* i‘ w 3 This is to certify that the thesis entitled BIOLOGY AND PATHOGENICITY FACTORS OF RUTSTROEMIA FLOCCOSUM AND THE EFFECTS OF LIGHTWEIGHT ROLLING ON DOLLAR SPOT DISEASE INCIDENCE IN CREEPING BENTGRASS PUTTING GREENS. presented by Paul Ryan Giordano has been accepted towards fulfillment of the requirements for the MS. degree in Plant Pathology M‘l/WM ‘6' Majo'r Preyéssoy's Signature flaw” 3‘ 20 /o J / Date MSU is an Affinnative ActionEqual Opportunity Employer LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE MAY 1 1 2012 042517 5/08 K:IProj/Acc&Pres/ClRC/DateDue.indd BIOLOGY AND PATHOGENICITY FACTORS OF RUTSTROEMIA FLOCCOSUM AND THE EFFECTS OF LIGHTWEIGHT ROLLING ON DOLLAR SPOT DISEASE INCIDENCE IN CREEPING BENTGRASS PUTTING GREENS. By Paul Ryan Giordano A THESIS Submitted to Michigan State University In partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Plant Pathology 2010 - llll ABSTRACT BIOLOGY AND PATHOGENICITY FACTORS OF RUTSTROEMIA FLOCCOSUM AND THE EFFECTS OF LIGHTWEIGHT ROLLING ON DOLLAR SPOT DISEASE INCIDENCE IN CREEPING BENTGRASS PUTTING GREENS. By Paul Ryan Giordano A study to investigate the effects of light-weight rolling on the reduction of dollar spot was conducted between 2008 and 2009. Treatments rolled in the afternoon exhibited similar disease reduction as treatments rolled in the morning when compared to the control. A twice day'l rolling treatment exhibited significantly less disease than all other treatments and resulted in greater rootzone soil volumetric water content (V WC) in both 2008 and 2009, when compared to the non-rolled control. Rolled treatments exhibited increases in fatty acid abundances associated with common bacteria when rootzone soil phospholipid fatty acids (PLF A) were analyzed. Furthermore, a general trend towards higher total bacterial PLFA abundances was present in rolled treatments. These results suggest that daily, season-long, light-weight rolling on putting greens may contribute to dollar spot reductions through the alteration of microbial populations in the upper rootzone. Additionally, afiemoon rolling treatments suggest mechanisms other than dew and guttation removal to be responsible for disease reductions. R. floccosum isolates were examined for oxalic acid (0A) and cell wall degrading enzymes (CWDE) production. Non-virulent isolate Sh12B displayed high levels of CWDES and 0A production to a lesser extent when compare to virulent isolates. This suggests that OA and CWDES may not be sole pathogenicity factors necessary for R. floccosum infection on turfgrass. To my loving parents, Patricia and Paul for all that you have done to help shape the person that I am today; I can never thank you enough. iii ACKNOWLEDGMENTS It would be impossible to thank every individual who has made a positive impact on my life in the past two years, but there are many people I owe a sincere thanks to for their aide and support during this time of my academic career. First and foremost, to my major professor, Dr. Joe Vargas, thank you for your inspiring wisdom and guidance; and for giving me the opportunity to learn and grow as an independent researcher. I owe tremendous thanks to my committee members Dr. Ray Hammerschmidt and Dr. Thom Nikolai for their keen advice and ideas, without which, none of this would be possible. I would like to thank my parents Patricia and Paul Giordano for your unconditional love and support, both personal and financial. My two sisters Lisa and Lori, thank you for all that you do. Much love to my wonderful fiancee Courtney Rose for your unyielding support and encouragement throughout this endeavor that keeps me grounded and driven. To my co-workers and friends, Nancy and Ron, thank you for always lending a helping hand and being there with listening ears and engaging ideas. To Liewei Yan, thank you for the insightful conversations and camaraderie. To Kevin Budsberg and Dr. Teri Balser who aided in the PLFA analysis. To Dr. Chris Vandervoort for her assistance with HPLC. To the whole MSU turf team, including everyone at the Hancock Turfgrass Research Center for being so helpful and setting a standard of excellence unmatched in this great industry. To my fellow graduate students and everyone in the department of Plant Pathology I extend sincere thanks and appreciation. iv TABLE OF CONTENTS Page LIST OF TABLES ................................................................................ vi LIST OF FIGURES ............................................................................... vii LITERATURE REVIEW .......................................................................... 1 Dollar Spot .......................................................................................... 1 Extracellular Enzyme and Acid Production by R. F loccosum ............................... 6 Lightweight Rolling .............................................................................. 10 Microbial Populations in Turfgrass Rootzones .............................................. 13 Literature Cited ................................................................................... 18 CHAPTER ONE DOLLAR SPOT REDUCTION THROUGH LIGHTWEIGHT ROLLING ON CREEPING BENTGRASS PUTTING GREENS Abstract ............................................................................................ 26 Introduction ....................................................................................... 27 Materials and Methods .......................................................................... 29 Results and Discussion .......................................................................... 32 Conclusions ....................................................................................... 48 Literature Cited ................................................................................... 53 CHAPTER TWO TURFGRASS ROOTZONE MICROBIAL RESPONSES TO ROLLING Abstract ............................................................................................ 55 Introduction ....................................................................................... 56 Materials and Methods .......................................................................... 59 Results and Discussion .......................................................................... 63 Conclusions ....................................................................................... 69 Literature Cited .................................................................................. 75 CHAPTER THREE ACID AND CELL WALL DEGRADIN G ENZYME PRODUCTION BY RUT STROEMI F LOCCOSUM Abstract ............................................................................................ 82 Introduction ....................................................................................... 83 Materials and Methods .......................................................................... 85 Results and Discussion .......................................................................... 90 Conclusions ....................................................................................... 97 Literature Cited ................................................................................. 102 Appendix .......................................................................................... 105 Table 1.01 Table 1.02 Table 2.01 Table 2.02 Table 2.03 Table 3.01 Table 3.02 Table 3.03 LIST OF TABLES Means and LSD comparisons for treatment effects of rolling on dollar spot disease incidence on different dates in 2008 ........................... 34 Means and LSD comparisons for treatment effects of rolling on dollar spot disease incidence at different dates in 2009 ............................ 35 Relative abundance (mol %) of PLFA indicators in turfgrass root zones .............................................................................. 65 PLFA markers used for taxonomic microbial groupings .................. 67 Bacterial organisms with antagonistic effects against common turfgrass pathogens ......................................................................... 72 Rutstroemiafloccosum isolates studied and their source of obtainment ........................................................................ 85 Disease ratings of R. floccosum isolates on swards of creeping bentgrass (Agrostis palustris Huds. cv. Crenshaw) ...................................... 91 Mean culture pH measurements of four R. floccosum isolates growing in potato dextrose broth for 15 days ............................................. 93 vi Figure 1.01 Figure 1.02 Figure 1.03 Figure 1.04 Figure 1.05 Figure 1.06 Figure 1.07 Figure 1.08 Figure 1.09 Figure 1.10 Figure 2.01 Figure 3.01 LIST OF FIGURES Seasonal dollar spot means among different rolling treatments on creeping bentgrass (Agrostis palustris Huds.) putting green plots in 2008 ............................................................................... 36 Seasonal dollar spot means among different rolling treatments on creeping bentgrass (Agrostis palustris Huds.) putting green plots in 2009 ............................................................................... 37 Quality ratings on creeping bentgrass (Agrostis palustris Huds) turf plots for rolling treatments on different dates in 2008 ............................ 40 Seasonal turfgrass quality ratings for rolled treatments on creeping bentgrass (Agrostis palustris Huds.) putting green plots in 2008. . . . .....41 Quality ratings on creeping bentgrass (Agrostis palustris Huds) turf plots for rolling treatments on different dates in 2009 ............................ 42 Seasonal turfgrass quality means for rolled treatments on creeping bentgrass (Agrostis palustris Huds.) putting green plots in 2009.. . ......43 Percent volumetric water content means for rolling treatments on creeping bentgrass (A grostis palustris Huds.) plots in 2008 ............... 44 Seasonal percent volumetric water content means for rolling treatments on creeping bentgrass (Agrostis palustris Huds.) plots in 2008 ............................................................................... 45 Percent volumetric water content means for rolling treatments on creeping bentgrass (A grostis palustris Huds.) plots in 2009 ............... 46 Seasonal percent volumetric water content means for rolling treatments on creeping bentgrass (A grostis palustris Huds.) plots in 2009 ........... 47 PLFA group abundance (mol %) in turfgrass root zones subjected to different rolling treatments ..................................................... 68 Mean media color change ratings from acid production of four R. floccosum isolates grown on bromophenol blue-amended potato dextrose agar for 15 days .................................................................. 92 vii Figure 3.02 Figure 3.03 Figure 3.04 Figure 3.05 Figure A.1 Figure A.2 Oxalic acid production of R. floccosum isolates grown in potato dextrose broth for 15 days ................................................................. 94 Oxalic acid production related to mycelial dry weight of R. floccosum isolates grown in potato dextrose broth for 15 days. ....................... 95 Pectin methylesterase activity of four R. floccosum isolates grown in liquid culture with creeping bentgrass (A grostis palustris Huds.) cell walls ............................................................................... 96 Cellulase activity of four R. floccosum isolates grown in liquid culture with creeping bentgrass (Agrostis palustris Huds.) cell walls ............ 97 Oxalate oxidase specific activity in creeping bentgrass (Agrostis palustris Huds.) tissue from different rolling treatments ................. 105 Peroxidase specific activity in creeping bentgrass (Agrostis palustris Huds.) tissue from different rolling treatments ............................ 