soluble sugars and ascorbic acid also were the most resistant to Glomerella cingulata (Stoneman) Spauld. & H. Schrenk (176). In grapes, the accumulation of antifungal proteins and sugars during fruit ripening is an important defense mechanism against the fungal pathogens Botrytis cinerea Pers.:Fr. and Guignardia bidwellii (Ellis) Viala & Ravaz (165, 185). Antimicrobial fruit volatiles have been investigated in relationship to anthracnose resistance. In strawberries, the effects of aldehydes, alcohols and esters on mycelial growth were investigated and (E)-Hex-2-enal was identified as the most biologically active. This compound altered the structure of the conidial cell wall and plasma membrane, causing disorganization and lysis of organelles, eventually, cell death (8). In blueberries, many of these compounds were also identified but the quantity of the volatile compounds (especially (E)-Hex-2-enal) was not correlated with anthracnose fruit rot resistance in the various blueberry cultivars (149, 150). The role of plant hormones such as ethylene also has been investigated in relationship to anthracnose resistance. In tobacco leaves, ethylene production increased between 24 and 48 hours after infection of C. destructivum O'Gara (34). In bananas, it has been shown that rot develops both in ethylene-inhibited and ripe banana fruits. Additionally, the researchers also showed that ethylene is not directly involved in the host–pathogen relationship but plays a role in infection in storage by ripening fruits prematurely (35, 36). Pre-formed antimicrobial compounds Several pre-formed compounds have been identified in resistant hosts in response to Colletotrichum infection. Resistance to C. gloeosporioides in unripe avocado fruit is correlated 19 with the presence of fungitoxic concentrations of the preformed antifungal compound 1-acetoxy2-hydroxy-4-oxoheneicosa-12,15-diene in the pericarp of unripe fruits (155). A second antifungal compound was subsequently purified from unripe avocado fruit and identified as 1acetoxy-2,4-dihydroxy-n-heptadeca-16-ene (158). When fruits ripen, the activity of the plantderived enzyme lipoxygenase increases causing the degradation of these preformed antifungal compounds and fruit gradually becomes more susceptible to infection (159). Interestingly, the lipoxygenase activity in avocado fruits is affected by the flavonoid epicatechin, which acts as a natural inhibitor (154). In green fruit, the concentration of this compound gradually decreases upon ripening until the fruit becomes completely susceptible. As previously mentioned, this compound is also capable of inhibiting the activity of pectin lyase produced by fungal pathogens. In unripe mangos, resistance to C. gloeosporioides has been associated with a mixture of antifungal compounds consisting of 5-12-cis-heptadecenyl resorcinol and 5-pentadecenyl resorcinol (58). These resorcinols occur at fungitoxic levels and they decrease to nontoxic levels at the same time that decay appears. Additionally, the concentration of the resorcinols decreases faster during ripening in disease-susceptible cultivars than in resistant cultivars. There is also evidence that mature fruit of green mango varieties with higher concentrations of these resorcinols are more resistant to C. gloeosporioides (86). Host resistance in other Colletotrichum fruit pathosystems is less well studied. In bananas, resistance has been attributed to dopamine and its oxidation products, which was isolated from the peel of unripe banana in concentrations that inhibited C. musae (Berk. & M.A. Curtis) Arx in vitro; it was therefore presumed to be a possible preformed antifungal compound. However, the concentration of the compound was not synchronized with changes in decay development (132). In blueberries, several studies have been carried out into the antifungal 20 properties of extracts from ripe blueberry fruit from wild highbush blueberry plants as they relate to fruit decay and herbivore preference (37-39). These studies indicated that the main antifungal compounds present in ripe blueberry fruit were water-soluble phenolics and acids. They also proposed that resistance in ripe blueberries may be due to an interaction between simple phenolic compounds and organic acids and not necessarily individual fungitoxic compounds. ACTIVE HOST DEFENSES Cell wall modifications Cell wall modifications in relationship to host defenses of fruit have been observed in various Colletotrichum-host plant interactions. Several studies have microscopically examined the infection process on attached and detached fruit such as almonds (53), avocados (41), blueberries (46, 201), citrus fruits (22), grapes (47), olives (75) and strawberries (43). In papaya fruit, resistance has been attributed to callose deposited beneath the fungal penetration peg (178). Resistant chili pepper fruit tissues showed evidence of structural modifications and increases in the thickness of the cuticle layer compared with those of susceptible and noninoculated fruit (101). Additionally, resistant cultivars had evidence of programmed cell death. Cell wall modifications have a direct benefit to the host by preventing pathogen ingress, and there is mounting evidence that cell wall fragments, particularly polysaccharides, can act as signaling molecules in plant-pathogen interactions (194). Induced antimicrobial compounds Limited research is available on inducible antimicrobial compounds in Colletotrichumfruit pathosystems. However, in avocados, the activity of the pre-formed antimicrobial diene 21 compound can be modulated by inducing its production and/or preventing its decrease during ripening (159). Research on C. sublineolum Henn. in sorghum has associated host plant resistance with 3-deoxyanthocyanidins, including luteolinidin, apigeninidin, and a caffeic acid ester of arabinosyl-5-O-apigeninidin (117). The researchers also found that these phytoalexins were concentrated in cells immediately below the infection site and that the majority of the antimicrobial compounds were highly localized (117). Other research on C. lindemuthianum on bean (Phaseolus vulgaris) has linked the phytoalexin phaseollin (a prenylated pterocarpan ) to resistance (12). Defense-related proteins A variety of defense mechanisms in other Colletotrichum-plant interactions have been observed, including the production of host-derived cell wall-degrading enzymes including chitinases and β-1-3-glucanases (31, 76, 110, 204). When the genes coding for these products are overexpressed in plant tissue, the result is often increased disease resistance. For example, in transgenic tobacco plants, the over expression of a chitinase gene led to broad resistance against the fungal pathogen Rhizoctonia solani J.G. Kühn, and the bacterial pathogen Pseudomonas syringae pv. tabaci (44). Upon infection of pepper fruits (Capsicum annuum L.), by C. gloeosporioides, defense-related proteins like cytochrome P450 (141), defensin, thionin-like protein (140), thaumatin-like protein (103) and esterase (107) are induced in incompatible interactions. Research has shown that C. acutatum infection is reduced significantly by a polygalacturonase inhibitor protein extracted from the fruit (78). This type of host plant resistance to fungi is found in apple (211), raspberry (93), and tomato (180) fruits. These 22 inhibitor proteins are predominantly expressed in the epidermal layers of the fruit and regulated in response to infection and wounding (211). A novel protein has been identified in pumpkin rinds is capable of inhibiting the growth of C. coccodes in vitro at 10-20 μM. This protein was found to be nonphytotoxic and heat-stable, and is proposed to be a possible natural antifungal agent (145). Reactive oxygen species Reactive oxygen species such as H2O2 are important in the resistance response in the C. coccodes-tomato fruit interaction. Expression appears to occur around 24 to 48 hours after inoculation, corresponding with melanized appressorium formation and attempted fungal penetration (122). In strawberries, H2O2 generation plays a role in restricting fungal penetration and inhibiting fungal invasion, leading to the hypersensitive response and triggering rapid necrosis at infection sites, or activating defense-related genes (25). GENETICS OF RESISTANCE In recent years there have been many successes in breeding resistant crop varieties against Colletotrichum species. In bean, resistance to anthracnose is primarily controlled by nine major independent genes. Most of these genes are dominant and behave in a Mendelian fashion and most likely exist in clusters (99). In other Colletotrichum-host pathosystems, host plant resistance has been shown to be controlled by a single gene, several genetic loci or a combination. In the C. acutatum-strawberry pathosystem, a single dominant gene (Rca2) has been shown to be responsible for high-level resistance, and minor resistance appeared to be quantitative and controlled by a number of minor 23 genes (50). In the C. acutatum-chili pepper pathosystem, resistance was mapped to a single recessive gene at the mature green fruit stage and a single dominant gene at the ripe fruit stage (119). In another species of pepper, resistance to C. capsici has been associated with a single recessive gene (144). Mapping of host plant resistance in Phaseolus vulgaris L. has revealed that two independent resistance genes within the same cluster confer resistance to different strains of C. lindemuthianum (72). Furthermore, resistance to C. higginsianum Sacc. has been shown to be localized to a single genetic locus RCH1 in the Arabidopsis ecotype Eil-0 (134). Quantitative resistance to plant pathogens is typically controlled by multiple genetic loci of small effect and is common in many Colletotrichum-host pathosystems (30, 50, 73, 89). While this phenomenon has been documented in many plant pathosystems it is poorly understood. In strawberries, different modes of inheritance of resistance to anthracnose have been suggested based on pathogenicity groups of Colletotrichum (51). Resistance to pathogenicity group 1 is quantitative. However, a single dominant gene, Rca2, controls resistance to pathogenicity group 2, although minor genes may also contribute to resistance in several cultivars (50). In previous studies, anthracnose fruit rot resistance in blueberries was not correlated with resistance to foliar infection (64) or the production of antimicrobial fruit volatiles (149). However, resistance has been associated with different infection strategies by C. acutatum (201).While there is information on relative anthracnose fruit rot resistance of many cultivars based on our own studies and those of Polashock et al. (2005), the information about inheritance of resistance in current breeding programs is largely unknown. Additionally, heritability of disease resistance is of particular interest in cultivated highbush blueberries due to decreasing heterozygosity because of increased intraspecific hybridization and the fact that the majority of genetic diversity is derived from only four wild selections (83). The effect of limited genetic 24 diversity on disease resistance and whether that resistance is dominant, recessive or quantitative is important in order to understand how resistance to C. acutatum is inherited. CONCLUSIONS The infection process of Colletotrichum spp. on fruit is complex and environmental factors play a significant role. The infection strategies that Colletotrichum spp. utilize on various hosts depends greatly on the tissue type and how specialized a pathogen is to a particular host. Host resistance against Colletotrichum spp. involves multiple mechanisms. Fruit physiological factors like sugar content and surface waxes can affect the growth of the fungus. Preformed antimicrobial compounds like antifungal dienes, resorcinols or flavonoids may be extremely important in unripe and ripe fruit resistance. Active defenses including cell wall modifications, induced antimicrobial compounds, oxidative bursts and defense-related proteins may also be important in the resistant response. Further study of how resistance in inherited might lead to the development of new anthracnose-resistant cultivars and allow us to better understand the basis of resistance on a genetic level. The objectives of this thesis are to: 1) investigate the minimum environmental factors (temperature, humidity, and wetness) required for infection in blueberries.2) identify differentially expressed genes of resistant blueberry fruits and the timing of their expression following pathogen inoculation, 3) investigate the role of secondary metabolites in the resistant interaction, 4) evaluate various inoculation techniques in several blueberry cultivars and identify significant correlations between fruit characteristics and resistance, and 5) investigate how resistance to blueberry anthracnose is inherited. 25 CHAPTER II: THE EFFECT OF ENVIRONMENTAL FACTORS ON INFECTION OF BLUEBERRY FRUIT BY COLLETOTRICHUM ACUTATUM ABSTRACT Anthracnose fruit rot of blueberries caused by Colletotrichum acutatum is a serious problem in humid blueberry-growing regions of North America. In order to develop a disease prediction model, the environmental factors that affect mycelial growth, conidial germination appressorium formation, and fruit infection by C. acutatum were investigated. Variables included temperature, wetness duration, wetness interruption, and relative humidity. The optimal temperature for mycelial growth was 26°C and little or no growth was observed at 5 and 35°C. The development of melanized appressoria was studied on parafilm-covered glass slides and infection was evaluated in immature and mature blueberry fruits. In all three assays, the optimal temperature for infection was identified as 25°C and infections increased up to a wetness duration of 48 hours. Three-dimensional Gaussian equations were used to assess to effect of 2 temperature and wetness duration on the development of melanized appressoria (R = 0.89) and 2 2 infection in immature (R = 0.86) and mature (R = 0.90) blueberry fruits. Interrupted wetness periods of different durations were investigated and models were fitted to the response of 2 2 2 melanized appressoria (R = 0.95) and infection in immature (R = 0.91) and mature (R = 0.85) blueberry fruits. Additionally, the development of melanized appressoria and fruit infection was 2 2 modeled in relationship to relative humidity (R = 0.99 and R = 0.97, respectively). Three summarized equations were then developed that incorporate the aforementioned variables. Our 26 results lay the ground work for the development of a disease prediction model for anthracnose fruit rot in blueberries. INTRODUCTION Blueberries are a popular fruit with nutraceutical benefits (48). The United States is the leading producer of blueberries in the world and Michigan is the number one blueberryproducing state. Anthracnose fruit rot, caused by Colletotrichum acutatum J. H. Simmonds, is a serious problem in most blueberry-growing regions, with reported annual yield losses as high as 10 to 20% (128). Various protectant and systemic fungicides are currently available that are effective against this disease (169). Sprays are usually initiated at pink bud or bloom and continued at 10 to 14 day intervals through fruit development and ripening (123). C. acutatum overwinters in infected twigs, fruit trusses and buds from which conidia are rain-splash dispersed during the growing season. Trapping of C. acutatum conidia in Michigan over several years has shown one or two peaks in conidial dispersal, one around bloom and the other around fruit ripening (202). Other research in Michigan has documented up to three peaks of C. gloeosporioides spore dispersal: at bloom, green fruit, and ripe fruit with spores present throughout the season (85). In that study, C. acutatum may have been misidentified as C. gloeosporioides since subsequent sampling in Michigan blueberry fields has confirmed C. acutatum to be the predominant pathogen (169). In addition to spore dispersal by rain, other environmental variables such as temperature, wetness duration and relative humidity are thought to play a role in the infection process. In other pathosystems, temperature and wetness duration had a direct effect on the development of melanized appressoria and fruit infection. In the C. coccodes-tomato 27 pathosystem, melanized appressoria can form on foliage and fruit in as early as 6 hours at optimal temperatures (16°C to 28°C) (27, 167). In strawberries, C. acutatum infection was investigated in both immature and mature fruit (206). In that study, disease incidence was generally higher in mature fruit than in immature fruit and increased with longer wetness durations (up to 48 hours) except at temperatures of ≤ 10°C and ≥ 30°C. The optimum temperature for infection was between 25 and 30°C. Similar temperatures also had a comparable effect on the sporulation of C. acutatum on strawberry fruit (104). Researchers have also looked at interrupted wetting periods and the role of relative humidity on the development of melanized appressoria and infection in Colletotrichum pathosystems. In Stylosanthes scabra Vogel, infection by C. gloeosporioides was relatively unaffected by brief interruptions (2 to 4 hours) in wetness when they occurred after the initial 12 hours and were followed by another 24 hours of continuous wetness (33). However, they did not investigate longer periods of dryness or at different stages of the infection process. The effect of relative humidity (RH) has also been investigated on the development of melanized appressoria of C. gloeosporioides on 1-month old mango leaves; conidia were able to germinate and form appressoria at RH values between 95 and 100% even though free water was only visible at 100% (56). In blueberries, a previous study indicated that temperatures of ≤ 5°C and ≥ 30°C inhibited mycelial growth of C. gloeosporioides on potato dextrose agar (PDA). Based on in vitro conidium germination studies, the authors suggested that 12 hours of continuous wetness at 15-27°C would be required for fruit infection (85). Another study indicated that the optimum temperature for anthracnose fruit rot infection in British Columbia was 20°C but infection was 28 possible at temperatures between 7°C and 30°C (192). These researchers then conducted further studies on temperature requirements by placing potted plants in a commercial field at weekly intervals with weather monitoring. Fruit was allowed to ripen naturally and evaluated for anthracnose infection. When comparisons were made between peak infection periods and weather data, regression values were not significant but infection occurred after a minimum of 10 hours of wetness at 11°C, 15°C, and 14°C in 2001, 2002, and 2003, respectively (193). In all the above cases, background infection by C. acutatum tended to be a complicating factor. A successful blueberry disease management program depends on the grower’s ability to recognize and anticipate problems. Knowing the critical periods when infection risk is highest can help in optimizing fungicide timing and effectiveness. The purpose of this study was to determine the role of temperature, wetness duration, interrupted wetness and relative humidity on infection parameters for C. acutatum on blueberries under controlled conditions and use the results to develop a disease prediction model. MATERIALS AND METHODS Fungal and plant material A single-conidium isolate of C. acutatum from blueberry fruit collected in Grand Junction, MI, USA in August 2006 (isolate #0001) was used for all experiments except the initial temperature wetness duration experiment on immature fruit. For this initial experiment, a singleconidium isolate from an infected sporulating blueberry bud was used. Both isolates were stored as previously described (125) and cultures were grown on potato dextrose agar (PDA) for a period of 14 days, after which conidia were harvested, re-cultured on one-quarter-strength PDA and allowed to microconidiate (produce conidia directly on conidia) for 3 to 4 days. For 29 inoculum production, sporulating cultures were flooded with 3 mL of sterile deionized water (SDW), and microconidia were dislodged using a sterilized L-shaped glass rod. The conidium concentration was determined using a haemocytometer, and the appropriate concentration was achieved via dilution with SDW. To study the development of melanized appressoria on parafilm 5 glass slides, a suspension of 1 × 10 conidia/ml was utilized. For all fruit infection experiments, 6 a suspension of 1 × 10 conidia was applied directly to the fruit surface. Several preliminary experiments were done on detached and attached immature and mature fruit but because of difficulty with background infections or suspected fungicide residues on the fruit, results from only a small subset were considered acceptable for further analysis. Also, to avoid background infection as much as possible, fruit was collected from multiple sources. All experiments presented here had little or no evidence of background infections. For the temperature and wetness duration assays, 20-cm-long detached ‘Jersey’ shoots bearing immature fruit were collected in June 2006 and 2007 from the Michigan State University (MSU) Horticulture Teaching and Research Center in East Lansing, MI. Detached mature ‘Jersey’ fruit were obtained from a commercial farm in Harrietta, MI in September 2006 and ‘Bluecrop’ fruit were obtained from a commercial farm in Traverse City, MI in September 2007. For interrupted wetness period experiments, immature ‘Jersey’ fruits were obtained from MSU’s Southwest Michigan Research and Extension Center in Benton Harbor, MI in July 2011, and detached mature ‘Bluecrop’ fruit were obtained from a commercial farm in Traverse City, MI in August 2011. For relative humidity assays, detached mature fruit were obtained from Sun Belle Berries (Miami, FL) (originating from Florida) and believed to be a southern highbush blueberry cultivar (southern-adapted V. corymbosum, usually with some introgressed V. darrowi Camp 30 and/or V. ashei Reade) based on appearance and harvest timing. In a preliminary experiment, the cultivar was susceptible to C. acutatum with similar symptoms to those of ‘Bluecrop’ or ‘Jersey’. Effect of temperature and wetness duration The optimal temperature for the growth of our isolate (#0001) was investigated by transferring 5.5-mm mycelial plugs to one-quarter strength PDA in 9-cm Petri dishes, and incubating the cultures in the dark at various temperatures for 10 days. Culture diameter (in mm) was then measured and results reported are the combined results of six experiments. Initially, 10, 15, 20, 25, and 30°C were tested and more detailed temperatures were evaluated in subsequent experiments. The following temperatures were tested across all experiments 5, 10, 15, 20, 22, 24, 25, 26, 28, 30, 31, 33, and 35°C. For each temperature, four replicate cultures were used to calculate the mean (n = 4) and each temperature was evaluated at least twice. To determine the effects of temperature and wetness duration on the formation of 6 melanized appressoria, 20-µl droplets of a conidial suspension (10 conidia/ml) were placed on parafilm-covered glass slides and placed in Petri dishes (90 mm diameter, 15 mm deep) with wet Whatman filter paper to create a humidity chamber. Plates were incubated at 10, 15, 20, 25, or 30°C in the dark with 4, 6, 8, 12, 16, 20, 24, 36, or 48 hours of wetness. For each temperature, 5 replicate droplets were used and 50 random conidia were counted per droplet (250 conidia in total) (n = 5). This experiment was conducted twice and experiments were normalized together using the highest mean melanized appressoria (%). No significant effect of experiment was observed during an ANOVA (P = 0.302), so therefore the average incidence of infected fruit was plotted in SigmaPlot for curve fitting. 31 For experiments with immature ‘Jersey’ fruit, detached fruit clusters were collected from mature field-grown plants in East Lansing, MI, and were placed in wet Florafoam® in 1 liter 6 plastic containers. The fruit was inoculated with C. acutatum (10 conidia/ml) and enclosed in a plastic bag. Fruit clusters wetted with SDW served as a control. Inoculated fruit clusters were incubated at 10, 15, 20, 25, or 30°C in the dark with 6, 12, 18, 24, 36, or 48 hours of wetness. Four twigs were used per treatment combination. At the end of the wetness period, ten green berries were removed from each twig, surface sterilized in 10% bleach, rinsed three times in SDW, cut in half and placed on potato dextrose agar in Petri dishes. Berries were observed for the presence of C. acutatum infection over a 2-week period. This experiment was conducted twice in 2006 and 2007 and experiments were normalized together using the highest mean infection rate (%). No significant effect of experiment was observed during an ANOVA (P = 0.287) so therefore, the average incidence of infected fruit was then plotted in SigmaPlot for curve fitting. For experiments with mature fruit, ten detached fruit per replicate (n = 4) were inoculated 6 by placing a 30-μl drop of C. acutatum (10 conidia/ml) in the calyx cup of each fruit. Fruit were then subjected to the same temperature-wetness period combinations as described for immature fruit. At the end of the wetness period, the fruit were surface sterilized, incubated at 100% RH, and observed for C. acutatum infection as described above. This experiment was conducted twice using ‘Jersey’ fruits in 2006 and ‘Bluecrop’ fruits in 2007 and experiments were normalized together using the highest mean infection rate (%). No significant effect of experiment was observed during an ANOVA (P = 0.416) so therefore, the average incidence of infected fruit was then plotted in SigmaPlot for curve fitting. 32 Effect of interrupted wetness periods To determine the effect of interrupted wetness on formation of melanized appressoria, 6 10-µl droplets of a conidial suspension of 10 conidia/ml were placed on parafilm-covered glass slides as described above. Slides were then placed in humidity chambers and incubated at room temperature (22 to 24°C) with wetness durations of 12, 24 and 48 hours. Interruptions of 1, 4, and 16 hours were then applied in the middle of each wetting period by removing parafilmcovered glass slides from humidity chambers and placing them in a laminar flow hood at room temperature (22 to 24°C and approximately 40% relative humidity). Following the interruption, 10-µl droplets of SDW were applied to the dried droplet location and the slides were returned to the humidity chambers. For each treatment, 4 replicate droplets were used and 50 conidia were counted per droplet (200 conidia in total) (n = 4). The mean was calculated for each treatment, average values were standardized against the uninterrupted control and data were plotted in SigmaPlot for curve fitting. This experiment was conducted twice. For experiments with immature and mature fruit, a 10-µl droplet of conidial suspension was placed on the side of detached ‘Jersey’ fruit placed equidistantly on wire-mesh screens. Fruit were then placed in humidity chambers and incubated at room temperature (22 to 24°C) with wetness durations of 12, 24 and 48 hours. Interruptions of 1, 4, and 16 hours were then applied in the middle of each wetting period by removing fruit from humidity chambers and placing them in a laminar flow hood at room temperature (22 to 24°C and approximately 40% relative humidity). Following the interruption, 10-µl droplets of SDW were applied to the sites of the dried droplets and the fruit were returned to the humidity chamber. At the end of the wetness period, fruit were removed, surface sterilized as described above, cut in half and placed on potato 33 dextrose agar in Petri dishes. For each treatment, four replicates of 10 fruit each were used (40 fruit in total per maturity level). The mean was calculated for each treatment, average values were standardized against the uninterrupted control and data were plotted in SigmaPlot for curve fitting. This experiment was conducted twice. Effect of relative humidity To determine the effect of relative humidity (RH) on formation of melanized appressoria, 6 10-µl droplets of a conidial suspension (10 conidia/ml) were placed on parafilm-covered glass slides and allowed to dry in a laminar flow hood as described above. Slides were then placed in chambers with varying RH levels using using different concentrations of glycerol (0, 10, 33, and 66%) in water. For each treatment, three replicates were used and 50 conidia were counted per replicate (150 conidia in total). Model 450 Watchdogs (Spectrum Technologies Inc., Plainfield, IL) were used to monitor relative humidity in the chambers. Relative humidity varied with temperature. The following RH values were achieved at 20°C: 100%, 96%, 86%, and 56%; at 25°C: 100%, 95%, 84%, and 54%; and at 30°C: 100%, 94%, 80%, and 50%. Slides were incubated for 3 days and monitored daily for the production of melanized appressoria. Melanized appressoria were observed after 2 days at 100% RH at all temperatures but were not observed at lower relative humidity values for 3 days. The percentage of appressorium formation was calculated for each treatment and average values were plotted in SigmaPlot for curve fitting. 6 For the mature fruit experiment, a conidial suspension (10 conidia/ml) was sprayed until runoff on the surface of detached mature fruit and allowed to dry as described above. Fruit were then placed in humidity chambers and incubated at various temperatures (20, 25 and 30°C) with 34 Table 2.1. Fitted-models for the development of melanized appressoria and infection level of immature and mature blueberry fruits by Colletotrichum acutatum using temperature (°C), wetness duration (h), wetness interruption (h), and relative humidity (%). Nine equations were developed based on curve fitting the raw data from a mixture of experiments and models were 2 examined for goodness of fit (R ), standard error of the estimate (SEE), the P value of the regression and the mean squared error (MSE) of the residuals (* denotes a chosen equation). 35 Table 2.1 (cont’d.) Fitted-model (corresponding equation) Planer Parabolic Gaussian* (Eq. A) Lorentzian 0.699 0.790 0.890 0.846 Wetness duration Temperature Immature infected fruit Planer Parabolic Gaussian* (Eq. D) Lorentzian Wetness duration Temperature Mature infected fruit Wetness interruption Wetness duration Relative melanized appressoria Independent/Dependent variables Adj. 2 R P-value Residual MSE 0.684 10.348 0.769 8.848 0.879 6.406 0.831 7.569 < 0.001 < 0.001 < 0.001 < 0.001 107.075 78.285 41.030 57.295 0.718 0.781 0.856 0.814 0.700 14.816 0.752 13.474 0.837 10.941 0.790 12.414 < 0.001 < 0.001 < 0.001 < 0.001 219.519 181.542 119.715 154.098 Planer Parabolic Gaussian* (Eq. G) Lorentzian 0.633 0.701 0.899 0.859 0.610 18.164 0.661 16.940 0.885 9.870 0.840 1.6413 < 0.001 < 0.001 < 0.001 < 0.001 329.921 286.951 97.421 135.559 Planer* (Eq. B) Parabolic 0.945 0.971 0.926 0.943 < 0.001 0.002 44.881 54.967 Gaussian Could not fit – exceeded maximum iterations Lorentzian Could not fit – exceeded maximum iterations Planer* (Eq. E) Parabolic 0.897 0.916 0.862 11.299 0.833 12.448 0.001 0.020 127.664 154.951 Gaussian 0.908 0.816 13.049 0.024 170.287 Lorentzian 0.794 0.589 21.148 0.110 447.241 Planer* (Eq. H) Parabolic 0.781 0.837 0.708 17.991 0.674 18.995 0.011 0.071 323.681 360.820 Gaussian 0.851 0.702 18.157 0.060 329.675 Lorentzian 0.851 0.702 18.157 0.060 329.675 Relative humidity Temperature Melanized appressoria Planer Parabolic Gaussian* (Eq. C) Lorentzian 0.183 0.002 12.467 0.402 155.430 0.494 0.204 11.132 0.252 123.914 0.996 0.993 1.022 < 0.001 1.045 Could not fit – exceeded maximum iterations Relative humidity Temperature Mature infected fruit Planer Parabolic Gaussian Lorentzian* (Eq. F,I) 0.663 0.588 14.114 0.008 199.191 0.907 0.854 8.411 0.001 70.736 Could not fit – exceeded maximum iterations 0.971 0.954 4.735 < 0.001 22.424 Wetness duration Temperature Melanized appressoria Wetness interruption Wetness duration Relative immature infected fruit Wetness interruption Wetness duration Relative mature infected fruit 2 R 36 SEE 6.699 5.913 varying levels of relative humidity (same levels as described above) for 3 days. For each temperature-RH combination, 10 fruits were used, replicated four times (40 fruits total per fruit maturity level). At the end of the incubation period, fruit were removed, surface sterilized as described above, placed equidistantly on wire mesh grates and incubated at 100% relative humidity. Fruit were observed for the presence of C. acutatum infection at 10 days post inoculation. Mean values were then calculated and plotted in SigmaPlot for curve fitting. Statistical analysis All statistical analyses were performed with SIGMAPLOT version 11 (SYSTAT Software, Chicago, IL). Regression analyses were used to develop quantitative models relating to temperature, wetness duration, interrupted wetness periods, and relative humidity. Threedimensional nonlinear regressions including Planer, Parabolic Gaussian, and Lorentzian were fit to the data from two experiments. Criteria for selecting the best statistical models included an 2 examination of normality, equality of variance, R values, and residual plots (Table 2.1). Equations were chosen and three models were proposed predicting the development of melanized appressoria and the infection levels of immature and mature blueberry fruits (Table 2.2). A summarized equation for each model was developed that best fit the data. RESULTS Temperature and wetness duration have an effect on the development of melanized appressoria and fruit infection Mycelial growth of our isolate of C. acutatum increased with increasing temperature from 10 to 26°C but then declined sharply between 27 and 30°C (Figure 2.1). Little to no 37 Colony diameter (mm) after 10 days 80 60 40 20 0 0 5 10 20 15 25 30 35 40 Temperature (°C) Figure 2.1. The effect of temperature on mycelial growth of Colletotrichum acutatum (isolate #0001) on quarter-strength potato dextrose agar. Colony diameter was measured in two perpendicular directions and averaged after a 10-day incubation at each temperature. Values are the average of three replications and error bars denote the standard error of the mean. 38 Table 2.2. Proposed model for the development of melanized appressoria and the infection level of immature and mature blueberry fruits by Colletotrichum acutatum. Nine equations were developed based on curve fitting the raw data from a mixture of experiments and these equations were used to make three combined models. Independent/Dependent Variables Equation Type X Y Z A Gaussian Wetness Temperature Melanized duration appressoria B Planer Interruption Wetness Relative duration Melanized appressoria C Gaussian Relative Temperature Melanized humidity appressoria D Gaussian Wetness Temperature Immature duration infection level E Planer Interruption Wetness Immature duration infection level F Lorentzian Relative Temperature Immature humidity infection level G Gaussian Wetness Temperature Mature duration infection level Mature H Planer Interruption Wetness duration infection level I Lorentzian Relative Temperature Mature humidity infection level 1 1 x0 48.6677 Coefficients y0 a 21.7020 68.5131 b 16.6815 c 6.8695 N/A 8.7148 -1.6989 1.3201 N/A 99.7093 23.8718 50.0223 2.2937 2.2990 46.7302 24.6465 94.4065 24.7335 10.5797 N/A 14.9471 -0.0945 1.8141 N/A 104.1017 25.8143 80.4083 10.7745 6.1264 42.5273 23.5491 101.6045 N/A 54.2761 -1.3035 1.7641 N/A 104.1017 25.8143 80.4083 10.7745 6.1264 17.4474 6.6446 2 Various equations where used to fit data as follows Planer: f (Z) = y0+a*x+b*y, Gaussian: f (Z)=a*exp(-0.5*( ((x-x0)/b) + ((y2 2 2 y0)/c) )), and Lorentzian: f (Z) = a/((1+((x-x0)/b) )*(1+((y-y0)/c) ))). . 39 2 R = 0.89 P < 0.001 80 60 40 20 40 30 0 30 25 20 20 10 15 10 Figure 2.2. The effect of temperature and wetness duration on the predicted development of melanized appressoria of Colletotrichum acutatum on parafilm-covered glass slides. A Gaussian equation (Table 2.2, Equation A) was used to fit the data. Shading intensity of the surface indicates changes in the percentage of melanized appressoria at 20% intervals starting at zero. 40 2 R = 0.86 R = 0.90 120 120 80 80 40 40 0 30 A 2 P < 0.001 25 20 15 10 40 30 20 10 0 30 25 B 20 P < 0.001 15 10 40 30 20 10 Figure 2.3. The effect of temperature and wetness duration on the predicted infection level of immature (A) and mature (B) blueberry fruits by Colletotrichum acutatum. Separate Gaussian equations were fit to the data for immature fruit (Table 2.2, Equation D) and mature fruit (Table 2.2, Equation G). Shading intensity of the surface indicates changes in the percentage of infected fruit at 20% intervals starting at zero. 41 Infected fruit (%) Infected fruit (%) Melanized appressoria (%) Wetness duration (h) 12 48 24 B C D E F 60 G H I 60 A 40 20 0 40 30 20 10 0 40 20 0 0 1 4 16 0 1 4 16 0 1 4 16 Interruption (h) Figure 2.4. The effect of interrupted wetness periods (0-, 1-, 4-, and 16-h of dry time in the middle of 12-, 24-, and 48-h wet periods) on the development of melanized appressoria (%) of Colletotrichum acutatum (A-C) and the percentage immature (D-F) and mature (G-I) blueberry fruit infected by C. acutatum. Values are the average of four replications and error bars denote the standard error of the mean. 42 growth was observed at 5°C and 35°C. The optimal temperature for melanized appressorium formation was 25°C and development was significantly reduced at 10 and 30°C. At 25°C, a minimum of 8 hours of continuous wetness was required for the development of melanized appressoria. However, 12 hours were required at 30°C and 24 hours were required at 10°C. 2 A Gaussian model was found to be the best fit for the raw data (R = 0.89) (Table 2.2, Equation A, and Figure 2.2). In the infection assays of immature and mature fruit, low levels of infection were seen after wetness durations of 6 hours (Figure 2.3). However, the infection percentage increased considerably after 12 to 18 hr of wetness at 20ºC to 25ºC, 18 to 24 hours at 30ºC, 24 hours at 15ºC, and 36 to 48 hours at 10ºC (Figure 2.3). The optimal temperature for fruit infection by C. acutatum was determined to be 25ºC based on infection incidence. For both immature and mature fruit experiments a Gaussian model was found to be the best fit for the 2 2 raw data sets (R = 0.86, R = 0.90, respectively) (Table 2.2, Equations D and G). In both immature and mature fruit studies presented here, no background infections were observed on uninoculated control fruit. Interrupted wetness periods reduce melanized appressorium formation and infection at shorter wetness durations In general, the longer the wetness period, the higher the percentage of melanized appressoria formed and infected fruit. However, interruption of the wetness period 43 2 R = 0.95 P < 0.001 80 60 40 20 16 14 12 10 0 8 45 40 6 35 30 4 25 20 15 2 Figure 2.5. The effect of wetness duration and interrupted wetness periods on the predicted development of melanized appressoria of Colletotrichum acutatum on parafilm-coated glass slides. Values are calculated as the percent melanized appressoria relative to the continuous wetness treatment. A Planer equation (Table 2.2, Equation B) was fit to the data. Shading intensity of the surface indicates changes with the percentage of melanized appressoria at 20% intervals starting at zero. 44 2 R = 0.90 2 P < 0.001 R = 0.78 120 120 80 80 40 0 45 A 40 16 12 0 8 35 25 P < 0.001 4 15 16 12 45 B 8 35 25 4 15 Figure 2.6. The effect of wetness duration and interrupted wetness periods on the predicted infection level of immature (A) and mature (B) blueberry fruit by Colletotrichum acutatum. Values are calculated as the percent infected fruit relative to the continuous wetness control. Planer equations (Table 2.2, Equation E) were fit to the data for immature fruit and mature fruit (Table 2.2, Equation H). Shading intensity of the surface indicates changes in the percentage of infected fruit at 20% intervals starting at zero. 45 significantly reduced successful appressorium formation and infection. The average percentage of melanized appressoria produced on parafilm-covered slides with no interruption in the wetness duration was 6%, 30%, and 63% after 12, 24, and 48 hours of wetness, respectively. For all three incubation times, the longer the wetness interruption, the greater the reduction in the number of melanized appressoria produced (Figure 2.4A-C). In order to compare the interrupted wetness treatments with the continuous wetness duration treatments, the control from each incubation time was used to standardize the amount of 2 melanized appressoria relative to no interruption. A Planer model best fit the raw data (R = 0.95) (Table 2.2, Equation B, and Figure 2.5). The percentage of infected immature fruit was 15%, 17%, and 30%, respectively, after 12, 24, and 48 hours of continuous wetness. At the 12- and 24-hour incubation times, interruptions of any duration dramatically reduced the percentage of infected fruit (Figure 2.4D and E). No reduction in the percentage of infected immature fruit was observed at the 48 hour incubation time (Figure 2.4F). In order to compare the interruption treatments with the continuous wetness duration treatments the control from each incubation time was used to standardize the amount of infected fruit relative to no interruption. A Planer model was 2 found to be the best fit for the raw data (R = 0.90) (Table 2.2, Equation E, and Figure 2.6A). The percentage of infected mature fruit was 25%, 28%, and 38%, repectively at 12, 24, and 48 hours of continuous wetness. At the 12-hour incubation time, there was a reduction in fruit infection but the duration of the interruption did not seem to matter (Figure 2.4G). No reduction in the percentage of mature fruit infection due to interrupted wetness was observed at the 24- and 48-h incubation times (Figure 2.4H and I). In order to compare 46 2 R = 0.99 2 R = 0.97 P < 0.001 80 P < 0.001 120 60 80 40 20 0 30 28 26 24 A 22 40 100 0 30 28 26 24 22 B 80 60 20 100 80 60 20 Figure 2.7. The effect of temperature and relative humidity on the predicted development of melanized appressoria (%) of Colletotrichum acutatum (A), and the predicted infection level of mature blueberry fruits (B) by C. acutatum. A Gaussian equation (Table 2.2, Equation C) was fit to the data for melanized appressoria and a Lorentzian equation (Table 2.2, Equations F and I) was used to fit the data for mature fruit. Shading intensity of the surface indicates changes in the percentage of melanized appressoria and percentage infected fruit at 20% intervals, respectively, starting at zero. 47 the interruption treatments with the continuous wetness duration treatments, the control from each incubation time was used to standardize our amount of infected fruit relative to no 2 interruption. A Planer model best fit the raw data (R = 0.78) (Table 2.2, Equation H, and Figure 2.6B). Relative humidity affects melanized appressorium development and fruit infection Conidia germinated and had formed melanized appressoria on parafilm-covered glass slides at 100% relative humidity at all temperatures tested after 48 h with an increase seen after 72 h of incubation. At 25°C, the percentage of melanized appressoria was 5%, 1%, and 0% at 95%, 84%, and 54% relative humidity, respectively. A Gaussian model best fit the raw 2 data (R = 0.99) (Table 2.2, Equation C, and Figure 2.7A). Mature fruit was infected at all temperatures, and increasing relative humidity led to higher infection incidence. At 20°C, the percentage of infected fruit was 43%, 20%, 7% and 3% at 100%, 96%, 86%, and 56% relative humidity, respectively. At 25°C, the percentage of infected fruit was 67%, 50%, 17% and 3% at 100%, 95%, 84%, and 54% relative humidity, respectively. At 30°C, the percentage of infected fruit was 47%, 30%, 13% and 0% at 100%, 94%, 80%, and 50% relative humidity, respectively. A Lorentzian model best fit the raw data 2 (R = 0.97) (Table 2.2, Equation F and I, Figure 2.7A). Since immature fruit was not tested, this model was included in the proposed comprehensive model for immature fruit infection (Table 2.2, Equation F). 48 Comprehensive models for appressorium development and infection of immature and mature fruit Three models were constructed based on three nonlinear regressions described above for temperature in relation to wetness duration and relative humidity, and wetness duration in relation to wetness interruption (Table 2.2). For the formation of melanized appressoria, each 2 component of the proposed model had R values greater than 0.89 (Table 2.1). For the 2 percentage of infected immature fruit, each component of the proposed model had R values greater than 0.86 (Table 2.2). For the percentage of infected mature fruit, each component of 2 the proposed model had R values greater than 0.78 (Table 2.3). For all models, an analysis of residuals did not reveal outliers or particular patterns. Therefore three general equations were developed for the development of melanized appressoria and infection in immature and mature fruit using chosen equations in Table 2.2. For the development of melanized appressoria: % For infection level of immature fruit: % For infection level of mature fruit: % 49 DISCUSSION Results of this study indicate that temperature, wetness duration, wetness interruptions and relative humidity have a direct effect on the development of melanized appressoria and infection of both immature and mature blueberry fruits by C. acutatum. The relationship between temperature and wetness conforms to previous studies of Colletotrichum spp. on other fruit crops (130, 168, 206). Temperature effects were also consistent with the mycelial growth assay. While the optimum temperature was found to be 26°C for mycelial growth, we did not investigate the effect of temperature in such small increments for appressorium formation and infection for which 25°C was found to be optimum when using 5°C increments. The development of melanized appressoria was observed as early as 8 hours after incubation on parafilm slides and infection as early as 6 hours after inoculation of immature and mature fruit at the optimum temperature of 25°C. Previous microscopy studies conducted on detached, ripe ‘Jersey’ fruit incubated at room temperature (22 to 24ºC) showed that 4 hours after inoculation, 10.6% of conidia had germinated (30). At 8 hours post inoculation, 41.5% of conidia had formed unmelanized appressoria and 11.4 % had formed melanized appressoria with an internal light spot, which is thought to indicate formation of a penetration peg (201). It is therefore possible that a small percentage of conidia could have successfully infected the fruit as early as 6 hours after inoculation under optimal conditions. Differences in the minimum environmental requirements for infection and the development of melanized appressoria on parafilm covered glass slides may be due to the presence of plant-derived waxes on the surface of blueberry fruits. The surface wax of avocado fruit has 50 been shown to induce germination and appressorium formation in the spores of Colletotrichum gloeosporioides (148). Additionally, disease incidence was generally higher in mature fruit than in immature fruit indicating greater tissue susceptibility, which is similar to other work in strawberries (206). This study also demonstrates that wetness interruptions play an inhibitory role in the infection process. Previous research on Cercospora kikuchii T. Matsumoto & Tomoy. on soybean leaves has shown that interruptions of different durations in 24-hour wetness periods have a significant effect on overall disease severity and the number of infections. Also, they found that the relative humidity during the interruption was important for disease severity in the sense that interruptions during which the relative humidity was high had less of an effect those during which the relative humidity was low (171). In our study, a greater inhibition of infection at lower initial wetness durations and longer interruptions was observed. This is likely due to the fact that the fungus has not yet developed protective structures or penetrated the host tissue and was therefore more vulnerable to desiccation. This study also identifies and quantifies relative humidity as an important environmental factor in the infection process of C. acutatum on blueberry. Our results agree with previous work on Colletotrichum gloeosporioides on mango which showed that appressoria are capable of developing between 95 and 100% relative humidity (56). Additionally, research on C. gloeosporioides on citrus fruits has shown that sustained high relative humidity for three days dramatically increases disease incidence (22). This component of our model is relevant in humid growing regions like Michigan where average daily relative humidity can remain above 90% for several days during the growing season. 51 High relative humidity may also reduce the impact of wetness interruptions on appressorium formation and infection. Many disease prediction models use regression equations, such as those based on polynomials (65, 170), logistic equations (26, 170), and complex three-dimensional response surfaces (21, 29, 207). Wilson et al. (1990) used a regression model with the logit of disease incidence of C. acutatum as the dependent variable which accurately described infection level as a function of wetness duration and temperature. In the C. coccodes-tomato pathosystem, fruit infection was correlated with the amount and duration of rain alone and in combination with other meteorological variables, accouting for 72% of the variation in anthracnose incidence. Infection was also negatively correlated with the number of hours during which no rainfall occurred within 4-day intervals that tomato fruit were exposed to field conditions (168). Another study used polynomial equations to describe the relationship between C. orbiculare infection on watermelon leaves in relationship to temperature and leaf wetness (130). Our models focus on the environmental requirements for infection; however, there are other epidemiological variables to consider such as pathogen dispersal and survival. Colletotrichum spp. have been classically identified as rain-splash-dispersed pathogens, and numerous studies link dispersal of conidia to rainfall (14, 210). Additionally, survival of Colletotrichum spp. over the winter in soil, host tissue and plant debris has been described in numerous pathosystems (49, 62, 68, 136, 212) while survival over the summer has been described on symptomless leaves of strawberry (112), which also is an important factor when predicting disease risk. While the assumption for our model is that inoculum is uniformly 52 distributed and unlimited, this may not be the case and inoculum availability needs to be taken into account when determining infection risk. This study provides useful information on temperature and wetness requirements for infection of blueberry fruit by Colletotrichum acutatum under controlled conditions. The absence of or suppression of background infection is very important to obtain accurate results. This information is the first step towards the development of a disease prediction model for anthracnose fruit rot in blueberries. Further research is needed to validate these results under field conditions. In the field, fruit clusters may stay wet longer than an exposed leaf wetness sensor of a weather station. Further studies are needed to determine how fruit wetness relates to behavior of electronic wetness sensors. ACKNOWLEDGEMENTS I would like to gratefully acknowledge funding from Michigan State University Project GREEEN (Generating Research and Extension to meet Economic and Environmental Needs) and the Michigan Blueberry Growers (the Blueberry People). I would also like to thank Christopher Woelk and Daniel Svoboda for technical assistance and Laura Avila for critical reading of the manuscript. 53 CHAPTER III: IDENTIFICATION OF DIFFERENTIALLY EXPRESSED GENES IN A RESISTANT VERSUS SUSCEPTIBLE CULTIVAR AFTER INFECTION BY COLLETOTRICHUM ACUTATUM Published in the peer-reviewed journal Molecular Plant Pathology (June 2011, Volume 12, Pages 463–477) ABSTRACT Anthracnose fruit rot, caused by the fungus Colletotrichum acutatum, is an important disease of blueberries worldwide. Cultivar Elliott is resistant, severely restricting fungal growth and sporulation compared to the susceptible cultivar Jersey. The objectives of this research were to: 1) Analyze pre-penetration events in ‘Elliott’ and ‘Jersey’ at different stages of fruit development, 2) Identify putative defense genes in ‘Elliott’ fruit, and 3) Monitor timing of the oxidative burst in both cultivars. Light microscopy revealed no differences in pre-penetration activities of C. acutatum on immature fruit of both cultivars. However, at fruit ripening, conidia germinated and formed appressoria faster on ‘Jersey’ than in ‘Elliott’ fruit. Using suppression subtractive hybridization, 37 differentially expressed sequence tags (ESTs) were detected in ‘Elliott’ versus ‘Jersey’ upon infection. Several of the ESTs had homology to known plant defense genes, such as a class II chitinase, pathogenesisrelated protein 10 (PR10), and β-1-3 glucanase. Two putative genes involved in oxidative stress were identified: a metallothionein-like protein and monodehydroascorbate reductase. ESTs of fungal origin were also detected. Many ESTs had no homology to known genes. Using semi-quantitative and quantitative RT-PCR, expression of most of the candidate genes 54 was detected earlier in ‘Elliott’ than in ‘Jersey’, some within 24 hours post infection (hpi). Monitoring of the oxidative burst showed that the overall H2O2 concentration was two to three times higher in ‘Elliott’ than in ‘Jersey’ at 24 hpi. Elucidation of the basis of resistance to C. acutatum in blueberries will facilitate the development of new anthracnose fruit rotresistant cultivars. INTRODUCTION Colletotrichum species are ubiquitous fungal pathogens, infecting numerous agronomically important plant species (2, 13, 16, 17, 67, 69, 189). Colletotrichum acutatum J.H. Simmonds causes fruit rots in a range of hosts, including almonds, apples, citrus, olives, stone fruits and strawberries (1, 131, 146). C. acutatum was first described as a separate species in 1965 (175). While initial studies referred to C. gloeosporioides (Penz.) Penz. & Sacc. as the causal agent of anthracnose fruit rot on highbush blueberries (46, 85, 128), C. acutatum is now considered the primary cause of the disease in temperate regions (146, 150, 169, 192, 213). The main symptom of anthracnose on blueberries is rotting of ripe fruit in the field before harvest and in storage after harvest (128). Initially, sunken areas develop on the fruit surface followed by the formation of acervuli exuding salmon-colored conidial masses. This disease can have a severe economic impact, with pre-harvest losses estimated at 10-20% and post-harvest losses as high as 100% in storage (128). Most blueberry cultivars are susceptible to anthracnose fruit rot. However, several resistant cultivars have been identified including ‘Elliott’, which displays strong resistance in the field and in laboratory inoculation studies (63, 150). 55 Host-pathogen interactions have been well characterized in only a few Colletotrichum-fruit pathosystems, including avocado (159), citrus (23, 189), mango (86), strawberry (8, 9, 25, 31) and pepper (7, 101, 102, 107, 119). Following direct penetration of host tissues, Colletotrichum species generally have two different host colonization strategies depending on the tissue or host being colonized: intracellular hemibiotrophy and subcuticular intramural necrotrophy. Intracellular hemibiotrophy is the direct invasion of the initial host cell by a primary infection vesicle, followed by the proliferation of primary and secondary hyphae; this strategy is common in Colletotrichum spp. infecting bean (137), pea (138), sorghum (199), and tobacco (173). In subcuticular intramural necrotrophy, Colletotrichum spp. will invade the plant superficially under the cuticle, generally producing necrotrophic hyphae sooner, proliferate intramurally, and only later invading the host cell intracellularly. This type of strategy is typical for infection by C. capsici on cowpea (151) and cotton (162). C. acutatum is known to use both infection strategies on almond (53), olive, (75) and strawberry (43). A recent study of the infection process on ripe blueberry fruit showed that C. acutatum exhibits different infection strategies depending on the cultivar being colonized (201). In the susceptible cultivar Jersey, the host-pathogen interaction was described as intracellular hemibiotrophy, and in the resistant cultivar Elliott as intramural necrotrophy. A build-up of phenolic compounds, degradation of fungal hyphae and an accumulation of anthocyanins were noticed in ‘Elliott’, suggesting an active resistance response (124, 201). In addition, conidium germination and appressorium formation were slower on ‘Elliott’ than ‘Jersey’ fruit (201). The infection process on immature fruit tissues of either cultivar was not investigated. 56 A variety of defense mechanisms in other Colletotrichum-plant interactions have been observed, including the production of reactive oxygen species (25), host-derived cell wall-degrading enzymes (31, 76, 110, 204), enzymes involved in secondary metabolism (7, 28, 111), host cell wall modification (101, 178), and pre-formed and induced antifungal compounds (58, 159). Upon infection of pepper fruits (Capsicum annuum L.) by C. gloeosporioides, many defense-related proteins like cytochrome P450 (141), defensin, thionin-like protein (140), thaumatin-like protein (103) and esterase (107) are induced in incompatible interactions. Since most of these defense responses can be monitored at the transcriptional level, a broad genetic screen can provide insight into the type of defense mechanism involved in a particular plant pathosystem. Defense responses in fruit to infection by Colletotrichum spp. have been studied mostly in immature fruit during the latent (or quiescent) phase of the infection (156), and are typical of ontogenic resistance. As avocado fruits mature, the pectin lyase activity of C. gloeosporioides increases as the inhibitor epicatechin slowly decreases in concentration. In addition, antifungal diene compounds are degraded by fruit lipoxygenases, and the fruit gradually become increasingly susceptible to infection (159). With the exception of the C. gloeosporioides-C. annuum system, to our knowledge, no research has been published on the defense mechanisms in ripe fruit at different stages of the infection process. Casado-Diaz et al. (2006) studied the resistance response of ripe strawberry fruit to C. acutatum by comparing naturally infected fruit to uninfected fruit without regard to infection timing. The C. acutatum-blueberry pathosystem is of interest because the defense response occurs in ripe fruit, and moreover, the infection strategy of the pathogen changes depending on host susceptibility (201). 57 A better understanding of the basis of resistance in the C. acutatum-blueberry pathosystem will aid the development of alternative management strategies for a major blueberry disease that is currently controlled primarily with fungicides (169). While anthracnose fruit rot resistance is an objective of blueberry breeding programs, no genotypic or phenotypic markers are currently available for rapid resistance screening of blueberry breeding lines. Evaluation of plants for anthracnose fruit rot resistance can only be accomplished when they bear fruit in sufficient quantities for inoculation, usually at 2 to 3 years of age. In previous studies, anthracnose fruit rot resistance in blueberries was not correlated with resistance to foliar infection (64) or production of antimicrobial fruit volatiles (149). Identification of anthracnose fruit rot resistance mechanisms and associated molecular markers could improve resistance screening procedures. The goal of this research was to compare the infection process by C. acutatum in fruit of the resistant cultivar Elliott versus the susceptible cultivar Jersey at different stages of fruit development, and to analyze the molecular mechanisms that underpin host plant resistance to anthracnose fruit rot in ‘Elliott’ blueberries. MATERIALS AND METHODS Fungal cultures and media A single-conidium isolate of Colletotrichum acutatum from blueberry fruit collected in Grand Junction, MI, USA in August of 2006 was used for all experiments. The isolate was stored as conidia in a nutrient solution (20% glycerol, 0.04% yeast extract, 0.1% malt extract, 0.04% glucose, 0.02% K2HPO4) at -80 °C. Cultures were grown on ¼-strength potato dextrose agar (PDA) for a period of 14 d, after which conidia were harvested, re-cultured on 58 ¼-strength PDA, and allowed to microconidiate for 3 to 4 d. This procedure was used to prepare conidial suspensions for storage and inoculum production and has been utilized in previous studies (177, 201, 203). For inoculum production, sporulating cultures were flooded with 3 mL of sterile deionized water (SDW), and microconidia were dislodged using a sterilized L-shaped glass rod. Conidia were counted using a hemacytometer, and the appropriate concentration was achieved via dilution with SDW. Plant material and inoculation procedures To confirm virulence of the pathogen and the host resistance phenotype observed in Wharton and Schilder (2008), ripe fruit were collected on 6 August 2006 from mature ‘Elliott’ and ‘Jersey’ bushes at the Southwest Michigan Research and Extension Center in Benton Harbor, MI, USA. Ten fruit of each cultivar were spray-inoculated until run-off with 6 a conidial suspension containing 10 conidia per milliliter, placed equidistantly on wire mesh over a layer of water in covered aluminum pans that acted as humidity chambers, and incubated for 8 d at 22-24°C. After 5 d, fruits were macroscopically examined daily for disease symptoms and signs. For scanning electron micrographs, ten fruit epidermal peels of (1 mm in diameter) each cultivar were dehydrated with ethanol, critical point dried, mounted on aluminum mounting stubs, and coated with gold using a gold sputter coater (EMSCOPE SC500 Sputter coater, Ashford, Kent, UK). Images were captured on a scanning electron microscope (SEM) (JEOL 6400V, Japan Electron Optics Laboratories, Tokyo, Japan). Micrographs were used to measure the diameter (in two perpendicular directions) of 50 randomly selected acervuli per cultivar. To quantify sporulation, inoculated fruit were incubated at 100% humidity on wire mesh screens in aluminum pans for 8 days (10 fruit per 59 replicate with five replicates). After incubation, ten inoculated fruit per replicate were shaken in 10 mL of sterile water for 5 minutes, and conidium concentration was determined using a hemacytometer. In order to monitor the pre-penetration activities of C. acutatum on fruit at different stages of development, 10-cm-long twigs with attached fruit clusters were collected from mature ‘Elliott’ and ‘Jersey’ bushes in Benton Harbor, Michigan, in 2008. The bushes had not been treated with fungicides for at least two months before removing the fruit clusters. Fruit clusters were collected at the following growth stages: petal fall (1 June), green fruit (16 June), fruit coloring (14 July for ‘Jersey’ and 28 July for ‘Elliott’), and ripe fruit (11 August). Fruiting twigs were placed with the cut ends in wet Oasis® Floral foam (SmithersOasis, Kent, OH, USA) and fruit were inoculated with 10-µL droplets of C. acutatum (1 x 5 10 conidia/mL) and incubated in round plastic containers (~950 mL) for 24, 48, 96, and 144 hours post inoculation (hpi) at 22–24 °C and 100% RH under continuous fluorescent light. After incubation, epidermal peels were collected from the inoculation sites, and were fixed and stained in accordance with procedures use by Wharton and Schilder (2008). A total of 100 conidia per replicate for five replicates were examined for germination and formation of unmelanized and melanized appressoria using light microscopy. All statistical analyses of the data were performed using a paired Student’s T-test (α=0.05) using the StatGraphics statistical computer program (StatPoint Inc., Warrenton, VA, USA) after checking for equality of variance. For the molecular experiments, ripe fruit of ‘Elliott’ and ‘Jersey’ (not treated with fungicides for at least 3 weeks prior to harvest) were provided by Hortifrut Chile S.A. (Santiago, Chile) and were equidistantly placed on wire mesh grates in humidity chambers 60 constructed from two Petri dish bottoms (15-cm diameter x 2.5-cm height) with moist filter paper on the bottom. Fruit were inoculated with 10-µL droplets of an aqueous suspension C. 6 acutatum (10 conidia/mL) or sterile deionized water (control) and incubated at 25°C in the dark. Fruit epidermal samples were collected prior to inoculation and at 24, 48, 96, and 144 hours after inoculation using a size 2 cork borer (5.5 mm in diameter) and forceps to remove the inoculated sections of the peel. The epidermal peels were placed in 100-mg aliquots (~5 to 6 peels) in microcentrifuge tubes and were flash frozen in liquid nitrogen. Ten aliquots were collected for each combination of cultivar, time interval after inoculation, and treatment (inoculation or water control). Samples were then stored at -80 °C until RNA extraction was performed. RNA extraction and cDNA library synthesis Total RNA was extracted from 100 mg of epidermal peels (five to six fruits) per cultivar (Elliott and Jersey) and time point (0, 24, 48-, 96- and 144-hpi) combination using a phenol-chloroform-based protocol that was adapted for use in a refrigerated microcentrifuge (166). The RNA extracts were then treated with a DNA-free kit (Applied Biosystems/Ambion, Austin, TX, USA) to remove residual genomic DNA. Extracted total RNA was denatured in glyoxal containing sample-loading dye (Ambion, Austin, TX, USA) at 50°C and was visualized on a 1.0%-agarose Tris-Acetate-EDTA (TAE) gel. Total RNA in each sample was quantified using a Nanodrop 2000 spectrophotometer (Thermo Scientific, Wilmington, DE, USA). For construction of the cDNA libraries for suppression subtractive hybridization, one compound sample per cultivar was created by 61 combining 0.2 µg RNA from each time point after inoculation with C. acutatum. cDNA synthesis was performed using 1.0 µg of RNA from each composite library in the Super SMART cDNA synthesis kit (Clontech Laboratories, Inc., Palo Alto, CA, USA) in accordance with the manufacturer’s protocols. Suppression subtractive hybridization, differential screening and sequencing Suppression subtractive hybridization (SSH) was performed with the cDNA from the Super SMART cDNA synthesis kit using 21 cycles to enrich the initial cDNA (Figure B.1) and the Suppression Subtractive Hybridization Kit (Clontech Laboratories, Inc.) in accordance with the manufacturer’s protocols (Figs. S2 and S3). A forward-subtracted ‘Elliott’ library (where ‘Elliott’ served as the tester and ‘Jersey’ as the driver) and a reversesubtracted ‘Jersey’ library (where ‘Jersey’ served as the tester and ‘Elliott’ as the driver) were constructed. The forward-subtracted cDNA library of ‘Elliott’ was cloned using the pGEM-T Easy Kit (Promega, Madison, WI, USA). Then 5 µL of the ligation reaction was used to transform DH5α ElectroMAX competent cells (Invitrogen, Carlsbad, CA, USA) by electroporation with the following parameters: 2.0 kV, 200 Ω, 25 μF. ‘Elliott’ cDNA clones were screened using a southern dot blot technique according to (183). To construct probes for each library, the PCR DIG Probe Synthesis Kit (Roche Applied Science, Mannheim, Germany) and 1 µL of the primary PCR product from the forward- and the reverse-subtracted libraries (derived from the SSH kit) were used as the template. The primers T7 (5'-AAT ACG ACT CAC TAT AG-3') and Sp6 (5'-ATT TAG GTG ACA CTA TAG-3') were used in a PCR reaction to amplify the insert from each of the clones (Figure 3.4A). For the blots, the DIG Luminescent Detection Kit for Nucleic Acids 62 (Roche Applied Science) was used and a 1-µL PCR product was loaded onto two separate 0.45-µ MAGNACHARGE nylon transfer membranes (GE Water and Process Technologies, Trevose, PA, USA) in a high-throughput 96-well system. One membrane was probed using 15 µL of the forward ‘Elliott’ library DIG PCR product and the other with 15 µL of the reverse ‘Jersey’ library DIG PCR product. Clones were considered differentially expressed if the diameter of the spot was more than 50% of that in the ‘Jersey’ reverse blot. Differentially expressed clones were independently screened three times, and a putative actin gene of -31 Vaccinium corymbosum with significant homology (E-value 7.0 x 10 using BLASTn) to a Prunus avium actin gene (GenBank accession no. FJ560908) served as an internal standard (Figure 3.4). The differentially expressed clones were sequenced in both directions using the PCR product from the blot screens and the primers T7 and Sp6. The sequences were combined to form contigs using Lasergene software (DNAStar, Madison, WI, USA) and analyzed using BLASTX and TBLASTX (11). For each contig, the nucleotide sequence with highest homology that had an annotated physiological role and an E-value of less than 0.01 using TBLASTX was chosen (Table 3.1). All DNA sequences were submitted to GenBank at the National Center for Biotechnology Information (http://blast.ncbi.nlm.nih.gov/) and grouped into the categories according to their putative physiological function, such as protein synthesis and processing, signal transduction and post-translational regulation, defenserelated, abiotic stimuli and development, hypothetical proteins, and no homology to known proteins (52). The isoelectric point (pI) was calculated for the homologous protein sequence of each putative plant EST using the ProtParam tool at the ExPASy Proteomic Server (http://ca.expasy.org/tools/protparam.html). 