106 viii LITERATURE REVIEW DOLLAR SPOT Causal Agent. Dollar spot is a disease of turfgrass found throughout the world (Smiley et al., 1992). It is caused by the fungal pathogen Rutstroemia floccosum syn. Sclerotinia homoeocarpa F .T. Bennet (Smiley et al., 1992) and is considered the most economically important disease on golf courses nationwide. More money is spent to manage dollar spot than any other pathogen on golf courses in the United States (Vargas, 2005). When grown in vitro, the fungus exhibits a characteristic white mat of fast- growing, fluffy mycelium, which turns to a gray or brown as the culture ages. After 15- 30 days of growth in culture, the white mycelium can be accompanied by planes of dark stroma (Smiley et al., 1992). Symptoms. The name dollar spot was given to the disease due to the symptoms it produces on the host turfgrass tissue, about the size of a silver dollar (Vargas, 2005). The disease appears on putting greens and low mown turf as small, straw colored patches, less than 6 cm, and commonly 1-3 cm, in diameter. In taller grass, like home lawns and athletic fields, the patches may reach 15 cm in diameter. As the disease progresses and conditions favorable for development persist, individual spots may often coalesce into irregularly shaped patches, destroying large areas of turf. Lesions on individual leaf blades are distinguished by pale, bleached, water soaked bands, bound by a tan or reddish-brown margin at the outer edge of the lesion in bentgrass, Ketucky bluegrass, fine-leaf fescue, zoysiagrass, and bermudagrass. During periods of leaf wetness, such as early morning hours when heavy dew and guttation fluid are present, the active mycelium is ofien visible as white, cottony or cobwebby growth (Smiley et al., 1992; Vargas, 2005; Beard and Tani, 1997). Mycelia disappear as the leaves dry. Disease Cycle. Rutstroemia floccosum can survive as dormant mycelia in previously infected tissue. The disease is spread primarily by mowers, maintenance equipment, and people carrying mycelia and infected plant tissue on shoes or other equipment. The mycelium rapidly invades healthy leaf tissue and extends into moisture- saturated air spaces surrounding infected leaves when the turfgrass environment favors activity of the fungus. Mycelia] infection is thought to occur through out ends of leaves or through stomata on leaf blades (Smiley et al., 1992). Although infertile aborted apothecial primordial have been produced by the fungus in culture and fertile apothecia have been observed from Festuca turf in field conditions (Baldwin and Newell, 1992), neither sexual nor asexual spores are considered important in the epidemiology of the disease (Fenstermacher, 1979; Jackson, 1973) Epidemiology. Dollar spot is known to infect many species of turfgrass (Smiley et al., 1992). In the northern US, it is most damaging to creeping bentgrass (Agrostis palustris Huds.), annual bluegrass (Poa annua), colonial bentgrass (Agrostis tenuis), and fine-leaf fescues (F estuca spp.) (Vargas, 2005). The disease can be prevalent on turfgrass throughout the US. from the latter part of the spring through the end of the fall. Most epidemics occur in July and again in late August through September (Smiley et al., 1992). Conditions that favor the development and infection of dollar spot are temperatures of 15-30 C. Many different vegetative compatibility groups exist that cause the disease at various temperature and humidity regimes (Vargas, 2005). Guttation fluid is thought to be an ideal food source for the firngus. Guttation water is a fluid rich in carbohydrates and amino acids that is exuded from the tip of the grass blade through hydathodes when turgor pressure builds up in the plant (Vargas, 2005). Dollar spot is more severe under conditions of low nitrogen fertility, dry soils, and/or water stress. The fungus can survive as mycelia and stroma on leaf tissue when conditions for infection are unfavorable. DOLLAR SPOT MANAGEMENT gm Dollar spot management is often an integrated plan implemented by golf course superintendents, and turf managers alike. Cultural, chemical, and biological control techniques are all used in the management of this disease. Methods of cultural control such as fertilization, irrigation, and cultivation are often effective in reducing the amount of diseased turfgrass, and can aide in the reduction of chemical inputs. Maintaining high nitrogen fertility, especially when applied at times of high disease pressure, reduces disease (Vargas, 2005). Maintaining proper soil moisture levels, thus avoiding plant drought stress, is thought to also contribute to disease control. Couch and Bloom (1960) found low soil moisture to be important in the development of dollar spot, and Howard and Smith reported more dollar spot in seasons with less rainfall (cited in Vargas, 2005). One of the most frequently used techniques in the cultural management of dollar spot is the removal of dew or plant exudates known as guttation fluid from the leaf blades in the early morning hours after mowing. Prolonged leaf wetness duration (LWD) increases the severity of dollar spot and other fungal diseases by providing a favorable environment for fungal penetration of leaf tissues (Huber and Gillespie, 1992; Williams et al., 1996; Gross et al., 1998; Walsh, 2000; Uddin et al., 2003). Walsh (2000) found that the minimum LWD for the development of dollar spot for one isolate of S. homoeocarpa on creeping bentgrass (Agrostis palustris Huds.) was 12 h at 175°C. A longer LWD was required for infection when the temperature was lower. Dew or guttation removal is widely implemented in the turfgrass industry, particularly on golf course putting greens, and is thought to inhibit fungal mycelium growth by reducing LWD or the amount of food source available for the fungus. Rolling, whipping, dragging, or syringing are among some of the techniques used by turf managers to accomplish adequate removal or dispersion of the moisture. Many studies have elucidated the benefits of early morning mowing, syringing, and other dew removal techniques in order to decrease LWD (Williams and Powell, 1995; Ellram et al, 2007) ultimately reducing dollar spot incidence. Williams and Powell (1995) noted that guttation droplets secrete from wound exudates and these droplets are rich in nutrients that pathogens may use during hyphal growth. Release of these exudates may be exacerbated in the early dawn hours due to a combination of a fresh wound being produced by mowing and that turgor pressure may be high at this time. Nikolai (2005) proposed the process of dew/guttation removal as an underlying mechanism behind dollar spot reduction among other hypotheses such as altered soil microbial populations, and plant phytoalexin defense responses (Nikolai, 2002) Chemical. While cultural management of dollar spot is relatively effective, the most efficient dollar spot management can be obtained when good cultural practices are combined with a sound fungicide program. Contact fungicides, like chlorothalonil, are effective at controlling dollar spot for short (7-14 days) periods of time. Resistance to contact fungicides has not been reported and is unlikely to develop due to the multi-site mode of action these fungicides exhibit (Golembiewski et al., 1995). Effective systemic fungicides include propiconazole, fenarimol, iprodione and vinclozolin, among others. Systemic fungicides are able to give longer periods of control than contact fungicides, but have a much greater likelihood of becoming ineffective due to pathogen resistance developing. Resistance can be defined as reduced efficacy and shortened control intervals of a previously sensitive fungal population. Since 1972, resistance has been reported to many major systemic fungicide classes, which include the benzimidizoles (Warren, 1974), dicarboximides (Detweiler et al., 1983), and the demethylation inhibitors (DMI’s) (Golembiewski et al., 1995). R. floccosum resistance to demethylation inhibitors was confirmed only 11 years after this family of fimgicides was introduced to manage this disease (Golembiewski et al., 1995). Today, demethylation inhibitors are ineffective against dollar spot on many golf courses where this fungicide family has been used extensively (Vargas, 2005). Biological. The introduction of the biological control agent Pseudomonas aureofaciens strain Tx-l sold as Spot-less (Turf Science Laboratories Inc., National City, CA) for use in managing dollar spot has been a successful addition to the cultural and chemical means available to manage dollar spot on golf courses (Dwyer, 1999). To date, P. aureofaciens has been only mildly successful as a biological agent used in field situations for the control of dollar spot. Satisfactory disease suppression during years of light to moderate dollar spot infection has been shown with composted materials containing Enterobacter cloaca (Nelson et al., 1991, Nelson et al., 1992) as well. EXTRACELLULAR ENZYME AND ACID PRODUCTION BY RUTSTROEMIA FLOCCOSUM While many products are available to control the dollar spot disease, much remains unknown about the true life cycle and epidemiology of the fungus. One key to disease control is having a keen understanding of the infection process of the pathogen of interest. While a great deal of effort and research has been dedicated to dollar spot control and temporal variations (i.e. vegetative compatibility groups), infection mechanisms are yet to be elucidated. Knowledge of these mechanisms and pathogenicity factors are important, and can play critical roles in the development of novel, effective methods of control. Infection Meflanisms. Plant pathogens must overcome many obstacles in order to successfully infect their host. Some bacteria, for example, enter their hosts through wounds or natural openings such as hydathodes in the plant. Other microorganisms may be able to infect by means of mechanical pressure created by specialized structures that aide in the facilitation of microbial entry (Agrios, 1988). Plant cells are surrounded by rigid walls in which a multitude of polysaccharides with specific structures are interconnected (Baur et al., 1973; Burke et al., 1974; Wilder and Albersheim, 1973). These interconnected polysaccharides form initial barriers which serve as primary defense mechanisms against invading pathogens. The evolutionary struggle between pathogen and host most likely gave rise to cell wall degrading enzymes being secreted by pathogens in order to aide in infection and overall virulence. Cell Wall Degrgdation. Many pathogens achieve host infection by disrupting the plant cell wall (Albersheim et al., 1969). Plant cell wall degradation results from the action of pathogen-produced enzymes that are capable of cleaving specific linkages in the wall matrix of their host (Albersheim et al., 1969). Extracellular proteins secreted by fungi and bacteria are often able to macerate plant tissues and degrade plant cell wall components. Pathogens must thus contain all of the enzymes corresponding to the types of glycosidic linkages present in the cell wall polysaccharides of their target host. Multiple enzymes can be sequentially secreted by a single fungal pathogen when grown on isolated host cell walls (English et al., 1970; Jones et al., 1972). This observation is logical when considering the array of polysaccharides associated with different plant cell walls. Some pathogens can control the concentration and variety of mono- and polysaccharide degrading enzymes they produce depending on the environment they are growing in and the amount and type of substrate available (Albersheim et al., 1969). Extensive research involving the elucidation of enzymes capable of degrading major constituents of creeping bentgrass and determining whether specificity exists within the dollar spot pathogen has not been conducted. Plant Cell Wall Comp_osition. The composition of plant cell walls varies significantly from one cell type to another, one species to another, and between accessions within species (Hazen et al., 2003). For instance, a typical dicotyledonous plant cell wall contains around 30% cellulose, 30% hemicellulose, 35% pectin and 1—5% structural protein. By contrast, a typical grass species contains around 25% cellulose, 55% hemicellulose and only 10% pectin (Cosgrove, 1997). The extent of cell wall disruption often varies among plant