63 Semi-quantitative RT-PCR and origin determination Total RNA was extracted from epidermal peels at each of the time points as described above and treated with the DNA-free kit. cDNA was then synthesized using SuperScript III Reverse Transcriptase (Invitrogen) and an oligo DT primer (20 bp). Two independent cDNA synthesis reactions were performed per sample. Eight expressed sequence tags (ESTs) were chosen based on their sequence similarities with genes or proteins related to disease resistance (EST01, EST03, EST04, EST05, EST08, EST10, EST11, and EST13), and primers were constructed for each EST using Primer Express Software (Applied Biosystems, Foster City, CA, USA) (Table 3.2). Primers were optimized by varying primer concentration and annealing temperature. A protocol using 25 cycles, a 58ºC annealing temperature, and 300 nM of each primer was found to be consistently optimal across the different primer sets. Semi-quantitative RT-PCR was carried out to determine the presence or absence of gene transcripts in both cultivars following inoculation. PCR reactions always contained cDNA reverse transcribed from 1.0 µg of total RNA as the template, which was extracted from 100mg of tissue aliquots (~5-6 epidermal peels). The amplification protocol included an initial denaturation step at 94ºC for 2 min followed by 25 cycles at 94°C for 1 min, 58ºC for 1 min, and 72°C for 1 min and 30 sec. The reaction was completed by a 10-min extension at 72ºC. PCR products were separated on a 2.0% agarose TAE gel. Experiments were repeated with two independent biological samples. For all time points, the putative actin protein described above was used as an internal loading standard. Semi-quantitative RT-PCR products were sequenced to ensure that the correct gene was being amplified from the template. 64 Additionally, the DNAse-treated total RNA of each sample was tested for evidence of contaminating genomic DNA by using it as a template in a PCR reaction with the putative actin gene primers. To confirm the plant or fungal origin of each EST, DNA was extracted from fungal mycelium according to (113) and from young leaves of ‘Elliott’ and ‘Jersey’ blueberry plants according to (57). To confirm the presence of genomic DNA, the putative actin gene and the fungal internal transcribed spacer region (ITS) were amplified using PCR with the primers ACT-F/ACT-R (Table 3.2) and ITS1F/ITS4 (71), respectively. Finally, PCR was performed as described above using extracted genomic DNA (using 35 instead of 25 cycles) and appropriate primers for semi-quantitative RT-PCR (Table 3.2). PCR products were visualized on a 2.0% agarose TAE gel. Quantitative RT-PCR For quantitative reverse transcriptase PCR (Q-RT-PCR) the same primers were used as described above in order to validate the previous techniques using two independent biological samples (Table 3.2). Q-RT-PCR was performed to detect differences in transcript levels of the selected ESTs and to look for an induction at particular time points following infection in both cultivars. For each sample, 1.0 µg of treated RNA was converted to cDNA using the protocol described above. Amplification was determined by using SYBR Green PCR Master Mix (Applied Biosystems) in the Applied Biosystems 7900HT sequence detection system (Research Technology Support Facility, Michigan State University, East Lansing, MI, USA). The appropriate primers were used at a concentration of 300 nM under the following conditions: 50°C for 2 min, 95°C for 10 min, and 40 cycles of 95°C for 15 sec 65 followed by 60°C for 1 min. (Table 3.2). PCRs were carried out in triplicate and were conducted at least twice, and the mean of the three replicates was used to generate an average concentration. Using a dilution series, cDNA from a representative sample (‘Elliott’ inoculated, 96 hpi) was used to generate standard curves and threshold values (CT) for each EST of interest. Standard curves were considered reliable if slopes were between -2.9 and 3.5. The quality of the amplification was assessed by dissociation curves, and product size was verified by gel electrophoresis. The concentration mean of each EST was then normalized to the endogenous putative actin control to obtain the normalized value for each gene. Samples were also tested for the presence of contaminating genomic DNA in the treated RNA samples prior to cDNA synthesis using the DNA-free kit mentioned above. H2O2 detection Green and ripe fruit were collected from the same location described above on 2 August 2010 and stored at 4°C. No fungicides were applied in the last 3 weeks prior to harvest. Fruit were inoculated with a 10-µl drop of an aqueous suspension containing 10 6 conidia/mL or water (untreated control) and then placed equidistantly on wire mesh grates in Petri dishes at 100% RH. After 0, 12, 18, 24, and 48 hpi, inoculated peel sections were excised with a cork borer (size 1, 4.5-mm diameter). H2O2 was detected using the Amplex® Red Hydrogen Peroxide/Peroxidase Kit (Invitrogen) as described in (4) with the following exceptions: epidermal peels of ~4.5mm in diameter were used instead of a fruit disk (10 mm diameter and 10 mm deep), and the reaction volume was 1 mL of phosphate buffered saline as opposed to 5 mL. A standard curve plot was used to calculate the H2O2 concentration 66 2 using a 10-fold-dilution series from 1 millimole to 1 nanomole of H2O2 (R = 0.95). Based on this curve, values in the 1-mL reaction ranged between 1 and 754 nanomoles and these 2 values were divided by the total area of the peels (3 peels x 15.9 mm per peel) to calculate 2 the concentration in nanomoles/mL/mm . Three independent experiments were conducted with five replicates per treatment. RESULTS Inoculation of C. acutatum at different stages of fruit development Detached fruit inoculation experiments demonstrated a host resistance phenotype that was previously described by Wharton and Schilder (2008). In the susceptible cultivar Jersey, conidial masses were seen as early as 5 d after inoculation, and at 8 d started to coalesce and cover the entire surface of the fruit (Figure 3.1A, C). Acervulus diameter ranged from 56–224 µm with an average of 126 µm (Figure 3.2A), and an average of 2.65 x 7 10 conidia were produced per berry (Figure 3.2B). In the resistant cultivar Elliott, berries remained firm and acervuli were not visible until 8 d after inoculation (Figure 3.1B). Acervuli were few and constricted, with small orange cirrhi. The average acervulus diameter was less than half of that on ‘Jersey’ fruit: 58 µm with a range of 26–106 µm (Figure 3.1D). 6 About tenfold fewer conidia (2.8 x 10 per berry) were produced in ‘Elliott’ compared to ‘Jersey’ (Figure 3.2B). Examination of pre-penetration activities of C. acutatum on immature fruit of both cultivars showed similar rates of conidium germination and appressorium formation. 67 A B C D Figure 3.1. Symptoms and signs of Colletotrichum acutatum infection of highbush blueberry fruit A) Fruit appearance 8 d after inoculation in the susceptible cultivar Jersey, and B) the resistant cultivar Elliott. Scanning electron micrograph of acervuli on the fruit surface of C) ‘Jersey’ and D) ‘Elliott’. Bar = 0.5 mm. 68 n = 50 A 6 120 35 b 100 80 60 Number of conidia per berry (10 ) Acervulus diameter (micrometers) 140 a 40 20 0 30 25 n=5 B b 20 15 10 5 a 0 Jersey Jersey Elliott Elliott Figure 3.2. A) Average diameter of acervuli of C. acutatum on fruit of the susceptible blueberry cultivar ‘Jersey’ (white bars) and resistant cultivar ‘Elliott’ (black bars) (n = 50). B) Quantity of C. acutatum conidia produced on ‘Jersey’ (open bars) and ‘Elliott’ (closed bars) after 8 d of incubation (n = 5). Error bars denote standard error of the mean. Means with the same letter are not significantly different from each other according to Student’s paired t-test (P < 0.05). 69 Melanized appressoria (%) 100 80 60 Jersey Elliott a b a a a a a a 40 20 0 Petal fall Green fruit Fruit coloring Ripe fruit Fruit development stage Figure 3.3. The percentage of C. acutatum conidia that formed melanized appressoria at 24 h post inoculation on blueberry fruit at different stages of development in the susceptible cultivar Jersey (white bars) and resistant cultivar Elliott (black bars). Error bars denote the standard error of the mean (n = 5). Within each development stage, means with the same letter are not significantly different from each other according to Student’s paired t-test (P < 0.05). 70 However, on ripe fruit, significantly more conidia had germinated and formed melanized appressoria by 24 hours post inoculation (hpi) on ‘Jersey’ than on ‘Elliott’ (Figure 3.3). Since the formation of melanized appressoria indicates attempted penetration of the plant epidermis, the infection process appears to advance more rapidly in ‘Jersey’ than in ‘Elliott’. However, by 48 hpi, 96 to 100% of germinated conidia had produced melanized appressoria regardless of fruit developmental stage or cultivar. Suppression subtractive hybridization of ‘Elliott’ and ‘Jersey’ cDNA libraries We constructed two representative subtracted cDNA libraries of ripe fruit from ‘Elliott’ and ‘Jersey’, each containing five discrete time points (0, 24, 48, 96, 144 hpi) following inoculation with C. acutatum. The pooled subtracted ‘Elliott’ cDNA library was cloned into the vector pGEM-T-Easy, and we screened 1,056 clones against the pooled subtracted ‘Jersey’ library. Using Southern blot analysis, 37 clones were found to be differentially expressed (Figure 3.4). DNA sequence analysis using TBLASTX identified significant homology of these clones with genes producing proteins involved in a variety of cellular processes. The 37 clones were grouped according to their putative physiological function, including defense-related (5 clones), abiotic stimuli and development responses (4 clones), protein synthesis and processing (3 clones), signal transduction and posttranslational regulation (1 clone), and hypothetical proteins (9 clones) (Tables 3.1 and B.1). As was the case in similar studies conducted to date, a large number (41%) of DNA sequences had no significant homology to known genes. Additionally, the theoretical isoelectric point (pI) of the most homologous protein was also calculated for each putative plant EST. A class II chitinase (EST01) and a beta-1,3 glucanase (EST08) were found to be 71 Table 3.1. Characteristics and predicted physiological function of differentially expressed sequence tags (ESTs) from ripe fruit of highbush blueberry cultivars Elliott versus Jersey after inoculation with Colletotrichum acutatum.The hypothetical function is based on homology to sequences in translated nucleotide databases (DDBJ/EMBL/GenBank) using TBLASTX. This table only displays -5 putative plant sequences with homologous E values lower than 1 x 10 . ID code a Size b (bp) Redundancy c Genbank Accession No. Homologous Gene (BLAST Hit Accession Number) Homologous Species Max Score Class II chitinase (AB465728) Beta-1,3-glucanase (X75946) Pathogenesis-related protein 10 (AM489568) Vaccinium corymbosum 353 5.0 x 10 Beta vulgaris 72.1 3.0 x 10 Actinidia deliciosa 71.2 2.0 x 10 Monodehydroascorbate reductase (EU327873) Metallothionein-like (AY857933) Hydroxyproline-rich glycoprotein precursor (U18791) Dihydroflavonol 4reductase (AY221249) Vaccinium corymbosum Gossypium hirsutum 127 7.0 x 10 100 3.0 x 10 Phaseolus vulgaris 40.9 6.0 x 10 Allium cepa 51.5 1.0 x 10 E value pI d Defense-related e EST01* 432 1 GW397252 EST08* 159 1 GW397259 EST10* 239 3 GW397261 -95 8.82 -11 9.37 -10 5.36 -28 5.78 -19 4.83 -5 9.70 -5 5.86 Abiotic stimuli and development EST03* 186 1 GW397254 EST05* 236 1 GW397256 EST07 138 1 GW397258 EST11* 389 1 GW397262 72 Table 3.1 (cont’d). Protein synthesis and processing Small nuclear EST06 223 1 GW397257 ribonucleoprotein E (EU974208) Signal transduction and post-translational regulation Ubiquitin-conjugating EST04* 220 1 GW397255 enzyme (NM_104180) a b 68.9 4.0 x 10 9.03 Arabidopsis thaliana -21 1.0 x 10 6.22 101 -19 Zea mays c d ID code = identification code; bp=number of base pairs; = number of times sequence was recovered; = theoretical isoelectric e point (pI) of the most homologous protein; = * denotes ESTs selected for further study. 73 Table 3.2. Primer DNA sequences, guanine-cytosine (GC) percentages, calculated primer melting temperatures (Tm), expected product size, product melting temperature, Q-RT-PCR standard curve plot slope, and calculated primer efficiency used in semiquantitative and quantitative RT-PCR expression analysis of differentially expressed sequence tags in blueberry fruit in response to C. acutatum inoculation. Primer Name Sequence (5’-3’) EST01-F EST01-R EST03-F EST03-R EST04-F EST04-R EST05-F EST05-R EST08-F EST08-R EST10-F EST10-R EST11-F EST11-R EST13-F EST13-R ACT-F ACT-R ATC CCC GGT AGT TCC AAA AC TGA GGA CTC TGG CAC TCC TT CAG ATT GGA GCT TTT GTT CTT ATG G GGA AGA AAT GCT TAT TCA GCC TAC A GGG CAG ACC TTA CCA CAA TCA TGG CGG TCT TCG AAT AAA CC GAG GTA CCG CAC TTG CAC TT CCA TTC ACA CCC AAG CTA CA TCC AGT CAA GAA GCA GTT GAC C CAT CCT TCC AAT GCC ATT GG ACC CTA CAA TGT CTT TCA CCA ACA CGG CCG AGG TAC TTT CCA C GTG TTC ACA TCG TCT GCT GGA TCG GTT GTT GGT GCT CTT GA ACC AGC CCG TCT TTA GTC CT TGA GCA CGT TGC CTG TTA CT TCA AGA GCC ACG TAT GCA AG TGC CCT CAT GAA GAT CCT TAC Primer GC (%) Primer Tm(°C) 50.0 55.0 40.0 40.0 52.4 50.0 55.0 50.0 50.0 50.0 41.7 63.2 52.0 50.0 55.0 50.0 50.0 47.6 54.8 57.9 59.0 59.0 57.1 55.2 57.7 55.0 57.1 54.3 56.5 58.2 58.0 59.0 58.1 56.9 55.2 54.2 74 Size (number of base pairs) Product Tm (°C) Q-RT-PCR Standard Curve Plot Slope Q-RT-PCR Primer Set Efficiency (%) 206 81.3 -3.30 100.9 98 74.1 -3.32 100.0 50 73.5 -3.53 91.9 225 83.2 -3.38 97.5 50 73.4 -3.21 104.9 50 74.0 -3.12 109.1 52 75.9 -3.24 103.5 197 79.8 -2.95 118.1 105 78.8 -3.47 94.3 1 A 2 3 4 (-) EV 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 1000 500 200 B C Figure 3.4. Suppression subtractive hybridization to detect differential expression of genes in ‘Elliott’ (resistant) compared to ‘Jersey’ (susceptible) blueberry fruit after inoculation with C. acutatum. A) Colony PCR products obtained from forward-subtracted ‘Elliott’ cDNA library with T7 and Sp6 primers used for dot blot analysis and visualized through gel electrophoresis. Lanes 1 and 24, 1-kb+ DNA ladder; lane 2, negative control; lane 3, EV=empty vector; lanes 4 through 23 PCR products from ‘Elliott’ forward-subtracted cDNA library. B) PCR products from the forward ‘Elliott’ library hybridized with the reverse ‘Jersey’ library DIG probe. C) PCR products from the forward ‘Elliott’ library hybridized with the forward ‘Elliott’ DIG probe. Dotted lines surround the actin PCR product which served as an internal standard. 75 fairly basic (pI = 8.82 and 9.37, respectively). However, other homologous proteins were found to be acidic such as metallothionein-like (EST05) protein and pathogenesis-related protein 10 (EST10) (pI = 4.83 and 5.36, respectively). Expression profiling and origin determination of selected ‘Elliott’ clones 5 Eight ESTs with significant homology (E value < 1.0 x 10- ) to known protein and gene sequences in the translated nucleotide database at the National Center for Biotechnology Information were further analyzed (Table 3.1). Specific primers were designed for analysis of temporal expression profiles through semi-quantitative RT-PCR (Table 3.2). Among these ESTs were a class II chitinase (EST01), monodehydroascorbate reductase (EST03), ubiquitinconjugating enzyme (EST04), metallothionein-like protein (EST05), beta-1,3-glucanase (EST08), pathogenesis-related protein 10 (EST10), dihydroflavonol 4-reductase (EST11) and a hypothetical protein from Mus musculus (EST13). EST13 was chosen because it had weak homology to a barley (Hordeum vulgare) cysteine proteinase inhibitor (accession number CAG38123, E value 4.3, maximum identity 27%) using BLASTX, and according to the CATH version 3.3 classification, it structurally contains a domain common in other plant cysteine proteinase inhibitors known as superfamily 3.10.450.10 (E value 1.4 x 10 -05 ) (142). In all cases, EST expression increased at an earlier time point in the host-pathogen interaction in ‘Elliott’ than in ‘Jersey’. Many of the ESTs in ‘Elliott’ were detected as early as 24 hpi, while others were detected at 48 hpi (Figure 3.5A). None of the ESTs was detected prior to inoculation, except dihydroflavonol 4-reductase (EST11), which was constitutively expressed. Also, the initial total RNA preparation after a DNAse treatment yielded no measurable amplification when visualized 76 A Jersey 0 24 48 B Elliott 96 144 0 24 48 96 144 Hours after inoculation EST01 EST01 EST03 EST03 EST04 EST04 EST05 EST05 EST08 EST08 EST10 EST10 EST11 EST11 EST13 EST13 Actin Actin ITS1F – ITS4 Figure 3.5. Temporal pattern of transcript accumulation and origin determination of putative genes in blueberry fruit of the resistant cultivar Elliott and the susceptible cultivar Jersey after inoculation with C. acutatum. A) Transcript accumulation of expressed sequence tags (Table 3.1) at 0, 24, 48, 96, and 144 hours post inoculation in semi-quantitative RT-PCR. B) The origin of the ESTs as determined by PCR using genomic DNA from ‘Elliott’, ‘Jersey’ and C. acutatum. In A and B, experiments were performed with specific primers constructed for each individual EST of interest (Table 3.2). Controls were performed with primers specific for actin (Table 3.2) and the fungal ITS region (ITS1F-ITS4). 77 EST/ actin expression 1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 B A Monodehydroascorbate reductase (EST03) Class II Chitinase (EST01) C D Ubiquitin-conjugating enzyme (EST04) Metallothionein-like (EST05) E F PR10 (EST10) Beta-1,3-glucanase (EST08) G H Dihydroflavonol 4-reductase (EST11) 0 20 40 Jersey-inoculated 60 Hypothetical protein (EST13) 80 100 120 140 0 20 40 Hours after inoculation Jersey-water Figure 3.6. 78 60 80 100 120 140 Elliott-inoculated Elliott-water Figure 3.6 (cont’d.). Ratio of transcript accumulation of expressed sequence tags (ESTs) relative to actin in quantitative RT-PCR at 0, 24, 48, 96, and 144 hours post inoculation of ripe blueberry fruit of a susceptible cultivar Jersey and a resistant cultivar Elliott with C. acutatum or water (control). Mean expression levels were normalized with mean endogenous actin levels, and an induced sample was used to generate the standard curve plots for each individual EST. ∆Rn thresholds varied between 0.15 and 0.75. Error bars denote standard error of the means of the replicates within a representative experiment (n = 3). 79 through gel electrophoresis, indicating that the genomic DNA was eradicated from the samples following the DNAse treatment. Additionally, sequencing of the semi-quantitative RT-PCR products indicated that the designed primers were amplifying their intended targets. The origin (plant or fungal) of the eight ESTs was then determined by PCR using the putative actin and the fungal internal transcribed spacer (ITS) as controls. PCR products for all eight of the ESTs were only seen when using the genomic DNA of the host plant as the template for the PCR reaction (Figure 3.5B). No PCR products were observed in the reaction with fungal genomic DNA, except for the control ITS1F-ITS4 reaction. This indicates that all of the selected ESTs originated from the host plant. Using quantitative RT-PCR, selected ESTs were also detected earlier in ‘Elliott’ than ‘Jersey’ (Figure 3.6). Throughout the experiments, the expression of putative actin remained relatively constant. For this reason the putative actin gene was treated as an endogenous control and the expression of the candidate genes was normalized with it. In general, normalized expression levels began to increase at 24 hpi (EST01, EST03, EST04, EST08, EST10, EST11 and EST13) and 48 hpi (EST05) in ‘Elliott’. If there was expression in ‘Jersey’, it visibly increased at 48 hpi (EST11) and 144 hpi (EST01, EST05, EST10, and EST13). There was virtually no expression of any of the selected ESTs in the water-treated control of either cultivar (Figure 3.6). Oxidative burst in green and ripe fruit after inoculation with C. acutatum Based on the identification and temporal expression of genes putatively involved in the oxidative burst response, we monitored the accumulation of H2O2, an indicator of reactive oxygen species, in inoculated epidermal peels of green and ripe fruit of both cultivars. Measurements were taken at 0, 12, 18, 24, and 48 hours after inoculation with C. acutatum. Overall, H2O2 levels were 80 2 H2O2 (nanomoles/mL/min ) 0.8 0.6 A Green fruit n=5 n=5 12 0.4 8 0.2 0.0 16 B Ripe fruit 4 0 10 20 30 40 50 Hours post inoculation Jersey-water Jersey-inoculated 0 0 10 20 30 40 50 Hours post inoculation Elliott-inoculated Elliott-water Figure 3.7. H2O2 accumulation in epidermal peels of green and ripe fruit of the susceptible blueberry cultivar Jersey and resistant cultivar Elliott inoculated with C. acutatum or water (control) at 0, 12, 18, 24 and 48 hours post inoculation (n=5). H2O2 was detected by the Amplex® Red Hydrogen Peroxide/Peroxidase Kit. Results from a representative experiment out of three independent experiments are displayed. 81 significantly lower in green fruit than in ripe fruit (Figure 3.7). An increase in H2O2 was observed at 18 h and a peak at 24 h after inoculation in both cultivars regardless of fruit maturity stage. However, H2O2 levels were two to three times higher in ‘Elliott’ than in ‘Jersey’. Even in the water control, the initial and maximum H2O2 concentrations were higher in ‘Elliott’ than in ‘Jersey, particularly in green fruit. DISCUSSION Following direct penetration of host tissues, Colletotrichum species have two different host colonization strategies depending on the host or tissue being colonized: intracellular hemibiotrophy or subcuticular intramural necrotrophy,(43, 53, 75, 137, 151). Wharton and Schilder (2008) linked the strategy of subcuticular intramural necrotrophy by C. acutatum on blueberry to host plant resistance and noted defense responses such as the production of amorphous phenolic globules around intracellular hyphae. They also observed a lower rate of conidium germination and appressorium formation on the resistant cultivar Elliott compared to the susceptible cultivar Jersey. However, they did not investigate the infection process on immature fruit. Our results suggest that differences in pre-penetration activities of C. acutatum are confined to ripe fruit, as no differences were observed on immature fruit of the two cultivars. The difference appears to be the result of a relative increase in the rate of appressorium formation on ‘Jersey’ as the fruit ripens, whereas the rate on ‘Elliott’ remains steady across fruit development stages. A possible reason for this phenomenon could be a change in the structure or composition of the waxy cuticle in ‘Jersey’ fruit that stimulates conidium germination and appressorium formation. In avocados, cuticular wax has 82 been shown to trigger conidium germination and appressorium formation of C. gloeosporioides (148). More research is needed to determine the cause of the observed differences. This study represents the first investigation of gene expression in a VacciniumColletotrichum interaction at different stages of the infection process and provides additional evidence for an active resistance response in the blueberry cultivar Elliott. It appears that this cultivar recognizes the pathogen at an earlier time point during infection than the susceptible cultivar Jersey, as expression of defense-related and other genes was detected earlier in ‘Elliott’ than in ‘Jersey’, often within 24 hours after inoculation. Given the scarcity of DNA sequence information for blueberry and Vaccinium spp. in general, the most limiting constraint on these experiments was the identification of homologous sequences. Many of the differentially expressed DNA sequences (41%) found in this study had no homology to known sequences. In addition, pooling of cDNA from multiple time points after inoculation into a single cDNA library for each cultivar for cost-efficient SSH screening may have reduced our ability to detect certain genes that were differentially expressed at specific time points. For example, defense genes expressed late in the infection process in the susceptible cultivar could have masked early over-expression in the resistant cultivar. In addition, cloning the cDNA library of ‘Elliott’ only could have limited our ability to detect specific plant genes that were upregulated in ‘Jersey’ as well as fungal genes that were upregulated in the compatible host-pathogen interaction. Nonetheless, this study provides valuable new insight in the molecular mechanisms underpinning fruit rot resistance in blueberries and also forms a basis for comparing defenseassociated genes of blueberry fruit tissue with those of other plant species. The selected ESTs may be useful in the development of markers for breeding new disease-resistant cultivars and as possible markers for the expression of systemic acquired resistance (SAR). 83 Limited research has been done on the molecular basis of host plant resistance to the fruit rot pathogen C. acutatum. In strawberries, researchers found that several defense-related proteins such as peroxidases and chitinases were induced in infected fruit. Interestingly, they also noticed reduced expression of defense-related genes in the susceptible cultivar as opposed to an induction in the resistant cultivar (31). In addition, expression was found to be tissue-dependent when comparing infected fruits and crowns. We investigated the expression profiles of three putative defense-related genes: a class II chitinase (EST01), beta-1,3-glucanase (EST08), and pathogenesisrelated protein 10 (EST10). These genes have been classically identified as pathogenesis-related proteins, i.e. they are induced upon pathogen inoculation and are also known to be important in the hypersensitive response (190). Furthermore, given the homology of EST08 and EST01 with other known proteins and their induction upon pathogen inoculation, they could also be referenced as PR-2 and PR-3, respectively (190). When these genes are overexpressed in plant tissue, the result is often increased disease resistance. For example, in transgenic tobacco plants, the overexpression of a chitinase gene led to broad resistance against the fungal pathogen Rhizoctonia solani and the bacterial pathogen Pseudomonas syringae pv. tabaci (44). Interestingly, the class II chitinase (EST01) that we identified has also been shown to accumulate in stems of blueberries at low temperatures and appears to be important in cold hardiness (100). Both EST01 and EST08 are homologous to genes coding for basic proteins, i.e. their expression is most likely linked to the ethylene/jasmonic acid pathway, and the proteins are likely not secreted into the apoplast. However, the most homologous protein to EST10 was in fact acidic, suggesting it may be linked to the salicylic acid signaling pathway, and could be secreted into the plant apoplast during infection (190). 84 In the group of ESTs related to abiotic stimuli and development, the dihydroflavonol 4reductase gene (EST11) was expressed constitutively, but expression was boosted by inoculation with C. acutatum in both cultivars, albeit earlier in the infection process in ‘Elliott’ than in ‘Jersey’. This enzyme typically catalyzes the reduction of leucoanthocyanidins to anthocyanidins (133). Previous research has shown an accumulation of anthocyanins in ‘Elliott’ in response to infection by C. acutatum (124). This gene possibly plays a role in the production of anthocyanins and maybe an antifungal compound detected in ‘Elliott’ fruit and capable of inhibiting the growth of C. acutatum in vitro (Miles and Schilder, unpublished). Other identified genes such as monodehydroascorbate reductase (EST03) and a metallothionein-like protein (EST05) are often associated with oxidative stress and protecting the plant from oxidative damage (108, 115). In the C. acutatum-blueberry fruit pathosystem, the timing of the increase in H2O2 is similar to that previously published in the C. coccodes-tomato fruit interaction (122) and seems to correlate well with the formation of melanized appressoria (201), indicating attempted penetration. However, reactive oxygen species may be plant or pathogen derived. If it is plant derived, the H2O2 may be important in the resistance response in ‘Elliott’ by preventing fungal penetration. On the other hand, if the H2O2 is pathogen derived this could indicate preferential necrotrophy of C. acutatum on ‘Elliott’ fruit. Since H2O2 can be a pathogenicity factor for necrotrophic pathogens, it may be important in the initial colonization of ‘Elliott’ fruit. However, because of the relatively short duration of the H2O2 boost and coincident timing with peak appressorium formation, it seems likely that it serves to prevent pathogen ingress. In either case, the above-mentioned genes are likely upregulated in ‘Elliott’ fruit to prevent oxidative damage to plant tissues. 85 The ubiquitin-conjugating enzyme or E2 ligase gene we identified (EST04) was present in both ‘Elliott’ and ‘Jersey’ but appears to be expressed only in ‘Elliott’. The ubiquitin-proteasome system is important in signal transduction in cellular processes including defense responses (181). E2 ligases enzymatically transfer ubiquitin and bind with E3 ligases, which target proteins for proteasomal degradation by ubiquitination. However, there is also evidence that E2 ligase genes can be upregulated in response to elicitors (184), UV light (160), and heat shock (66). The identified E2 ligase may play a role in the resistance response in ‘Elliott’ fruit. Our results clearly show a time delay between inoculation and measurable gene expression. Consequently, there was a significant lag period before a defense response was detectable, which would correlate well with previous microscopy studies outlined in Wharton and Schilder (2008). This could be due to the amount of time required by the host for measurable expression of defenserelated genes or delayed recognition of the invading pathogen by the host. In support of the latter hypothesis, a similar result was observed in sorghum mesocotyls when inoculated with C. sublineolum (118). While the expression of PR-10 was detected as early as 4 h after inoculation with the fungus Cochliobolus heterostrophus, the authors did not record significant expression of PR-10 until 36 h and peak expression at 48 h after inoculation with C. sublineolum (118). Therefore, the lag period in the C. acutatum-blueberry pathosystem is most likely due to the amount of time required by the fungus to gain entry and be recognized by the host. Our results add new insight into the host responses of blueberry fruit to infection by C. acutatum at the molecular level and suggest that pathogen ingress into the host is required for the activation of resistance. A more detailed investigation of gene expression during the early stages of infection, including pre-penetration events, will help to pinpoint when the host first recognizes that it is being attacked by C. acutatum and initiates the resistance response. Furthermore, chemical 86 analysis and studies on the genetic inheritance of resistance will complement molecular research in elucidating the basis of anthracnose fruit rot resistance in blueberries. ACKNOWLEDGEMENTS I would like to gratefully acknowledge funding from Michigan State University Project GREEEN (Generating Research and Extension to meet Economic and Environmental Needs). I would also like to thank Dr. James Hancock and Hortifrut Chile S.A. for providing fruit, Dr. Miaoying Tian for technical training, Christine Bates for technical assistance, and Dr. Paul Goodwin (University of Guelph, Canada) and Dr. Raymond Hammerschmidt for critical review of the manuscript. 87 CHAPTER IV: CHARACTERIZATION AND BIOLOGICAL ACTIVITY OF FLAVONOIDS FROM RIPE FRUIT OF AN ANTHRACNOSE-RESISTANT BLUEBERRY CULTIVAR ABSTRACT Anthracnose fruit rot, caused by Colletotrichum acutatum, is among the most important diseases of blueberries. Most cultivars are susceptible but ‘Elliott’ is resistant. Our objective was to identify possible antifungal compounds that play a role in the resistance response. Chemical fractions from freeze-dried, ripe fruit of ‘Elliott’ and the susceptible cultivar Jersey were extracted with methanol and ethyl acetate. Extracts were screened on solid media for suppression of microconidiation of C. acutatum. The methanolic extract was fractionated and the soluble methanolic fraction from ‘Elliott’ was the most biologically active. This fraction was dried, dissolved in water, and screened in vivo by pre-treating ripe ‘Jersey’ fruit with 0.5, 1, 2, and 4% solutions (w/v) and subsequently inoculating the fruit with C. acutatum. An 88% reduction in infection incidence was observed after 12 days with the 4% solution. Anthocyanins and other flavonoids were then quantified in fruit of the two cultivars using HPLC-MS. ‘Elliott’ fruit contained more anthocyanins (4.87 mg/g of freeze-dried tissue) than ‘Jersey’ (3.27 mg/g of freezedried tissue); however, the same compounds were found in both cultivars. ‘Elliott’ fruit also contained more non-anthocyanin flavonoids (0.18 mg/g of freeze-dried tissue) than ‘Jersey’ (0.12 mg/g of freeze-dried tissue), including two unique compounds in ‘Elliott’. The non-anthocyanin flavonoid fractions of both ‘Elliott’ and ‘Jersey’ significantly decreased the growth of C. acutatum in a liquid bioassay, and the effect was somewhat more pronounced in the ‘Elliott’ fraction. MS/MS analysis of the two unique ‘Elliott’ compounds was also conducted. These results provide 88 new insights into the role of anti-fungal compounds in the resistance response in ripe ‘Elliott’ blueberries. INTRODUCTION Anthracnose fruit rot caused by the fungus Colletotrichum acutatum J.H. Simmonds is a major postharvest disease of blueberries (Vaccinium corymbosum L.). Most blueberry cultivars are susceptible to anthracnose fruit rot but some resistant blueberry cultivars exist (150). The blueberry cultivar Elliott has been consistently identified as resistant to C. acutatum, whereas the cultivar Jersey has been identified as susceptible (84, 150). However, the mechanism of disease resistance is not known. Screening attempts have found no correlation between resistance to fruit and foliar infection (64) or the production of antimicrobial fruit volatiles (149). Host-pathogen interactions have only been well characterized in a few Colletotrichum-fruit pathosystems, including avocado, citrus, and mango (23, 189), and most of the information on resistance mechanisms in fruit comes from studies of C. gloeosporioides on avocado (153, 159). A microscopy study of infection of ripe blueberry fruit by C. acutatum showed different infection strategies depending on the cultivar being colonized. In addition, an accumulation of amorphous phenolic globules was observed around the site of infection in the resistant cultivar (201). More recent studies have identified an oxidative burst following infection (125) and higher sugar content in resistant blueberry cultivars (126). However, the role of antifungal compounds in the resistance response also needs to be investigated. The biochemical composition, particularly anthocyanin content, of ripe blueberry fruit has been investigated from a nutraceutical perspective (90, 95, 97, 152). However, the role of chemical constituents in host plant defense in blueberries to plant pathogens is largely unknown. Several 89 studies have been carried out on the antifungal properties of extracts of ripe fruit from wild highbush blueberry plants as they relate to fruit decay and herbivore preference (37-39). Although these studies did not investigate cultivated blueberries or identify individual compounds, they indicated that the main antifungal compounds present in ripe blueberry fruit were water-soluble phenolics and acids. They also proposed that resistance to fungal decay in ripe blueberries may be due to an interaction between simple phenolic compounds and organic acids and not necessarily individual fungitoxic compounds. Demand for blueberries and other fruit crops continues to rise as the health benefits are becoming increasingly apparent with respect to cardiovascular and neurodegenerative diseases in humans (94, 96, 116, 147). Blueberries are rich in antioxidants in the form of anthocyanins and other phenolic compounds (96). Blueberries contain many different flavonoids: five main classes of anthocyanins (70), three classes of flavonols (80), and one class of flavan-3-ols (10). In fact, anthocyanins (the predominate class of flavonoids in blueberries) have been shown to aid in obesity prevention (91), cardiovascular health (48), act as an anti-inflammatory (196) and have anti-cancer effects (18, 172, 179) . Chemical compounds have been implicated in host plant resistance to Colletotrichum spp. in various crops. For instance, in avocado, antifungal dienes (154) and the flavonoid epicatechin (79) play a role in ontogenic resistance of unripe fruit to infection by C. gloeosporioides. A reduction in these compounds as fruit ripens is correlated with increased susceptibility to infection. Researchers have also identified constitutive alk(en)ylresorcinols in fruit of Colletotrichum-resistant mango varieties (86). In sorghum, 3-deoxyanthocyanidin accumulated in response to infection by C. sublineolum and this compound exhibited higher fungitoxicity than other phytoalexin components (117). Based on microscopic observations in blueberry and chemical resistance mechanisms in 90 other fruit crops mentioned above, the potential role of chemical constituents in the defense against anthracnose fruit rot in blueberry merits investigation as well. Therefore, the objectives of this study were to: (1) assess chemical extracts of blueberry fruit for biological activity against C. acutatum, and (2) identify and quantify specific compounds in the active fractions. MATERIALS AND METHODS Plant and fungal material Ripe fruit of blueberry (Vaccinium corymbosum L.) cultivars Elliott and Jersey were harvested in August 2010 from mature bushes at the Michigan State University Southwest Michigan Research and Extension Center in Benton Harbor, MI, USA and a commercial field in Traverse City, MI, USA. From each cultivar, a composite sample of approximately 10 kg of blueberries was harvested and transported to the laboratory. For analysis, 0.5 kg of fruit was stored overnight at -20°C and freeze-dried. Freeze-dried material was stored at -20°C until chemical extraction. A single-conidium isolate of Colletotrichum acutatum (#0001) from blueberry fruit collected in Grand Junction, MI, USA in August of 2006 was used for all experiments. The isolate was stored as conidia, and cultured in accordance with previous studies (125, 201). For inoculum production, sporulating cultures were flooded with 3 ml of sterile deionized water (SDW), and conidia were dislodged using a sterilized L-shaped glass rod. Conidia were counted using a 6 hemacytometer, and a concentration of 10 conidia/ml was achieved via dilution with SDW and used for all experiments. 91 Chemical extraction and fractionation Initially, fruits of ‘Elliott’ and ‘Jersey’ (prior to inoculation and 4 days post inoculation) were homogenized for 5 min with a Sorvall tissue homogenizer (Thermo Fisher Scientific, Waltham, MA, USA) in 80% acidified methanol (pH = 2.0) and centrifuged at 5,000 x g. The extraction was dried using a rotory evaporator, and the remaining water portion was lyophilized. The lyophilized material was dissolved with acidified methanol to a specific concentration (5 g fresh wt./ml). Insoluble material was then further extracted with ethyl acetate and centrifuged at 5,000 x g. This extraction was then dried down and dissolved in acidified methanol as described above. Methanolic extracts (5 mg of fresh weight tissue per μl) from ripe fruits of ‘Elliott’ and ‘Jersey’ before and 4 days after inoculation with C. acutatum conidia were loaded onto a 250-µm cellulose thin layer chromatography (TLC) plate (Analtech Inc., Newark, DE) (50 µl per sample) and compounds were separated with butanol:acetic acid:water (4:1:5 top phase). Plates were allowed to dry overnight and then sprayed with an aqueous conidial suspension of C. acutatum 6 (10 conidia/ml). Plates were incubated at room temperature (22 to 24°C) for 60 hours in a 100% relative humidity chamber and subsequently stained using the gas phase created from iodine crystals. Areas of inhibition of fungal growth were measured. On a replicate plate, those areas were removed, dissolved in methanol, and scanned with a UV/Vis spectrophotometer (200 to 600 nm) (Figure C.1). Since no significant differences were observed between inoculated and uninoculated ‘Elliott’ extracts, further extractions were conducted on fresh, non-inoculated fruit. An exhaustive extraction with water, methanol, and ethyl acetate from fresh whole fruits was used (Figure C.2). As described above fresh tissue was homogenized for 5 min with a Sorvall tissue homogenizer in water (pH = 2.0) and centrifuged at 5,000 x g. The residue was then further extracted in methanol 92 and centrifuged at 5,000 x g. Then the residue was further extracted with ethyl acetate and centrifuged at 5,000 x g. All fractions were dried using a rotory evaporator or lyophilized and stored at -20°C for bioassays. Initial screenings on PDA plates (described below) indicated that the methanolic extract had the most biological activity. Then, a second extraction with methanol was used on lyophilized ripe fruits of ‘Elliott’ and ‘Jersey’ (Figure C.3). Residue from the second extraction was further extracted with ethyl acetate. Methanol-soluble and -insoluble fractions were obtained by stirring the dried methanolic extract in methanol. Anthocyanin and non-anthocyanin flavonoid fractions were obtained from the methanolic extract by standard laboratory protocols (163). Antimicrobial screening of extracts and in vivo activity of the methanol-soluble fractions 6 For the bioassay, 400-µl aliquots of an aqueous suspension of 10 conidia/ml were applied to the entire surface of full-strength potato dextrose agar plates and allowed to dry. Extracts and fractions were screened for antimicrobial activity by applying extracts dissolved in dimethyl sulfoxide (DMSO) directly onto the media (10-µl droplets) at various concentrations (1000, 500, 250, 125, 63, 31, and 16 µg). Plates were incubated without parafilm for 48 hours at 25°C in the dark and monitored for the inhibition of C. acutatum microconidiation. Extracts or fractions were considered biologically active if microconidiation was absent. Each extract or fraction was tested at least twice and all experiments had at least 2 replicates. For in vivo inhibition, the methanolsoluble fraction was dissolved in water at varying concentrations and ‘Jersey’ fruits were pretreated by immersing fruit in the extract prior to inoculation with C. acutatum using five 93 replicates and 10 fruits per replicate. Fruits were incubated at 100% relative humidity, and disease incidence was rated at 12 days post inoculation Identification and quantification of flavonoids Anthocyanins and non-anthocyanin flavonoids were separated on a 150 x 4.6-mm Symmetry Shield C18 column with a 5-µm particle size (Waters Corporation, Milford, MA, USA) at a flow rate of 0.5 ml/min and a column temperature of 25°C using a 2695 Separator Module, with a photodiode array (PDA) detector (#2996) and a Micromass ZQ mass spectrometer (Waters Corporation). To obtain anthocyanins, compounds were eluted with a gradient of 10% formate and 1% formic acid in acetonitrile at a flow rate of 0.5 ml/min over a 55-minute period with a gradient from 5% to 35% acetonitrile. The PDA and MS detectors scanned ions (250-750 m/z in positive ion mode) at 200- to 550-nm wavelengths in order to identify compounds. Compounds were identified according to Additionally, published research on retention times, masses, and absortion spectra was used to identify many of the compounds (152). Non-anthocyanin flavonoid compounds were eluted with a gradient of 1% formate and 1% formic acid in acetonitrile at a flow rate of 0.5 ml/min over a 40-minute period with a gradient from 10% to 35% acetonitrile. The PDA detector scanned ions (250 to 750 m/z in positive ion mode with scan time of 0.3 seconds, an interscan delay of 0.1 seconds and a source temperature of 140°C) at 200 to 550 nm wavelengths and in order to identify compounds The standards quercetin-3-O-rhamnoside (Sigma-Aldrich, St Louis, MO, USA) and quercetin-3-O-glucoside (Sigma-Aldrich) were included. To quantify anthocyanins and non-anthocyanin flavonoids, 5 g of freeze-dried blueberries per cultivar were extracted separately and standardized to yield equal volumes of extract per gram of fresh weight (FW) of fruit. For non-anthocyanin flavonoids, anthocyanins were removed prior 94 to quantification using a SepPak C18 cartridge (Waters Corporation) according to standard protocols (163). Each extract was then analyzed with duplicate injections. The mean peak areas from the duplicate analyses were used to read the concentration of anthocyanin from the standard curve. Data were collected for five replicate samples per cultivar and averaged to determine the quantity of different compounds. For anthocyanins, a wavelength of 520 nm and a standard curve plot of cyanidin 3-O-glucoside (Polyphenols Laboratories, Sandnes, Norway) were used. For nonanthocyanin flavonoids, a wavelength of 255 nm and a standard curve plot of quercetin-3-Orhamnoside were used. The identities of non-anthocyanin flavonoid compounds 08F and 09F (Table 4.2) were further confirmed by using MS/MS. Analysis was accomplished using a QTRAP 3200 mass spectrometer (Applied Biosystems/MDS Sciex, Foster City, CA, USA) coupled to a UFLC LC20AD system (Shimadzu Corporation, Kyoto, Japan). The mass spectrometer was operated in the positive ion mode with a TurboIon Spray source. A daughter ion scan was used using the parent ions 449 (compound 08F) and 493 (compound 09F) with an initial ionization of 20 V and a collision voltage of 30 V. The other ionization parameters were as follows: curtain gas (psi), 10; ion source gas 1 12; ion source gas 2 30; source temperature 400°C; entrance potential 10 V; collision-activated dissociation high; ion spray voltage 5,500 V. The mass spectrometer and the HPLC system were controlled by Analyst 1.4.2 software (Applied Biosystems/MDS Sciex). Liquid bioassay of specific extracts Minimal media consisting of 0.7% KH2PO4, 0.4% KNO3, 0.3% Na2HPO4, 0.1% MgSO4, 0.03% CaCl2, 1.5% BACTO-agar, and 4% D-glucose and 4% D-fructose (5) was used to grow C. 95 acutatum in the presence of various extracts and fractions (50 μg per 100 μl of culture). Extracts were dissolved in methanol and applied to wells in 96-well culture plates and allowed to dry. 6 Minimal medium (90 μl) was then applied followed by an addition of 10 µl of 10 washed conidia/ml. Conidia were washed by centrifuging at 500 x g for 5 minutes and re-suspending the pellet with an equal volume of water 3 times. Cultures were incubated at 25°C in the dark. Growth was quantified daily using the change in optical density (λ 590 nm) from day 0. Light absorbance was read with an EL 800 Universal Microplate Reader (BioTek Instruments, Winooski, VT, USA). Experiments were repeated twice with three replicate wells per treatment. RESULTS AND DISCUSSION Extracts and fractions from the resistant cultivar Elliott suppress microconidiation and blueberry fruit infection by Colletotrichum acutatum Extracts and fractions from the resistant cultivar Elliott were more effective at inhibiting microconidiation of C. acutatum on agar plates than the susceptible cultivar Jersey. In particular, the methanol, methanol-soluble, and non-anthocyanin fractions were the most biologically active (Table C.1). The non-anthocyanin flavonoid fraction of ‘Elliott’ was the most effective at inhibiting microconidiation (activity observed at an amount of 125 µg when applied directly to a microconidiating C. acutatum culture). In the ‘Elliott’ anthocyanin fraction, activity was only observed at 1000 µg and no activity was observed in the ‘Jersey’ anthocyanin fraction. No activity was observed in the ethyl acetate extract of either cultivar (Table C.1). Since, the most active fraction was the methanolic extract of ‘Elliott’ and higher activity was observed in the nonanthocyanin fraction, this indicates that antifungal compounds were soluble in methanol and were able to be eluted from a SepPak column with ethyl acetate and are therefore non-polar in nature. 96 80 Infected fruit (%) 60 40 20 0 Treatments Figure 4.1. Anthracnose infection incidence in fruit of the susceptible cultivar Jersey pretreated with methanol-soluble fractions from ‘Jersey’ and ‘Elliott’ followed by inoculation with Colletotrichum acutatum. Bars denote the standard error of the mean (n = 5). From left to right: Negative control = uninoculated fruit with no extract applied, Positive control = inoculated fruit with no extract applied, J.M.S. = ‘Jersey’ methanol-soluble fraction, and E.M.S. = ‘Elliott’ methanol-soluble fraction. 97 Non-anthocyanin flavonoids appear to be the primary antimicrobial compounds because the most active fractions always contained non-anthocyanin flavonoids. Previous research has demonstrated the antimicrobial activity of non-anthocyanin flavonoids against Colletotrichum species, including methoxylated flavones (6), isoflavonoids (60), and 3-deoxyanthocyanidins (117). An 88% reduction in anthracnose infection incidence in ‘Jersey’ fruits was observed when they were pretreated with a 4% solution of the methanol soluble fraction from ‘Elliott’ (containing mainly anthocyanins and non-anthocyanin flavonoids) (Figure 4.1). A dose response was evident. Activity was also observed using ‘Jersey’ fractions but it was not as distinct as ‘Elliott’ and the dose response was not as distinct. This showed that the extracted compounds also had antifungal activity in planta. The resistant cultivar Elliott contains more anthocyanins and non-anthocyanin flavonoids than the susceptible cultivar Jersey Overall, ‘Elliott’ fruits contained 4.87 mg anthocyanins per gram of freeze-dried tissue of and ‘Jersey’ fruits contained 3.27 mg of anthocyanins per gram of freeze-dried tissue. However, ‘Jersey’ had a more diverse anthocyanin profile than ‘Elliott’ (Figure 4.2, Table 4.1). Additionally, anthocyanidins in ‘Elliott’ tended to be glycosylated with arabinose and galactose, and anthocyanidins in ‘Jersey’ were more commonly glycosylated with glucose. The ratio of the different anthocyanins to themselves is similar in ‘Jersey’ to other cultivars including; Coville (98), Blueray (90), and the Vaccinium ashei Rabbiteye cultivar Tifblue (152), whereas 57% of the anthocyanins in ‘Elliott’ comprised two compounds: malvidin-3-O-galactoside and malvidin-3-Oarabinoside. No acylated anthocyanins were observed in either cultivar. 98 λ = 520 nm 0.8 AU 0.6 0.4 0.2 A 0.0 0.8 AU 0.6 0.4 0.2 0.0 0 B 5 10 15 20 25 30 Minutes 35 40 45 50 55 Figure. 4.2. HPLC chromatograms of anthocyanins from the anthracnose-resistant blueberry cultivar Elliott (A) and susceptible cultivar Jersey (B) using 10-μl injections at λ = 520 nm. 99 60 Table 4.1. Anthocyanins from the resistant blueberry cultivar Elliott and the susceptible blueberry cultivar Jersey identified in HPLC/MS analysis and quantified using HPLC/PDA. Compounds were putatively identified based on previous research, available spectra, standards, retention times and m/z ratios. Anthocyanins were quantified at a wavelength of 520 nm using a standard curve plot of cyanidin 3-O-glucoside. Compound Compound a b ID Name RT (M+H) m/z (Total, aglycone) c d Elliott ± SE (µg/g) d Jersey ± SE (µg/g) 01A Delphinidin 3O-galactoside 15.5 465, 303 583.45 ± 83.83 346.10 ± 64.29 02A Delphinidin 3O-glucoside 18.2 465, 303 31.13 ± 27.27 201.67 ± 45.08 03A Cyanidin 3-Ogalactoside 20.1 449, 287 106.77 ± 32.64 68.68 ± 28.46 04A Delphinidin 3O-arabinoside 22.0 435, 303 406.36 ± 66.87 300.88 ± 56.78 05A Cyanidin 3-Oglucoside 23.6 449, 287 24.56 ± 24.56 42.63 ± 27.61 06A Petunidin 3-Ogalactoside 24.4 479, 317 443.85 ± 111.96 209.08 ± 44.86 27.5 419, 287 65.94 ± 28.87 198.36 ± 40.08 27.5 479, 317 ++ ++ 07A 08A Cyanidin 3-Oe arabinoside Petunidin 3-Oe glucoside 09A Peonidin 3-Ogalactoside 29.8 463, 301 64.54 ± 29.19 37.05 ± 26.34 10A Petunidin 3-Oarabinoside 32.2 449, 317 256.31 ± 47.46 163.53 ± 40.61 100 Table 4.1 (con’t.). 11A 12A Malvidin 3-Of galactoside Peonidin 3-Oglucoside f 33.5 493, 331 1330.57 ± 153.76 452.26 ± 68.51 33.5 463, 301 ++ ++ 13A Malvidin 3-Oglucoside 37.0 493, 331 71.90 ± 30.64 467.01 ± 67.37 14A Peonidin 3-Oarabinoside 38.3 433, 301 28.87 ± 28.87 24.56 ± 24.56 15A Malvidin 3-Oarabinoside 39.5 463, 301 1454.57 ± 240.91 763.07 ± 99.66 4868.82 3274.87 Total a Compounds are listed in elution order, and A denotes an anthocyanin. b c RT = retention time. d e Compounds were identified based on UV/Vis spectra, m/z ratios, and previous research. Data are expressed as µg per gram of lyophilized tissue. Fruit tissue contained 79 to 82% water. Cyanidin-3-O-arabinoside (419/287) coelutes with petunidin-3-O-glucoside (479/317) (27.5 min). Petunidin-3-O-glucoside predominates. f Malvidin-3-O-galactoside (493/331) coelutes with peonidin-3-O-glucoside (463/301) (33.5 min). Malvidin-3-O-galactoside predominates. 101 0.08 λ = 255 nm AU 0.06 0.04 0.02 0.00 A 0.08 AU 0.06 0.04 0.02 0.00 B 20 22 24 26 28 30 Minutes 32 34 36 38 40 Figure 4.3. HPLC chromatograms of non-anthocyanin flavonoids from the anthracnose-resistant blueberry cultivar Elliott (A) and susceptible cultivar Jersey (B) using 10-μl injections at λ 255 nm. Arrows denote compounds selected for further analysis. 102 Table 4.2. Non-anthocyanin flavonoids from the resistant blueberry cultivar Elliott and the susceptible blueberry cultivar Jersey identified in HPLC/MS analysis and quantified using HPLC/PDA. Compounds were putatively identified based on previous research, available spectra, standards, retention times and m/z ratios. Non-anthocyanin flavonoids were quantified at a wavelength of 255 nm using a standard curve plot of quercetin 3-O-rhamnoside. d d RT (M+H) m/z (Total, aglycone) Myricetin rutinoside 23.9 627, 319 3.93 ± 0.29 9.72 ± 0.34 02F Quercetin rutinoside 26.1 611, 303 10.52 ± 0.41 5.38 ± 0.73 03F Myricetin pentoside 27.0 451, 319 0.29 ± 0.69 0.27 ± 0.88 04F Myricetin methyl pentoside 27.6 465, 319 10.61 ± 0.87 5.39 ± 0.64 05F Quercetin hexoside 28.1 465, 303 16.19 ± 2.29 56.32 ± 10.39 05F Quercetin 3-O-glucoside 28.3 465, 303 5.18 ± 0.10 0.45 ± 0.74 06F Quercetin pentoside 30.9 435, 303 14.53 ± 0.90 28.28 ± 0. 01 07F Quercetin pentoside 31.8 435, 303 30.78 ± 3.25 12.28 ± 0.31 08F Quercetin 3-Orhamnoside 32.7 449, 303 61.28 ± 7.70 2.11 ± 1.50 09F Dimethylmyricetin methyl pentoside 36.5 493, 347 26.54 ± 2.79 1.06 ± 0.74 179.86 123.28 Compound a ID Compound 01F b c Total a Jersey ± SE (µg/g) Compounds are listed in elution order, and F denotes a non-anthocyanin flavonoid. b c Elliott ± SE (µg/g) Compounds were identified based on UV/Vis spectra, m/z ratios, standards and previous research. RT = retention time. d Data is expressed as µg per gram of lyophilized tissue. Fruit tissue contained 79 to 82% water. 103 Furthermore, ‘Elliott’ fruits contained more non-anthocyanin flavonoids (0.18 mg/g of freeze-dried tissue) than ‘Jersey’ (0.12 mg/g of freeze-dried tissue) (Table 4.2). ‘Elliott’ also had a more diverse chemical profile than ‘Jersey’, and the majority of non-anthocyanin flavonoids in ‘Jersey’ were made up of compound 05F, likely a quercetin hexoside (Figure 4.3). Some specific non-anthocyanin flavonoids were almost completely exclusive to ‘Elliott’, including compound 08F (61 µg/g of freeze-dried tissue in ‘Elliott’ versus 2 µg/g of freeze-dried tissue in ‘Jersey’) and compound 09F (27 µg/g of freeze-dried tissue in ‘Elliott’ and 1 µg/g of freeze-dried tissue in ‘Jersey’). These two compounds were further characterized using MS/MS. Daughter ions for compound 08F (parent M+H ion 449) included 303, 147 and 129, indicating an aglycone mass consistent with quercetin and glycosylated with a methyl pentose (Figure 4.4A). Using a standard and additional UV/Vis information (peaks at 256 and a shoulder at 350 nm) this compound is likely quercetin-3-O-rhamnoside (Figure 4.4B). Daughter ions for compound 09F (parent M+H ion 493) included 347, 147 and 129, indicating an aglycone mass consistent with a dimethylmyricetin and glycosylated with a methyl pentose (Figure 4.4A). Additional UV/Vis information yielded peaks at 256 and a shoulder at 350 nm and taken with them MS data this compound is likely dimethylmyricetin methyl pentoside (Figure 4.4B). Dimethylmyricetin methyl pentoside or plausibly Syringetin-rhamnoside which has been reported in Bog bilberries (Vaccinium uliginosum L.) (109). Given that the two unique compounds in the resistant blueberry cultivar contain a methylated sugar and the latter compound is methoxylated twice, the chemical contribution to the resistance response in ‘Elliott’ fruit might be due to the increased fungitoxicity of specific compounds. Researchers have found that methoxylated flavones can inhibit the growth of Colletotrichum gloeosporioides from citrus fruits (6). Researchers have also found that 104 303 100 AU % 0.03 129 0.01 147 A 0 0.02 0.00 B 347 100 AU % 0.02 129 147 C 0 100 200 D 300 400 500 m/z [M+H] 600 0.01 0.00 250 350 450 Wavelength (nm) 550 Figure 4.4. MS/MS and UV/Vis spectra of two flavonoid compounds unique to the anthracnoseresistant cultivar Elliott. (A) Daughter ions detected when selecting for the parent ion of quercetin 3-O-rhamnoside (compound 08F in Table 4.2). (B) UV/Vis spectrum of quercetin 3-O-rhamnoside (compound 08F in Table 4.2). (C) Daughter ions detected when selecting for the parent ion of dimethylmyricetin methyl pentoside (compound 09F in Table 4.2). (D) UV/Vis spectrum of dimethylmyricetin methyl pentoside (compound 09F in Table 4.2). 105 methyoxylation may also be important in imparting a certain degree of lipophilicity to the flavonoid and may be required for the compounds to be biologically active (197). Non-anthocyanin flavonoids suppress the growth of Colletotrichum acutatum in a liquid medium In our liquid assay, methanolic extracts of ‘Elliott’ and ‘Jersey’ inhibited the hyphal growth of C. acutatum at the concentrations tested (50 µg per 100 μl of culture); however, the reduction in growth was more pronounced with the ‘Elliott’ methanolic extract. The anthocyanin standard cyanidin-3-O-glucoside increased the overall growth of C. acutatum. Additionally, anthocyanin fractions from which non-anthocyanin flavonoids were removed also increased the growth of C. acutatum (Figure 4.5A). This indicates that anthocyanins have little antifungal activity on the growth of C. acutatum, while they are significantly more abundant in both ‘Elliott’ and ‘Jersey’ tissues than other flavonoids, they do not seem to play a direct role in the resistance response. However, ‘Elliott’ fruits are known to produce an oxidative burst of H2O2 that is more than two times greater than in ‘Jersey’ fruits following the inoculation of C. acutatum (125). Therefore, the main contribution of anthocyanins might be to scavenge reactive oxygen species following oxidative burst in order to prevent damage to host tissues. The standards quercetin-3-O-glucoside and quercetin-3-O-rhamnoside decreased the overall growth of C. acutatum, and the non-anthocyanin flavonoid fractions of ‘Elliott’ and ‘Jersey’ significantly decreased the growth of C. acutatum; the effect was again slightly more pronounced in the ‘Elliott’ fraction (Figure 4.5B). This result matches our initial biological activity screenings and demonstrates that the non-anthocyanin flavonoids have the most significant effect on the growth of C. acutatum. Other researchers have found quercetin-3-O-rhamnoside to be the 106 ∆ Optical density (590 nm) 1 A 0.4 Positive control Cy-gluc E.M.E. J.M.E. E.A.F. J.A.F. 0.5 0 Positive control Q-gluc Q-rham E.M.E. J.M.E. E.N.A.F. J.N.A.F. 0.2 0 2 4 6 B 0 0 2 4 6 Time (days) Time (days) Figure 4.5. Change in the optical density (λ 590 nm) of liquid medium inoculated with conidia of Colletotrichum acutatum over time in the presence of various fractions (50 μg per 100 μl of culture) from ripe fruit of the anthracnose-resistant blueberry cultivar Elliott and susceptible cultivar Jersey. (A) Anthocyanin-containing extracts and fractions from ‘Elliott ‘and ‘Jersey’ as well as the standard cyanidin 3-O-glucoside. (B) Flavonol-containing extracts and fractions from ‘Elliott’ and ‘Jersey’ as well as the standards quercetin 3-O-glucoside and quercetin 3-Orhamnoside. The positive control denotes the fungus alone in the medium. Bars denote the standard error of the mean (n = 3). Note , Postive control = methanol only, Cy-gluc = cyanidin 3glucoside, Q-gluc = quercetin-3-O-glucoside, Q-rham = quercetin3-O-rhamnoside, J.M.E. = ‘Jersey’ methanol extract, E.M.E. = ‘Elliott’ methanol extract, E.A.F. = ‘Elliott’ anthocyanin fraction, J.A.F. = ‘Jersey’ anthocyanin fraction, E.N.A.F. = ‘Elliott’ non-anthocyanin flavonoid fraction, and J.N.A.F = ‘Jersey’ non-anthocyanin flavonoid fraction. 107 most antimicrobial when compared against other glycosides; interestingly this compound is almost entirely unique to ‘Elliott’ and may be important in the resistance response (195). Most of the compounds we identified in these fractions were flavonol glycosides; however, when exposed to water, flavonoids can oxidize and become deglycosylated. In onions, levels of quercetin glucosides in dried brown areas were much lower than of the quercetin glucosides in fleshy areas, whereas levels of the aglycone quercetin were high in dried brown areas. This suggests that quercetin was formed by deglucosidation and oxidation of quercetin glucosides (182). Deglycosylation tends to increase the activity of flavonoids (208), and further oxidation can even lead to the creation of quinones (143) which are antimicrobial against Colletotrichum species (121). CONCLUSION The resistant blueberry cultivar Elliott has more anthocyanins and non-anthocyanin flavonoids than the susceptible cultivar Jersey, and there are two almost unique flavonoid compounds present. Anthocyanins do not seem to play a direct role in the resistance response but may play an indirect one by protecting host tissues from oxidative damage. The non-anthocyanin flavonoid fraction from the resistant cultivar ‘Elliott’ seems to play a key role in suppressing microconidiation and reducing growth of C. acutatum. Interestingly, this fraction contains two unique compounds; quercetin-3-O-rhamnoside, and another flavonol glycoside which may play a more specific role in the resistance response because of increased biological activity. Non-anthocyanin flavonoids likely complement other aspects of 'Elliott' fruit rot resistance that have been previously described, such as higher levels of sugar content (126) and an oxidative 108 burst following initial fungal penetration (125). Further investigation into how this process is regulated at the molecular level might provide new insight into the interaction. ACKNOWLEDGEMENTS I would like to gratefully acknowledge funding from MSU Project GREEEN (Generating Research and Extension to meet Economic and Environmental Needs). I also would like to thank Dr. Ray Hammerschmidt for help with developing the liquid bioassay, Dr. Christine Vandervoort for helping with the chemical separation using HPLC, and Dr. Daniel Jones for helping with the MS/MS analysis. 