pathogens and relies upon the secretion of extracellular enzymes. Soft rot pathogens, for instance, are able to separate individual cells causing severe tissue maceration (Mount et al., 1970; Stephens and Wood, 1975). Other pathogens are limited to mycelia penetration through the cell wall, which aides in the infection process (Weinhold and Motta, 1973). Oxalic Acid. Many Sclerotim‘a species are capable of producing compounds and enzymes which may contribute to the virulence on their particular host. Oxalic acid (0A) is thought to play multiple roles in the virulence of Sclerotinia species and has been shown to be required for the pathogenicity of a multitude of species which infect a wide range of hosts (Maxwell and Lumsden, 1970; Noyes and Hancock, 1981; Marciano et al., 1983; Godoy et al., 1990; Button and Evans, 1996; Zhou and Boland, 1999). Since many of the pathogens in this genus produce cell wall degrading enzymes, which are optimally active at a low pH, 0A is thought to play a major role in lowering the pH of the apoplast during the infection process (Bateman and Beer, 1965). Direct toxicity of 0A in the plant is also thought to be a factor in weakening the host, facilitating greater invasion of the pathogen (Noyes and Hancock, 1981). Other theories as to what role 0A plays in pathogen virulence include the chelation of Ca2+ by the oxalate anion, which is thought to compromise the integrity of the host cell wall as well as inhibit Cay-dependent defense responses (Bateman and Beer, 1965). In the case of Sclerotinia sclerotiorum, the oxidative burst of the host plant is suppressed by 0A through the inhibition of H202 production, most likely via the blocking of a signaling step in the oxidase assembly/activation stream, in tobacco and soybean cells (Cessna et al., 2000). Beaulieu (2008) showed that the dollar spot pathogen R. floccosum (Sclerotinia Homoeocarpa) indeed produces oxalic acid by analysis of secretions via gas chromatography. However, concentrations of secreted 0A were substantially lower than other Sclerotinia species such as S. Sclerotiorum. While the dollar spot pathogen may no longer be considered a member of the Sclerotinia genus (Powell, 1998) the production of oxalic acid is a commonality that cannot be ignored with regard to the elucidation of infection mechanisms and virulence. Hypovirulence. Sclerotinia species infect a wide range of hosts and are responsible for extensive economic loss due to their widespread distribution. Hypovirulence has been reported to occur in species such as S. sclerotiorum, S. minor, and S. homoeocarpa (R. floccosum) (Boland, 2004). Hypovirulence refers to the reduced ability of selected isolates within a population of a fungal plant pathogen to infect, colonize, kill and (or) reproduce on susceptible host tissues (Elliston, I982). Often times these isolates may be associated with various phenotypic characteristics such as reduced growth rate or sporulation, and altered colony morphology or color. Hypovirulence in Sclerotinia spp. is often due to the infection by fungal viruses and associated double— stranded RNA elements (Boland, 2004). In the case of S. homoeocarpa (R. floccusom), the causal agent of dollar spot, hypovirulence has been associated with the presence of double-stranded ribonucleic acid (dsRNA) (Zhou and Boland, 1997). However, some isolates detected as hypovirulent have been variable in expression of the phenotype, or were not associated with detectable concentrations of dsRNA. Presence of the hypovirulence associated virus Ophiostoma mitovirus 3a (OMV3a) in numerous strains of hypovirulent S. homoeocarpa has been reported, however many isolates testing positive for the presence of the virus do not display the hypovirulent phenotype, leading to the presumption of possible latent infection by the OMV3a virus (Melzer et al., 2003). Conclusive evidence has been difficult to ascertain due to inconsistencies in the determination of the true mechanism contributing to hypovinrlence in the dollar spot fungus. Hypovirulent isolates of various plant pathogens have been exploited to better understand pathogenicity factors and virulence. Pathogenesis by fungi in the genus Sclerotinia has been associated with the production of aforementioned cell wall degrading enzymes (Hancock, 1967; Lumsden 1976, 1979; Lumsden, 1969) as well as oxalic acid (Maxwell and Lumsden 1970). Oxalic acid has been confirmed as a pathogenicin determinant in S. sclerotiorum by using oxalic-acid-deficient mutants (Godoy et a1. 1990). Double stranded RNA-associated hypovirulence in isolates of S. scerotiorum was shown to be associated with reduced and delayed production of oxalic acid, in comparison to virulent isolates (Zhou and Boland, 1999). Mechanisms associated with dsRNA hypovirulent strains of the dollar spot fungus in relation to pathogenicity factors have yet to be elucidated, particularly with regard to oxalic acid and extracellular enzyme production. LIGHTWEIGHT ROLLING Turfgrass as a commodity is subject to unique agricultural management practices compared to most cash crops. Many cultural techniques implemented in the maintenance of high quality turfgrass are not applicable with other families of managed plants. The ability of some turfgrasses to withstand not only frequent (daily) mowing, but low (< 1 inch) mowing heights, as well as heavy traffic volumes are some of the defining characteristics of this classification of plant. On a typical golf course, the putting greens 10 are the most intensely managed areas. Golf course superintendents and turf managers alike have been experimenting with different techniques to improve the quality of their greens for centuries. One commonly used cultural practice that has been found to be beneficial in numerous aspects of management is rolling. Lightweight rolling is most often used to increase ball roll speed and distance on highly maintained putting greens. The rollers, weighing from 200 to 1000 lbs (91 to 454 kg), are manufactured in many different shapes and sizes and can have a significant affect on the playability of a putting green. Green Speeds. The Stirnpmeter is a device used to measure the speed of a golf course putting green by applying a known force to a golf ball and measuring the distance travels in feet. Designed by golfer Edward Stimpson, Sr. in 1935, the Stimpmeter, is an angled track that releases a golf ball at a known velocity so that the distance it rolls on a putting green's surface can be measured. The further a golf ball rolls, the “faster” the green is considered. It was first used by the USGA during the 1976 US. Open at Atlanta and was made available to golf course superintendents in 1978 (Radko, 1977). Since the inception of the Stimpmeter, golf course superintendents have used a variety of cultural and chemical practices in an attempt to obtain firmer, faster greens; rolling being one of them (Throssell, 1986). The implementation of these practices is often thought to be detrimental to the health and quality of the turfgrass on the putting green (Kussow, 1998; Stier, 2006), especially during times of heat stress. Rolling Histog. In the 1920's, numerous publications addressed roller weight, frequency, compaction and soil texture (Harban, 1922; Piper and Oakley, 1921; Anonymous, 1926) without coming to any valid conclusions. Rolling had been a ll common practice on most golf courses at that time, mainly for the purpose of improving surface uniformity. Shortly thereafter, the practice of frequent rolling ceased, as turfgrass research showed a link between high levels of soil compaction and turf root grth (DiPaola and Hartwiger, 1994). The negative connotations around rolling persisted within the turfgrass industry until the 1990’s (Hartwiger, 1996). Some turfgrass scientists believed that rolling may be causing unwanted damage to a healthy stand of turf, in addition to the near century old concerns regarding soil compaction. These same scientists felt the need to investigate the potential for above ground turfgrass problems associated with continual season-long turf rolling and the possibility that pathogens may invade crushed tissues, leading to diseased turf (Beard, 1994). Disease Suppression. With demand for faster putting green speeds increasing, with continued emphasis on turfgrass quality, rolling has become a major component of putting green maintenance. During a rolling experiment designed to compare green speed, Nikolai et a1. (2001) also noticed that morning rolling (3 times per week) not only increased green speeds but decreased the incidence of dollar spot significantly on creeping bentgrass maintained at putting green height. After conducting the study for several years, with all results yielding significant dollar spot reduction, many theories began to arise as to why and how rolling was inhibiting the development of this prolific turfgrass disease. Many of the theories revolved around the belief that rolling, which was conducted immediately after morning mowing, was removing excess dew or guttation water exuded by the plant and serving as a nutrient source for R. floccosum. The benefits of rolling have proven to be numerous and because experiments involving rolling have shown to reduce dollar spot disease on creeping bentgrass greens 12 (Nikolai et al., 2001), many hypotheses to explain this phenomenon have been proposed. Dispersion of dew and guttation fluid (Nikolai et al., 2001), enhancement of phytoalexin production, increased surface water holding capacity, and microbial changes in the upper rootzone have all been proposed as possible factors responsible for reduced dollar spot incidence, although the actual mechanism(s) remains unknown (Nikolai, 2005). Routine rolling can produce a more prostrate turf canopy and limit the gradual elevation of plant crowns at the thatch—soil interface during the growing season (Beard, 2002). These effects could reduce the amount of leaf blade and leaf sheath tissue removed or damaged at low mowing heights. This could also enhance photosynthetic capacity because the youngest leaf blades, which would be most often removed by mowing, are the most photosynthetically active (Youngner, 1969). Additionally, maintaining the position of crowns lower in the mat layer may reduce plant expOsure to high temperature stress because temperatures are often greatest just below the surface of dense, short-mowed turf (Beard, 1973). MICROBIAL POPULATIONS IN TURF GRASS ROOTZONES The rhizosphere (as well as the phyllosphere) is an infection epicenter where soil borne plant pathogens encounter their host and establish a parasitic relationship. The rhizosphere is also where a complex community of both soil microflora and microfauna can interact with pathogens, as well as the host, and influence the outcome of infection of the plant. The rhizosphere can be considered a battlefield, where complex communities of microbial players interact with one another, an environment where pathogenic and 13 beneficial microorganisms constitute groups that have a major influence on plant grth and development (Lynch, 1990). Agricultural management practices can have significant impacts on the size and activity of soil microbial communities (Bolton et al., 1985; Fraser et al., 1988; Kirchner et al., 1993; Powlson et al., 1987). Microorganisms in soil are critical to the maintenance of soil function in both natural and managed agricultural soils because of their involvement in key processes such as soil structure formation, decomposition of organic matter, toxin removal, and nutrient cycling. In addition, many microorganisms have been found to play key roles in suppressing soilbome plant diseases, promoting plant growth, and changing natural vegetation (Doran et al., 1996). Suppressive Soils. A suppressive soil can be defined as one with a broad array of antibiosis-related firnctions, some of which associated with the suppression of plant pathogens (van Elsas et al., 2008) The mechanisms by which soils are suppressive to different pathogens, although not well understood, can often involve biotic and/or abiotic factors. These factors may often vary with the pathogen as well as host plant in the environment. The main agents responsible for soil suppressiveness are thought to be microbial in nature, due to the observation that sterilization by autoclaving, stem pasteurization, and irradiation has rendered soils more conducive to pathogens (Malajczuk, 1983). The mechanisms by which these microorganisms make soil suppressive can be divided into several categories: nutrient competition, microbial antagonism, parasitism, and systemic induced resistance (Raaijamkers et al., 2009). Commtition. Competitive colonization of the rhizosphere and successful establishment in the root zone is a prerequisite for effective biocontrol, regardless of the 14 mechanism(s) involved (Weller, 1988; Raaijmakers et al 1995). Competition for nutrients can often times be an effective biocontrol mechanism, particularly competition for organic compounds which are necessary for reactivation of propagules and subsequent proliferation of rhizosphere-dwelling pathogens (Paulitz et a1 1992; Van Dijk and Nelson 2000; Fravel et a1. 2003). Competition for micronutrients, essential for grth and activity of the pathogen, can take place as well. Ant_agonism. Representatives of a range of bacterial (Pseudomonas, Burkholderia, Bacillus, Serratia, many Actinomycetes) and fungal (T richoderma, Penicillium, Gliocladium, Sporidesmium, nonpathogenic F usarium spp.) genera have been identified as antagonists of soilbome plant pathogens. One of the most important groups containing antagonistic microorganisms is the group containing fluorescent Pseudomonads. Several antibiotic-producing Pseudomonas spp. were isolated from soils suppressive to diseases such as take-all of wheat, black rot of tobacco and F usarium wilt (Keel et al. 1996; Tamietti et a1. 1993; Weller et a1. 1988). Naturally occurring root- associated fluorescent Pseudomonas spp. producing the antibiotic 2,4-DAPG were numerously populated in take-all-suppressive soil and are key components of specific suppression of Gaeumanonmyces graminis var. tritici (Raaijmakers and Weller, 1998; Raaijmakers et al., 1997). This suppressive activity was lost when 2,4-DAPG-producing Pseudomonas spp. were eliminated and, conversely, conducive soil gained suppressiveness to take-all when 2,4-DAPG-producing Pseudomonas strains were introduced. Biocontrol microorganisms may adversely affect the population density, dynamics, and metabolic activities of pathogens. 15 Ingrced Systemic Resistance. Along with biocontrol activity of microorganisms in the rhizosphere, several microbial groups can have a direct positive effect on plant growth and health. Phytostimulatory and biofertilizing microbes can promote plant health by making the plant “stronger”. Many rhizosphere microorganisms can induce a systemic response in the plant as well, resulting in the activation of plant defense mechanisms (Pieterse et al., 2003). Induced systemic resistance (ISR) does not necessarily confer complete resistance, but rather protects the plant from various types of pathogen infection. The capacity to convey this type of ISR has been identified in a wide range of bacteria (Van loon et al., 1998; Haas and Defago, 2005), not only in greenhouse experiments, but under field conditions as well (Zehnder et al., 2001; Pieterse et al., 2003). Microbial Populations in Soil. Several studies have identified trends within the microbial activity of soil related to and responsible for the suppression of pathogens. For example, van Os & van Ginkel (2001) showed a clear relationship between the suppression of Pythium root rot in bulbous Iris and soil microbial biomass and activity. Their findings showed that high microbial biomass and activity induced the suppression of Pythium growth and development in the soil. Relationships have been found between microbial diversity and root disease suppression as well (Nitta, 1991; Workneh and van Bruggen, 1994). Rovira & Wildennuth (1981) indicated that the microbiota in a “rich” soil tends to reduce the severity of attack by many soilbome plant pathogens, or, in other words, soils higher in microbiota content and diversity tend to Show trends towards general disease suppression. Tippett (1978) provided an example of the importance of the soil microbiota on the level of suppressiveness. By adding soil containing large 16 quantities of microorganisms to microbially deficient soil, they were able to eliminate P. cinnamoni. Many antagonistic microorganisms are naturally present in soil and exert a certain degree of biological control over plant pathogens, regardless of human activities. However, this level of natural control is often insufficient for consistent, reliable disease- free cropping. Researchers are, therefore, attempting to enhance the effectiveness of antagonists, thus increasing suppressiveness (Hoitink et al., 2003). Management of the biotic and abiotic properties of a soil is an important approach in promoting the activity and diversity of beneficial microorganisms. Cultural practices have been proposed as means to decrease soil inoculum levels, or increase suppressiveness, thus limiting the densities and activities of rhizosphere pathogens. Disease suppression has been obtained through crop rotation (Cook et al., 2002), as well as other practices such as intercropping (Schneider et al., 2003), residue destruction (Baird et al., 2003), organic amendments (Tilston et al., 2002), and tillage management practices (Sturz et al., 1997; Pankhurst et al., 2002). With public perception and environmental stewardship at the forefront of turfgrass management concerns, alternative options for management of diseases are becoming highly desired in the turfgrass industry. Dollar spot is considered to be the most important disease on turfgrass, particularly in the northeastern regions of the United States (Vargas, 2005). If used as a tool in the management of dollar spot, rolling could potentially provide major economical benefits to golf course superintendents as well as boast significant conservational benefits. However, literature regarding R. floccosum suppression relating to rolling is sparse, particularly regarding mechanisms involved in 17 disease reductions. A multi-faceted approach was taken in an attempt to elucidate particular hypotheses regarding the rolling effect on dollar spot. R. floccosum infection mechanisms, including enzyme, and toxin production were studied. Understanding specific plant-pathogen interactions and mechanisms related to pathogen virulence is important in elucidating novel methods of disease control. These have gone relatively uninvestigated with regard to the dollar spot pathogen on turfgrass. Rolling hypotheses and R. floccosum infection and virulence are discussed in the chapters that follow. LITERATURE CITED Agrios, G. N. 1988. Plant Pathology, 3rd edition. Academic Press, New York. Albersheim, P., Jones, T. M., and English, P. D. 1969. Biochemistry of the cell wall In relation to infective processes. Annu. Rev. Phytopathol. 7: 171-194. Anonymous, 1926. Rolling the fairways and putting greens. Bulletin of the Green Section of the US. Golf Association. 6: 59. Baird, R. E., Watson, C. E., and Scruggs, M. 2003. Relative longevity of Macrophomina phaseolina and associated mycobiota on residual soybean roots in soil. Plant Dis. 87: 563-566. Baldwin, N. A., and Newell, A. J. 1992. Field production of fertile apothecia by Sclerotinia homoeocarpa in F estuca turf. J. Sports Turf Re. Inst. 68: 73-76. Bateman, D. F. and Beer, S. V. 1965. Simultaneous production and synergistic action of oxalic acid and polygalacturonase during pathogenesis by Sclerotiorum rolfsii. Phytopathology. 55: 204-211. Baur, W. D., Talmadge, K. W., Keegstra, K., and Albersheim, P. 1973. The structure of plant cell walls, 11. The hemicellulose of the walls of suspension-cultured sycamore cells. Plant Physiol. 51: 174-187. Beard, J. B. 1973‘. Turfgrass: Science and culture. Prentice Hall, Englewood Cliffs, NJ. 18 Beard, J. B. 1994. Turf rolling. Grounds Maintenance 29: 44-52. Beard, J. and Tani, T. 1997. Color Atlas of Turfgrass Diseases, pp. 141-147. Ann Arbor Press. Chelsea, MI. Beard, J. B. 2002. Turf management for golf courses. 2nd ed. Ann Arbor Press, Chelsea, MI. Beaulieu, R. A. 2008. Oxalic acid production by Sclerotinia homoeocarpa: the causal agent of dollar spot. Senior Honors Thesis. Ohio State University. Columbus, OH, USA. Boland, G. J. 2004. Fungal viruses, hypovirulence, and biological control of Sclerotinia species. Can. J. Plant Pathol. 26: 6-18. Bolton, J., Elliot L. F ., Papendickc, P. R., and Bezdiccek, D. F. 1985. Soil microbial biomass and selected soil enzyme activities; effect of fertilization and cropping practices. Siol Biol Biochem. 17: 297-302. Burke, K., Kaufman P., MCNeil, M., and Albersheim, P. 1974. A survey of the walls of suspension-cultured monocots. Plant Physiol. 54: 109-115. Cessna, S. G., Sears, V. E., Dickman, M. B., and Low, P. S. 2000. Oxalic acid, a pathogenicity factor for Sclerotinia sclerotiorum, suppresses the oxidative burst of the host plant. Plant Cell. 12: 2191—2199. Cook, R. J ., Schillinger, W. F., and Christensen, N. W. 2002. Rhizoctonia root rot and take-all of wheat in diverse direct-seed spring cropping systems. Can. J. Plant Pathol. 24: 349-358. Cosgrove, D. J. 1997. Assembly and enlargement of the primary cell wall in plants. Annu. Rev. Cell Dev. Biol. 13: 171—201. Couch, H. B. and Bloom, J. R. 1960. Influence of environment on diseases of turf grasses. II. Effect of nutrition, pH and soil moisture on Sclerotinia dollar spot. Phytopathology. 50: 761—763. Detweiler, A. R., Vargas, J. M., and Danneberger, T. K. 1983. Resistance of Sclerotinia homoeocarpa to Iprodione and Benomyl. Plant Dis. 67: 627-630. DiPaola, J. M. and C. R. Hartwiger. 1994. Green speed, rolling and soil compaction. Golf Course Manag. 62: 49-51,78. Doran, J. W., Sarrantonio, M., and Liebig, M. A. 1996. Soil health and sustainability. Adv. Agron. 56: 2—54. 19 Dutton, M. V., and Evans, C. S. 1996. Oxalate production by fungi: Its role in pathogenicity and ecology in the soil environment. Can. J. Microbiol. 42: 881—895. Dwyer, P. J. 1999. Field efficacy, persistence and antibiotic production of Pseudomonas aureofaciens. Dissertaiton for the degree of MS. Michigan State University East Lansing, MI. Elliston, J. E. 1982. Hypovirulence. Adv. Plant Pathol. 1: 1-33. Ellram, A., Horgan, B., and Hulke, B. 2007. Mowing strategies and dew removal to minimize dollar spot on Creeping bentgrass. Crop Sci. 49: 2129-2137. Fenstermacher, J. M. 1979. Certain features of dollar spot disease and its causal organism, Sclerotinia homoeocarpa. Pages 49-53 in: Adv. in Turf. Path., P. O. Larsen and B. G. Joyner, eds. HBJ, Duluth, MN. English, P. D., Jurale J. B., and Albersheim, P. 1970. Host pathogen interactions 11. Parameters affecting plysaccaride-degrading enzyme secretion by Colletotrichum lidemuthianum grown in culture. Plant Physiol. 47: 1-6. Fraser, D. G., Doran, J. W., Sahs, W. W., and Lesoing, G. W. 1988. Soil microbial populations and activities under conventional and organic management. J. Environ. Qual. ’ 17: 585-590. Fravel, D., Olivain, C., and Alabouvette, C. 2003. F usarium oxysporum and its biocontrol. New phytol. 157: 493-502. Godoy, G., Steadman, J. R., Dickman, M. B., and Dam, R. 1990. Use of mutants to demonstrate the role of oxalic acid in pathogenicity of Sclerotinia sclerotiorum on Phaseolus vulgaris. Physiol. Molec. Plant Path. 37: 179-191. Golembiewski, R. C., Vargas, J. M., Jones, A. L., and Detweiler, A. R., 1995. Detection of Demethylation Inhibitor (DMI) Resistance in Sclerotinia homoeocarpa Populations. Plant Dis. 79: 491-493. Haas, D. and Defago, G. 2005. Biological control of soil-bome pathogens by fluorescent pseudomonads. Nat. Rev. Microbiology. 3: 307-319. Hancock, J. G. 1967. Hemicellulose degradation in sunflower hypocotyls infected with Sclerotinia sclerotiorum. Phytopathology. 57: 203-206. Harban, W. S. 1922. The effect of trampling and rolling on turf. Bulletin of the Green Section of the US Golf Association. 