109 CHAPTER V: EVALUATION OF SCREENING METHODS AND FRUIT COMPOSITION IN RELATION TO ANTHRACNOSE FRUIT ROT RESISTANCE IN BLUEBERRIES Published in the peer-reviewed journal Plant Pathology (Early View - Doi: 10.1111/j.13653059.2011.02541.x) ABSTRACT Anthracnose fruit rot is an important disease of blueberries and losses are common in humid growing regions. Most commercial cultivars are susceptible and the disease is usually managed with fungicides. However, a few cultivars are considered resistant. The objectives of this study were to: (1) Compare different inoculation techniques for anthracnose fruit rot resistance screening, (2) Screen ripe fruit from a range of blueberry cultivars using selected techniques, and (3) Investigate the role of fruit characteristics in anthracnose fruit rot resistance. The following inoculation methods were evaluated on ripe fruit of a susceptible and resistant cultivar using a conidial suspension: spray, droplet, and injection inoculation of whole fruit; and droplet inoculation of the open surface of cut fruit. All whole-fruit inoculations yielded similar results. Despite the removal of the epidermis, resistance was also expressed in cut fruit but relatively fewer conidia were produced. The cut-fruit assay required substantially less time and half the amount of fruit to accomplish than whole-fruit assays. Detached ripe fruit from 24 cultivars in 2008 and 26 cultivars in 2009 were screened for resistance. Results from the cut-fruit assay correlated best with published resistance ratings. To determine the possible role of fruit characteristics in resistance, fruit pH, titratable acidity, sugar content, and firmness were regressed against various fruit rot resistance measures. Fruit rot resistance was positively correlated with sugar content. On defined 110 media, mycelial growth was restricted as sugar concentration increased and pH decreased, suggesting that fruit composition may play a role in the resistance phenotype. INTRODUCTION Anthracnose fruit rot, caused by the fungus Colletotrichum acutatum J.H. Simmonds, is a major disease of highbush blueberries (Vaccinium corymbosum L.) in humid growing regions. The disease is characterized by sunken areas on infected fruit that become covered by salmon-colored conidial masses. Infections may occur as early as fruit set and remain latent until fruit ripening (128). Most commercial cultivars are susceptible, and the disease is usually managed with fungicides (123). The use of resistant cultivars with desirable horticultural characteristics would be a cost-effective and environmentally safe way of managing fruit rots. One of the difficulties in breeding for anthracnose fruit rot resistance is the labor and amount of fruit required for sufficient screening, which requires natural infection or artificial inoculation of fruit. Through a screening effort that took more than ten years to complete, anthracnose fruit rot resistance profiles have been generated for the majority of older blueberry cultivars (150). In that study, the authors artificially inoculated immature fruit on potted plants with multiple strains of C. acutatum and rated disease incidence once the fruit ripened. Many of the newer cultivars were not included but were screened for resistance to post-harvest fruit rot in another study that relied on natural field infection rather than inoculation (84). Fruit rot resistance may be manifested in two ways, namely reduced infection incidence (150) and reduced symptom severity or sporulation (201). Not much is known about the mechanism of anthracnose fruit rot resistance in blueberries. In previous studies, resistance to fruit infection by C. acutatum was not correlated with resistance to foliar infection (64) or with the 111 production of antimicrobial fruit volatiles (149). Recent studies have shown an active resistance response in ripe fruit of cv. Elliott leading to a restriction of fungal growth in fruit epidermal and subepidermal tissues (201). An accumulation of phenolic compounds (201), an oxidative burst following infection (125), and the presence of putative antifungal compounds (124) are thought to play a role in anthracnose fruit rot resistance in cv. Elliott. In our initial studies, ‘Elliot’ had a lower sugar content and higher pH than the susceptible cultivar Jersey, and we hypothesized that fruit composition may play a role in resistance, either directly by affecting fungal growth or indirectly by modulating fungal or plant enzyme activity (124). Fruit composition, i.e. sugar content, pH and titratable acidity, is an important determinant of flavor and sensory quality of blueberry fruit (42, 84). Ripe blueberries typically contain between 10% and 14% sugar (48% D-glucose, 49% D-fructose, and 3% sucrose) by fresh weight and have a pH of 2.5 to 3.5, depending on the cultivar and study (42, 84, 120). A better understanding of how these traits affect resistance to C. acutatum will provide further information about the nature of the resistance response and aid in the development of resistant cultivars. Although disease resistance is an objective of many blueberry breeding programs, no genotypic or phenotypic markers are currently available for rapid screening of blueberry breeding lines for anthracnose fruit rot resistance. Resistance screening of plants can only be accomplished when they bear fruit in sufficient quantities for evaluation, usually at 2 to 3 years of age. Relying on natural field infection is less labor intensive than using artificial inoculation but would require observations over multiple locations and/or years due to variability between sites and growing seasons. Furthermore, it would be difficult to compare cultivars directly due to widely varying fruit maturation dates, chilling requirements, and cold hardiness levels which prevent some cultivars from being grown in certain locations. Unless anthracnose fruit rot resistance is a specific priority 112 of a breeding program, new cultivars may not be evaluated for anthracnose fruit rot resistance prior to their release. The infection process of Colletotrichum spp. on fruits has been studied in a number of plant pathosystems (36, 131, 159, 206). In general, as fruits start to ripen they become increasingly susceptible to infection. During fruit ripening, many physiological changes occur, such as a reduction in fruit firmness, changes in pH and cell wall composition, and an increase in soluble sugars and secondary metabolites, such as anthocyanins (20, 164). In avocado, several factors have been associated with increased fruit susceptibility to infection by Colletotrichum gloeosporioides (Penz.) Penz. & Sacc. as fruit ripens, including an increase in fruit pH (159) and a decrease in preformed antimicrobial compounds (155) and pathogenicity factor inhibitors, such as epicatechin (79). Other studies have shown that ammonia secretion by the fungus increases the pH of host tissues and is important for pathogenicity on almond (55), apple (157), avocado (157) and tomato (3, 157). Soluble sugars may also play a role in defense responses during ripening. Guava cultivars that contained high levels of soluble sugars and ascorbic acid also were the most resistant to Glomerella cingulata (Stoneman) Spauld. & H. Schrenk (176). In grapes, the accumulation of antifungal proteins and sugars during fruit ripening is an important defense mechanism against the fungal pathogens Botrytis cinerea Pers.:Fr. and Guignardia bidwellii (Ellis) Viala & Ravaz (165, 185). In blueberries, resistance to anthracnose fruit rot is expressed in green as well as in ripe fruit (125, 201). It is not known whether the same mechanism governs resistance at both developmental stages. Since immature fruit of all blueberry cultivars is resistant to anthracnose fruit rot unless physically damaged, our focus is on resistance expressed in ripe fruit. To improve 113 techniques for resistance screening in blueberry breeding and evaluate fruit characteristics as possible phenotypic markers of resistance, we conducted a study with the following objectives: 1) Compare different inoculation techniques for anthracnose fruit rot resistance screening, 2) Screen ripe fruit from a range of blueberry cultivars using selected techniques, and 3) Investigate the role of fruit characteristics in anthracnose fruit rot resistance. The overall goal of this project is to facilitate the development of anthracnose fruit rot-resistant blueberry cultivars. MATERIALS AND METHODS Plant and fungal material Ripe blueberry fruit was collected from the Michigan Blueberry Growers Association (MBG) variety trial field in Grand Junction, MI, USA on 27 July and 7 August in 2008 and 2009 and stored at 4°C for no more than 3 days prior to testing. Approximately 50% of the fruit of selected cultivars was ripe at the time of harvest. Germplasm types were categorized into five groups: northern highbush (Vaccinium corymbosum L.), southern highbush (adapted V. corymbosum with minimal chilling requirement, some crossed with V. darrowi Camp and/or V. ashei Reade), half high (V. corymbosum × V. angustifolium Ait.), and intermediate (between southern and northern highbush blueberries in chilling requirement with genetic contributions from southern highbush) (Table 5.1). The cultivars included in this study are commonly grown in Michigan and account for more than 95% of the cultivated blueberries in the state (105). A singleconidium isolate of C. acutatum isolated from blueberry fruit in Grand Junction, MI, USA in August 2006 was used for inoculations. This isolate (CA001) was the most virulent of 25 isolates in a preliminary test and was used for all experiments. Fungal cultures of C. acutatum were grown and stored in accordance with Miles et al. (2011). For inoculum production, sporulating cultures 114 were flooded with 3 mL of sterile deionized water (SDW), and conidia were dislodged using a 6 sterilized L-shaped glass rod. Conidia were counted using a hemacytometer, and diluted to 1 x 10 conidia per milliliter with SDW. Inoculation methods Ripe fruit of cvs. Elliott (resistant) and Jersey (susceptible) was inoculated with a conidial 6 suspension (10 conidia per milliliter) using four different methods: 1) applying a 10-µL droplet into the calyx cup, 2) spraying the berries until runoff, 3) injecting 50 µL with a syringe into the interior of the fruit, and 4) applying a 50-µL droplet to the open surface of a cut fruit. For the cutfruit assay, berries were cut in half longitudinally with a sterile scalpel and placed cut side up. There were five replicates per cultivar with 5 fruits per replicate. Fruit in all treatments was incubated on wire mesh screens at 22-24°C and 100% humidity for 10 days post inoculation (dpi), except cut fruit which was incubated for 3 dpi. After incubation, conidium production was quantified by placing five inoculated fruit in 5 mL of sterile water and gently inverting the tube for 5 minutes. Conidium concentration per milliliter was determined using a hemacytometer as the average of three separate counts (Table 5.1). To further evaluate whole- and cut-fruit inoculation methods, a range of blueberry cultivars (24 in 2008 and 26 in 2009) was screened for anthracnose fruit rot resistance. Ripe whole fruit were spray inoculated as described above followed by evaluation of infection incidence (proportion of fruit infected) and severity (percentage of the fruit surface supporting sporulation) after 10 days. Cut fruit were drop inoculated and sporulation was measured with a hemacytometer as the number of conidia produced per fruit half. 115 To determine the utility of spectrophotometry for measuring sporulation, conidia were harvested from a 4-day-old microconidiating culture of C. acutatum. A dilution series from 1.0 x 5 7 10 to 1 x 10 conidia per ml was used to develop a standard curve for conidium concentration based on optical density of the conidial suspension. The absorbance was read with an EL 800 Universal Microplate Reader (BioTek Instruments, Winooski, VT, USA) at a wavelength of 590 nm. The actual conidium concentration of each sample was determined microscopically with a hemacytometer. All measurements were made on triplicate subsamples. The same procedures were used to quantify the number of conidia produced on cut fruit surfaces of seven blueberry cultivars (Bluecrop, Blueray, Elliott, Jersey, Liberty, Nelson and Rubel) in 2008. In addition, fruit from an F-1 population resulting from a cross of cvs. Draper (resistant) and Jewel (susceptible) was used to validate the spectrophotometric procedure. Plants were grown in Interlachen, FL, USA and fruit was collected in April and May, 2011 when approximately 50% of the fruit on a bush was ripe. Fruit was cooled immediately, shipped within three days of the harvest, and inoculated and evaluated as described above. Assessment of fruit characteristics Ripe fruit of different blueberry cultivars was harvested at appropriate times from early July to late August over 4 years (2005 to 2008) from the MBG variety trial in Grand Junction, MI, USA. Sugar content, pH, and titratable acidity (TA) were measured in juice extracted from 25 berries per cultivar (five replicates of five berries) that were blended at high speed in a tissue homogenizer (Ultra Turrax T25; Janke and Kunkel Co., Staufen, Germany). Sugar content was determined using a handheld refractometer (Westover model RHB-32; Southwest United Industries, Tulsa, OK, USA). Results are reported as percent sugar content (wt/wt) on a fresh116 weight basis. TA was determined using 10 mL of juice diluted to 100 mL with SDW, titrated with 0.1 N sodium hydroxide to pH 8.2, and expressed as percent citric acid (wt/wt) on a fresh-weight basis. Fruit firmness was determined on a sample of 50 fruit per cultivar per year using a portable 2 firmness measuring instrument (188). Data were reported as force (newtons/mm ) required to deform the surface of the fruit. These results represent an expansive data set including more cultivars and years of data collection than previously reported by (84). Effect of sugar concentration on mycelial growth in vitro To determine the effect of sugar concentration on fungal growth, a minimal medium was utilized as described by (5), consisting of 0.7% KH2PO4, 0.4% KNO3, 0.3% Na2HPO4, 0.1% MgSO4, 0.03% CaCl2, 1.5% BACTO-agar, and varying concentrations of sugar (ranging from 4 to 20%). The sugar utilized in the media was 100% D-glucose, 100% D-fructose, or a 50/50 mixture of both sugars. The medium was adjusted to a final pH of 6.5 and autoclaved at 121°C and 100 kPA for 30 minutes. Hyphal plugs (~5.5 mm in diameter) were transferred from a 7-day-old potato dextrose agar (PDA) culture of C. acutatum to Petri plates containing the defined minimal medium. There were three replicate plates per treatment. Plates were incubated without Parafilm for 7 days in the dark at 25°C. After incubation, the diameter of each colony was measured in two perpendicular directions, and the average diameter was calculated after subtracting the diameter of the initial hyphal plug. The experiment was conducted three times. 117 Effects of pH and sugar concentration on mycelial growth in liquid media To determine the effects of pH and the interaction of pH and sugar concentration on mycelial growth, liquid media were used, namely potato dextrose broth containing 0.4% potato starch and 2% dextrose, and the minimal medium described above containing 4% D-glucose and 4% D-fructose without the addition of agar. Both media were adjusted to a final pH of 2.0, 2.5, 3.0, 3.5, 4.0, 5.0, and 6.5 using dilute hydrochloric acid as measured with an Accumet AB15 pH meter (Thermo Fisher Scientific, Waltham, MA, USA). The media were sterilized with a 0.45-μ syringe filter (Millipore, Billerica, MA, USA) and placed in 2.95-mL aliquots in 14-mL polypropylene Falcon tubes (Becton Dickinson Labware, Franklin Lakes, NJ, USA). A 50-µL conidial suspension 6 (10 conidia/mL) from a 3- to 7-day-old culture was added to each tube and cultures were shaken at 300 rpm at 22-24°C. After 7 days, the mycelial suspensions were filtered through pre-weighed, 9-cm-diam. Whatman filter paper disks in a Büchner funnel by applying a vacuum. Filter disks were allowed to air dry for 4 days and the dried mycelium was weighed. To study the interaction of pH and sugar concentration, the minimal medium as described above was used with sugar concentrations of 8, 10, 12, 14, and 16% (50/50 D-glucose/D-fructose) at three pH levels (3.0, 4.0, and 5.5). Each treatment was replicated three times and the experiments were conducted twice. Statistical analyses All statistical analyses were performed with the StatGraphics statistical computer program (StatPoint Inc., Warrenton, VA, USA) and SigmaPlot version 11 (SYSTAT Software, Chicago, IL, USA). For the different inoculation techniques, the data were statistically analyzed using a paired Student’s t-test (α = 0.05) after checking for normality and equality of variance. For the cultivar resistance screening experiments, statistical differences were determined by one-way ANOVA and 118 Fisher’s Protected LSD test (P = 0.05). Since the factor “year” had no significant effect, we pooled the data from 2008 and 2009 to calculate means and standard errors. For fruit composition analysis, all experiments were analyzed by ANOVA according to a completely randomized design. Previously reported resistance ratings using proportion decayed fruit from Polashock et al., (2005) were utilized for regression analyses. Linear and power (log-log) regressions were run using SigmaPlot for obtaining P and r values. For investigating the combined effect of pH and sugar, a two-factor ANOVA was used to analyze main and interaction effects of sugar content and pH (α = 0.05). RESULTS Inoculation method does not affect the resistance phenotype The three whole-fruit inoculation methods (droplet, spray, and injection) resulted in similar infection phenotypes with abundant sporulation on the fruit surface 10 days after inoculation 6 (Figure 5.1). In each method, ‘Jersey’ fruit yielded similar numbers of conidia (23.2 x 10 to 26.5 6 6 x 10 conidia per milliliter), whereas conidium production was much lower in ‘Elliott’ (2.8 x 10 6 to 4.0 x 10 conidia per milliliter) (Figure 5.2). The variance of the data was highest in the injection and lowest in the spray inoculation procedure. In the cut-fruit method, fungal sporulation was evident on the cut fruit surface of both cultivars after 3 days. However, ‘Jersey’ showed more darkening and maceration of the internal fruit tissues than ‘Elliott’ (Figure 5.1D and 1I). No significant necrosis or sporulation was observed in the water control of either cultivar (Figure 5.1E and 1J). Overall, the cut-fruit method yielded about 10-fold fewer conidia per berry than the whole-fruit inoculation methods, but the 119 A B C D E F G H I J Figure 5.1. Signs and symptoms of fruit infection by Colletotrichum acutatum on two different blueberry cultivars, Jersey (A to E), and Elliott (F to J) after inoculation with a conidial suspension using different techniques, such as applying a 10-µL droplet into the calyx cup (A and F), spraying the berries until runoff (B and G), injecting 50 µL into the interior of the fruit with a syringe (needle still visible) (C and H), and applying a 50-µL droplet to the open surface of a cut fruit (D and I). Fifty microliters of sterile deionized water served as a control in the cut fruit experiments (E and J). All treatments were incubated for 10 days after inoculation, except the cut-fruit treatments (3 days). 120 Jersey 6 No. of conidia produced per berry (10 ) 40 Elliott 30 20 10 0 Drop Spray Syringe Cut surface Inoculation methods Figure 5.2. Conidium production on fruit of blueberry cultivars Jersey and Elliott after inoculation with a Colletotrichum acutatum conidial suspension using different inoculation methods (applying a 10-µL droplet into the calyx cup, spraying the berries until runoff, injecting 50 µL into the interior of the fruit with a syringe, and applying a 50-µL droplet to the open surface of a cut fruit). All treatments were incubated for 10 days after inoculation, except the cut-fruit treatment (3 days). Bars denote the standard error of the mean (n = 5 with 10 fruits per replicate). 121 Table 5.1. Anthracnose fruit rot resistance profiles of different blueberry cultivars after artificial inoculation with Colletotrichum acutatum as measured by infection incidence (proportion of fruit infected), infection severity (the percentage of the fruit surface supporting sporulation), and sporulation capacity (number of conidia produced on the cut surface of a half berry). Ripe fruit was collected from a field planting in Grand Junction, MI, USA in July and August of 2008 and 2009 (n = 10 unless otherwise noted and SE = standard error of the mean). Cultivar Sporulation capacity Infection severity Germplasm Infection incidence 6 (proportion ± SE) (percentage ± SE) (no. of conidia x 10 ± SE) a c type (2008 and 2009) (2009 only) (2008 and 2009) Aurora NHB 0.64 ± 0.05 3.9 ± 0.5 0.56 ± 0.02 Berkeley NHB 0.92 ± 0.03 8.8 ± 0.8 4.56 ± 0.20 Bluechip NHB 0.75 ± 0.13 - d 14.97 ± 3.25 Bluecrop NHB 0.98 ± 0.02 21.6 ± 2.0 12.30 ± 0.90 Bluehaven NHB 0.88 ± 0.06 7.7 ± 1.1 0.51 ± 0.07 Bluejay NHB 0.63 ± 0.04 14.4 ± 3.4 3.12 ± 1.75 Blueray NHB 1.00 ± 0.00 27.2 ± 3.6 10.81 ± 2.75 Bounty NHB 0.98 ± 0.02 26.9 ± 2.8 7.60 ± 1.60 Brigitta NHB 0.66 ± 0.03 7.5 ± 1.7 2.15 ± 0.67 Caroline NHB 0.65 ± 0.13 d 1.90 ± 1.15 Collins NHB 0.98 ± 0.01 22.7 ± 2.1 1.68 ± 0.60 Darrow NHB 0.98 ± 0.01 15.0 ± 2.4 7.62 ± 1.80 Denise NHB 0.86 ± 0.06 9.6 ± 1.9 3.34 ± 0.65 Draper NHB x SHB 0.68 ± 0.07 c 8.4 ± 2.3 1.05 ± 0.80 Duke NHB 0.70 ± 0.04 c 14.9 ± 1.5 7.25 ± 0.60 Earliblue NHB 0.70 ± 0.04 36.3 ± 6.4 2.87 ± 0.55 Elliott NHB 0.47 ± 0.06 2.4 ± 0.2 0.29 ± 0.08 b b - 122 c c Table 5.1 (cont’d.) Jersey NHB 0.98 ± 0.02 21.6 ± 1.4 3.35 ± 0.31 Lateblue NHB 0.50 ± 0.06 d 3.00 ± 0.55 Liberty NHB 0.52 ± 0.04 4.3 ± 0.75 1.00 ± 0.33 Nelson NHB 0.70 ± 0.06 d 14.10 ± 5.00 Northblue HH 0.70 ± 0.07 9.9 ± 0.9 8.04 ± 0.10 Northland NHB 0.81 ± 0.05 8.9 ± 0.7 0.98 ± 0.48 Rose NHB 0.57 ± 0.08 6.1 ± 1.3 4.13 ± 1.10 Rubel NHB 0.98 ± 0.02 14.3 ± 1.3 2.07 ± 0.80 Toro NHB 0.81 ± 0.04 5.4 ± 0.9 5.62 ± 2.15 b - b - a Abbreviations for germplasm type as previously defined in materials and methods. NHB = northern highbush; SHB = southern highbush; HH = half high (V. corymbosum x V. angustifolium); NHB x SHB = intermediate. b Data were collected only in 2008 and therefore n = 5. c Data were collected only in 2009 and therefore n = 5. d No data collected. 123 difference between the cultivars was proportionally the same. However, if the cut fruit was incubated longer than 3 days, significant background infection became apparent, since these were field-collected berries (data not shown). A linear regression yielded the best fit for a standard curve of conidium concentration 2 against optical density of an aqueous suspension of C. acutatum conidia from culture (R > 0.99) (Figure 5.3A). Conidia harvested from inoculated cut fruit surfaces from the Draper x Jewel 2 population in 2011 also showed a strong positive linear correlation with optical density (R = 0.90), although there was an occasional outlier (Figure 5.3B). A linear function also described the standard curve for conidia from cut fruit surfaces of selected blueberry cultivars in 2008. However 2 the relationship was not as strong as for the previous curves (R = 0.81) (Figure 5.3C). Fruit resistance results correlate with previously published resistance ratings The results of screening different blueberry cultivars for anthracnose fruit rot resistance in the whole- and cut-fruit assays were consistent from 2008 to 2009, as there were no significant effects of year or cultivar × year. P-values were 0.32 (whole fruit) and 0.51 (cut fruit) for the effect of year, and 0.13 (whole fruit) and 0.21 (cut fruit) for cultivar × year. Data were therefore reported as an average of the two years (Table 5.1), except for infection severity which was only rated in 2009. Water controls in the whole-fruit inoculations showed evidence of background infection, primarily in the susceptible cultivars. Of the resistance measures evaluated, sporulation capacity in the cut-fruit assay was most strongly correlated with cultivar resistance ratings published by Polashock et al. (2005) 124 6 No. of conidia per milliliter (10 ) No. of conidia per milliliter (10 ) No. of conidia per milliliter (10 ) 6 A 6 B C 16 2 R = 0.99 12 P < 0.001 8 4 0 16 12 2 R = 0.90 P < 0.001 8 4 0 16 12 2 Bluecrop R = 0.81 P = 0.006 Nelson 8 Blueray Liberty 4 Jersey Elliott Rubel 0 0.00 0.25 0.50 0.75 Optical density (590 nm) 1.00 Figure 5.3. Relationship between the concentration of an aqueous suspension of Colletotrichum acutatum conidia and optical density at 590 nm. A) Conidia produced on PDA. B) Conidia from cut fruit surfaces from a Draper x Jewel F1 population. C) Conidia produced on cut fruit surfaces of selected blueberry cultivars. 125 Table 5.2. Pearson correlation coefficients (r) and statistical significance (P) for regressions between different measures of anthracnose fruit rot resistance after artificial inoculation of a range of blueberry cultivars with Colletotrichum acutatum. Infection incidence (proportion of fruit infected), infection severity (the percentage of the fruit surface supporting sporulation), sporulation capacity (number of conidia produced on the cut surface of a half berry) and previously published resistance ratings (Polashock et al., 2005) were subjected to linear regression. All values were log transformed prior to regression. R-values in boldface are statistically significant at α < 0.05 Dependent Variable (Values from Table 5.1) Independent Variable Infection incidence Infection severity Sporulation capacity Number of cultivars in common r value P value Infection severity 22 0.69 < 0.001 Sporulation capacity 26 0.40 0.044 Proportion decayed (Polashock et al., 2005) 18 0.31 0.208 Sporulation capacity 22 0.63 0.002 Proportion decayed (20) 15 0.64 0.010 Proportion decayed (20) 18 0.86 < 0.001 126 (r = 0.86; P < 0.01). Infection incidence in the whole-fruit assay was positively correlated with infection severity on whole fruit (r = 0.69, P < 0.01) and sporulation capacity on cut fruit (r = 0.40, P = 0.04) but only weakly correlated with resistance ratings by Polashock et al. (2005) (r = 0.33, P = 0.21). Infection severity on whole fruit also was correlated with sporulation capacity on cut fruit (r = 0.63, P = 0.002) and resistance ratings by Polashock et al. (2005) (r = 0.64, P = 0.01) (Table 5.2). Anthracnose resistance is correlated with certain fruit characteristics Average weight per berry ranged from 1.10 g to 3.07 g, sugar content from 10.3% to 13.9%, pH from 2.53 to 3.28, titratable acidity from 0.4% to 1.2%, and fruit firmness from 262 to 2 519 newtons per mm (Table 5.3). Anthracnose fruit rot resistance as measured by sporulation capacity on cut fruit, infection severity on whole fruit, and proportion decayed from Polashock et al. (2005) was negatively correlated in a linear fashion with percent soluble sugar in the fruit (r = 0.53, P = 0.02; r = -0.44, P = 0.09; r = -0.62, P < 0.01, respectively), i.e. the higher the sugar content, the more resistant the fruit. The correlation between infection incidence on whole fruit and sugar content showed the same trend but was not significant (r = -0.24, P = 0.34). Additionally, infection severity on whole fruit showed marginally significant negative correlations with berry weight (r = -0.44, P = 0.09) and titratable acidity (r = -0.45, P = 0.08) (Table 5.4). Correlations with fruit firmness were not significant (Table 5.4). Mycelial growth is reduced at high sugar content and low pH Colonies of C. acutatum grew faster on solid media with D-glucose than with D-fructose (Figure 5.4D). Growth on the medium with a 50/50 mixture of D-glucose and D-fructose was more 127 Table 5.3. Characteristics of fruit of different blueberry cultivars collected from a field planting in Grand Junction, MI, USA from 2005 to 2008. Values shown are averages and standard errors over 4 years. Five berries were used per replicate. a b Berry weight (g) Sugar content (%) Titratable Firmness 2c acidity (%) (N/mm ) Cultivar Type Years tested Aurora NHB 05, 06, 07, 08 Berkeley NHB Bluecrop NHB 05, 06, 07, 08 2.0 ± 0.1 11.2 ± 0.3 2.8 ± 0.1 0.9 ± 0.1 354.6 ± 67.9 Bluegold NHB 05, 06, 07, 08 1.9 ± 0.2 12.6 ± 0.3 2.8 ± 0.1 1.0 ± 0.1 342.7 ± 66.1 Bluehaven NHB 1.9 ± 0.3 11.8 ± 0.2 2.7 ± 0.0d 0.8 ± 0.0d 146.2 ± 95.5 Bluejay NHB 05, 06, 07, 08 1.8 ± 0.1 12.5 ± 0.3 3.0 ± 0.1 0.7 ± 0.0d 299.6 ± 54.5 Blueray NHB 05, 06, 08 2.3 ± 0.1 10.5 ± 0.8 3.0 ± 0.1 0.7 ± 0.1 365.8 ± 89.3 Bluetta NHB 05, 06, 08 1.4 ± 0.1 10.4 ± 0.2 2.8 ± 0.1 0.8 ± 0.2 396.8 ± 99.4 Brigitta NHB 05, 06, 07, 08 2.6 ± 0.3 12.1 ± 0.6 2.8 ± 0.1 0.9 ± 0.0d 463.4 ± 90.1 Chanticleer NHB 06, 07, 08 3.1 ± 0.2 10.5 ± 0.5 2.8 ± 0.1 Darrow NHB 05, 06 2.6 ± 0.1 12.2 ± 0.9 2.6 ± 0.0d 1.2 ± 0.0d 519.2 ± 211.7 Draper X 05, 06, 07, 08 2.2 ± 0.2 12.2 ± 0.9 2.8 ± 0.1 1.1 ± 0.1 519.2 ± 124.6 Duke NHB 05, 06, 07, 08 1.9 ± 0.0d 10.3 ± 0.2 3.0 ± 0.1 0.7 ± 0.1 572.5 ± 148.6 Earliblue NHB Elliott Jersey 07 only 06, 07 2.0 ± 0.2 13.4 ± 0.5 2.8 ± 0.1 1.1 ± 0.0d 415.6 ± 85.9 2.0 12.3 2.8 0.6 227.7 1.1 ± 0.1 276.4 ± 79.9 1.2 ± 0.0 11.1 ± 0.7 2.9 ± 0.0d 0.7 ± 0.1 361.7 ± 80.9 NHB 05, 06, 07, 08 1.8 ± 0.2 12.9 ± 0.9 2.8 ± 0.1 NHB 05, 06, 07, 08 1.6 ± 0.2 13.4 ± 0.4 3.2 ± 0.2 0.5 ± 0.0d 319.9 ± 64.3 05, 06, 07 d pH 128 1.1 ± 0.1 341.8 ± 69.7 Table 5.3 (cont’d.) Jewel SHB 08 only 2.5 Lateblue NHB 06, 07, 08 2.1 ± 0.1 12.9 ± 0.5 2.8 ± 0.1 0.9 ± 0.1 167.1 ± 54.6 Legacy X 05, 06, 07, 08 2.2 ± 0.1 13.2 ± 0.6 3.1 ± 0.1 0.6 ± 0.1 373.2 ± 75.1 Liberty NHB 05, 06, 07, 08 2.4 ± 0.3 13.3 ± 0.7 2.9 ± 0.1 0.9 ± 0.0d 518.8 ± 114.6 Nelson NHB 05, 06, 07, 08 2.1 ± 0.3 12.2 ± 0.2 2.8 ± 0.1 1.0 ± 0.0d 318.8 ± 53.4 O’Neal SHB 06, 08 1.7 ± 0.1 12.5 ± 0.2 3.28 ± 0.4 0.5 ± 0.0d 121.2 ± 89.8 X 05, 06, 07, 08 2.2 ± 0.1 11.3 ± 0.3 2.9 ± 0.1 Patriot NHB 07 only 2.2 Rubel NHB 05, 06, 07 1.1 ± 0.1 Sapphire SHB 08 only 1.8 Spartan NHB 05, 06, 07, 08 2.3 ± 0.1 11.4 ± 0.1 3.0 ± 0.1 0.6 ± 0.0d 396.2 ± 75.4 Star SHB 1.5 ± 0.2 10.8 ± 0.4 3.1 ± 0.3 Toro NHB 05, 06, 07, 08 2.6 ± 0.0d Ozark Blue 06, 08 11.5 11.9 2.6 2.5 13.9 ± 0.3 2.9 ± 0.1 13.1 3.0 0.7 211.1 0.8 ± 0.1 354.0± 66.5 1.0 310.3 0.9 ± 0.1 424.9 ± 110.2 0.4 262.0 0.6 ± 0.1 161.2 ± 125.5 11.1 ± 0.3 2.8 ± 0.1 0.9 ± 0.0d 374.0 ± 73.2 a Abbreviations for germplasm type as previously defined. NHB = northern highbush; SHB = southern highbush; HH = half high (V. corymbosum x V. angustifolium); X = intermediate. b To calculate means, 5 berries per replicate (25 in total) were used for each year of study. All values displayed in the table correspond to the mean of the yearly means. c Force required to indent blueberry fruit. d Indicates standard error values less than 0.05. 129 Table 5.4. Pearson correlation coefficients (r) for regressions of various measures of anthracnose fruit rot resistance after artificial inoculation of blueberry fruit with Colletotrichum acutatum (Table 5.1) against fruit characteristics (Table 5.3). Infection incidence (proportion of fruit infected), infection severity (percentage of the fruit surface supporting sporulation), sporulation capacity (number of conidia produced on the cut surface of a half berry) and previously published resistance ratings (150) were subjected to linear regression against fruit variables. Statistically significant r-values are indicated in boldface. Independent variable Dependent variable Berry Weight (g) Sugar content (%) pH Titratable acidity (%) Firmness (Newton per 1 2a mm ) r-value Infection incidence -0.11b -0.24 0.16 0.28 0.06 Infection severity -0.44* -0.44* 0.42 -0.45* -0.07 0.17 -0.53** 0.10 -0.03 0.03 0.29 -0.62** -0.30 0.26 0.18 Sporulation capacity Proportion decayed (150) a Force required to indent fruit surface. b *Indicates statistical significance at P ≤ 0.10, ** Indicates statistical significance at P ≤ 0.05 130 similar to growth with D-glucose than with D-fructose. However, mycelial growth was reduced to a similar extent as concentrations of both sugars increased. Average colony diameter was reduced by 37% and 35%, respectively, as D-glucose and D-fructose concentrations increased from 4% to 20%. On the medium with a 50/50 mixture of D-glucose and D-fructose, the largest reduction in growth occurred between 12% (Figure 5.4B) and 16% sugar content (Figure 5.4C), reducing average colony diameter by about 1 cm (Figure 5.4D). In a more detailed experiment using a 50/50 mixture of D-glucose and D-fructose within the physiological range of blueberries, the biggest reduction in mycelial growth was seen between 11% and 13% sugar content (Figure 5.4E). When investigating the effect of initial medium pH on mycelial growth of C. acutatum in liquid media, a similar pattern was observed on both media with slightly more growth in potato dextrose broth than in the minimal medium. In the latter, mycelial growth was significantly reduced below an initial pH of 4.0 and no growth had occurred at an initial pH of 2.5 after 7 days of incubation. In potato dextrose broth, a similar reduction was observed but in this case, no growth occurred at a pH of 2.0. An initial pH of 4.0 appeared optimal for mycelial growth (Figure 5.5A). If the fungus grew, the pH was modified in the culture to 5.6 to 6.5 by day 7 depending on the initial pH (data not shown). The combined effect of initial pH and sugar content in the physiological range of blueberries showed that the fungus was able to grow under all conditions, but mycelial growth was most curtailed by high sugar content combined with a low pH (Figure 5.5B). Unfortunately, we could not test pH levels lower than 3 due to the fungus’ inability to grow at low pH in the minimal medium, but by extrapolation, mycelial growth would have been even more reduced and in some cases virtually absent at the more acidic pH levels typical of blueberries, i.e. 2.5 to 3.2 (Table 5.3). 131 Colony diameter (mm) C B A 60 54 50 50 40 46 30 42 2 D 6 10 14 18 Sugar content (%) D-Glucose 22 9 E D-Fructose 10 11 12 13 Sugar content (%) 14 15 50/50 D-Glucose/D-Fructose Figure 5.4. Mycelial growth of Colletotrichum acutatum on minimal medium with different sugar (50/50 D-glucose/D-fructose) concentrations: 8% (w/v) (A), 12% (w/v) (B), and 16% (w/v) (C) after 7 days. D) The effect of sugar content in minimal medium on mycelial growth of C. acutatum after 7 days using D-glucose, D-fructose and a 50/50 D-glucose/D-fructose mixture. E) The effect of sugar content in minimal medium on mycelial growth of C. acutatum after 7 days using a 50/50 D-glucose/D-fructose mixture in the physiological range for blueberries. In D and E, error bars denote standard error of the mean. 132 0.05 Potato dextrose broth Minimal medium Dry mycelial weight (g) 0.04 0.03 0.02 0.01 0.00 2 A 3 4 5 6 Initial pH 0.04 Dry mycelial weight 0.04 g 0.03 g 0.02 g 0.01 g 0.00 g 0.03 0.02 0.01 0.00 8 5.0 10 4.5 12 4.0 3.5 B 14 3.0 16 Figure fdsf 5.5. 133 Figure 5.5 (cont’d.). A) The effect of initial pH in potato dextrose broth and minimal medium on the amount of mycelial growth (dry weight) of Colletotrichum acutatum after 7 days in 3ml cultures. Initial pH is indicated here since the pH changed as cultures grew. Error bars denote standard error of the mean. B) The effect of initial pH and sugar content (using a 50/50 Dglucose/D-fructose mixture) on mycelial growth (dry weight) of C. acutatum after 7 days. Different degrees of shading indicate different mycelial weights, with darker shades indicating less growth . 