2: 148-150. Hartwiger, C. 1996. The ups and downs of rolling putting greens. USGA Green Section Record. 34(4): 1-4. 20 Hazen, S. P., Hawley, R. M., Davis, G. L., Henrissat, B., and Walton, J. D. 2003. Quantitative trait loci and comparative genomics of cereal cell wall composition, Plant Physiol. 132: 263—271. Hoitink H. A. J. and McSpadden-Gardener B. 2003. Disease suppression through manipulation of microbial communities in composts and soils. Int. Congr. Plant Pathol., 8th, New Zealand. Huber, L. and Gillespie, T. J. 1992. Modeling leaf wetness in relation to plant disease epidemiology. Phytopathology. 30: 553—577. Jackson, N. 1973. Apothecial production of Sclerotinia homoeocarpa F. T. Bennett. Pages 353-357 in: Proc. Int. Turfgrass Res. Conf. American Society of Agronomy, Madison, WI. Jones, T. M., Anderson, A. J., and Albersheim, P. 1972. Host-pathogen interactions IV. Studies on the polysaccharide-degrading enzymes secreted by F usarium oxysporum t. sp. Lycopersici. Physiol. Plant Pathol. 2: 153-166. Keel, C., Weller, D. M., Natsch, A., De'fago, G., Cook, R. J., and Thomashow, L. S. 1996. Conservation of the 2,4-diacetylphloroglucinol biosynthesis locus among fluorescent Pseudomonas strains from diverse geographic locations. Appl. Environ. Microbiology. 62: 552—563. Kirchner, M.J., Wollum II, A. F ., and King, L.D. 1993. Soil microbial populations and activities in reduced chemical input agro-ecosystems. Soil Sci. Soc. Amer. J. 57: 1289- 1295. Kussow, W. R. 1998. Putting green management systems. Wisc. Turf Res. Reports XV: 85-93. Lumsden, R. D. 1969. Sclerotinia sclerotiorum infection of bean and the production of cellulose. Phytopathology. 59: 653-657. Lumsden, R. D. 1976. Pectolytic enzymes of Sclerotinia sclerotiorum and their localization in infected bean. Can. J. Bot. 54: 2630-3641. Lumsden, R. D. 1979. Histology and physiology of pathogenesis in plant disease caused by Sclerotinia sclerotiorum. Can. J. Bot. 54: 2630—2641. Lynch, J. 1990. The Rhizosphere. Wiley, London, UK, p. 458. Malajczuk, N. 1983. Microbial Antagonism to Phytophthora, Its Biology, Taxonomy, Ecology and Pathology. APS Press. St Paul, MN. 21 Marciano, P., DiLenna, P., and Magro, P. 1983. Oxalic acid, cell wall degrading enzymes and pH in pathogenesis and their significance in the virulence of two Sclerotinia sclerotiorum isolates on sunflower. Physol. Plant Pathol. 22: 339-345. Maxwell, D. P. and Lumsden, R. D. 1970. Oxalic acid production by Sclerotinia sclerotiorum in infected bean and in culture. Phytopathology. 60: 1395-1398. Melzer, M. S., Deng, F., and Boland, G. J. 2003. Latent infection and distribution of a hypovirulence-associated double-stranded RNA (OmaMV3 a) in populations of Sclerotinia homoeocarpa. Can. J. Plant Pathol. 25: 430 (Abstr.) Mount, M. S., Bateman, D. F., and Basharn, H. G. 1970. Induction of electrolyte loss, tissue maceration and cellular death of potato tissue by an endypolygalacturonase transeliminase. Phytopathology. 60: 924-931. Nelson, E. B. and Craft, C. M. 1991. Introduction and establishment of strains of Enterobacter cloacae in golf course turf for biological control of dollar spot. Plant Dis. 75: 510-514. Nelson, E. B. and Craft, C. M. 1992. Suppression of dollar spot on creeping bentgrass and annual bluegrass turf with compost-amended topdressings. Plant Dis. 76: 954-958. Nikolai, T. A. 2002. Effects of rolling and fertility on putting green root zone mixes. Dissertation for Doctor of Philosophy. Michigan State University. E. Lansing, MI. Nikolai, T. A. 2005. The superintendent’s guide to controlling putting green speed. John Wiley & Sons. Hoboken, NJ. Nikolai, T. A., Rieke, P. E., Rogers, J. N. 111, and Vargas, J. M. Jr. 2001. Turfgrass and soil responses to lightweight rolling on putting green root zone mixes. Int. Turf. Soc. Res. J. 9(Part 2): 604-609. Nitta T. 1991. Diversity of root firngal floras: its implications for soil-borne diseases and crop growth. Jpn. Agric. Res. 25: 6—1 1. Noyes, R. D. and Hancock, J. G. 1981. Role of oxalic acid in the Sclerotinia wilt of sunflower. Physiol Plant Pathol. 7: 123-132. Pankhurst, C. E., McDonald, H. J. B., Hawke, G., and Kirkby, C. A. 2002. Effect of tillage and stubble management on chemical and microbiological properties and the development of suppression towards cereal root disease in soils from two sites in NSW Australia. Soil Biol. Biochem. 34: 833-840. Paulitz, T. C., Anas, 0., and Fernando, D. G. 1992. Biological control of Pythium damping-off by seed treatment with Pseudomonas putida: relationship with ethanol production by pea and soybean seeds. Biocontrol Sci. Technol. 2: 193 -201. 22 Pieterse, C. M. J ., van Pelt, J. A., Verhagen, B. W. M., Torr, J ., van Wees, S. C. M. Leon- Kloosterziel, K. M., and van Loon, L. C. 2003. Induced systemic resistance by plant growth-promoting rhizobacteria. Symbiosis. 35: 39-54. Piper, C. V., and R. A. Oakley. 1921. Rolling the turf. Bulletin of the Green Section of the US. Golf Association. 1: 36. Powell, J. F. 1998. Seasonal variation and taxonomic clarification of the dollar spot pathogen: Sclerotinia homeocarpa. Dissertation for Doctor of Philosophy. Michigan State University, East Lansing, MI. Powlson, D. S., Brookes, P. C., and Christensen, B. T. 1987. Measurement of soil microbial biomass provides an early indication of changes in total soil organic matter due to straw incorporation. Soil Biol. Biochem. 19: 159-164. Raaijmakers, J. M., Leeman, M., Van Oorschot, M. M. P., Van der Sluis, I., Schippers, B., and Bakker, P. A. H. M. 1995. Dose-response relationships in biological control of Fusarium wilt of radish by Pseudomonas spp. Phytopathology. 85: 1075-1081. Raaijmakers, J. M., Paulitz, T. C., Steinberg, C., Alabouvette, C., and Moenne-Loccoz, Y. 2009. The rhizosphere: a playground and battlefield for soilbome pathogens and beneficial microorganisms. Plant and Soil. 321: 341-361. Raaijmakers, J. M. and Weller D. M. 1998. Natural plant protection by 2,4- diacetylphloroglucinol-producing Pseudomonas spp. in take-all decline soils. Mol. Plant Microbe Interact 11:44—52. Raaijmakers J. M., Weller D. M., and Thomashow L. S. 1997. Frequency of antibiotic producing Pseudomonas spp. in natural environments. Appl. Environ. Microbiol 63: 881—87. Radko, A. M. 1973. Refining green section specifications for putting green construction. Proc. Sec. Int. Turfgrass Research Con. 287-297. Rovira, A. D. and Wildermuth. G. B. 1981. The nature and mechanisms of suppression, p. 385-415. In M. J. C. Asher and Shipton, P. (ed.), Biology and control of take-all. Academic Press, Inc. (London), Ltd., London, UK. Schneider, 0., Aubertot, J. N., Roger-Estrade, J., and Dore, T. 2003. Analysis and modeling of the amount of oilseed rape residues left at the soil surface after different soil tillage operations. 7th International Conference on Plant Pathology, Tours France, 3-5 December 2003. Smiley, R. W., Demoeden, P. H., and Clark, B. B. 1992. Infectious foliar disease. Comp. of Turf. Dis., 2"d ed. Amer. Phytopath. Soc., St. Paul, MN. pp 11-37. 23 Stephens, G. J. and Wood, R. K. S. 1975. Killing of protoplasts by soft rot bacteria. Physiol. Plant Pathol. 5: 165-181. Stier, J. 2006. Shorter mowing heights are hazardous to summer health. The Grass Roots. 35: 4-5,7,9. Sturz, A.V., Carter, M. R., and Johnston, H. W. 1997. A review of plant disease pathogen interactions and microbial antagonism under conservation tillage in temperate humid agriculture. Soil Tillage Res. 41: 169-189. Throssell, C. 1981. Management practices affecting putting green speed. Science in Agriculture. 28: 9. Tippett, J. 1978. Interaction between Phytophthora cinnamoni and plant roots. Dissertation for Doctor of Philosophy. Monash University, Australia. pp. 217. Tilston, E., Pitt, L. D., Groenhof, A. C. 2002. Composted recycled organic matter suppresses soil-bome diseases of field crops. New Phytol. 154: 731-740. Tamietti, G., Ferraris, L., Matta, A., and Abbattista Gentile, I. 1993. Physiological responses of tomato plants grown in Fusarium suppressive soil. J. Phytopathology. 138: 66—76. Uddin, W., Serlemitsos, K., and Viji, G. 2003. A temperature and leaf wetness duration based model for the prediction of gray leaf spot of perennial ryegrass. Phytopathology. 93: 336—343. van Dijk, K. and Nelson, E. B. 2000. Fatty acid competiton as a mechanism by which Enterobacter cloacae suppresses Pythium ultmum sporangium germination and damping off. Appl. Environ. Microbiol. 66: 5340-5347. van Elsas, J. D., Costa, R., Jansson, J., Sjoling, S., Bailey, M., Nalin, R., Vogel, T. M., and van Overbeek, L. 2008. The metagenomics of disease-suppressive soils-experiences from the METACONTROL project. Trends in Biotechnology. 26: 591-601. van Loon, L. C., Bakker, P. A. H. M., and Pieterse, C.M.J. 1998. Systemic resistance induced by rhizosphere bacteria. Ann. Rev. Phytopathology. 36: 453-483. van Os, G. J. and van Ginkel J. H. 2001. Suppression of Pythium root rot in bulbous Iris in relation to biomass and activity of the soil microflora. Soil Biol. Biochem 32: 1447— 54. Vargas, J. M., Jr. 2005. Fungal diseases of turfgrass, I: diseases primarily occurring on golf course turfs. Pages 15-32 in Management of Turfgrass Diseases, 3rd ed. CRC Press, Inc., Boca Raton, FL. 24 Walsh, B. K. 2000. Epidemiolgy and disease forcasting system for dollar spot caused by Sclerotinia homoeocarpa F.T. Bennet. Dissertation for Doctor of Philosophy. University of Guelph, Guelph, ON, Canada. Warren, C. G., Sanders, P., and Cole, H. 1974. Sclerotinia homoeocarpa tolerance to benzimidizole configuration fungicides. Phytopathology. 64: 1139-1142. Weinhold, A. R. and Motta, J. 1973. Initial host responses in cotton to infection by Rhizoctonia solam'. Phytopathology. 63: 157-162. Weller, D. M. 1988. Biological control of soilbome pathogens in the rhizosphere with bacteria. Annu. Rev. Phytopathology. 26: 379-407. Weller, D. M., Howie, W. J., and Cook, R. J. 1988. Relationship between in vitro inhibition of Gaeumannomyces graminis var. tritici and suppression of take-all of wheat by fluorescent pseudomonads. Phytopathology. 78: 1094—1100. Wilder, B. M. and Albersheim, P. 1973. A structural comparison of the wall hemicellulose of cell suspension cultures of sycamore and of red kidney bean. Plant Physiol. 51: 889-893. Williams, D. W. and Powell, A. J. 1995. Dew removal and dollar spot on creeping bentgrass. Golf Course Management. 63: 49-52. Williams, D. W., Powell, A. J., Vincelli, P., and Doughtery, P. J. 1996. Dollar spot on bentgrass influenced by displacement of leaf surface moisture, nitrogen, and clipping removal. Crop Sci. 36: 1304—1309. Workneh F. and van Bruggen A. II. C. 1994. Microbial density, composition, and diversity in organically and conventionally managed rhizosphere soil in relation to suppression of corky root of tomatoes. Appl. Soil Ecol. 1: 219-30. Younger, V. B. 1969. Growth and development. In A. A. Hanson and F. V. Juska, eds. Turfgrass. Agronomy Monograph No. 14. Madison, WI: American Society of Agronomy. 187-216. Zhou, T. and Boland, G. J. 1997. Hypovirulence and double-stranded RNA in Sclerotinia homoeocarpa. Phytopathology. 87: 147-153. Zhou, T. and Boland, G. J. 1999. Mycelia] grth and production of oxalic acid by virulent and hypovirulent isolates of Sclerotinia sclerotiorum. Can. J. of Plant Path. 21: 93—99 Zehnder, G. W., Murphy, J. F., Sikora, E. J., and Kloepper, J. W. 2001. Application of rhizobacteria for induced resistance. Euro. J. of Plant Pathol. 107: 39—50. 25 CHAPTER ONE DOLLAR SPOT REDUCTION THROUGH LIGHTWEIGHT ROLLING ON CREEPING BENTGRASS PUTTING GREENS ABSTRACT A study to investigate the effects of light-weight rolling on the reduction of dollar spot was conducted at the Hancock Turfgrass Research Center on the campus of Michigan State University in 2008 and 2009. Creeping bentgrass (Agrostis palustris) plots subjected to daily (5 days week") rolling treatments resulted in significantly less disease when compared to a non-rolled control. Morning rolling treatments and afternoon rolling treatments were implemented to investigate the effects of dew and guttation fluid removal and dispersal. Treatments rolled in the p.m. (after dew and guttation naturally dissipated) exhibited similar disease reduction as treatments rolled in the am. (while dew and guttation fluid were present) when compared to the control. Cumulative effects of rolling were investigated with treatments rolled twice (2x) day]. The 2x rolling treatment exhibited significantly less disease than all other treatments and resulted in Significantly higher upper rootzone soil volumetric water content means in both 2008 and 2009, when compared to the control. These results suggest that mechanisms other than dew and guttation dispersal may be responsible for disease reduction due to the observed dollar spot inhibition from afternoon rolling. Additionally, data suggests that season-long lightweight rolling twice day'1 may have an effect on upper rootzone volumetric water content in sand based creeping bentgrass putting greens. 26 INTRODUCTION Turfgrass is subject to unique agricultural management practices compared to those of other agronomically important crop systems. Many cultural techniques implemented in the maintenance of high quality turfgrass are not applicable with other species of managed plants, even within the family Gramineae. One commonly used cultural practice on golf course turf is lightweight rolling. Lightweight rolling is most often implemented to increase ball roll speed and distance during putting as well as to enhance surface uniformity on intensively maintained putting greens. The rollers, typically weighing from 200 to 1000 lbs (91 to 454 kg), are manufactured in many different shapes and sizes, and can have a significant affect on the playability of a putting green thus having dramatic impacts on the game of golf. The rolling of putting greens has been common on most golf courses for decades, and prior research has led to even more questions regarding its effects on turfgrass and underlying soil. Many negative associations with rolling have persisted within the turfgrass industry (Hartwiger, 1996). Theories that rolling may be causing unwanted damage to a healthy stand of turf, as well as concerns regarding soil compaction issues, have existed since the 1920’s. Turf scientists have debated over the need to investigate the potential for above ground turfgrass problems associated with continual, season-long turf rolling and the possibility that pathogens may invade crushed tissues, leading to diseased turf (Beard, 1994). Nikolai et al. (2001) conducted a multiple year rolling study at Michigan State University, aimed at the effects of rolling and fertility on putting green speed, disease incidence, and soil properties in different root zone mixes. This study gave rise to 27 surprisingly significant data regarding numerous aspects of plant management. Morning rolling (3 times per week) was found to dramatically decrease the incidence of the fungal disease dollar, spot caused by Rutstroemia floccosum, on creeping bentgrass maintained at putting green height, thus improving overall turfgrass quality. These results were observed in several years of the study. Since that time, others have observed similar results with lightweight rolling. Inguagiato et. al. (2009) observed slight reductions in turfgrass anthracnose, caused by Colletotrichum cereale, on plots rolled every other day. These results were contrary to the previous suggestions that rolling enhances disease severity (Demoeden, 2000; Smiley et. al., 2005). The findings of decreased disease occurrence through rolling spurred many hypotheses and theories with regard to possible suppression mechanisms. Many of the theories stem from the belief that rolling, which typically is conducted immediately after morning mowing, removes excess dew or guttation fluid exuded by the plant. This moisture is thought to serve as a nutrient rich food source for foliar feeding pathogens like R. floccosum (Williams and Powell, 1995). Dew removal and reduced leaf wetness duration (LWD) are widely accepted techniques used to reduce dollar spot incidence on turfgrass. Many studies have revealed the benefits of early morning mowing, syringing, and other dew removal techniques in order to decrease LWD (Williams and Powell, 1995; Ellram et al, 2007), ultimately reducing dollar spot incidence. If used as a tool in the management of dollar spot, rolling could have substantial economic and environmental benefits. However, the literature is sparse on rolling studies, particularly on the topic of disease suppression mechanisms. 28 The objective of this study was to test specific hypotheses regarding rolling and its effect on dollar spot incidence on creeping bentgrass putting greens. Different rolling treatments were implemented for the duration of this study in order to generate data used to draw conclusions related to each hypothesis. The first objective was to investigate the hypothesis that morning rolling contributed to the removal or dispersal of secondary dew/guttation fluid after mowing, and that the reduction of disease is directly related to this practice. Other hypotheses related to soil microbial populations and plant defense responses induced by rolling are addressed in forthcoming chapters. Phenomena related to lightweight turfgrass rolling, dew/guttation removal, and dollar spot reductions were examined. Investigations into possible cumulative effects of repeated daily rolling were also undertaken in order to elucidate the potential for expedited disease reduction. MATERIALS AND METHODS Field research was conducted from June through October in 2008 and 2009 at the Hancock Turfgrass Research Center on the campus of Michigan State University, East Lansing, Michigan (42°43'48" N, 84°28'35" W). A research putting green, constructed in 2005, consisting of a mixed sward of creeping bentgrass (Agrostis stolonifera cv ‘Independence ’) and annual bluegrass (Poa annua L.) was used. A 60 x 60 ft (18.3 x 18.3 m) research area was divided into twelve 7 x 12 ft (2.13 x 3.65 m) evenly spaced plots. Prior to initiating the study, each plot was randomly assigned a rolling treatment. Alleys of three feet separated each plot so that the roller had adequate space to stop without impeding into neighboring plots. All plots were mowed in the morning between 6 and 7 am. before morning rolling treatments were 29 implemented. Rolling treatments were carried out 5 days week'1 (Monday-Friday) and were as follows: 1) control (no rolling), 2) rolled once (1x) in the am. immediately after mowing, 3) rolled once (1x) in the pm. when turf was dry or dew and guttation water had dissipated, and 4) rolled twice (2x) in the am. immediately after mowing. The morning rolling treatments were implemented between 7:00 and 8:30 am. and the afternoon rolling between 1:00 and 2:00 pm. Rolling of plots was conducted with a Tru-Turf ride- on greens roller, model RS48-11B (Tm-Turf Pty. Ltd, Molendinar, Queensland, AUS), with a 39 inch (99 cm) roll swath weighing 562 pounds (255 kg) without an operator. A single rolling treatment consisted of rolling across the plot using multiple passes in opposite directions to ensure complete coverage of the plot with minimal overlap of rolled areas. Once a single rolling pass was made, the process was repeated immediately on plots rolled twice per day. To ensure study uniformity, the rolling treatment was the only practice that differed among plots. All other cultural and chemical practices remained constant among treatments for the duration of the study. The root zone soil consisted of a sand based 80:20 (sandzpeat v/v) mix constructed to USGA recommendations (U SGA, 1993). Sand topdressing was applied to the entire research area on a light, frequent (bi/tri-weekly) basis throughout the growing season in order to simulate typical golf course putting green maintenance practices. No vertical mowing or core cultivation occurred on the research plots during the course of the study in order to minimize turfgrass and soil disruption. Fungicides were not applied to any of the research plots during the duration of the study in order to encourage dollar spot disease infection. Insecticides and herbicides were applied as needed, and were applied uniformly over the entire study area. 30 Nitrogen fertility was applied at a rate of 24.4 kg N ha_1 mo 4 (0.5 lbs N 1000 ft. sq.‘l mo'l) during the growing season from April to September of each year. Irrigation was applied via four Rain Bird Eagle irrigation heads, model 750 (Rain Bird Distribution Co., Azusa, CA), which were located at the comers of the research area. Irrigation was applied as needed in order to keep the turf stand healthy and free of wilting symptoms. Plots were mowed at a height of 0.156 inches (3.96 mm) six days week'1 with a walk- behind Toro Greensmaster 1000 greens mower (Toro Co., Bloomington, MN). and clippings were collected throughout the study. Dollar spot disease counts were taken when disease was active and symptoms occurred (generally once per week or bi-weekly) by counting the number of individual infection centers per plot. Spots greater than 3 cm in diameter were counted as one infection center. Larger, coalescing spots were broken down into smaller spots when rating and considered to be multiple infection centers. All ratings were taken in the morning before mowing, when conditions were conducive to counting infection centers. Quality ratings were taken on a regular basis throughout the growing seasons (generally once per month from May — October). Quality was rated on a scale of l- 9 where l=dead/necrotic, 6=acceptable, and 9=excellent. Quality ratings were based on a combination of characteristics such as color, density, unifOrmity, and playability. Ratings of 6 and above were regarded as acceptable turf for a creeping bentgrass putting green. Percent volumetric water content (%VWC) was measured using a F ieldScout TDR 300 Soil Moisture Meter (Spectrum Technologies Inc., Plainfield, IL) with probe rods at a depth of 1.5 inches (3.81 cm). Twenty measurements were taken at random locations in each plot and averaged in order to obtain a representative %VWC for each 31 plot on each measurement date. All volumetric water content measurements were taken at least one full day (24 hours) after a rain event, with-holding irrigation, in order to ensure consistent VWC ratings. The experimental design was a randomized block design with three replications. Analysis of variance was performed on observational measurements in order to determine significant effects, followed by Fisher’s protected Least Significant Difference (LSD) if differences were found at the probability level (P) < 0.05. Treatment differences were analyzed using the Proc GLM procedure of the Statistical Analysis System (SAS Institute, 2009), and when appropriate, means were separated using Fisher’s LSD procedure at the 0.05 level of probability unless otherwise noted. RESULTS AND DISCUSSION Dollar spot disease incidence data were collected on a total of 15 dates between 2008 and 2009 (Tables 1.01 and 1.02, respectively). Three dates (18-Sep, 7-Oct, 20-Oct) in 2008 resulted in significant treatment effects on dollar spot, with the twice rolled (2x) a.m. treatment having less dollar Spot than the control. Treatments rolled 1x day], either in the morning or afternoon, were not significantly different from one another on any of the dates recorded (P > 0.05). In 2008, no significant treatment effect for either 1x rolling treatments were found compared to the control; and neither of the 1x treatments were significantly different from the 2x treatment according to Fischer’s LSD (P > 0.05). While differences were not significant on individual measuring dates, combined seasonal dollar spot means from 2008 indicated a significant rolling effect on dollar spot incidence in both the 1x pm. and 2x a.m. treatments (Figure 1.01). The 1x am. 32 treatment was not significantly different from any other treatment (P > 0.05). This lack of statistical separation is likely due to one replication within the 1x a.m. treatment having uncharacteristically higher dollar spot incidence than the other two replications. The high dollar spot counts throughout the season on that particular plot resulted in a much higher seasonal mean, thus rendering the 1x a.m. treatment statistically similar to the control. In 2009, dollar spot ratings were taken on nine separate occasions (Table 1.02). Disease pressure was much higher in 2009 than in 2008, resulting in an average increase in disease incidence of 2.72 times between the two yearly means. Dollar spot incidence occurred earlier in the season and at much higher levels throughout the season as indicated by the number of infection centers recorded. The 2x a.m. treatment had significantly less dollar spot incidence than the control on all but the first two rating dates in 2009 (Table 1.02). When dollar spot incidence was most severe (September and October), all rolling treatments resulted in significantly lower disease counts compared to the control. July 31, was the only rating date where significant differences between 2x am. and either of the 1x rolling treatments were observed (Table 1.02). However, the 2x a.m. rolling treatment showed a trend towards lower dollar spot incidence. The 1x am. and 1x p.m. treatments were not significantly different from one another on any observation date, but both had a significant effect on dollar spot when compared to the control on four of the nine rating dates (25-Aug, 3-Sep, 22-Sep, 17-Oct) (Table 1.02). 