134 At the typical sugar content of ripe blueberries (10.4 to 13.9%; Table 5.3) moderate growth suppression would be expected. DISCUSSION This study represents a broad approach to screening blueberry cultivars for anthracnose fruit rot resistance, comparing and improving screening methods, and correlating resistance ratings to a previous study. When we inoculated the susceptible cultivar Jersey and the resistant cultivar Elliott, the quantity of conidia produced was similar in fruit inoculated by syringe injection compared to droplet and spray inoculation. We also obtained a difference of similar proportion by drop-inoculating the open surface of cut fruit. This suggests that anthracnose fruit rot resistance in ‘Elliott’ fruit is expressed not only in the skin but also in the flesh of the blueberry. While Colletotrichum studies generally do not involve inoculation of wounded or cut fruit because wounding is not required for infection, similar resistance levels were found in five olive cultivars whether the fruit was wounded or not prior to inoculation (131). Since the majority of soluble phenolic compounds in blueberries occur in the epidermal and subepidermal layers of the fruit peel, e.g., anthocyanins and flavonol glycosides (161), mechanisms other than the presence of soluble phenolic substances may play a role in the resistance response. Under natural conditions, infection usually takes place on green fruit followed by a latent period until the fruit ripens, at which time the infection progresses until most of the fruit is covered with acervuli (123). Infection incidence tends to increase rapidly during fruit ripening, with later harvests having poorer fruit quality (128). Due to the labor and space involved in inoculating green fruit on plants in the field or greenhouse, we chose to inoculate detached ripe fruit collected from field-grown plants. We compared three different disease evaluations, including disease incidence 135 and severity on whole fruit and sporulation capacity on cut fruit. The whole-fruit inoculations did not seem to correlate well with the other data sets, which is likely due to background infections as indicated by the water controls. In our experience it is difficult to avoid background infection in the field unless plants are grown in dry regions or protected from precipitation. Therefore, using whole ripe fruit from field-grown plants may not provide accurate resistance ratings. Furthermore, due to the variation in flowering and ripening times of different cultivars, it is possible that some escape infection or, conversely, are exposed to higher-than-average doses of inoculum. While an escape of infection due to timing of flowering or fruit ripening may lead to “field resistance”, this would not be a reliable indicator of cultivar performance across regions. The cut-fruit inoculation assay for blueberries showed the strongest correlation with previously reported anthracnose fruit rot resistance ratings by Polashock et al. (2005) in New Jersey and appears to be an efficient method for fruit rot resistance screening. Furthermore, due to the relatively short incubation period, interference from background infections is mostly eliminated as acervuli from natural infections on the fruit epidermis are not able develop to any significant extent during the 3-day incubation period. Since local pathogen strains were used in both studies, the strong correlation suggests that strain variation is not a big issue, and that a cultivar that is resistant in one region can also be expected to be resistant in another region. It should be pointed out that Polashock et al. (2005) rated disease incidence (proportion of fruit decayed) whereas our cut-fruit assay measured the quantity of spores produced per half berry, which could be interpreted as a measure of disease severity. Disease incidence and severity are often correlated but not necessarily congruent. In addition to a direct relationship between inoculum abundance and measures of disease incidence and severity, sporulation and fruit 136 softening must be sufficiently severe to be visible to the naked eye against the dark background of a blueberry fruit, thereby affecting incidence ratings. The cut-fruit assay is an improvement over conventional anthracnose fruit rot resistance screening assays in that testing can be done on younger bushes that produce small amounts of fruit, facilitating screening earlier in the breeding process. This way, moderately to highly susceptible genotypes can be excluded before much effort has been expended in their selection and maintenance. It does not preclude conducting other screening assays at a later stage. Since microscopic quantification of conidia requires training and is somewhat tedious, a possible improvement is the use of spectrophotometric methods in quantification of sporulation. We found a linear correlation between optical density and conidium concentration in several experiments, with the strongest correlation observed for conidia harvested from culture. Since pigments and debris released from injured blueberry fruit may affect optical density readings, it is advisable to develop a standard curve for cut fruit from representative cultivars prior to conducting a largescale screening assay. . While use of optical density for quantifying conidia from cut fruit may be slightly less accurate, it is more time efficient and requires less training than microscopic observation. A positive linear correlation between fruit sugar content and anthracnose fruit rot resistance suggests that soluble sugars may play a role in the resistance response. However, it most likely is an additive effect as indicated by the relatively low r values and the fact that some moderately susceptible cultivars had fairly high soluble sugar concentrations. Diseases have been classified as high- and low-sugar diseases (88, 191). Low-sugar diseases are characterized by heightened resistance to pathogens when tissues contain more sugar, a concept termed “high-sugar resistance” (88, 191). Soluble sugars have been found to repress 137 photosynthetic genes as well as induce a number of defense-related genes and this shift in expression could contribute to the overall resistance response (87). We found that high sugar concentrations in artificial media had a negative impact on hyphal growth of C. acutatum, presumably by increasing osmotic stress. This reduction was more pronounced with D-glucose than with D-fructose. Our results suggest that internal sugar content in fruit may play a role in slowing the growth of C. acutatum during the colonization of the fruit and should be investigated further. Since sugar has been shown to induce the expression of pathogenesis-related proteins in Arabidopsis thaliana, a similar response may occur in blueberries (187). It is interesting that we did not see a correlation between pH and anthracnose fruit rot resistance ratings despite the fact that hyphal growth of C. acutatum was significantly inhibited in artificial media at pH values between 2.5 and 3.0, which are similar to natural pH levels in ripe blueberries. Our method of measuring the pH of blueberry fruit was fairly crude, however, in that it involved the maceration of the entire fruit. By doing so, we might have missed potential pH variation in different fruit tissues, such as the exocarp and mesocarp (159). Since the epidermis is the first physical barrier encountered by the pathogen, the specific pH of this tissue may be critical for the establishment of C. acutatum. Based on other pathosystems, it is likely that the fungus is able to modify the pH in its immediate surroundings, thereby reducing its effect (55). Modulation of pH by Colletotrichum spp. has been shown to play an important role in the colonization of other plant hosts (157). Our observation of a pH change a period of growth by C. acutatum in artificial media supports this assumption. The role of berry weight and titratable acidity in fruit rot resistance appears to be limited as indicated by relatively low correlation coefficients and marginal statistical significance. However, if the rate of colonization is the same, the fungus would be expected to take longer to colonize 138 large than small berries, which may be manifested in proportionally less surface area colonized in heavier (larger) berries. Percent titratable acidity may be related to berry size and therefore not be directly correlated with infection severity. Since no significant correlations were found between fruit firmness and any of the resistance ratings, it is presumed not to play a role in fruit rot resistance. In this study, we have identified a relatively rapid and simple screening assay for anthracnose fruit rot resistance that can be used early in the selection process of blueberry germplasm. This would not preclude later screening under greenhouse or field conditions. Relying on natural infection only may introduce confounding effects of weather and inoculum availability as flowering and ripening times differ among blueberry cultivars. Artificial inoculation of fruit with no background infection is most desirable to identify an accurate resistance phenotype. Despite evidence for the role of sugar content in anthracnose fruit rot resistance, this characteristic is not suitable as a single marker due to the moderate strength of the correlation. However, selection for high sugar content may inadvertently yield anthracnose fruit rot-resistant cultivars as an added benefit. More investigation is required to elucidate the molecular and biochemical mechanism(s) of anthracnose fruit rot resistance in blueberry in order to identify specific markers for resistance screening and novel strategies for disease management. ACKNOWLEDGEMENTS I would like to gratefully acknowledge funding from Michigan State University Project GREEEN (Generating Research and Extension to meet Economic and Environmental Needs). I would also like to thank the Michigan Blueberry Growers Association for allowing access to their 139 blueberry variety planting and Christine Bates and Daniel Svoboda for technical assistance. We also thank Jerri Gillett and Laura Avila for critical reading of the manuscript. 140 CHAPTER VI: INHERITANCE OF RESISTANCE TO COLLETOTRICHUM ACUTATUM IN HIGHBUSH BLUEBERRY (VACCINIUM CORYMBOSUM L.) FRUIT ABSTRACT Anthracnose fruit rot, caused by Colletotrichum acutatum, is a major disease of highbush blueberries. The inheritance of fruit rot resistance to C. acutatum was investigated in crosses of parents with varying susceptibility. Three cultivars with known resistance profiles (Bluecrop, Elliott, and Jersey) and progeny from sixteen crosses of parents with varying levels of susceptibility were screened. Fruit of field-grown bushes was inoculated when immature, harvested when ripe, incubated, rated after 5, 8 and 12 days and area under the disease progress curve (AUDPC) values were calculated. AUDPC values varied from 117 to 417 in 2010 and 100 to 545 in 2011. These values were then compared with the actual and predicted resistance values from two previous studies. One study focused on disease incidence and the AUDPC values were 2 2 significantly correlated with our values in 2010 (R = 0.73) and 2011 (R = 0.71). The other study utilized disease severity and once again the AUDPC values were significantly correlated with our 2 2 values in 2010 (R = 0.63) and 2011 (R = 0.65). These findings have important implications for anthracnose resistance breeding, and provide strong evidence that anthracnose resistance is highly heritable in highbush blueberries. Additionally, this research provides benchmark values so future breeding selections can be evaluated for their resistance to C. acutatum. INTRODUCTION Anthracnose fruit rot is caused by the fungus Colletotrichum acutatum. The main symptom of anthracnose on blueberries is rotting of ripe fruit in the field before harvest and in 141 storage after harvest (128). Infections occur as early as fruit set, but remain latent until fruit ripening, which complicates detection of the disease. Initially, sunken areas develop on the fruit surface, followed by the formation of sporulating structures (acervuli) exuding salmon-colored spores (conidia). This disease can have a severe economic impact, with preharvest losses estimated at 10% to 20% and post-harvest losses as high as 100% in storage (128). Most blueberry cultivars are susceptible to anthracnose fruit rot including popular cultivars such as Bluecrop, Bluegold, Duke, Jersey, Nelson and Ozarkblue. However, several resistant cultivars have been identified, including Elliott, Brigitta and Legacy, display strong resistance in the field and in laboratory inoculation studies (63, 126, 150). The development of alternative management strategies is necessary because anthracnose is primarily controlled by fungicides (169) at this time. Some of these fungicides are suspected carcinogens (e.g., chlorothalonil), whereas others are prone to fungicide resistance development (e.g., azoxystrobin). In all likelihood, more fungicide sprays are used than necessary because of the difficulty in optimizing spray timing, incorrect disease diagnoses and the possible development of fungicide resistance. In other Colletotrichum plant pathosystems, host resistance has been shown to be localized to a single gene, several genetic loci or a combination of these. In the C. acutatum-strawberry pathosystem, a single dominant gene (Rca2) has been shown to be responsible for high-level resistance, and weaker resistance appeared to be quantitative and controlled by a number of minor genes (50). In the C. acutatum-chili pepper pathosystem, resistance was mapped to a single recessive gene at the mature green fruit stage and a single dominant gene at the ripe fruit stage (119). In another species of pepper, resistance to C. capsici has been associated with a single recessive gene (144). Mapping of host resistance in Phaseolus vulgaris has revealed two 142 independent resistance genes within the same cluster that confer resistance to different strains of C. lindemuthianum (72). Resistance to C. higginsianum has been shown to be localized to a single genetic locus RCH1 in the Arabidopsis ecotype Eil-0 (134). In previous studies, anthracnose fruit rot resistance in blueberries was not significantly correlated with resistance to foliar infection (64) or the production of antimicrobial fruit volatiles (149). However, resistance has been associated with different infection strategies by C. acutatum (201), high levels of reactive oxygen species (125), accumulation of phenolic compounds (124), antimicrobial compounds (127) and high sugar content of blueberry fruit (126). While we have information on the relative anthracnose fruit rot resistance of many cultivars based on our own studies and those of Polashock et al. (2005), the inheritance of resistance in current breeding programs is largely unknown. Additionally, the heritability of disease resistance is of particular interest in cultivated highbush blueberries due to decreasing heterozygosity because of increased interspecific hybridization and the fact that the majority of genetic diversity is derived from only four wild selections (83). An understanding of how resistance to C. acutatum is inherited is important in making strategic breeding decisions as well as estimating durability of resistance. The objectives of this study were to: 1) determine the anthracnose fruit rot susceptibility of daughters from crosses of parents with varying susceptibility and 2) correlate their susceptibility with actual and predicted resistance values from previous studies. MATERIALS AND METHODS Plant material Ten to 20 progeny of crosses between parents varying in susceptibility to anthracnose were selected that were interspersed within two breeding selection blocks located at Michigan State 143 University’s Southwest Michigan Research and Extension Center (Benton Harbor, MI). The selected families were planted together at 0.7 m spacing within rows and 3 m between rows. The plantings were established in 2006 and 2008. Parents with varying susceptibility were identified based on the previous screenings by Polashock et al. (2005) and Miles et al. (2011) (Table 6.1). The most susceptible cultivars according to the two previous studies are ‘Bluecrop’, ‘Duke’, ‘Jersey’ and ‘Nelson’. While the most consistently resistant cultivars between the studies are ‘Brigitta’ and ‘Elliott’ (Table 6.1). Sixteen crosses were selected, including ‘Aurora’ x ‘Legacy’, ‘Aurora’ x ‘Ozarkblue’, ‘Bluegold’ x ‘Elliott’, ‘Bluegold’ x ‘Nelson’, ‘Brigitta’ x ‘Draper’, ‘Brigitta’ x ‘Duke’ , ‘Brigitta’ x ‘Ozarkblue’, ‘Draper’ x ‘Legacy’, ‘Draper’ x ‘Nelson’, ‘Draper’ x ‘Ozarkblue’, ‘Liberty’ x ‘Legacy’, ‘Liberty’ x ‘Nelson’, ‘Liberty x Ozarkblue’, ‘Nelson’ x ‘Ozarkblue’, ‘Ozarkblue’ x ‘Elliott’, and ‘Ozarkblue’ x ‘Legacy’. Additionally within the field, three reference cultivars (Bluecrop, Elliott, and Jersey) were present. Twenty individual daughter plants were selected within each family and the same plants were screened in 2010 and 2011. Fungal material and inoculation methods A single-conidium isolate of C. acutatum isolated (#0001) from blueberry fruit in Grand Junction, MI, USA in August 2006 was used for inoculations. This isolate was the most virulent of 25 isolates in a preliminary test and was in two previous studies (125, 126). Fungal cultures of C. acutatum were grown and stored in accordance with Miles et al. (2011a). For inoculum production, sporulating cultures were flooded with 3 mL of sterile deionized water (SDW), and conidia were dislodged using a sterilized L-shaped glass rod. Conidia were counted using a 6 hemacytometer and diluted to 1 x 10 conidia per milliliter with SDW. 144 Table 6.1. Previous anthracnose fruit rot resistance profiles of blueberry cultivars used in this study using proportion decayed values (150) and sporulation capacity (126). Both studies used artificial inoculation, however, Polashock et al. (2005) inoculated green fruit and rated disease incidence (proportion decayed) and Miles et al. (2011b) inoculated the cut surface of ripe fruit and rated the quantity of conidia produced on a cut fruit surface (sporulation capacity). Cultivar Aurora Bluecrop Bluegold Brigitta Draper Duke Elliott Jersey Legacy Liberty Nelson Ozarkblue Proportion decayed (150) 0.78 0.54 0.18 0.44 0.15 0.38 0.10 0.50 0.55 145 Sporulation capacity 6 (no. of conidia x 10 per berry) (126) 0.56 12.30 2.15 1.05 7.25 0.29 3.35 1.00 14.10 - Pea-sized green fruit was inoculated on each daughter plant by using a hand-pump sprayer and spraying conidial suspensions until run-off, covering the whole plant with a 114-liter, clear plastic bag for approximately 12 hours overnight. Bags were removed the following morning. In 2010, plants were inoculated on 21 and 22 of June with nighttime temperatures ranging from 20 to 27°C. In 2011, plants were inoculated on 16 and 17 of June when nighttime temperatures ranging from 15 to 24°C. Disease rating Fruit from inoculated plants was harvested in individual pint-sized clamshells when ripe (fully blue) on four different dates in 2010 (7, 14, 21, and 28 of July) and in 2011 (6, 13, 20, and 27 of July). Fruit were immediately cooled, transported to the laboratory, and placed equidistantly on wire mesh over a layer of water in covered aluminum pans that acted as humidity chambers, and incubated for 12 days at 22 to 24 °C. Fruits were then monitored at 5, 8 and 12 days for incidence of C. acutatum in the various progeny populations and cultivars. Between 25 and 200 fruits were harvested from each individual daughter plant, and the area under the disease progress curve (AUDPC) was calculated for each individual. AUDPC values were calculated by summing the area between 5 to 8 days and 8 to 12 days. Due to the fact that plants were small and it was not always possible to procure enough fruit from each daughter, some families were sampled more extensively than others. 146 Statistical analyses Data were analyzed using an unbalanced analysis of variance and least square means differences by Tukey’s HSD’s multiple range test for mean separation procedures using the statistical algorithms and GLIMMIX procedure of SAS version 6.04 (SAS Institute, Cary, NC) and SIGMAPLOT version 11 (SYSTAT Software, San Jose, CA). There was a significant effect of year (P = 0.008) and therefore the data from 2010 and 2011 were analyzed separately. Average AUDPC values for each family and cultivar were correlated against the previously reported resistance ratings using proportion of decayed fruit from Polashock et al. (2005) and Miles et al. (2011). We correlated the average AUDPC values of the progeny of each family with the average the previously determined value of the parents (mid-parent value). Expected midparent values from the two previous studies were compared with the mean AUDPC of each 2 cross family and the R value from this linear correlation was used to estimate heritability. RESULTS Daughters of cross families show a wide variation in susceptibility to C. acutatum Fruit rot caused by C. acutatum significantly increased between 5 and 12 days after inoculation. The area under the disease progress curve (AUDPC) was decided as the best method to measure differences in disease incidence among the cross families and cultivars (Figure 6.1). The analysis of variance showed many statistically significant differences among the cross families and reference cultivars (P < 0.001 for 2010 and P < 0.001 for 2011). In 2010, AUDPC values across the families and cultivars averaged 220 and ranged from 117 and 417. In 2011, the average was 262 and ranged from 100 and 545 (Table 6.2). AUDPC values were highest for 147 0.5 2010 Total proportion decayed 0.4 2011 0.3 0.2 0.1 0.0 5 8 12 Time (days) Figure. 6.1. The proportion of all blueberry fruit infected over time by Colletotrichum acutatum in 2010 and 2011 at 5, 8 and 12 days post incubation averaged over all of the cross families and cultivars. 148 Table 6.2. Area under the disease progress (AUDPC) values of the incidence of Colletotrichum acutatum on various blueberry cultivars and cross families screened by artificial inoculation in 2010 and 2011 at the Southwest Michigan Research and Extension Center (Benton Harbor, MI). Fruit of 4 to 7-year old bushes were inoculated when immature, harvested when ripe, incubated, and rated at 5, 8 and 12 days. A statistically significant effect of year (P = 0.008) so data from the two years was analyzed separately. a b a Cultivar/Cross N Bluecrop Elliott Jersey Aurora x Legacy Aurora x Ozarkblue Bluegold x Elliott Bluegold x Nelson Brigitta x Draper Brigitta x Duke Brigitta x Ozarkblue Draper x Legacy Draper x Nelson Draper x Ozarkblue Liberty x Legacy Liberty x Nelson Liberty x Ozarkblue Nelson x Ozarkblue Ozarkblue x Elliott Ozarkblue x Legacy 5 5 5 14 19 19 19 8 20 20 13 20 13 14 20 11 20 13 19 b a 2010 417.4 ± 34.8 159.6 ± 20.0 280.8 ± 29.8 117.9 ± 33.0 277.2 ± 25.0 132.0 ± 21.5 342.7 ± 24.1 127.3 ± 50.3 190.1 ± 24.1 163.8 ± 22.1 188.7 ± 26.6 206.4 ± 33.2 226.4 ± 33.1 127.2 ± 18.4 219.7 ± 36.3 242.1 ± 25.5 311.5 ±31.4 192.9 ± 48.7 262.3 ± 20.9 b N abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg 5 5 5 3 8 16 18 2 20 13 16 12 8 2 12 6 18 0 13 2011 545.2 ± 70.0 138.2 ± 16.1 281.5 ± 22.9 162.4 ± 2.5 148.2 ± 19.4 183.6 ± 21.2 366.3 ± 40.2 186.1 ± 38.9 288.1 ± 26.0 319.2 ± 31.7 138.1 ± 31.7 165.5 ± 29.1 339.0 ± 31.6 99.9 ± 27.9 331.8 ± 47.2 155.7 ± 11.2 440.1 ± 26.5 abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg abcdefg Abcdefg Abcdefg Abcdefg Abcdefg Abcdefg Abcdefg Abcdefg Abcdefg abcdefg c Nd 423.6 ± 41.6 abcdefg N = Number of individual daughter plants per cross family. Values in columns followed by the same letter are not significantly different according to a Tukey’s HSD mean separation test (α = 0.05). c No data collected. 149 Nelson x Ozarkblue Bluegold x Elliott 10 2010 A Frequency 8 6 4 2 0 10 2011 B Frequency 8 6 4 2 0 AUDPC Range Figure 6.2. Distribution of anthracnose fruit rot resistance (expressed as area under the disease progress curves (AUDPC)) in a susceptible x moderate cross (‘Nelson’ x ‘Ozarkblue’) and a resistant x resistant cross (‘Bluegold’ x ‘Elliott’) in 2010 and 2011. A) ‘Nelson’ x ‘Ozarkblue’ in 2010, B) ‘Nelson’ x ‘Ozarkblue’ in 2011, C) ‘Bluegold’ x ‘Elliott’ in 2010, and D) ‘Bluegold’ x ‘Elliott’ in 2011. Discrete levels were defined every 50 AUDPC values.. 150 ‘Bluecrop’ (417 in 2010 and 545 in 2011), moderate for ‘Jersey’ (281 in 2010 and 282 in 2011), and lowest for ‘Elliott’ (160 in 2010 and 138 in 2011). Families with the lowest average AUDPC values were ‘Aurora’ x ‘Legacy’ (118 in 2010 and 162 in 2011), ‘Bluegold’ x ‘Elliott’ (132 in 2010 and 184 in 2011), ‘Brigitta’ x ‘Draper’ (127 in 2010 and 186 in 2011), ‘Draper’ x ‘Legacy’ (189 in 2010 and 138 in 2011), and ‘Liberty’ x ‘Legacy’ (127 in 2010 and 100 in 2011). Those that showed relatively high overall AUDPC values included ‘Bluegold’ x ‘Nelson’ (343 in 2010 and 366 in 2011), ‘Draper’ x ‘Ozarkblue’ (226 in 2010 and 339 in 2011), and ‘Nelson’ x ‘Ozarkblue’ (312 in 2010 and 440 in 2011) (Figure 6.2). Midparent regressions show a correlation between predicted and actual disease resistance values suggesting heritability of resistance AUDPC values for families and cultivars were significantly correlated with the values derived from previous studies of anthracnose resistance including; proportion decayed which 2 2 focuses on disease incidence (R = 0.73, P = 0.001 for 2010, and R = 0.71, P = 0.004 for 2011) 2 (150) (Figure 6.3) and sporulation capacity which focuses on disease severity (R = 0.63, P = 2 0.033 for 2010, and R = 0.65, P = 0.029 for 2011) (126) (Figure 6.4). The predicted midparent values for the proportion decayed for the various families averaged 0.391 and ranged from 0.310 and 0.525, using the parental values generated by Polashock et al. (2005). The predicted midparent values for sporulation capacity of the various families averaged 5.36 and ranged between 1.60 and 7.58, based on the parental calculated by Miles et al. (2011). 151 600 2 R = 0.734 500 P = 0.001 AUDPC Bluecrop 400 BGxN 300 OxLE 200 Elliott A Jersey BRxDU BGxE 100 600 NxO OxE BRxO 2 R = 0.711 500 P = 0.004 Bluecrop NxO AUDPC OxLE 400 BRxO 300 BRxDU BGxN Jersey 200 B 100 0.1 Elliott BGxE 0.3 0.5 0.7 0.9 Proportion decayed (derived from Polashock et al. 2005) Figure 6.3. Correlation between the average anthracnose fruit rot incidence on various blueberry cultivars and cross families expressed as area under the disease progress curves (AUDPC) against actual and predicted proportion decayed values from Polashock et al. (2005) in (A) 2010 and (B) 2011. Abbreviations for parent cultivars: BG = Bluegold; BR = Brigitta; DU = Duke; E = Elliott; LE = Legacy; N = Nelson; O = Ozarkblue. 152 600 2 R = 0.632 500 P = 0.033 AUDPC Bluecrop 400 Jersey 300 200 Elliott A BRxDR 100 600 2 Bluecrop R = 0.650 P = 0.029 500 AUDPC BRxDU 400 Jersey 300 BRxDU 200 BRxDR B LIxN DRxN 100 LIxN DRxN Elliott 0 2 4 6 8 10 12 14 6 Sporulation capacity (Number of conidia x 10 ) (derived from Miles et al. 2011) Figure 6.4. Correlation between the average anthracnose fruit rot incidence on various blueberry cultivars and cross families expressed as area under the disease progress curves (AUDPC) against actual and predicted sporulation capacity values from Miles et al. (2011b) in (A) 2010 and (B) 2011. Abbreviations for parent cultivars: BR = Brigitta; DU = Duke; DR = Draper; E = Elliott; LI = Liberty; N = Nelson. 153 DISCUSSION There were significant differences in the relative disease susceptibilities of the various families, as well as between the three reference cultivars. No family or cultivar exhibited complete resistance; however this was not expected as previous studies have never identified a cultivar that was immune to anthracnose fruit rot. Additionally, for the three cultivars screened, our results mirrored two previous studies and identified ‘Bluecrop’ as highly susceptible, ‘Jersey’ as intermediate, and ‘Elliott’ as highly resistant to anthracnose fruit rot (126, 150). Additionally, the overall AUDPC values for the families were significantly correlated with the predicted values based on previous screens of the parents for disease incidence and severity. This suggests that anthracnose resistance is a highly heritable trait, which is strongly dependent on parental susceptibility. Furthermore, this inheritance appears to be quantitative in nature, as a continuous pattern of variability was observed within and between families. We can also make preliminary estimate of heritability to anthracnose fruit rot resistance in blueberry as ranging between 0.61 to 2 0.73 (H ), which is fairly high. A study on C. graminicola resistance in corn reported values between 26 and 70% using an ordinary least squares regression (30). Quantitative resistance to plant pathogens is typically controlled by multiple loci of small effect and is common in many Colletotrichum pathosystems (30, 50, 73, 89). While this phenomenon has been documented in many systems it is poorly understood. In strawberry different modes of inheritance for resistance to anthracnose have been suggested based on the pathogenicity groups of Colletotrichum (51). Resistance to pathogenicity group 1 is quantitative, however, a single dominant gene, Rca2, controls the resistance to pathogenicity group 2 although minor genes may also contribute to this resistance in several cultivars (50). 154 Understanding how resistance is inherited might provide insight into the mechanism of resistance. Multiple mechanisms of blueberry anthracnose resistance have been identified such as increased sugar content within the fruit (126), a number of pathogensis related-proteins (125), antimicrobial compounds (127) and an oxidative burst (125). A variety of defense mechanisms in other Colletotrichum–plant interactions has been observed, including the production of reactive oxygen species (25), host-derived cell wall-degrading enzymes (31, 76, 110, 204) and preformed and induced antifungal compounds (159). Future work to study segregating populations of specific crosses will make it possible to identify loci that are involved in anthracnose fruit rot resistance in blueberries. In strawberries sequenced characterized amplified region (SCAR) markers have been developed for the Rca2 gene and it has been demonstrated to be highly predictive of anthracnose resistance (114). Identification of quantitative trait loci (QTL) will be useful for developing a marker-assisted selection protocol, which will greatly facilitate future screening for anthracnose fruit rot resistance early in the breeding and selection process. ACKNOWLEDGEMENTS I would like to gratefully acknowledge funding from Michigan State University Project GREEEN (Generating Research and Extension to meet Economic and Environmental Needs). I would also like to thank Christopher Woelk, Kevin Messing, and Chelsea Reynolds for technical assistance and Laura Avila for critical reading of the manuscript. 155 CONCLUSIONS AND FUTURE DIRECTIONS 156 Anthracnose fruit caused by Colletotrichum acutatum J. H. Simmonds is the most important postharvest disease of blueberries (Vaccinium corymbosum). This thesis focused on three main aspects of this disease; 1) host defense responses, 2) the inheritance of resistance and 3) the environmental factors that are important in the infection process. Our investigation of the environmental requirements to C. acutatum infection indicated that temperature, wetness duration, wetness interruptions and relative humidity have a direct effect on the development of melanized appressoria and infection of both immature and mature blueberry fruits by C. acutatum. The relationship between temperature and wetness conforms to previous studies of Colletotrichum spp. on other fruit crops (130, 168, 206). This research provides useful information on temperature and wetness requirements for infection of blueberry fruit by C. acutatum under controlled conditions and this represents the first step towards the development of a disease prediction model for anthracnose fruit rot in blueberries. Further research is needed to validate these results under field conditions. In the field, fruit clusters may stay wet longer than an exposed leaf wetness sensor of a weather station would indicate. This needs to be taken into account when applying a prediction model. We also investigated differentially expressed genes in the resistant blueberry cultivar Elliott, this research represents the first profiling of gene expression in a VacciniumColletotrichum interaction at different stages of the infection process and provides additional evidence for an active resistance response in ‘Elliott’. We identified differentially expressed ESTs in ‘Elliott’ versus ‘Jersey’ that were involved in defense, abiotic stimuli and development and oxidative stress. Our results clearly showed a time delay between inoculation and measurable gene expression. Consequently, there was a significant lag period before any defense response was detectable, which would correlate well with previous microscopy studies outlined in Wharton and 157 Schilder (2008). These results add new insight into the host responses of blueberry fruit to infection by C. acutatum at the molecular level and suggest that pathogen ingress into the host is required for the activation of resistance. A more detailed investigation of gene expression during the early stages of infection, including pre-penetration events, will help to pinpoint when the host first recognizes C. acutatum and initiates the resistance response. Furthermore, chemical analysis and studies on the genetic inheritance of resistance will complement molecular research in elucidating the basis of anthracnose fruit rot resistance in blueberries. Consequently, we looked at antimicrobial compounds present in ripe ‘Elliott’ fruits in response to C. acutatum, and we found that ‘Elliott’ has more anthocyanins and non-anthocyanin flavonoids than ‘Jersey’. Anthocyanins do not seem to play a direct role in the resistance response but may play indirectly protect host tissues from oxidative damage. The non-anthocyanin flavonoid fraction from ‘Elliott’ plays a key role in suppressing microconidiation and reducing growth of C. acutatum. This fraction contained two unique flavonol glycosides; quercetin-3-Orhamnoside, and a dimethylmyricetin glycoside which may play a more specific role in the resistance response because of its increased biological activity. The role of non-anthocyanin flavonoids in the resistance response likely complements other aspects of ‘Elliott’ fruit rot resistance that have been previously described, such as, higher levels of sugar content and an oxidative burst following initial fungal penetration. Further investigation into how these processes are regulated at the molecular level might provide new insights into the interaction. In addition to that we investigated screening methods for anthracnose fruit rot resistance breeding and we have developed a relatively rapid and simple screening assay that can be used early in the selection process of blueberry germplasm. This would not preclude later screening under greenhouse or field conditions but would certainly provide an initial resistance profile. 158 Relying on natural infection only may introduce confounding effects of weather and inoculum availability as flowering and ripening times differ among blueberry cultivars. Artificial inoculation of fruit with no background infection is most desirable to identify an accurate resistance phenotype. Despite evidence for the role of sugar content in anthracnose fruit rot resistance, this characteristic is not suitable as a single marker due to the moderate strength of the correlation. However, selection for high sugar content may inadvertently yield anthracnose fruit rot-resistant cultivars as an added benefit. Future investigation is required to elucidate the molecular and biochemical mechanism(s) of anthracnose fruit rot resistance in blueberry in order to identify specific markers for resistance screening and novel strategies for disease management. Lastly, we investigated the inheritance of C. acutatum resistance in highbush blueberries. There were significant differences in the relative disease susceptibilities of the various cross families, as well as between the three reference cultivars (Bluecrop, Elliott, and Jersey). No family or cultivar exhibited complete resistance; however this was not expected as previous studies have never identified a cultivar that was immune to anthracnose fruit rot. Additionally, the overall AUDPC values for the families were significantly correlated with the predicted values based on previous screenings of the parents for disease incidence and severity. This suggests that anthracnose resistance is a highly heritable trait, meaning that it is strongly dependent on parental susceptibility. Furthermore, this inheritance appears to be quantitative in nature, as a continuous pattern of variability was observed within and between families. We can also make preliminary 2 estimates of resistance in blueberry as ranging between 0.61 to 0.73 (H ), which is fairly high. Future work on segregating populations of specific crosses will make it possible to identify loci that are involved in anthracnose fruit rot resistance in blueberries. In strawberries SCAR markers have been developed for such loci and they have been demonstrated to be highly predictive of 159 anthracnose resistance (114). Identification of quantitative trait loci (QTL) will be useful for developing a marker-assisted selection protocol, which will greatly facilitate future screening for anthracnose fruit rot resistance early in the breeding and selection process. 