33 .God A .3 Own FREE..— 8 36808 Enact? 8.385%? 8: 2a 5:28 a 5 3:2 2:8 2: ,3 332.8 882 E. 2.03838. .32 E3395 Batman; 36 05 8 EnocEwfi 28 282.2%: “oz .1. ._.. .m Z 8.8 8.8 8.8 8.8 8.2 8.2 8.853 .. ._. .. m2 m2 m2 880888 .82 82 88.8 88 m2 8.m aim 88.8 88.8 88.8 8.: 8A 88. :8: 88.8 88.8 88.8 8.8 8.2 8s .35 8; «8.8 8.8 8.8 8.: 8.: 88882382 88 :2 :8 8 8: :8 asses: wooN 8888 5:83 8% 8:5 Lo .82 .woom E 83% Eobbfi co 85205 888% 8% 8:3 so mam—3H mo macho 28.585 .8.“ 38585.8 mm; 98 €on .5.— 038,—. 34 .586 :86 . Ed .36 uoEo E 232 b25395 .35. 323305 332 one 35862 + ”.93. bzfmnoa .832 8 35:8 Eat Eobbfi E :82 + .390 A A: amt. {28... 2 wEEoooa Caveat—o baa—SEE: 8: 0.8 5:28 a 5 5:2 2:8 05 3 330:8 832 5. 3.038232 .32 bzfimnoa Rotate; mod 2: 8 EmocEwfi Ea ESE—Em .oZ .. .mz £3 33 3.3 8.8— 3% 2.: E? 8.3 ES A83 a3 a * ._. * a * ._. m2 m2 8fioa§m +3 2.: +3 8.3 in 3% B can 3 3.: £3 a m3 m3 m3 :8 CAN +3 Ewe +3 SE 2 mama £92: 8 SE 8 gm 8 82 8% SR 8a 5 3 m3: +3 $62 +3 8&2 a MS : a $2 8 3% 8 33 Sam 38 .88 5 32a 32% 33% £33 £32 88.8 3mm 8.: 3% 2858338582 ES 88 Q8 Em 5+ NE LE 88 as 8888; 88 888 888: 8% 8:8 Lo .82 doom E 888 EobbE «a 35205 8.8qu Comm 8:8 co m£=8 mo Beebe E2535 com 385988 mm; 93 2802 No; 2an 35 3.5 c m 2x a.m. . 17.72 b l l | Y 26 44 ab 1! r} A l 1 I A 35.67 a * easonal dollar spot meansz among different rolling treatments on creeping bentgrass (A grostis palustris Huds.) putting green plots in 2008. :0 VII! l3on1 srods .reuoqm on Figure 1.0 . 36 lxpm. Rolling Treatment 1x a.m. Control " Treatment means followed by the same letter are not significantly different according to Fisher’s LSD (P > 0.05). 2 . . . Values are means of 6 different measurement dates and three rephcatrons per date. .83. A A: Gm: Pair.— 9 96.508 820$: bEeoCEwfi 8: 8a .532 2:8 05 .3 330:8 2.8:. 3:258; ._ .83 no: 2238332 3:: :5 mean: Eva—2:808 25.6%: a mo 23:. Pa mo:_a> N 3253:. ”3:3— .Eé mu .5: m— .Eé x— .8250 _ _ _ c YW.....§. p .. . . ... .... .. “I TCN a 3.8 . . . - a. m. m. A l cg p r - m. 1 z , 11}. T a: e a 2:3 3 m E. .m - I 2: m N - . .- :2 m... - a: . a 3.3: :3 doom 5 $2: :0on mega 3.35 atonéem mtaexmwv 389:3 weapons no 35832”. mam—:8 “sebum: means %508 8% 8:8 Enemeom .84 2:9..— 37 These data support earlier observations reported by Nikolai et al., (2001) who found that dollar spot severity was reduced on rolled plots with every dollar spot outbreak that was recorded over a six-year period. Combined 2009 seasonal dollar spot means indicate a significant treatment effect on dollar spot in all rolling treatments (Figure 1.02). Additionally, the 2x a.m. treatment resulted in significantly less dollar spot compared to all other treatments (P < 0.05). Both the 1x am. and 1x p.m. treatment resulted in significantly lower dollar spot incidence compared to the control (P < 0.05), and no significant differences were observed among the two 1x rolling treatments ((1 = 0.05). In 2008, seven quality ratings were taken and are reported in Figure 1.03. Turfgrass quality was rated on a scale of 1-9, with 9 signifying excellent quality, 6 and above acceptable, and 1 indicating necrotic or dead turf. All rating dates resulted in a significant effect on quality for the 2x a.m. rolling treatment in 2008 (P < 0.05) compared to the control. The 1x rolling treatments (am. and p.m.) had significant effects on turfgrass quality when compared to the control at three observation dates (S-Jul, 26-Jul, lO-Oct). No differences in turfgrass quality were observed among the two 1x rolling treatments at any date in 2008 (P > 0.05). The 2x a.m. treatment had significantly better turfgrass quality than the 1x a.m. treatment on three dates (5-Jul, 26-Jul, lO-Oct) in 2008 (P < 0.05). Additionally the 2x a.m. treatment resulted in significantly better turfgrass quality compared to the 1x p.m. treatment on two dates (5-Jul, 26-Jul) in 2008 (P < 0.05). When a seasonal mean was calculated for 2008 turfgrass quality, both 1x rolling treatments were significantly better than the control while the 2x a.m. treatment, which 38 averaged a rating of 7.7, resulted in significantly better turfgrass quality than all other treatments (Figure 1.04). In 2009, quality ratings were taken on 7 different dates and are reported in Figure 1.05. The 2x a.m. treatment resulted in significantly better turfgrass quality compared to the control on six of the seven rating dates. The 1x a.m. treatment was significantly better than the 1x p.m. treatment on only one occasion early in the grong season (19- Jun); the rest of the rating dates resulted in statistically similar ratings for both 1x rolling treatments. Both 1x rolling treatments resulted in significant treatment effects on the final four rating dates in 2009 (Figure 1.05). Quality ratings from all rating dates in 2009 were averaged in order to get seasonal turfgrass quality means for treatments, which are reported in Figure 1.06. Both 1x treatments resulted in significantly better turfgrass quality compared to the control (P < 0.05), but were not significantly different from one another (P > 0.05). The 2x a.m. treatment resulted in significantly better turfgrass quality compared to all other treatments (P < 0.05). Percent volumetric water content (%VWC) of the top 1.5 inches (3.81 cm) of soil was measured in 2008 and 2009 on six and four occasions, respectively. In 2008, significant treatment effects were observed on three reading dates, data is reported in Figure 1.07. The 2x a.m. treatment had significantly higher %VWC compared to the control on three different occasions in 2008 (13-Jul, lO-Sep, and l7-Oct). The 1x rolling treatments (1x am. and 1x pm.) were not significantly different from one another on any rating date, and only resulted in a significant treatment effect on one measurement date. 39 .God A A: Oma— m..§_mE 3 wag—88 avenge bassoficwfi 8: 0.3 532 2:8 05 .3 332.8 £808 £83 wnufi 55:5 b. f3 0388: 8 coo—u n _ 28 633988 N 0 f3 E2388 u a 08:3 038 958 a; a no woman 5:95 a 3.5 usual woomS :o— woomRE woon Cw wooQomk. woomkk. ”GOQNNB «CONEE (6-1) Ammo Wyn]. a MS .Ed xN E .Ed 5 8 .Ed x. a 75:00 E .wcom 3 3:8 #:88me so $5.53.: wax—2 5% 303 E 385 MahazBQ “.28.“va mag—macs waives no mauve“ mum—«=0 do.— 2=wE 4O Figure 1.04. Seasonal turfgrass quality ratings for rolled treatments on creeping bentgrass (Agrostis palustris Huds.) putting green plots in 2008. 0 O. o O O. O %" .?0 V3; ’ 90 O O O. 90. ;p¢;v¢ ooeooo 5‘5... coc€§ fi’fl 0&5 v 0 ‘o'a'o'o O oo$U¥¢5 9. 'V'VOV 00.006 00.0090; v' 00 ’9 a .909 O fic”¢”’¢” 09. 000.096. 9000900090 ”00090. 0' . V§Vvv¢ob ’O..... 5%3flhfl%%% roooooooo .0. 500 90 O %‘5fl%fl%fl$5gg§f%fl‘ ' ar’avv033¢k.fi 9 9 0 O ¥%%%%%%%%%%%”&”,.O.. OOOOOOCOOODO.“ 9'9 e‘o o'o‘o'O‘o'o’e'o o o o o . . .. VVVVVVVVof' OOOOQOQOOOOQOOV 0.9.0.0009. O 090VQOOoootocooooooooooovoo a 9 O 9 O 0 9 o O 9 O 9 o 9 9 o o o 0.0 0.:‘o.¢‘o.o. .o. .9 o o .. §' .. .... 9. .0... 9.0.0.0900 O ' oJ3¢?§’oo€%ofifififififiyfiéfifi}fifl 3 ooeooooocooooooooo .00....‘0....O0..I 9%fififlh%%%%%%%%% 0990009090.... %yflfififiy”foog” 9650 o¢$§VV3oo eoodkoooooooo 9 J”%. 93”. 0" 33". a f. ”a 00990 f .. o O 353&%%§%$ 'o'o'c'o'o'o’c'o'o5.0333333a'e 9'9...o.o.o 0.0.0.0. . Q 0 o O o 9 ”33335903. 999.090. 0 0%. .J¢f. .O.%%&’Q.‘.0.0‘OOQQQOO 00. ‘fOOO 0% 0 000.090.00.00 0.0.. .00... “OO. 900‘... :.:. to a . ‘c% 000 o .oo....... 00 O o o . 9 —q>-V‘—______ .____ 8 6- mm swfim 41 1x p.m. 1x a.m. Control acceptable, and l = dead or necrotic turf. ,6: Rolling Treatment excellent " Quality based on a 1-9 rating scale where 9 T Within rating dates, means followed by the same letter are not significantly different according to Fisher’s LSD (P > 0.05). Amoo A .5 am»— m..§_m_m 3 map—88 “=9.an basaomzcwfi 8: 98 .832 088 05 .3 332—8 338 £83 maze.— 553» 5. f3 3850: .6 Eco n _ 23 633308 N o «:5 28:85 N o 80:3 28m 958 a; a no woman $.30 .. 8.5 «are «cows _ \o _ woomRR woom\m _ \w wooQoNk. woonF wooQNva woom\©\© (6'1) him-o mm a MS .Ed xN 3 .Ed x. N .E.w x_ Z 35:00 D .ooom E 32% «cobbmo no 358235 w£=8 new $03 E 335 fitmécn. utmoswwv 3.39:3 wfiooouo co mwawfi 5:30 .34 239m 42 .God A 5: Own. mzonmfi 9 wasp—88 Bogotmu 3885qu 8: 2a .532 089. 05 .3 330:8 83E $.28 wczfl 5.23 + ..tE 0:28: .5 cue—u u _ 93 63838.“ n 0 ..t3 «5:083 n a 22.3 038 mats. a; a no woman 3:95 .. Eofiuaouh uni—cm .Ed 5 .Ed 5 35:00 .. “. ...‘...“ O “ .1“. ‘ 1‘ ‘ ‘ . ' I n ‘ ‘..“‘ .“ ““ fl . a? 99 9999999 9909 a? . 99 c o 9999.999‘99 .1“ A ..9999479Q999cqu939 fiufi‘ofloufi‘o. ‘ 9x99999wufio9999om9999 999. .nv 0 9 9 .3» 9 . .mf aura“. 9‘ .omzv nxso»9nmoon“a. t omaad cxov . O .maflT O anVOow O O 9 . 99 o9 9 99v... 99 . .. 699999. c. 9. .9 99999 9 . . 99999 «NW 9 9 9 9 W9»... “90 9 .99 99. . I .1 9, .%9ofoo.f. 99999 4 9w 9 . .. . O Q ‘ ‘ ‘ A . . ‘ 9 9’99? 9949 . v 94? o 9 99.9 9“. o. } 0 o 96” 9 999 v! 0 99991 O . . . . , . o . 9.99 9 . . . ..9 999+ 99 .9999 9.999999 ’9 $9. . ,3. . 9..9 .199“ (a 999949 99999 99 . ‘ . O ... 0909 v o. O") 96! 099099909999 9. 9. A 99 99.9.9.9999 9.9.9.79 9. . 9 . vQ'. .....‘Q. ‘Q ...".‘C,, “ . . . 9"...” 99’3“. 9 999999 99 999 9999. 999 9 no. 9 99 I. v 999999999999 99. . o 9 9 .9. l 1‘! 9999. ||||ll l l ‘ .99 9999999 9. 1 I ‘ I, ll ‘ . . 9 .. . ‘ T. 9 . no 9 9 9 o o 9 9 9 9 9H 9 . . t9 9. .99 m3. .3? .ucacmm9.uu .c3 9 ct ow . n A . \. ... 9999 999 9o9999.»99%o99»9% no. . . _ ‘H . .u .. n ..msmAmejmor 99mg. .vw .mumcav .uwo. . ..a‘. . . .U n .9999. 999 9999999. 99 9H9 9 9 999V999. . . . . ’9 . . .x .1 .u.99999’99‘9999voouou9uu99é99nlr9 on . ., IA . 9.1?» ..m9 909994n099999N99’9 9 . . 999. V. AAT 9 OWV’OO .900 0’ g 009 .99 9 . 9 ...999 I 999 3.999999% 90 99 9 . 9 9 59999.. . 9%; .. .9 990999 9'99 999%. O o 99 9’99. 5 .. .99 9’99. 9.9 I 9 . 9999 99999909 .. . . .. . . .. .. 9 9 9 9 9 99.999999 99 .9 .. . I 99999 9 9 49.90.99,. ‘3112 ( . '9 9.999 of ’9 .... 99 fiofioéod , ‘ . . , X .y.. 0 . ..H. III | [I‘ll ’ | .9 O 90,90000 .6. 00 l I II I I I l H ‘ . . £0 0 O O. 9.! n x . VI ... a u .v . h o o 9 ya 9 o Luv ‘ . “,5 my. 9“. NY . . .. x . v o 99 099 , o . . . ... .V . .. . fl 0 . .. :a . 9' v ' 0;! '9 ob 6.3 f. «w o 9 9” am 0 K x O '.".' -_' fig: . ‘ 6 9% 95:; o 9 9 . ... .9. ’9 .39 9 9 9 O. O z; 9 ‘9 9 .’ 9 9 h C ‘ , v ‘0 9 9.9:. 9 9 4;». 9 9’9 9:» o 9:9:9 9,9,9 9 9 «:4 v '3 9 '9 9 9 Q o O 9 9 V O, I\ l --lll .99909999 9.9% 9 ...9 999 999999 .99 ’9 9 9 u 0 O. 9 6 0V! 0 «9 . 9 99 £9” 9 O O O O O O 0. O 99.... 9.9 9 v “9‘ . . . O 0' .000 ’0 ’0 . ...: ... 9 3%” .wa {*9 9 9mm“ ..,.. 2,. v 0 O . I... w. Nwflw kw». . ; E \O In (6-1) fluent) Wm a .88 a 3% :0on mafia «63$ fibuécm umhackwv‘v 383:3 waives no 938305 3:8 8% 388 bass: mag—mg Enemcom do; onE 43 .Amodv Om; £27.:— 3 @5508 East? 385%; 8: 93 3:2 oEam 05 3 832.8 v.52: $23 was”. 55$» ... 45:53.5 hog 2.28232 ooh: 98 JOE ..oa wwEEE 032x. 83 mo mamas; 98 £82: 2258: ... 039 “588382 woomkg: woonmE wooQoCa woomRCw wooQomR woon—R 2; 8a c3 23 2; 2:: cod“ 8.: 8.2 5:; m a 2.2 .Edifi cm 3“: Z: “.11: «no. . oodN .de A 92 . _% 93: 3.5 comm Bun—GOO D a Wmmnm SN ooém (mdep "1! 