160 APPENDICES 161 APPENDIX A. RAW DATA FIGURES FROM CHAPTER II 80 60 40 20 40 30 0 30 20 25 20 10 15 10 Figure A.1. The effect of temperature and wetness duration on the development of melanized appressoria of Colletotrichum acutatum on parafilm slides. Discrete transitions indicate levels of melanized appressoria every 20% starting at zero. 162 120 120 80 80 40 40 0 30 A 20 10 40 30 20 10 0 30 B 20 10 40 30 20 10 Figure A.2. The effect of temperature and wetness duration on the infection level of immature (A) and mature (B) blueberry fruits by Colletotrichum acutatum. Discrete transitions indicate levels of infected fruit every 20% starting at zero. 163 80 60 40 20 16 14 12 10 0 45 8 40 35 6 30 4 25 20 15 2 Figure A.3. The effect of wetness duration and interrupted wetness periods on the development of melanized appressoria of Colletotrichum acutatum on parafilm slides. Values are calculated as the melanized appressoria (%) relative to the no interruption control. Discrete transitions indicate levels of melanized appressoria every 20% starting at zero. 164 160 160 120 120 80 80 40 16 12 0 45 40 0 8 35 25 4 15 A 16 12 45 8 35 25 4 15 B Figure A.4. The effect of wetness duration and interrupted wetness periods on the infection level of immature (A) and mature (B) blueberry fruits by Colletotrichum acutatum. Values are calculated as the infected fruit (%) relative to the no interruption control. Discrete transitions indicate levels of infected fruit every 20% starting at zero. 165 80 125 60 100 75 40 20 0 30 28 26 24 A 22 50 100 25 0 30 28 26 24 B 22 80 60 20 100 80 60 20 Figure A.5. The effect of temperature and relative humidity on the development of melanized appressoria (%) of Colletotrichum acutatum (A), and the infection level of mature blueberry fruits (B) by C. acutatum. Discrete transitions indicate levels of melanized appressoria every 20% starting at zero and in mature fruit levels are indicated every 20% starting at zero. 166 APPENDIX B. SUPPLEMENTAL TABLES AND FIGURES FROM CHAPTER III Table B.1. Characteristics and predicted physiological function of differentially expressed sequence tags (ESTs) from ripe fruit of highbush blueberry cultivar Elliott versus cultivar Jersey after inoculation with Colletotrichum acutatum. The hypothetical function is based on homology to sequences in translated nucleotide databases (DDBJ/EMBL/GenBank) using TBLASTX. Some ESTs are likely of fungal origin based on homology with fungal genes. ID code a Size b (bp) Redundancy c Genbank Accession No. Homologous Gene (BLAST Hit Accession Number) Homologous Species Small subunit ribosomal RNA (AY157626) 12S small subunit ribosomal RNA (FJ190616) Penicillium expansum Petriella setifera Score E value -5 Putative fungal proteins (E values < 1x 10 ) EST02 262 1 GW397253 EST09 251 1 GW397260 -5 -1 -52 209 7.0 x 10 47.3 7.0 x 10 -12 Hypothetical proteins (E values > 1x 10 and < 1x10 ) EST12 215 1 GW397263 Hypothetical protein (AF241472) 290 3 GW397264 Hypothetical protein (AC135360) 122 1 GW397265 EST16 300 1 EST17 183 EST18 EST19 EST13* EST14 d Penicillium waksmanii Mus musculus -3 32.2 2.0 x 10 35.9 5.6 x 10 Hypothetical protein (XM001907104) Mus musculus 31.8 1.0 x 10 GW397267 Hypothetical protein (AC135360) Mus musculus 37.3 2.5 x 10 1 GW397268 Hypothetical protein (AC098883) 40.0 1.5 x 10 319 1 GW397269 34.5 1.6 x 10 137 1 HO762505 Mus musculus Actinidia Hypothetical protein (EF530146) deliciosa Ricinus Hypothetical protein (XM002519631) communis 42.8 2.3 x 10 167 -2 -1 -2 -1 -1 -2 Table B.1. (cont’d.) -1 a Unknown (E values > 1 x 10 ) EST15 80 1 EST20 272 1 EST21 99 1 EST22 272 1 EST23 233 1 EST24 63 1 EST25 120 1 EST26 150 1 EST27 273 1 EST28 188 1 EST29 174 1 EST30 290 1 EST31 332 1 EST32 311 1 EST33 322 1 b GW397266 HO762506 HO762507 HO762508 HO762509 HO762510 HO762511 HO762512 HO762513 HO762514 HO762515 HO762516 HO762517 HO762518 HO762519 No homology No homology No homology No homology No homology No homology No homology No homology No homology No homology No homology No homology No homology No homology No homology c N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A d N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A ID code = identification code; bp=number of base pairs; = number of times sequence was recovered; = * denotes ESTs selected for further study. 168 Elliott Jersey 1 2 3 4 5 6 7 8 9 10 11 12 13 14 Figure B.1. Optimization of cDNA amplification using various PCR amplification cycles (1530), for C. acutatum-inoculated fruit of blueberry cultivars Jersey (lanes 2-7) and Elliott (lanes 914) (5 µL loaded per well). Lanes 1 and 8, 1-kb+ DNA ladder; lanes 2 and 9, 15 cycles; lanes 3 and 10, 18 cycles; lanes 4 and 11, 21 cycles; lanes 5 and 12, 24 cycles; lanes 6 and 13, 27 cycles; lanes 7 and 14, 30 cycles. Twenty one cycles was found to be optimal for both libraries and those products were used for digestion and adapter ligation. 169 Elliott Jersey 1 2 3 4 5 6 Figure B.2. Restriction enzyme digestion of pooled cDNA libraries of the blueberry cultivars Jersey and Elliott after inoculation with C. acutatum for the generation of shorter, blunt-end fragments (necessary for adaptor ligation and subtraction). Lanes 1 and 4, 1-kb+ DNA ladder; lanes 2 and 5, samples from ‘Jersey’ and ‘Elliott’ before digestion, respectively; lanes 3 and 6, samples from ‘Jersey’ and ‘Elliott’ after digestion, respectively. 170 1 2 3 4 5 6 7 8 9 10 Figure B.3. Primary (lanes 2-5) and secondary (lanes 7-10) PCR products of pooled cDNA libraries in subtracted and nonsubtracted samples of blueberry cultivars Elliott and Jersey after inoculation with C. acutatum. These products were used for the construction of DIG probes and for cloning of the subtractive libraries for differential screening. Lanes 1 and 6, 1-kb+ DNA ladder; lanes 2 and 3, ‘Elliott’ and ‘Jersey’ subtracted primary PCR products, respectively; lanes 4 and 5, ‘Elliott’ and ‘Jersey’ nonsubtracted primary PCR products, respectively; lanes 7 and 8, ‘Elliott’ and ‘Jersey’ subtracted secondary PCR products, respectively; lanes 9 and 10, ‘Elliott’ and ‘Jersey’ nonsubtracted secondary PCR products, respectively. 171 APPENDIX C. SUPPLEMENTAL FIGURES FROM CHAPTER IV Table C.1. Bioactivity of extracts and fractions from the anthracnose-resistant blueberry cultivar Elliott and susceptible cultivar Jersey as measure by their ability to inhibit microconidiation of Colletotrichum acutatum on solid media (+ = activity, - = no activity). Type of extract/fraction Amount of extract/fraction (μg) Cultivar Methanolic Elliott 1000 + 500 + 250 - 125 - 63 - 31 - 16 - Ethyl Acetate Elliott - - - - - - - Soluble methanol Elliott + + + - - - - Insoluble methanol Elliott + - - - - - - Anthocyanin Elliott + - - - - - - Non-anthocyanin flavonoid Elliott + + + + - - - Methanolic Jersey + - - - - - - Ethyl Acetate Jersey - - - - - - - Soluble methanol Jersey + + - - - - - Insoluble methanol Jersey - - - - - - - Anthocyanin Jersey - - - - - - - Non-anthocyanin flavonoid Jersey + + + - - - - 172 1 2 3 4 1.0 0.8 1 2 AU 0.6 3 4 0.4 0.2 A 0.0 200 B 300 400 500 Wavelength (nm) 600 Figure C.1. TLC plate assay showing inhibition of microconidiation of Colletotrichum acutatum and UV/Vis spectra of methanolic extracts from ripe fruit of the anthracnose-resistant blueberry cultivar Elliott and susceptible cultivar Jersey. (A) Cellulose TLC plate after inoculation with conidia of C. acutatum and stained with iodine crystals (gas phase). (B) UV/Vis spectra of the boxed area in the TLC plate. (Note: 1 = methanolic extract of uninoculated ‘Elliott’ fruit, 2 = methanolic extract of ‘Elliott’ fruit 4 days after inoculation, 3 = methanolic extract of uninoculated ‘Jersey’ fruit, and 4 = methanolic extract of ‘Jersey’ fruit 4 days after inoculation. 173 50g of frozen blueberry tissue* - Grind tissue - Add 150 ml of H2O - Lyophilize Residue H2O Extract - Add 150 ml of MeOH - Evaporate liquid Residue Methanol Extract - Add 150 ml of EtAc - Evaporate liquid Ethyl Acetate Extract Waste *C. acutatum inoculated and uninoculated blueberry tissue from ‘Elliott’ and ‘Jersey’ was analyzed. Figure C.2. Schematic presentation of the extraction procedures for biochemical analysis of blueberry fruit using fresh fruit material and the solvents water, methanol, and ethyl acetate. 174 50g of lyophilized blueberry tissue* - Add 150 ml of MeOH - Evaporate liquid Methanol Extract - Add 20 ml MeOH for 1 min Residue - Evaporate liquid Methanol Soluble Fraction - Add 150 ml of EtAc Methanol Insoluble Fraction - Evaporate liquid Ethyl Acetate Extract *C. acutatum inoculated and uninoculated blueberry tissue from ‘Elliott’ and ‘Jersey’ was analyzed. Waste Figure C.3. Schematic presentation of the extraction procedures for biochemical analysis of blueberry fruit using lyophilized fruit material and the solvents methanol, and ethyl acetate. 175 APPENDIX D. PATHOGENICITY OF VARIOUS ISOLATES OF COLLETOTRICHUM ACUTATUM ON ‘JERSEY’ BLUEBERRIES Table D.1. Pathogenicity of various isolates of Colletotrichum acutatum on detached fruit of the susceptible blueberry cultivar Jersey. Isolates were obtained from various hosts including highbush blueberry (Vaccinium corymbosum), currant (Ribes spp.), cranberry (Vaccinium macrocarpon), raspberry (Rubus spp.), blackberry (Rubus spp.), strawberry (Fragaria x ananassa), and grape (Vitis spp.). Ripe ‘Jersey’ fruits were harvested from a field planting in 6 Harrietta, MI in August 2006, spray-inoculated with 10 conidia per ml, incubated at 100% relative humidity, and rated for incidence of disease 10 days post inoculation (n = 5 and SE = standard error of the mean). Isolate number Strain ID Location (County) Tissue type Host 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 Cola 103-193 BB 599-81b BB 599-42f BB 599-122f BB 599-72a BB 599-1b BB 599-26b BB 599-143c BB 599-121c 13-a-a 1c-a 1a-a 4a-a None 6a Colac 01-083 6a 4a 1f-b 10b-b Colsp 02-034b Berrien Van Buren Allegan Newago Unknown Van Buren Allegan Muskegon Newago Unknown Unknown Berrien Ingham Kent Ingham Unknown Van Buren Van Buren Unknown Unknown Ingham Fruit Fruit Fruit Fruit Fruit Fruit Fruit Fruit Fruit Fruit Fruit Fruit Fruit Fruit Fruit Fruit Fruit Fruit Fruit Fruit Fruit Blueberry Blueberry Blueberry Blueberry Blueberry Blueberry Blueberry Blueberry Blueberry Raspberry Blackberry Blackberry Blackberry Currant Cranberry Cranberry Cranberry Cranberry Raspberry Raspberry Raspberry 176 Proportion decayed ± SE 0.90 ± 0.06 0.90 ± 0.06 0.70 ± 0.06 0.70 ± 0.06 0.85 ± 0.10 0.70 ± 0.13 0.65 ± 0.17 0.70 ± 0.10 0.75 ± 0.15 0.40 ± 0.08 0.55 ± 0.05 0.15 ± 0.05 0.75 ± 0.05 0.45 ± 0.10 0.70 ± 0.10 0.40 ± 0.08 0.80 ± 0.00 0.60 ± 0.08 0.65 ± 0.10 0.45 ± 0.05 0.50 ± 0.17 Mean a separation abcdefghij abcdefghij abcdefghij abcdefghij abcdefghij abcdefghij abcdefghi abcdefghij abcdefghij abcdefghi abcdefghi abcdefghi abcdefghij abcdefghi abcdefghij abcdefghi abcdefghij abcdefghi abcdefghi abcdefghi abcdefghi Table D.1. (cont’d.) 22 23 24 25 26 a 1a Colac 03-109 Colac 03-110 Colsp 04-133 0001 Leelanau Ingham Ingham Ingham Van Buren Fruit Fruit Fruit Leaf Fruit Strawberry Strawberry Strawberry Grape Blueberry 0.80 ± 0.14 0.40 ± 0.08 0.60 ± 0.08 0.30 ± 0.13 0.95 ± 0.05 abcdefghij abcdefghi abcdefghi abcdefghi abcdefghij Values in columns followed by the same letter are not significantly different according to a Fisher’s LSD mean separation test (α = 0.05). 177 APPENDIX E. SCREENING DAUGHTER BLUEBERRY PLANTS FROM A DRAPER x JEWEL CROSS FOR RESISTANCE TO COLLETOTRICHUM ACUTATUM Table E.1. Anthracnose resistance ratings of detached blueberryfruits from several daughters of a Draper x Jewel cross. Fruits were harvested in Interlachen, FL between 26 April and 12 May in 2011, stored, and shipped overnight to Michigan State University and inoculated with 10 6 spores/ml by spraying (whole fruit assays) or by a 50 µl droplet on a cut fruit surface (cut fruit assays). Fruits incubated at 100% relative humidity at 22-24°C and rated at 3 days (for cut fruit assays) or 10 days ( for whole fruit assays). For each daughter 5 fruits were screened per replicate (25 in total, and n = 5) per type of inoculation. Cultivar/ Daughter number Jewel 1 2 3 4 5 7 9 10 12 13 14 15 16 17 18 19 20 21 22 25 Whole fruit assays Infection severity (percentage ± SE) 59.6 ± 12.0 34.8 ± 22.3 26.8 ± 6.8 38.8 ± 6.8 44.4 ± 12.8 47.6 ± 14.3 70.4 ± 21.5 61.0 ± 21.0 33.2 ± 13.3 50.0 ± 30.8 28.5 ± 25.5 26.4 ± 14.6 58.4 ± 33.2 55.3 ± 19.6 16.3 ± 15.5 48.4 ± 27.3 62.8 ± 27.1 40.8 ± 17.5 40.9 ± 27.4 22.6 ± 25.0 70.4 ± 35.0 Cut fruit assays Sporulation capacity Optical density 6 (λ 590 nm ± SE) (no. of conidia x 10 ± SE) 12.8 ± 7.9 0.19 ± 0.10 15.1 ± 19.7 0.24 ± 0.28 2.5 ± 4.1 0.04 ± 0.03 14.1 ± 6.3 0.16 ± 0.09 4.0 ± 7.2 0.06 ± 0.07 1.8 ± 2.1 0.04 ± 0.03 5.8 ± 4.7 0.11 ± 0.05 64.4 ± 11.0 0.48 ± 0.09 1.0 ± 2.4 0.02 ± 0.02 1.5 ± 1.7 0.03 ± 0.02 18.6 ± 8.7 0.24 ± 0.11 2.2 ± 1.5 0.03 ± 0.02 2.1 ± 3.6 0.05 ± 0.03 7.9 ± 8.0 0.10 ± 0.07 1.8 ± 1.6 0.11 ± 0.05 4.5 ± 1.9 0.07 ± 0.01 1.9 ± 2.5 0.04 ± 0.03 11.8 ± 4.5 0.17 ± 0.09 16.2 ± 2.6 0.20 ± 0.01 12.4 ± 12.8 0.14 ± 0.09 1.4 ± 1.5 0.04 ± 0.02 178 Table E.1. (cont’d.) 27 28 30 31 32 33 34 35 37 38 39 40 41 42 43 44 45 48 49 50 51 52 53 54 55 56 58 59 60 61 62 65 66 67 69 72 73 75 76 77 78 79 83 66.4 ± 31.5 36.4 ± 22.7 45.9 ± 14.0 35.6 ± 19.6 54.4 ± 28.5 24.0 ± 17.9 71.2 ± 12.9 38.0 ± 34.4 36.0 ± 29.5 47.6 ± 24.6 48.4 ± 23.5 71.0 ± 19.1 76.4 ± 17.7 26.8 ± 26.0 50.0 ± 10.5 93.2 ± 19.1 42.8 ± 17.8 42.4 ± 30.6 40.8 ± 32.3 54.0 ± 24.7 48.0 ± 27.4 51.6 ± 15.0 46.0 ± 48.2 59.2 ± 26.4 33.2 ± 20.2 46.8 ± 27.0 67.2 ± 24.6 69.3 ± 25.7 52.4 ± 8.6 82.8 ± 22.0 27.2 ± 10.0 51.2 ± 22.5 50.8 ± 21.6 88.8 ± 27.1 37.6 ± 16.6 64.4 ± 21.1 53.6 ± 8.0 58.0 ± 31.5 44.0 ± 11.4 30.8 ± 23.8 42.0 ± 17.0 57.2 ± 8.1 26.0 ± 13.0 3.6 ± 2.3 1.0 ± 1.3 5.1 ± 3.8 2.2 ± 2.6 7.2 ± 3.5 10.5 ± 2.4 7.8 ± 4.3 12.8 ± 13.5 0.4 ± 0.4 41.7 ± 29.0 1.5 ± 1.9 11.5 ± 6.0 4.5 ± 2.2 13.7 ± 9.7 4.7 ± 2.6 19.0 ± 16.6 7.7 ± 4.3 9.8 ± 6.3 16.9 ± 15.8 1.5 ± 2.3 2.6 ± 3.9 26.9 ± 15.4 3.7 ± 4.4 22.0 ± 11.2 4.8 ± 3.7 2.5 ± 1.5 22.6 ± 20.0 8.0 ± 6.6 11.5 ± 6.5 10.0 ± 7.6 10.7 ± 11.0 41.8 ± 14.9 6.8 ± 6.1 11.7 ± 8.7 7.0 ± 1.4 56.7 ± 5.2 5.4 ± 3.1 44.1 ± 11.1 2.1 ± 1.8 1.1 ± 1.9 0.6 ± 0.8 7.2 ± 5.1 2.7 ± 4.7 179 0.06 ± 0.05 0.05 ± 0.04 0.13 ± 0.09 0.25 ± 0.25 0.11 ± 0.04 0.11 ± 0.03 0.09 ± 0.03 0.16 ± 0.11 0.03 ± 0.02 0.35 ± 0.27 0.04 ± 0.03 0.17 ± 0.05 0.07 ± 0.05 0.12 ± 0.08 0.09 ± 0.15 0.14 ± 0.09 0.11 ± 0.05 0.11 ± 0.06 0.20 ± 0.10 0.07 ± 0.03 0.06 ± 0.04 0.27 ± 0.09 0.06 ± 0.05 0.20 ± 0.07 0.10 ± 0.06 0.05 ± 0.02 0.24 ± 0.19 0.09 ± 0.07 0.14 ± 0.04 0.09 ± 0.05 0.16 ± 0.32 0.40 ± 0.13 0.11 ± 0.04 0.19 ± 0.09 0.10 ± 0.07 0.53 ± 0.11 0.14 ± 0.25 0.37 ± 0.10 0.03 ± 0.01 0.03 ± 0.01 0.04 ± 0.04 0.09 ± 0.07 0.06 ± 0.05 Table E.1. (cont’d.) 85 86 87 88 89 90 93 94 99 101 64.0 ± 20.7 29.6 ± 14.3 59.6 ± 27.1 80.8 ± 43.4 46.4 ± 9.7 42.8 ± 14.4 32.0 ± 16.7 58.8 ± 19.1 66.8 ± 19.9 78.8 ± 12.1 34.8 ± 22.3 7.0 ± 5.4 17.4 ± 8.4 5.7 ± 6.6 31.7 ± 14.4 6.2 ± 4.4 15.1 ± 8.8 29.2 ± 8.9 4.2 ± 3.0 20.7 ± 15.3 180 0.32 ± 0.11 0.09 ± 0.02 0.17 ± 0.08 0.09 ± 0.08 0.30 ± 0.12 0.09 ± 0.08 0.15 ± 0.08 0.32 ± 0.07 0.08 ± 0.06 0.20 ± 0.13 Table E.2. Anthracnose resistance ratings of detached blueberryfruits from several daughters of a Draper x Jewel cross. Fruits were harvested in Corvallis, OR between 20 July and 11 August in 2011, stored, and shipped overnight to Michigan State University and inoculated with 10 6 spores/ml by a 50 µl droplet on a cut fruit surface. Fruits incubated at 100% relative humidity at 22-24°C and rated at 3 days. For each daughter 5 fruits were screened per replicate (25 in total, and n = 5) per type of inoculation. Optical density (λ 590 nm ± SE) 0.07 ± 0.01 0.18 ± 0.01 0.17 ± 0.01 0.12 ± 0.02 0.12 ± 0.04 0.20 ± 0.01 0.11 ± 0.03 0.11 ± 0.01 0.08 ± 0.01 0.12 ± 0.01 0.09 ± 0.01 0.07 ± 0.01 0.18 ± 0.00 0.09 ± 0.01 0.11 ± 0.01 0.17 ± 0.03 0.13 ± 0.01 0.14 ± 0.01 0.17 ± 0.00 0.13 ± 0.01 0.20 ± 0.01 0.10 ± 0.01 0.08 ± 0.01 0.07 ± 0.01 0.12 ± 0.01 0.14 ± 0.03 0.09 ± 0.01 0.13 ± 0.01 0.16 ± 0.01 0.07 ± 0.01 0.09 ± 0.02 Cultivar/Daughter number Draper 1 8 10 11 13 14 15 16 18 19 20 21 22 24 25 27 28 29 30 31 32 33 34 35 39 40 41 42 43 44 181 Table E.2. (cont’d.) 45 50 54 56 60 61 65 75 77 79 83 85 99 0.10 ± 0.02 0.13 ± 0.02 0.18 ± 0.02 0.13 ± 0.02 0.18 ± 0.01 0.08 ± 0.01 0.30 ± 0.01 0.32 ± 0.02 0.13 ± 0.01 0.10 ± 0.02 0.12 ± 0.01 0.22 ± 0.02 0.09 ± 0.02 182 0.35 Optical density (Table E.2 values) r = 0.744 0.30 P < 0.001 0.25 0.20 0.15 0.10 0.05 0.0 0.1 0.2 0.3 0.4 0.5 Optical density (Table E.1 values) Figure E.1. Relationship between the concentration of an aqueous suspension of Colletotrichum acutatum from anthracnose resistance cut fruit screenings (590 nm) (Table E.1 and E.2 values) of detached blueberry fruits from several daughters of a Draper x Jewel cross. 183 APPENDIX F. THE ROLE OF ETHYLENE IN BLUEBERRY ANTHRACNOSE FRUIT ROT RESISTANCE 0.16 Ethylene levels (ppm) per 25g of blueberries 0.14 n=4 0.12 0.10 0.08 0.06 0.04 0.02 0.00 Figure F.1. External differences in ethylene concentration in ‘Elliott’ and ‘Jersey’ fruits after inoculation of Colletotrichum acutatum or sterile deionized water. Fruits were spray-inoculated 6 with 10 conidia/ml, incubated at 100% relative humidity for 48 hours at 22-24°C, placed in quart sized mason jars, ethylene concentration was measured using then was measured 2 hours later using previously described methods (1). Error bars denote standard error of the mean. 184 1.6 Ethylene concentration (ppm) 1.4 n=3 1.2 1.0 0.8 0.6 0.4 0.2 0.0 Elliott Jersey Figure F.2. Internal differences in ethylene concentration within ‘Elliott’ and ‘Jersey’ fruits. Fruits were immersed in a saturated solution of KCl, a vaccum was applied and headspace analysis was used to measure the concentration of ethylene in accordance with previous methods (1, 2). Error bars denote standard error of the mean. 185 Ethylene concentration (ppm) 6 No. of conidia produced per berry (10 ) 30 n=5 25 20 15 10 5 0 1.2 1.0 n=3 0.8 0.6 0.4 0.2 0.0 0 A 2 4 6 8 10 12 14 16 Time (days) B Figure F.3. Effect of ethylene on anthracnose symptom development in ‘Elliott’ and ‘Jersey’ fruits. A) Amount of conidia produced per berry in ‘Elliott’ and ‘Jersey’ fruits in each chamber (ethylene or air). B) Levels of ethylene in both chambers throughout a 14 day period. Two 30gallon tanks were used to simulate an environment with 1ppm ethylene or air. Fruits were spray6 inoculated with 10 conidia/ml, incubated at 100% relative humidity for 14 days at 22-24°C, ethylene concentration was measured in accordance with previous methods (1). Closed circles represent the ethylene chamber, open circles represent the air chamber. Error bars denote standard error of the mean. References: (1) S. Jayanty et al. J. Amer. Soc. Hort. 127:998-1005, 2002. (2) L. Sun et al. Euphytica. 165:55-67, 2009. 186 APPENDIX G. FIRST REPORT OF GRAPE ROOT ROT CAUSED BY ROESLERIA SUBTERRANEA IN MICHIGAN Published in the peer-reviewed journal Plant Disease (July 2009, Volume 93, Page 765) In September of 2008, declining grapevines were observed in two vineyards (Vitis interspecific hybrids ‘Canada Muscat’ and ‘Chardonel’) in Fennville, MI. Affected vines were stunted with shortened internodes and yellow leaves; others had dead cordons or were entirely dead. The grower reported that vines were losing vigor and collapsing during a period of 2 years. Renewal trunks would collapse during the second season of growth. Several symptomatic vines showed signs of root decay. On one vine, distinctive fruiting bodies (mazaedia) were apparent on the roots below the soil line and resembled those of Roesleria subterranea (Weinm.) Redhead (2,3,4). The mazaedia were 4 to 5 mm tall and 1 mm in diameter with white-to-tan stipes and powdery, gray-to-greenish hemispherical heads. Ascospores were hyaline to light grayish green, disk shaped, and 4 to 6 μm in diameter. This fungus, also known as R. hypogaea Thüm & Pass., has been previously reported to cause grape root rot, vine decline, and replant problems in North America and Europe (2,3,4). The fungus was cultured from ascospores on potato dextrose agar (PDA). Colonies grew slowly (approximately 2 mm per day at 22 to 24°C) and were green in the center. No spores were produced. DNA was extracted, and internal transcribed spacer (ITS) sequences obtained by PCR were compared with known sequences using BLASTn (1). Our isolate had 100% ITS sequence homology to an isolate from Germany, Roesleria subterranea strain IB (Accession No. EF060304.1). To test for pathogenicity, the fungus was grown in potato dextrose broth for 14 days at 22 to 24°C. An aqueous suspension (0.1 g of fungus per ml) was 187 prepared by blending mycelia with sterile deionized water (SDW) in a food processor. Five twonode, rooted ‘Chardonnay’ cuttings (45 days old) were placed in the suspension. Five other cuttings were placed in SDW (control). After 3 h, plants were removed and repotted in fresh BACTO soil (Michigan Peat Company, Houston, TX) and kept in a growth chamber at 23°C with a 16/8-h light/dark cycle. After 25 days, inoculated plants were, on average, 63% smaller with 77% lower fresh-root weight and fewer fine roots than the control plants. The pathogen was recovered from surface-disinfested roots of all five inoculated plants but not from the control plants. Cultures appeared similar to the original isolate and their identity was confirmed by ITS sequencing. To our knowledge, this is the first report of R. subterranea on grapes in Michigan and the Midwest. This fungus needs to be recognized as a potential cause of vine decline and replant problems in Midwestern vineyards. References: (1) S. F. Altschul et al. J. Mol. Biol. 215:403, 1990. (2) W. Gärtel. Page 40 in: Compendium of Grape Diseases. R. C. Pearson and A. C. Goheen, eds. The American Phytopathological Society, St. Paul, MN, 1988. (3) M. Kirchmair et al. Mycol. Res. 112:1210, 2009. (4) J. R. Liberato et al. Pest and Diseases Image Library. Online publication, 2009. 188 APPENDIX H. FIRST REPORT OF PYTHIUM STERILUM CAUSING ROOT ROT OF BLUEBERRY IN THE UNITED STATES Published in the peer-reviewed journal Plant Disease (May 2011, Volume 95, Page 614) In September 2009, ~40 declining blueberry plants (Vaccinium corymbosum L. ‘Jersey’) were observed in a poorly drained area of a 30-year-old field near Fennville, MI. The stunted bushes had yellow leaves and defoliation; others were completely dead. The grower reported that the bushes had been declining over several years. Root samples tested positive in a Phytophthora ELISA test (Agdia Inc., Elkhart IN). Twenty root pieces (2 cm long and 2 to 3 mm in diameter) were surface disinfested and plated on Rye A agar; five yielded fungal-like colonies that were subcultured on potato dextrose agar (PDA). One isolate was white and grew slowly (3 to 4 mm/day at 22 to 24°C). Three isolates were white and grew faster (10 to 12 mm/day at 22 to 24°C) in a chrysanthemal pattern. The fifth was a Fusarium sp. DNA of the white colonies was extracted and the internal transcribed spacer (ITS) region was sequenced using ITS1 (5′TCCGTAGGTGAACCTGCGG-3′) and ITS4 (5′-TCCTCCGCTTATTGATATGC-3′) primers. The slow-growing morphotype had 99% identity to Phytophthora sp. isolate 92-209C (Accession No. EU106591) in GenBank but failed to induce symptoms in multiple inoculation tests. The fast-growing morphotype (Accession No. HQ398249) had 98% identity to Pythium sterilum UASWS0265 from declining alder stands in Poland (Accession No. DQ525089). Sequencing of the COXII gene using the FM66/FM58 primer set (3) yielded a product (Accession No. HQ721468) with 100% identity to P. sterilum GD32a from forest soil in Poland (Accession No. EF421185). Hyphae were hyaline, coenocytic, and 4 to 7 μm wide with some swellings at the 189 tips (7 to 9 μm wide). No sporangia, oogonia, or antheridia were observed. Mycelium tested positive in the ELISA test described above. According to Agdia Inc., 10 of 19 tested Pythium spp. have shown similar cross reactivity. Pythium spp. are known to cause root rot of blueberries in Oregon (2). In British Columbia, P. sterilum was commonly isolated from roots of declining blueberry bushes (4). P. sterilum Belbahri & Lefort only reproduces asexually (1). Our isolate was similar but did not produce sporangia in water or on PDA, V8 juice agar, Rye A agar, or water agar. Roots of 10 2-month-old ‘Bluecrop’ cuttings were placed in an aqueous suspension of rinsed mycelium (0.1 g/ml) from 21-day-old cultures grown in V8 broth or in sterile deionized water (control). After 1 h, plants were potted in peat moss/perlite (2:1) or autoclaved sand (five each) and placed in a glasshouse at 25°C. After 7 days, inoculated plants in both soil types had wilted or collapsed with significant necrosis on the roots and primary shoot. Control plants showed no symptoms. In a similar experiment with 6-month-old plants in sand, symptoms appeared after 10 to 12 days. The pathogen was recovered from surface-disinfested root and stem sections of all inoculated plants but not control plants and its identity was confirmed by sequencing of the ITS region. To our knowledge, this is the first report of P. sterilum on blueberries in the United States. While this disease appears to be uncommon in Michigan, it is a potential cause of plant decline, the diagnosis of which may be complicated by cross reactivity in ELISA testing. References: (1) L. Belbahri et al. FEMS Microbiol. Lett. 255:209, 2006. (2) D. R. Bryla and R. G. Linderman. HortScience 43:260, 2008. (3) F. N. Martin. Mycologia 92:711, 2000. (4) S. Sabaratnam. BC Plant Health Fund Final Report. B.C. Retrieved from http://www.agf.gov.bc.ca/cropprot/phf_final_report.pdf, 2008. 190 REFERENCES 191 REFERENCES 1. Adaskaveg, J., and R. Hartin. 1997. Characterization of Colletotrichum acutatum isolates causing anthracnose of almond and peach in California. Phytopathology 87:979-987. 2. Afanador-Kafuri, L., D. Minz, M. Maymon, and S. Freeman. 2003. Characterization of Colletotrichum isolates from tamarillo, passiflora, and mango in Colombia and identification of a unique species from the genus. Phytopathology 93:579-587 doi:doi:10.1094/PHYTO.2003.93.5.579. 3. Alkan, N., R. Fluhr, A. Sherman, and D. Prusky. 2008. Role of ammonia secretion and pH modulation on pathogenicity of Colletotrichum coccodes on tomato fruit. Mol. PlantMicrobe Interact. 21:1058-1066. 4. Alkan, N., O. Davydov, M. Sagi, R. Fluhr, and D. Prusky. 2009. Ammonium secretion by Colletotrichum coccodes activates host NADPH oxidase activity enhancing host cell death and fungal virulence in tomato fruits. Mol. Plant-Microbe Interact. 22:1484-1491 doi:doi:10.1094/MPMI-22-12-1484. 5. Allen, E., and J. Kuc. 1968. α-Solanine and α-chaconine as fungitoxic compounds in extracts of Irish potato tubers. Phytopathology 58:776-781. 6. Almada-Ruiz, E., M. A. Martinez-Tellez, M. M. Hernandez-Alamos, S. Vallejo, E. Primo-Yufera, and I. Vargas-Arispuro. 2003. Fungicidal potential of methoxylated flavones from citrus for in vitro control of Colletotrichum gloeosporioides, causal agent of anthracnose disease in tropical fruits. Pest Manage. Sci. 59:1245-1249. 7. Anand, T., R. Bhaskaran, T. Raguchander, R. Samiyappan, V. Prakasam, and C. Gopalakrishnan. 2009. Defense responses of chili fruits to Colletotrichum capsici and Alternaria alternata. Biologia Plantarum 53:553-559. 192 8. Arroyo, F. T., J. Moreno, P. Daza, L. Boianova, and F. Romero. 2007. Antifungal activity of strawberry fruit volatile compounds against Colletotrichum acutatum. J. Agric. Food Chem. 55:5701-5707 doi:10.1021/jf0703957. 9. Arroyo, F. T., J. Moreno, G. García-Herdugo, B. D. Santos, C. Barrau, M. Porras, C. Blanco, and F. Romero. 2005. Ultrastructure of the early stages of Colletotrichum acutatum infection of strawberry tissues. Can. J. Bot. 83:491-500. 10. Arts, I. C. W., B. van de Putte, and P. C. H. Hollman. 2000. Catechin contents of foods commonly consumed in The Netherlands. 1. Fruits, vegetables, staple foods, and processed foods. J. Agric. Food Chem. 48:1746-1751. 11. Atschul, S., W. Gish, W. Miller, E. Myers, and D. Lipman. 1990. Basic local alignment search tool. J. Mol. Biol 215:403-410. 12. Bailey, J. A., and B. J. Deverall. 1971. Formation and activity of phaseollin in the interaction between bean hypocotyls (Phaseolus vulgaris) and physiological races of Colletotrichum lindemuthianum. Physiol. Plant Pathol. 1:435-449. 13. Batson, W., and K. Roy. 1982. Species of Colletotrichum and Glomerella pathogenic to tomato fruit. Plant Dis. 66:1153-1155. 14. Bellaire, L. d., M. Chillet, C. Dubois, and X. Mourichon. 2000. Importance of different sources of inoculum and dispersal methods of conidia of Colletotrichum musae, the causal agent of banana anthracnose, for fruit contamination. Plant Pathol. 49:782-790. 15. Ben-Daniel, B., D. Bar-Zvi, and L. Tsror. 2011. Pectate lyase affects pathogenicity in natural isolates of Colletotrichum coccodes and in pelA gene disrupted and gene overexpressing mutant lines. Mol. Plant Pathol. DOI: 10.1111/j.1364-3703.2011.00740.x. 193 16. Bergstrom, G., and R. Nicholson. 2000. The biology of Colletotrichum graminicola and maize anthracnose. Pages 374-394. in: Colletotrichum: Host Specificity, Pathology, and Host-Pathogen Interaction D. Prusky, S. Freeman, and M. B. Dickman, eds. APS Press, St. Paul, MN. 17. Bernstein, B., E. Zehr, R. Dean, and E. Shabi. 1995. Characteristics of Colletotrichum from peach, apple, pecan, and other hosts. Plant Dis. 79:478-482. 18. Bobe, G., B. Wang, N. Seeram, M. Nair, and L. Bourquin. 2006. Dietary anthocyaninrich tart cherry extract inhibits intestinal tumorigenesis in APC(Min) mice fed suboptimal levels of sulindac. J Agric Food Chem 54:9322 - 28. 19. Bowen, J., M. Templeton, K. Sharrock, R. Crowhurst, and E. Rikkerink. 1995. Gene inactivation in the plant pathogen Glomerella cingulata: three strategies for the disruption of the pectin lyase gene pnIA. Mol. Gen. Genet. 246:196-205. 20. Brady, C. 1987. Fruit ripening. Annu. Rev. Plant Physiol. 38:155-178. 21. Broome, J., J. English, J. Marois, B. Latorre, and J. Aviles. 1995. Development of an infection model for Botrytis bunch rot of grapes based on wetness duration and temperature. Phytopathology 85:97-102. 22. Brown, G. E. 1975. Factors affecting postharvest development of Colletotrichum gloeosporioides in citrus fruits. Phytopathology 65:404-409. 23. Brown, G. E. 1977. Ultrastructure of penetration of ethylene-degreened Robinson tangerines by Colletotrichum gloeosporioides after ethylene treatment. Phytopathology 67:700-706. 24. Brown, G. E. 1978. Hypersensitive response of orange-colored Robinson tangerines to Colletotrichum gloeosporioides after ethylene treatment. Phytopathology 68:700-706. 194 25. Brown, S. H., O. Yarden, N. Gollop, S. Chen, A. Zveibil, E. Belausov, and S. Freeman. 2008. Differential protein expression in Colletotrichum acutatum: changes associated with reactive oxygen species and nitrogen starvation implicated in pathogenicity on strawberry. Mol. Plant Pathol. 9:171-190. 26. Bulger, M., M. Ellis, and L. Madden. 1987. Influence of temperature and wetness duration on infection of strawberry flowers by Botrytis cinerea and disease incidence of fruit originating from infected flowers. Phytopathology 77:1225-1230. 27. Byrne, J. M., M. K. Hausbeck, and R. Hammerschmidt. 1997. Conidial germination and appressorium formation of Colletotrichum coccodes on tomato foliage. Plant Dis. 81:715718. 28. Campos, Â., A. Ferreira, M. Hampe, I. Antunes, N. Brancão, E. Silveira, J. Silva, and V. Osório. 2003. Induction of chalcone synthase and phenylalanine ammonia-lyase by salicylic acid and Colletotrichum lindemuthianum in common bean. Brazil. J. Plant Physiol. 15:129-134. 29. Carisse, O., G. Bourgeois, and J. Duthie. 2000. Influence of temperature and leaf wetness duration on infection of strawberry leaves by Mycosphaerella fragariae. Phytopathology 90:1120-1125. 30. Carson, M., and A. Hooker. 1981. Inheritance of resistance to stalk rot of corn caused by Colletotrichum graminicola. Phytopathology 71:1190-1196. 31. Casado-Diaz, A., S. Encinas-Villarejo, B. Santos, E. Schiliro, E. M. Yubero-Serrano, F. Amil-Ruiz, M. I. Pocovi, F. Pliego-Alfaro, G. Dorado, and M. Rey. 2006. Analysis of strawberry genes differentially expressed in response to Colletotrichum infection. Physiol. Plant. 128:633-650. 195 32. Centis, S., I. Guillas, N. Séjalon, M.-T. Esquerré-Tugayé, and B. Dumas. 1997. Endopolygalacturonase genes from Colletotrichum lindemuthianum: cloning of CLPG2 and comparison of its expression to that of CLPG1 during saprophytic and parasitic growth of the fungus. Mol. Plant-Microbe Interact. 10:769-775 doi:doi:10.1094/MPMI.1997.10.6.769. 33. Chakraborty, S., D. Ratcliff, and F. McKay. 1990. Anthracnose of Stylosanthes scabra: effect of leaf surface wetness on disease severity. Plant Dis. 74:379-384. 34. Chen, N., P. H. Goodwin, and T. Hsiang. 2003. The role of ethylene during the infection of Nicotiana tabacum by Colletotrichum destructivum. J. Exp. Bot. 54:2449-2456. 35. Chillet, M., O. Hubert, and L. Bellaire. 2006. Relationship between ripening and the development of banana anthracnose caused by Colletotrichum musae (Berck. and Curt.) Arx. J. Phytopathol. 154:143-147. 36. Chillet, M., O. Hubert, and L. de Lapeyre de Bellaire. 2007. Relationship between physiological age, ripening and susceptibility of banana to wound anthracnose. Crop Prot. 26:1078-1082. 37. Cipollini, M., and E. Stiles. 1992. Antifungal activity of ripe ericaceous fruits: Phenolicacid interactions and palatability for dispersers. Biochem. Syst. Ecol. 20:501-514. 38. Cipollini, M., and E. Stiles. 1992. Relative risks of fungal rot for temperate ericaceous fruits: effects of seasonal variation on selection for chemical defense. Can. J. Bot. 70:1868-1877. 39. Cipollini, M., and E. Stiles. 1993. Fruit rot, antifungal defense, and palatability of fleshy fruits for frugivorous birds. Ecology 74:751-762. 196 40. Cline, W. O. 1997. Postharvest infection of blueberries during handling. Acta Hortic. 446:319-326. 41. Coates, L., I. Muirhead, J. Irwin, and D. Gowanlock. 1993. Initial infection processes by Colletotrichum gloeosporioides on avocado fruit. Mycol. Res. 97:1363-1370. 42. Connor, A. M., J. J. Luby, J. F. Hancock, S. Berkheimer, and E. J. Hanson. 2002. Changes in fruit antioxidant activity among blueberry cultivars during cold-temperature storage. J. Agric. Food Chem. 50:893-898. 43. Curry, K., M. Abril, J. Avant, and B. Smith. 2002. Strawberry anthracnose: Histopathology of Colletotrichum acutatum and C. fragariae. Phytopathology 92:10551063. 44. Dana, M. M., J. A. Pintor-Toro, and B. Cubero. 2006. Transgenic tobacco plants overexpressing chitinases of fungal origin show enhanced resistance to biotic and abiotic stress agents. Plant Physiol. 142:722. 45. Daugrois, J., C. Lafitte, J. Barthe, and A. Touze. 1990. Induction of B-1, 3 glucanase and chitinase activity in compatible and incompatible interactions between Colletotrichum lindemuthianum and bean cultivars. J. Phytopathol. 130:225-234. 46. Daykin, M., and R. Milholland. 1984. Infection of blueberry fruit by Colletotrichum gloeosporioides. Plant Dis. 68:948-950. 47. Daykin, M., and R. Milholland. 1984. Histopathology of ripe rot caused by Colletotrichum gloeosporioides on muscadine grape. Phytopathology 74:1339-1341. 48. de Pascual-Teresa, S., D. A. Moreno, and C. García-Viguera. 2010. Flavanols and anthocyanins in cardiovascular health: a review of current evidence. Int. J. Mol. Sci. 11:1679-1703. 197 49. DeMarsay, A., and P. V. Oudemans. 2003. Colletotrichum acutatum infections in dormant highbush blueberry buds. Phytopathology 93:S20. 50. Denoyes-Rothan, B., G. Guérin, E. Lerceteau-Kohler, and G. Risser. 2005. Inheritance of resistance to Colletotrichum acutatum in Fragaria × ananassa. Phytopathology 95:405412. 51. Denoyes, B., and A. Baudry. 1995. Species identification and pathogenicity study of French Colletotrichum strains isolated from strawberry using morphological and cultural characteristics. Phytopathology 85:53-57. 52. Die, J. V., M. A. Dita, F. Krajinski, C. I. González-Verdejo, D. Rubiales, M. T. Moreno, and B. Román. 2007. Identification by suppression subtractive hybridization and expression analysis of Medicago truncatula putative defense genes in response to Orobanche crenata parasitization. Physiol. Mol. Plant Pathol. 70:49-59. 53. Dieguez-Uribeondo, J., H. Forster, A. Soto-Estrada, and J. E. Adaskaveg. 2005. Subcuticular-intracellular hemibiotrophic and intercellular necrotrophic development of Colletotrichum acutatum on almond. Phytopathology 95:751-758. 54. Diéguez-Uribeondo, J., H. Förster, and J. E. Adaskaveg. 2003. Digital image analysis of internal light spots of appressoria of Colletotrichum acutatum. Phytopathology 93:923930 doi:doi:10.1094/PHYTO.2003.93.8.923. 55. Diéguez-Uribeondo, J., H. Förster, and J. E. Adaskaveg. 2008. Visualization of localized pathogen-Induced pH modulation in almond tissues infected by Colletotrichum acutatum using confocal scanning laser microscopy. Phytopathology 98:1171-1178 doi:doi:10.1094/PHYTO-98-11-1171. 198 56. Dodd, J. C., A. B. Estrada, J. Matcham, P. Jeffries, and M. J. Jeger. 1991. The effect of climatic factors on Colletotrichum gloeosporioides, causal agent of mango anthracnose, in the Philippines. Plant Pathol. 40:568-575. 57. Doyle, J. J., and J. L. Doyle. 1990. Isolation of plant DNA from fresh tissue. Focus 12:13-15. 58. Droby, S., D. Prusky, B. Jacoby, and A. Goldman. 1986. Presence of antifungal compounds in the peel of mango fruits and their relation to latent infections of Alternaria alternata. Physiol. Mol. Plant Pathol. 29:173-183. 59. Drori, N., H. Kramer-Haimovich, J. Rollins, A. Dinoor, Y. Okon, O. Pines, and D. Prusky. 2003. A combination of external pH and nitrogen assimilation affects secretion of the virulence factor pectate lyase by Colletotrichum gloeosporioides. Appl. Environ. Microbiol 69:3258-3262. 60. Durango, D., W. Quiñones, F. Torres, Y. Rosero, J. Gil, and F. Echeverri. 2002. Phytoalexin accumulation in Colombian bean varieties and aminosugars as elicitors. Molecules 7:817-832. 61. Durrands, P. K., and R. M. Cooper. 1988. The role of pectinases in vascular wilt disease as determined by defined mutants of Verticillium albo-atrum. Physiol. Mol. Plant Pathol. 32:363-371. 62. Eastburn, D., and W. Gubler. 1992. Effects of soil moisture and temperature on the survival of Colletotrichum acutatum. Plant Dis. 76:841-842. 63. Ehlenfeldt, M. 2003. 'Elliott' highbush blueberry. J. Am. Pomol. Soc. 57:2-6. 199 64. Ehlenfeldt, M. K., J. J. Polashock, A. W. Stretch, and M. Kramer. 2006. Leaf disk infection by Colletotrichum acutatum and its relation to fruit rot in diverse blueberry germplasm. HortScience 41:270-271. 65. Evans, K., W. Nyquist, and R. Latin. 1992. Model based on temperature and leaf wetness duration for establishment of Alternaria leaf blight of muskmelon. Phytopathology 82:890-895. 66. Feussner, K., I. Feussner, I. Leopold, and C. Wasternack. 1997. Isolation of a cDNA coding for an ubiquitin-conjugating enzyme UBC1 of tomato--the first stress-induced UBC of higher plants. FEBS letters 409:211. 67. Freeman, S., T. Katan, and E. Shabi. 1998. Characterization of Colletotrichum species responsible for anthracnose diseases of various fruits. Plant Dis. 82:596-605 doi:doi:10.1094/PDIS.1998.82.6.596. 68. Freeman, S., Z. Shalev, and J. Katan. 2002. Survival in soil of Colletotrichum acutatum and C. gloeosporioides pathogenic on strawberry. Plant Dis. 86:965-970. 69. Freeman, S., D. Minz, E. Jurkevitch, M. Maymon, and E. Shabi. 2000. Molecular analyses of Colletotrichum species from almond and other fruits. Phytopathology 90:608614. 70. Gao, L., and G. Mazza. 1994. Quantitation and distribution of simple and acylated anthocyanins and other phenolics in blueberries. J. Food Sci. 59:1057-1059. 71. Gardes, M., and T. D. Bruns. 1993. ITS primers with enhanced specificity for basidiomycetes - application to the identification of mycorrhizae and rusts. Mol. Ecol. 2:113-118. 200 72. Geffroy, V., M. Sévignac, P. Billant, M. Dron, and T. Langin. 2008. Resistance to Colletotrichum lindemuthianum in Phaseolus vulgaris: a case study for mapping two independent genes. Theor. Appl. Genet. 116:407-415. 73. Geffroy, V., M. Sévignac, J. C. F. De Oliveira, G. Fouilloux, P. Skroch, P. Thoquet, P. Gepts, T. Langin, and M. Dron. 2000. Inheritance of partial resistance against Colletotrichum lindemuthianum in Phaseolus vulgaris and co-localization of quantitative trait loci with genes involved in specific resistance. Mol. Plant-Microbe Interact. 13:287296. 74. Gillett, J. M., and A. C. Schilder. 2009. Environmental requirements for infection of blueberry fruit by Colletotrichum acutatum. Acta Hortic. 810:355-360. 75. Gomes, S., P. Prieto, P. Martins-Lopes, T. Carvalho, A. Martin, and H. Guedes-Pinto. 2009. Development of Colletotrichum acutatum on tolerant and susceptible Olea europaea L. cultivars: A microscopic analysis. Mycopathologia 168:203-211. 76. Goodwin, P. H., R. P. Oliver, and T. Hsiang. 2004. Comparative analysis of expressed sequence tags from Malva pusilla, Sorghum bicolor, and Medicago truncatula infected with Colletotrichum species. Plant Science 167:481-489. 77. Grayer, R. J., and J. B. Harborne. 1994. A survey of antifungal compounds from higher plants, 1982-1993. Phytochemistry 37:19-42. 78. Gregori, R., M. Mari, P. Bertolini, J. A. S. Barajas, J. B. Tian, and J. M. Labavitch. 2008. Reduction of Colletotrichum acutatum infection by a polygalacturonase inhibitor protein extracted from apple. Postharvest Biol. Technol. 48:309-313. 79. Guetsky, R., I. Kobiler, X. Wang, N. Perlman, N. Gollop, G. Avila-Quezada, I. Hadar, and D. Prusky. 2005. Metabolism of the flavonoid epicatechin by laccase of 201 Colletotrichum gloeosporioides and its effect on pathogenicity on avocado fruits. Phytopathology 95:1341-1348 doi:doi:10.1094/PHYTO-95-1341. 80. Häkkinen, S. H., and A. R. Törrönen. 2000. Content of flavonols and selected phenolic acids in strawberries and Vaccinium species: influence of cultivar, cultivation site and technique. Food Res. Int. 33:517-524. 81. Hammerschmidt, R., and J. Kuc. 1982. Lignification as a mechanism for induced systemic resistance in cucumber. Physiol. Plant Pathol. 20:61-71. 82. Hammerschmidt, R., E. Nuckles, and J. Kug. 1982. Association of enhanced peroxidase activity with induced systemic resistance of cucumber to Colletotrichum lagenarium. Physiol. Plant Pathol. 20:73-82. 83. Hancock, J. F., and J. H. Siefker. 1982. Levels of inbreeding in highbush blueberry cultivars. HortScience 17:363-366. 84. Hancock, J. F., P. Callow, S. Serce, E. J. Hanson, and R. Beaudry. 2008. Effect of cultivar, controlled atmosphere storage, and fruit ripeness on the long-term storage of highbush blueberries. HortTechnology 18:199-205. 85. Hartung, J., C. Burton, and D. Ramsdell. 1981. Epidemiological studies of blueberry anthracnose disease caused by Colletotrichum gloeosporioides. Phytopathology 71:449453. 86. Hassan, M. K., E. K. Dann, D. E. Irving, and L. M. Coates. 2007. Concentrations of constitutive alk(en)ylresorcinols in peel of commercial mango varieties and resistance to postharvest anthracnose. Physiol. Mol. Plant Pathol. 71:158-165. 202 87. Herbers, K., P. Meuwly, W. B. Frommer, J. P. Metraux, and U. Sonnewald. 1996. Systemic acquired resistance mediated by the ectopic expression of invertase: possible hexose sensing in the secretory pathway. Plant Cell 8:793-803 doi:10.1105/tpc.8.5.793. 88. Horsfall, J. G., and A. E. Dimond. 1957. lnteractions of tissue sugar, growth substances, and disease susceptibility. Z. Pflanzenkr. Pflanzenschutz 64:415-421. 89. Iamsupasit, N., S. Chakraborty, D. Cameron, and S. Adkins. 1993. Components of quantitative resistance to anthracnose (Colletotrichum gloeosporioides) in tetraploid accessions of the pasture legume Stylosanthes hamata. Aust. J. Exp. Agric. 33:855-860. 90. Ichiyanagi, T., C. Tateyama, K. Oikawa, and T. Konishi. 2000. Comparison of anthocyanin distribution in different blueberry sources by capillary zone electrophoresis. Biol. Pharm. Bull. 23:492-497. 91. Jayaprakasam, B., S. K. Vareed, L. K. Olson, and M. G. Nair. 2005. Insulin secretion by bioactive anthocyanins and anthocyanidins present in fruits. J. Agric. Food Chem. 53:2831. 92. Ji, C., and J. Kuc. 1995. Purification and characterization of an acidic B-1, 3-glucanase from cucumber and its relationship to systemic disease resistance Induced by Colletotrichum lagenarium and Tobacco Necrosis Virus. Mol. Plant Microbe Interact. 8:899-905. 93. Johnston, D., B. Williamson, and G. McMillan. 1994. The interaction in planta of polygalacturonases from Botrytis cinerea with a cell wall-bound polygalacturonaseinhibiting protein (PGIP) in raspberry fruits. J. Exp. Bot. 45:1837. 203 94. Joseph, J. A., N. A. Denisova, D. Bielinski, D. R. Fisher, and B. Shukitt-Hale. 2000. Oxidative stress protection and vulnerability in aging: putative nutritional implications for intervention. Mech. Ageing Dev. 116:141-153. 95. Kader, F., B. Rovel, M. Girardin, and M. Metche. 1996. Fractionation and identification of the phenolic compounds of highbush blueberries (Vaccinium corymbosum, L.). Food Chem. 55:35-40. 96. Kahkonen, M. P., A. I. Hopia, and M. Heinonen. 2001. Berry phenolics and their antioxidant activity. J. Agric. Food Chem. 49:4076-4082. 97. Kalt, W., and J. E. McDonald. 1996. Chemical composition of lowbush blueberry cultivars. J. Am. Soc. Hortic. Sci. 121:142-146. 98. Kalt, W., J. McDonald, R. Ricker, and X. Lu. 1999. Anthocyanin content and profile within and among blueberry species. Can. J. Plant Sci. 79:617-624. 99. Kelly, J. D., and V. A. Vallejo. 2004. A comprehensive review of the major genes conditioning resistance to anthracnose in common bean. HortScience 39:1196-1207. 100. Kikuchi, T., and K. Masuda. 2009. Class II chitinase accumulated in the bark tissue involves with the cold hardiness of shoot stems in highbush blueberry (Vaccinium corymbosum L.). Scientia Horticulturae 120:230-236. 101. Kim, K.-H., J.-B. Yoon, H.-G. Park, E. W. Park, and Y. H. Kim. 2004. Structural modifications and programmed cell death of chili pepper fruit related to resistance responses to Colletotrichum gloeosporioides infection. Phytopathology 94:1295-1304 doi:doi:10.1094/PHYTO.2004.94.12.1295. 204 102. Kim, S., J. Yoon, J. Do, and H. Park. 2008. A major recessive gene associated with anthracnose resistance to Colletotrichum capsici in chili pepper (Capsicum annuum L.). Breed. Sci. 58:137-141. 103. Kim, Y. S., J. Y. Park, K. S. Kim, M. K. Ko, S. J. Cheong, and B. J. Oh. 2002. A thaumatin-like gene in nonclimacteric pepper fruits used as molecular marker in probing disease resistance, ripening, and sugar accumulation. Plant Mol. Biol. 49:125-135. 104. King, W. T., L. V. Madden, M. A. Ellis, and L. L. Wilson. 1997. Effects of temperature on sporulation and latent period of Colletotrichum spp. infecting strawberry fruit. Plant Dis. 81:77-84. 105. Kleweno, D. 2007. Michigan Fruit Inventory 2006-2007. http://www.nass.usda.gov/Statistics_by_State/Michigan/Publications/Michigan_Rotation al_Surveys/mi_fruit07/blueberry.pdf. 106. Kleweno, D. 2010. Michigan Agricultural Statistics 2009-2010. Michigan Department of Agriculture and USDA National Agricultural Statistics Service. Lansing, MI. 107. Ko, M. K., W. B. Jeon, K. S. Kim, H. H. Lee, H. H. Seo, Y. S. Kim, and B. J. Oh. 2005. A Colletotrichum gloeosporioides-induced esterase gene of nonclimacteric pepper (Capsicum annuum) fruit during ripening plays a role in resistance against fungal infection. Plant Mol. Biol. 58:529-541. 108. Kumari, M. V. R., M. Hiramatsu, and M. Ebadi. 1998. Free radical scavenging actions of metallothionein isoforms I and II. Free Radical Res. 29:93-101. 109. La tti, A. K., L. Jaakola, K. R. Riihinen, and P. S. Kainulainen. 2009. Anthocyanin and flavonol variation in bog bilberries (Vaccinium uliginosum L.) in Finland. J. Agric. Food Chem 58:427-433. 205 110. Lafitte, C., J. Barthe, X. Gansel, G. Dechamp-Guillaume, C. Faucher, D. Mazau, and M. Esquerré-Tugayé. 1993. Differential induction by endopolygalacturonase of β-1, 3glucanases in Phaseolus vulgaris isoline susceptible and resistant to Colletotrichum lindemuthianum race β. Mol. Plant-Microbe Interact. 6:628-634. 111. Lahey, K. A., R. Yuan, J. K. Burns, P. P. Ueng, L. W. Timmer, and K. R. Chung. 2004. Induction of phytohormones and differential gene expression in citrus flowers infected by the fungus Colletotrichum acutatum. Mol. Plant-Microbe Interact. 17:1394-1401. 112. Leandro, L. F. S., M. L. Gleason, F. W. Nutter Jr, S. N. Wegulo, and P. M. Dixon. 2001. Germination and sporulation of Colletotrichum acutatum on symptomless strawberry leaves. Phytopathology 91:659-664. 113. Lee, S., M. Milgroom, and J. Taylor. 1988. A rapid, high yield mini-prep method for isolation of total genomic DNA from fungi. Fungal Genet. Newsl. 35:23-24. 114. Lerceteau-Köhler, E., G. Guérin, and B. Denoyes-Rothan. 2005. Identification of SCAR markers linked to Rca2 anthracnose resistance gene and their assessment in strawberry germplasm. Theor. Appl. Genet. 111:862-870 doi:10.1007/s00122-005-0008-1. 115. Leterrier, M., F. J. Corpas, J. B. Barroso, L. M. Sandalio, and L. A. Del Río. 2005. Peroxisomal monodehydroascorbate reductase. Genomic clone characterization and functional analysis under environmental stress conditions. Plant Physiol. 138:2111. 116. Liu, M., X. Li, C. Weber, C. Lee, J. Brown, and R. Liu. 2002. Antioxidant and antiproliferative activities of raspberries. J. Agric. Food Chem. 50:2926-2930. 117. Lo, S. C., K. De Verdier, and R. L. Nicholson. 1999. Accumulation of 3deoxyanthocyanidin phytoalexins and resistance to Colletotrichum sublineolum in sorghum. Physiol. Mol. Plant Pathol. 55:263-273. 206 118. Lo, S. C., J. D. Hipskind, and R. L. Nicholson. 1999. cDNA cloning of a sorghum pathogenesis-related protein (PR-10) and differential expression of defense-related genes following inoculation with Cochliobolus heterostrophus or Colletotrichum sublineolum. Mol. Plant-Microbe Interact. 12:479-489 doi:doi:10.1094/MPMI.1999.12.6.479. 119. Mahasuk, P., P. W. J. Taylor, and O. Mongkolporn. 2009. Identification of two new genes conferring resistance to Colletotrichum acutatum in Capsicum baccatum. Phytopathology 99:1100-1104 doi:doi:10.1094/PHYTO-99-9-1100. 120. Mathews, R. H., P. R. Pehrsson, and M. Farhat-Sabet. 1987. Sugar content of selected foods: Individual and total sugars. USDA Home Economics Research Report Number 48:7. 121. Meazza, G., F. E. Dayan, and D. E. Wedge. 2003. Activity of Quinones on Colletotrichum Species. J. Agric. Food Chem 51:3824-3828 doi:10.1021/jf0343229. 122. Mellersh, D., I. Foulds, V. Higgins, and M. Heath. 2002. H2O2 plays different roles in determining penetration failure in three diverse plant–fungal interactions. Plant J. 29:257268. 123. Miles, T. D., and A. M. C. Schilder. 2008. Anthracnose fruit rot (ripe rot). Michigan State University Extension Bulletin E-3039:4 pages. 124. Miles, T. D., P. S. Wharton, and A. C. Schilder. 2009. Cytological and chemical evidence for an active resistance response to infection by Colletotrichum acutatum in 'Elliott' blueberries. Acta Hortic. 810:361-368. 125. Miles, T. D., B. Day, and A. C. Schilder. 2011. Identification of differentially expressed genes in a resistant versus a susceptible blueberry cultivar after infection by Colletotrichum acutatum. Mol. Plant Pathol. 12:463-77. 207 126. Miles, T. D., J. Hancock, P. Callow, and A. C. Schilder. 2011. Evaluation of screening methods and fruit composition in relation to anthracnose fruit rot resistance in blueberries. Plant Pathol. (in press). 127. Miles, T. D., M. Nair, C. Vandervoort, and A. M. C. Schilder. 2011. Antifungal compounds in ripe fruit from a resistant blueberry cultivar suppress infection by Colletotrichum acutatum. Phytopathology 101:S120. 128. Milholland, R. D. 1995. Anthracnose fruit rot (ripe rot). Pages 17. in: Compendium of Blueberry and Cranberry Diseases F. L. Caruso, and D. C. Ramsdell, eds. APS Press, St. Paul, MN, USA. 129. Miyara, I., H. Shafran, M. Davidzon, A. Sherman, and D. Prusky. 2010. pH regulation of ammonia secretion by Colletotrichum gloeosporioides and its effect on appressorium formation and pathogenicity. Mol. Plant Microbe Interact. 23:304-316 doi:10.1094/mpmi-23-3-0304. 130. Monroe, J., J. Santini, and R. Latin. 1997. A model defining the relationship between temperature and leaf wetness duration, and infection of watermelon by Colletotrichum orbiculare. Plant Dis. 81:739-742. 131. Moral, J., K. Bouhmidi, and A. Trapero. 2008. Influence of fruit maturity, cultivar susceptibility, and inoculation method on infection of olive fruit by Colletotrichum acutatum. Plant Dis. 92:1421-1426. 132. Muirhead I.F., and B. J. Deverall. 1984. Evaluation of 3,4-dihyroxybenzaldehyde, dopamine and its oxidation products as inhibitors of Colletotrichum musae (Berk. & Curt.) Arx in green banana fruits. . Aust. J. Bot. 32:575–82. 208 133. Nakajima, J., Y. Tanaka, M. Yamazaki, and K. Saito. 2001. Reaction mechanism from leucoanthocyanidin to anthocyanidin 3-glucoside, a key reaction for coloring in anthocyanin biosynthesis. J. Biologic. Chem. 276:25797. 134. Narusaka, Y., M. Narusaka, P. Park, Y. Kubo, T. Hirayama, M. Seki, T. Shiraishi, J. Ishida, M. Nakashima, A. Enju, T. Sakurai, M. Satou, M. Kobayashi, and K. Shinozaki. 2004. RCH1, a locus in Arabidopsis that confers resistance to the hemibiotrophic fungal pathogen Colletotrichum higginsianum. Mol. Plant-Microbe Interact. 17(7):749-762 doi:doi:10.1094/MPMI.2004.17.7.749. 135. Nicholson, R., S. Van Scoyoc, E. Williams, and J. Kuc. 1977. Host-pathogen interactions preceding the hypersensitive reaction of Malus sp. to Venturia inaequalis. Phytopathology 67:108-114. 136. Norman, D., and J. Strandberg. 1997. Survival of Colletotrichum acutatum in soil and plant debris of leatherleaf fern. Plant Dis. 81:1177-1180. 137. O'Connell, R., J. Bailey, and D. Richmond. 1985. Cytology and physiology of infection of Phaseolus vulgaris by Colletotrichum lindemuthianum. Physiol. Plant Pathol. 27:7598. 138. O'Connell, R. J., A. B. Uronu, G. Waksman, C. Nash, J. P. R. Keon, and J. A. Bailey. 1993. Hemibiotrophic infection of Pisum sativum by Colletotrichum truncatum. Plant Pathol. 42:774-783. 139. O’Connell, R., S. Perfect, B. Hughes, R. Carzaniga, J. Bailey, and J. Green. 2000. Dissecting the cell biology of Colletotrichum infection processes. Pages 57-76. in: Colletotrichum: Host Specificity, Pathology, and Host-Pathogen Interaction D. Prusky, S. Freeman, and M. B. Dickman, eds. APS Press, St. Paul, MN, USA. 209 140. Oh, B. J., M. K. Ko, I. Kostenyuk, B. Shin, and K. S. Kim. 1999. Coexpression of a defensin gene and a thionin-like gene via different signal transduction pathways in pepper and Colletotrichum gloeosporioides interactions. Plant Mol. Biol. 41:313-319. 141. Oh, B. J., M. K. Ko, Y. S. Kim, K. S. Kim, I. Kostenyuk, and H. K. Kee. 1999. A cytochrome P450 gene is differentially expressed in compatible and incompatible interactions between pepper (Capsicum annuum) and the anthracnose fungus, Colletotrichum gloeosporioides. Mol. Plant-Microbe Interact. 12:1044-1052. 142. Orengo, C., A. Michie, S. Jones, D. Jones, M. Swindells, and J. Thornton. 1997. CATH-a hierarchic classification of protein domain structures. Structure 5:1093-1109. 143. Osman, A., D. P. Makris, and P. Kefalas. 2008. Investigation on biocatalytic properties of a peroxidase-active homogenate from onion solid wastes: An insight into quercetin oxidation mechanism. Process Biochem 43:861-867. 144. Pakdeevaraporn, P., S. Wasee, P. W. J. Taylor, and O. Mongkolporn. 2005. Inheritance of resistance to anthracnose caused by Colletotrichum capsici in Capsicum. Plant Breeding 124:206-208 doi:10.1111/j.1439-0523.2004.01065.x. 145. Park, S.-C., J.-Y. Kim, J.-K. Lee, I. Hwang, H. Cheong, J.-W. Nah, K.-S. Hahm, and Y. Park. 2009. Antifungal mechanism of a novel antifungal protein from pumpkin rinds against various fungal pathogens. J. Agric. Food Chem. 57:9299-9304 doi:10.1021/jf902005g. 146. Peres, N. A., L. W. Timmer, J. E. Adaskaveg, and J. C. Correll. 2005. Lifestyles of Colletotrichum acutatum. Plant Dis. 89:784-796 doi:doi:10.1094/PD-89-0784. 210 147. Pinhero, R., and G. Paliyath. 2001. Antioxidant and calmodulin-inhibitory activities of phenolic components in fruit wines and its biotechnological implications. Food Biotechnol. 15:179-192. 148. Podila, G. K., L. M. Rogers, and P. E. Kolattukudy. 1993. Chemical signals from avocado surface wax trigger germination and appressorium formation in Colletotrichum gloeosporioides. Plant Physiol. 103:267-272. 149. Polashock, J., R. Saftner, and M. Kramer. 2007. Postharvest highbush blueberry fruit antimicrobial volatile profiles in relation to anthracnose fruit rot resistance. J. Am. Soc. Hortic. Sci. 132:859-868. 150. Polashock, J. J., M. K. Ehlenfeldt, A. W. Stretch, and M. Kramer. 2005. Anthracnose fruit rot resistance in blueberry cultivars. Plant Dis. 89:33-38. 151. Pring, R., C. Nash, M. Zakaria, and J. Bailey. 1995. Infection process and host range of Colletotrichum capsici. Physiol. Mol. Plant Pathol. 46:137-152. 152. Prior, R. L., S. A. Lazarus, G. Cao, H. Muccitelli, and J. F. Hammerstone. 2001. Identification of procyanidins and anthocyanins in blueberries and cranberries (Vaccinium spp.) using high-performance liquid chromatography/mass spectrometry. J. Agric. Food Chem. 49:1270-1276. 153. Prusky, D. 1996. Pathogen quiescence in postharvest diseases. Annu. Rev. Phytopathol. 34:413-434. 154. Prusky, D., and N. Keen. 1993. Involvement of preformed antifungal compounds in the resistance of subtropical fruits to fungal decay. Plant Dis. 77:114-119. 211 155. Prusky, D., N. Keen, and I. Eaks. 1983. Further evidence for the involvement of a preformed antifungal compound in the latency of Colletotrichum gloeosporioides on unripe avocado fruits. Physiol. Plant Pathol. 22:189-198. 156. Prusky, D., R. A. Plumbley, and I. Kobiler. 1991. The relationship between antifungal diene levels and fungal inhibition during quiescent infection of unripe avocado fruits by Colletotrichum gloeosporioides. Plant Pathol. 40:45-52. 157. Prusky, D., J. McEvoy, B. Leverentz, and W. Conway. 2001. Local modulation of host pH by Colletotrichum species as a mechanism to increase virulence. Mol. Plant-Microbe Interact. 14:1105-1113. 158. Prusky, D., I. Kobiler, Y. Fishman, J. Sims, S. Midland, and N. Keen. 1991. Identification of an antifungal compound in unripe avocado fruits and its possible involvement in the quiescent infections of Colletotrichum gloeosporioides. J. Phytopathol. 132:319-327. 159. Prusky, D., I. Koblier, R. Ardi, D. Beno-Moalem, N. Yakoby, and N. Keen. 2000. Resistance mechanisms of subtropical fruits to Colletotrichum gloeosporioides. Pages 232-244. in: Colletotrichum: Host Specificity, Pathology, and Host-Pathogen Interaction D. Prusky, S. Freeman, and M. B. Dickman, eds. APS Press, St. Paul, MN, USA. 160. Ri-He, P., Y. Quan-Hong, X. Ai-Sheng, F. Hui-Qin, L. Xian, and P. You-Liang. 2003. Ubiquitin-conjugating enzyme (E2) confers rice UV protection through phenylalanine ammonia-lyase gene promoter unit. Acta Botanica Sinica 45:1351-1358. 161. Robards, K., P. D. Prenzler, G. Tucker, P. Swatsitang, and W. Glover. 1999. Phenolic compounds and their role in oxidative processes in fruits. Food Chem. 66:401-436. 212 162. Roberts, R., and J. Snow. 1984. Histopathology of cotton boll rot caused by Colletotrichum capsici. Phytopathology 74:390-397. 163. Rodriguez-Saona, L. E., and R. E. Wrolstad. 2001. Extraction, isolation, and purification of anthocyanins. Current protocols in food analytical chemistry UNIT F1.1. 164. Sacher, J. A. 1973. Senescence and postharvest physiology. Annu. Rev. Plant Physiol. 24:197-224. 165. Salzman, R. A., I. Tikhonova, B. P. Bordelon, P. M. Hasegawa, and R. A. Bressan. 1998. Coordinate accumulation of antifungal proteins and hexoses constitutes a developmentally controlled defense response during fruit ripening in grape. Plant Physiol. 117:465-472 doi:10.1104/pp.117.2.465. 166. Salzman, R. A., T. Fujita, K. Zhu-Salzman, P. M. Hasegawa, and R. A. Bressan. 1999. An improved RNA isolation method for plant tissues containing high levels of phenolic compounds or carbohydrates. Plant Mol. Biol. Rep. 17:11-17. 167. Sanogo, S., R. E. Stevenson, and S. P. Pennypacker. 2003. Appressorium formation and tomato fruit infection by Colletotrichum coccodes. Plant Dis. 87:336-340 doi:doi:10.1094/PDIS.2003.87.4.336. 168. Sanogo, S., S. P. Pennypacker, R. E. Stevenson, and A. A. MacNab. 1997. Weather variables associated with infection of tomato fruit by Colletotrichum coccodes. Plant Dis. 81:753-756 doi:10.1094/pdis.1997.81.7.753. 169. Schilder, A., J. Gillett, and J. Woodworth. 2002. The kaleidoscopic nature of blueberry fruit rots. Acta Hortic. 574:81-83. 213 170. Schuh, W. 1991. Influence of temperature and leaf wetness period on conidial germination in vitro and infection of Cercospora kikuchii on soybean. Phytopathology 81:1315-1318. 171. Schuh, W. 1993. Influence of interrupted dew periods, relative humidity, and light on disease severity and latent infections caused by Cercospora kikuchii on soybean. Phytopathology 83:109-113. 172. Seeram, N., L. Adams, Y. Zhang, R. Lee, D. Sand, H. Scheuller, and D. Heber. 2006. Blackberry, black raspberry, blueberry, cranberry, red raspberry, and strawberry extracts inhibit growth and stimulate apoptosis of human cancer cells in vitro. J Agric Food Chem 54:9329 - 39. 173. Shen, S., P. Goodwin, and T. Hsiang. 2001. Hemibiotrophic infection and identity of the fungus, Colletotrichum destructivum, causing anthracnose of tobacco. Mycol. Res. 105:1340-1347. 174. Shin, W., S. Park, and E. Kim. 2006. Protective effect of anthocyanins in middle cerebral artery occlusion and reperfusion model of cerebral ischemia in rats. Life sciences 79:130137. 175. Simmonds, J. 1965. A study of the species of Colletotrichum causing ripe fruit rots in Queensland. Qld. J. Agric. Sci. 22:437-459. 176. Singh, J., and S. Sharma. 1981. Screening and chemical basis of resistance in guava varieties to anthracnose (Glomerella cingulata). Haryana J. Hortic. Sci. 10:155-157. 177. Smith, B. J., J. B. Magee, and C. L. Gupton. 1996. Susceptibility of rabbiteye blueberry cultivars to postharvest diseases. Plant Dis. 80:215-218. 214 178. Stanghellini, M., and M. Aragaki. 1966. Relation of periderm formation and callose deposition to anthracnose resistance in papaya fruit. Phytopathology 56:444-450. 179. Stoner, G., L. Wang, N. Zikri, T. Chen, S. Hecht, C. Huang, C. Sardo, and J. Lechner. 2007. Cancer prevention with freeze-dried berries and berry components. Semin Cancer Biol 17:403 - 10. 180. Stotz, H. U., J. J. A. Contos, A. L. T. Powell, A. B. Bennett, and J. M. Labavitch. 1994. Structure and expression of an inhibitor of fungal polygalacturonases from tomato. Plant Mol Biol 25:607-617. 181. Sullivan, J., K. Shirasu, and X. Deng. 2003. The diverse roles of ubiquitin and the 26S proteasome in the life of plants. Nature Rev. Genet. 4:948-958. 182. Takahama, U., and S. Hirota. 2000. Deglucosidation of quercetin glucosides to the aglycone and formation of antifungal agents by peroxidase-dependent oxidation of quercetin on browning of onion scales. Plant Cell Physiol. 41:1021-1029. 183. Takahara, H., K. Toyoda, G. Tsuji, Y. Kubo, Y. Inagaki, Y. Ichinose, and T. Shiraishi. 2005. Identification of genes expressed during spore germination of Mycosphaerella pinodes. J. Gen. Plant Pathol. 71:190-195. 184. Takai, R., N. Matsuda, A. Nakano, K. Hasegawa, C. Akimoto, N. Shibuya, and E. Minami. 2002. EL5, a rice N acetylchitooligosaccharide elicitor responsive RING H2 finger protein, is a ubiquitin ligase which functions in vitro in co operation with an elicitor responsive ubiquitin conjugating enzyme, OsUBC5b. Plant J. 30:447-455. 185. Tattersall, D. B., R. van Heeswijck, and P. B. Hoj. 1997. Identification and characterization of a fruit-specific, thaumatin-like protein that accumulates at very high 215 levels in conjunction with the onset of sugar accumulation and berry softening in grapes. Plant Physiol. 114:759-769 doi:10.1104/pp.114.3.759. 186. Templeton, M. D., R. A. Dixon, C. J. Lamb, and M. A. Lawton. 1990. Hydroxyprolinerich glycoprotein transcripts exhibit different spatial patterns of accumulation in compatible and incompatible interactions between Phaseolus vulgaris and Colletotrichum lindemuthianum. Plant Physiol. 94:1265-1269 doi:10.1104/pp.94.3.1265. 187. Thibaud, M., S. Gineste, L. Nussaume, and C. Robaglia. 2004. Sucrose increases pathogenesis-related PR-2 gene expression in Arabidopsis thaliana through an SAdependent but NPR1-independent signaling pathway. Plant Physiol. and Biochem. 42:8188. 188. Timm, E., G. Brown, P. Armstrong, R. Beaudry, and A. Shirazi. 1996. Portable instrument for measuring firmness of cherries and berries. Appl. Eng. Agric. 12:71-77. 189. Timmer, L., and G. Brown. 2000. Biology and control of anthracnose diseases of citrus. Pages 300-316. in: Colletotrichum: Host Specificity, Pathology, and Host-Pathogen Interaction D. Prusky, S. Freeman, and M. B. Dickman, eds. APS Press, St. Paul, MN, USA. 190. van Loon, L. C., and E. A. van Strien. 1999. The families of pathogenesis-related proteins, their activities, and comparative analysis of PR-1 type proteins. Physiol. Mol. Plant Pathol. 55:85-97. 191. Vanderplank, J. E. 1984. Sink-induced loss of resistance. Pages 107-116. in: Disease Resistance in Plants 2nd edition J. E. Vanderplank, ed. Academic Press, London. 216 192. Verma, N., L. MacDonald, and Z. K. Punja. 2006. Inoculum prevalence, host infection and biological control of Colletotrichum acutatum: causal agent of blueberry anthracnose in British Columbia. Plant Pathol. 55:442-450. 193. Verma, N., L. MacDonald, and Z. K. Punja. 2007. Environmental and host requirements for field infection of blueberry fruits by Colletotrichum acutatum in British Columbia. Plant Pathol. 56:107-113. 194. Vorwerk, S., S. Somerville, and C. Somerville. 2004. The role of plant cell wall polysaccharide composition in disease resistance. Trends Plant Sci. 9:203-209. 195. Waage, S. K., and P. A. Hedin. 1985. Quercetin 3-O-galactosyl-(1Æ6)-glucoside, a compound from narrowleaf vetch with antibacterial activity. Phytochemistry 24:243-245. 196. Wang, H., M. G. Nair, G. M. Strasburg, Y. C. Chang, A. M. Booren, J. I. Gray, and D. L. DeWitt. 1999. Antioxidant and antiinflammatory activities of anthocyanins and their aglycon, cyanidin, from tart cherries. J. Natural Prod. 62:294-296. 197. Wang, Y., M. Hamburger, J. Gueho, and K. Hostettmann. 1989. Antimicrobial flavonoids from Psiadia trinervia and their methylated and acetylated derivatives. Phytochemistry 28:2323-2327. 198. Wattad, C., D. Kobiler, A. Dinoor, and D. Prusky. 1997. Pectate lyase of Colletotrichum gloeosporioides attacking avocado fruits: cDNA cloning and involvement in pathogenicity. Physiol. Mol. Plant Pathol. 50:197-212. 199. Wharton, P. S., and A. M. Julian. 1996. A cytological study of compatible and incompatible interactions between Sorghum bicolor and Colletotrichum sublineolum. New Phytol. 134:25-34. 217 200. Wharton, P. S., and J. Diéguez-Uribeondo. 2004. The biology of Colletotrichum acutatum. An. Jard. Bot. Madr. 61:3-22. 201. Wharton, P. S., and A. M. C. Schilder. 2008. Novel infection strategies of Colletotrichum acutatum on ripe blueberry fruit. Plant Pathol. 57:122-134. 202. Wharton, P. S., J. S. Dickman, and A. M. C. Schilder. 2002. Timing of spore release by Colletotrichum acutatum in Michigan blueberry fields. Phytopathology 92:S86. 203. Wharton, P. S., A. Iezzoni, and A. L. Jones. 2003. Screening cherry germ plasm for resistance to leaf spot. Plant Dis. 87:471-477 doi:doi:10.1094/PDIS.2003.87.5.471. 204. Wijesundera, R., J. Bailey, R. Byrde, and A. Fielding. 1989. Cell wall degrading enzymes of Colletotrichum lindemuthianum: Their role in the development of bean anthracnose. Physiol. Mol. Plant Pathol. 34:403-413. 205. Wijesundera, R. L. C., J. A. Bailey, and R. J. W. Byrde. 1984. Production of pectin lyase by Colletotrichum lindemuthianum in culture and in infected bean (Phaseolus vulgaris) tissue. J. Gen. Microbiol. 130:285-290 doi:10.1099/00221287-130-2-285. 206. Wilson, L., L. Madden, and M. Ellis. 1990. Influence of temperature and wetness duration on infection of immature and mature strawberry fruit by Colletotrichum acutatum. Phytopathology 80:111-116. 207. Wu, L., J. Damicone, J. Duthie, and H. Melouk. 1999. Effects of temperature and wetness duration on infection of peanut cultivars by Cercospora arachidicola. Phytopathology 89:653-659. 208. Xu, H., and S. Lee. 2001. Activity of plant flavonoids against antibiotic resistant bacteria. Phytother Res. 15:39-43. 218 209. Yakoby, N., I. Kobiler, A. Dinoor, and D. Prusky. 2000. pH regulation of pectate lyase secretion modulates the attack of Colletotrichum gloeosporioides on avocado fruits. Appl. Environ. Microbiol. 66:1026-1030 doi:10.1128/aem.66.3.1026-1030.2000. 210. Yang, X., L. Wilson, L. Madden, and M. Ellis. 1990. Rain splash dispersal of Colletotrichum acutatum from infected strawberry fruit. Phytopathology 80:590-595. 211. Yao, C., W. S. Conway, R. Ren, D. Smith, G. S. Ross, and C. E. Sams. 1999. Gene encoding polygalacturonase inhibitor in apple fruit is developmentally regulated and activated by wounding and fungal infection. Plant Mol Biol 39:1231-1241. 212. Yoshida, S., and A. Shirata. 1999. Survival of Colletotrichum dematium in soil and infected mulberry leaves. Plant Dis. 83:465-468. 213. Yoshida, S., and T. Tsukiboshi. 2002. Shoot blight and leaf spot of blueberry anthracnose caused by Colletotrichum acutatum. J. Gen. Plant Pathol. 68:246-248. 214. You, B.-J., and K.-R. Chung. 2007. Phenotypic characterization of mutants of the citrus pathogen Colletotrichum acutatum defective in a PacC-mediated pH regulatory pathway. FEMS Microbiol. Lett. 277:107-114 doi:10.1111/j.1574-6968.2007.00951.x. 219