9'1) DMA % .88 5 3% «635 $5,323 QEEMS 389:3 mangoes :o 35:53.: w5=8 8% 282: 63% 38:8 86>» 050829, :5ch .84 Bin 44 t Figure 1.08. Seasonal percent volumetric water content (%VWC) means for rolling treatments on creeping bentgrass '(Agrostis palustris Huds.) plots in 2008. m o WEN”, , wt 7 ..o <3 . We V F! I 2&4 o 1 ~ 1- a , '— cc 0\ v—t : an r— 7 I I f V I I I I I I N O 00 \O V N O N N v—l v—t v-It v—i F-l O O o (de . ) /0 2x a.m. 1x p.m. Control Rolling Treatment "' Seasonal %VWC treatment means are the average of six different measurement dates in 2008. 1' Treatment means followed by the same letter are not 5 'stically different according to Fisher’s LSD (P > 0.05). Amoo A as am:— mzonmE 8 @6325 East? £89.38.“ 8: 0.8 .532 2:8 05 3 332.8 28:. £33 wave.— 553’ ... .Eogeob Ba meow—“£32 025 one .83 on $532 om mo «omega 3a 5808 Eon—82... ... open “CoEoosmmoz moonCA: moonCw aoomhk mooQNNG 006 com ooé coo % oo.w 2:: W comm U... coo; m. 8.3 m my w o. . .fl—‘n.m xNa ”a.m.. . 6N. ~—.h— OCWfi /\ Andy: § 3.: «mm: oodN . . on mom . E m a: % com oo NN 35:00 E m w.mm ooém + .ooom 5 32a «633% 9.333% SEEMS mmfiwfion wfimoouo no $5.53.: was—o.— .Sm ...maaoE 38:8 83>» 050829» :5ch do; 2:me 46 .Amoo A Ac amu— mLofiE 8 way—38 2.20:6 Encamzsm «on 98 .832 2:3 05 .3 330:8 288 2.258; ... .ooom E moat 20605308 820E“. Boo mo ounce; 05 2a 2:88 «5:535 032x. 338$ .. Eofiflmofi. waged .Ed US Add o3 1.2...“ a... 5. we: 3.950 (qxbp u; 9'1):m/\% , ,,.--1il M: E 2: I - II 1 cm i! I I mm I iI em .88 5 3% mass: “.333 3.8%3 $89qu mfiaoouo no 358305 wfize 8o 8.88 GER; 52:8 883 050839» 3083 Enemaom .o_ ._ onE O 47 The final measurement date (17-Oct) resulted in a significant treatment effect on %VWC for all rolled treatments. Treatment means for individual dates were combined to obtain a seasonal %VWC mean for treatments in 2008 and are reported in Figure 1.08. The 2x a.m. rolling treatment resulted in a higher seasonal %VWC compared to the control and was the only treatment with a significant effect on %VWC (P < 0.05). Two of the four dates in 2009 in which VWC measurements were taken resulted in a significant treatment effect on %VWC in the 2x a.m. treatment, (90-Aug and13-Oct); results are reported in Figure 1.09. No significant treatment effects on VWC were observed with either of the 1x rolling treatments at any measurement date in 2009. Treatment means from each measurement date were averaged in order to obtain a seasonal %VWC mean for treatments; results are reported in Figure 1.10. Similar to 2008, only the 2x a.m. rolling treatment was significantly different from the control (P<0.05), resulting in a 3.35% increase in %VWC. Neither of the 1x rolling treatments had a significant effect on %VWC (P. 0.05). CONCLUSIONS It is widely accepted that the removal and dispersal of dew from turfgrass by early morning mowing plays a significant role in reducing dollar spot disease (Williams and Powell, 1995; Ellram et al., 2007). However, results from this study indicate alternative mechanisms in disease control. Many questions remain regarding lightweight rolling, particularly concerning the mechanisms involved in dollar spot reductions. A common theory revolves around dollar spot incidence being reduced on rolled plots due to the removal or dispersal of excess moisture on the turfgrass canopy after initial mowing. The 48 moist, humid environment is thought to not only serve as a microclimate conducive for fungal growth and infection, but guttation exuded from freshly mown turfgrass is thought to supply a nutrient-rich food source to the dollar spot pathogen, Rutstroemia floccosum (Williams and Powell, 1995). In an attempt to challenge the notion that dollar spot reduction is solely due to the removal of dew and guttation fluid by morning rolling, one rolling treatment was implemented in the morning hours, while dew was still present on the plant canopy, and one in the aftemoon hours, after surface moisture had naturally evaporated. Both morning and afternoon rolling treatments resulted in significant reductions in disease. No significant differences were found between the 1x a.m. treatment and the 1x p.m. treatment in either year of the study. These results confirm the observation that daily rolling of turfgrass plays a role in disease suppression but challenge the hypothesis that dew and guttation removal from early morning rolling is the underlying mechanism contributing to dollar spot reduction on creeping bentgrass putting greens. Afternoon rolling treatments proved to have a statistically equal treatment affect on dollar spot reduction as morning rolling treatments. Additionally, no statistical differences existed between the aftemoon and morning treatments. This suggests that, aside from the physical removal/dispersal of surface moisture, possible biological or physiological mechanisms could be involved in the reduction of dollar spot on creeping bentgrass putting greens when daily lightweight rolling is taking place. Rolling five days week'l, regardless of the time of day, consistently resulted in lower disease incidence, as well as superior turfgrass quality ratings, in 2008 and 2009. In order to investigate whether or not rolling has cumulative effects on dollar spot 49 reduction, the 2x a.m. treatment was implemented. This 2x day'l rolling treatment resulted in consistently lower dollar spot observations compared to both the control and the 1x rolled treatments, as well as higher turfgrass quality ratings in both years of the study. Greater reductions in dollar spot counts, significantly better turfgrass quality, and high probabilities of treatment effects on plots rolled 2x day'l, particularly at the conclusion of the second year, are indications of a possible cumulative effect rolling may be having on disease suppression and turfgrass health. These results are consistent with Nikolai et al. (2001) in the respect that greater differences with regard to disease occurrence, existed between rolled and non-rolled plots as the study progressed for multiple years. Contrary to previous rolling research, Horvath et al. (2009) observed no differences in dollar spot severity on rolled and non-rolled plots of creeping bentgrass. However, rolling treatments were implemented 2-3 times week'1 and results were based on six weeks of data collection. Cumulative effects of rolling that have been observed herein as well as in Nikolai et al. (2001) are based on multiple years of data collection, significant treatment effects were typically a result of season long rolling, not observed until later observation dates in the season, or until seasonal averages were obtained. Average volumetric water content was significantly higher in the 2x a.m. treatment in both years of the study, and both 1x rolling treatments trended towards higher %VWC when compared to the control. These observations not only suggests that rolling may be contributing to to greater water holding capacity in the upper root zone of the turfgrass canopy, but support observations by Couch and Bloom (1960) and Liu et al. (1995), where higher soil moisture resulted in reduced dollar spot development and incidence. Reports by Danneberger (1989), Hamilton et al. (1994), and Hartwiger et al. 50 (2001) indicated rolling did not increase soil compaction of putting greens constructed with a high sand content root zone. While this may be true, rolling may be contributing to a decrease in pore size in the top 1.5 in. (3.81 cm) of the root zone. Smaller pores equate to a greater attractive force by which water can be held. Additionally, volumetric V 6 = l water content can be defined as VT, where V.,. is the volume of water and V T is the total volume associated with the soil (i.e. soil volume + water volume + void space). If pore size is decreased due to slight compaction in the upper (top 3.8] cm) root zone, a reduction in void space takes place, thus lowering VT in the equation. This may be responsible for increasing 9 or the total volumetric water content measurements in rolled treatments. The phenomenon by which rolling inhibits disease incidence was investigated. Rolling had a significant effect on dollar spot incidence in both years of this study. Morning and afiemoon rolling treatments had similar effects on dollar spot reduction compared to the control. These similarities between morning and aftemoon rolling suggest mechanisms other than secondary dew/guttation removal may be responsible for dollar spot reductions. Many theories still remain, and future rolling research should be conducted in order to investigate the potential biological or physiological impacts associated with turfgrass rolling. This study also revealed a trend towards cumulative effects of rolling on turfgrass health and quality. Multiple rolling applications per day resulted in substantially better turfgrass quality with less disease when compared to treatments rolled once per day. Continued research investigating this effect, particularly in aftemoon or dew/guttation free settings should be pursued. The mechanism by which rolling is detrimental to dollar spot incidence is most likely multi-faceted, involving 51 aspects of soil physics, microbial ecology, plant physiology, and plant pathology or epidemiology. In the forthcoming chapter, one particular mechanism regarding microbial populations is discussed. 52 LITERATURE CITED Beard, J. B. 1994. Turf rolling. Grounds Maintenance. 29: 44-52. Couch, H. B. and Bloom, J. R. 1960. Influence of environment on diseases of turfgrasses. II. Effect of nutrition, pH, and soil moisture on Sclerotinia dollar spot. Phytopathology. 50: 761-763. Danneberger, K. 1989. No speed limit. Landscape Manage. 29: 66—70. Demoeden, P. H. 2000. Creeping bentgrass management: Summer stresses, weeds, and selected maladies. John Wiley & Sons, Hoboken, NJ. pp. 10-12, 47. DiPaola, J. M. and Hartwiger, C. R. 1994. Green speed, rolling and soil compaction. Golf Course Management. 62(9): 49-51, 78. Ellram, A., Horgan, B., and Hulke, B. 2007. Mowing strategies and dew removal to minimize dollar spot on creeping bentgrass. Crop Sci. 49: 2129-2137. Hamilton, G. W. Jr., Livingston, D. W., and Grover, A. E. 1994. The effects of lightweight rolling on putting greens. Science and Golf II. pp. 425-430. Harban, W. S. 1922. The effect of trampling and rolling on turf. Bulletin of the Green Section of the US. Golf Association. 2: 148-150. Hartwiger, C. 1996. The ups and downs of rolling putting greens. USGA Green Section Record. 34: 1-4. Hartwiger, C. E., Peacock, C. H., and DiPaola, J. M. 2001. Impact of lightweight rolling on putting green performance. Crop Sci. 41: 1179-1184. Horvath, B. J., Nichols, A. 13., and Cutulle, M. A. 2009. The effects of mowing height and rolling on ball speed, quality and disease severity of creeping bentgrass. USGA Turf. and Env. Res. Online. 8: 1-5. Inguagiato, J. C., Murphy, J. A., and Clarke, B. B. 2009. Anthracnose disease and annual bluegrass putting green performance affected by mowing practices and lightweight rolling. Crop Sci. 49: 1454-1462. Kussow, W. R. 1998. Putting green management systems. Wisc Turf Res. Reports. XV: 85-93. Liu, L. X., Hsian. T., Carey, K., and Eggens, J. L. 1995. Microbial populations and suppression of dollar spot disease in creeping bentgrass with inorganic and organic amendments. Plant Dis. 79: 144-147. 53 Nikolai, T. A., Rieke, P. E., Rogers, J. N. III, Vargas, J. M. Jr. 2001. Turfgrass and soil responses to lightweight rolling on putting green root zone mixes. International Turfgrass Society Research Journal. 9: pp. 604-609. Piper, C. V. and Oakley, R. A. 1921. Rolling the turf. Bulletin of the Green Section of the US. Golf Association. 1: 36. Radko, A. M. 1977. How fast are your greens? USGA Green Section Record. 15: 10-11. Richards, J., Karcher, D., Nikolai, T., Richardson M., Patton, A., and Landreth, J. 2008. Mowing height, frequency, and rolling frequency affect putting green speed. Arkansas Turfgrass Report 2007. Ark. Ag. Exp. Stn. Res. Ser. 557: 52-56. Smiley, R. W., Demoeden, P. H., and Clarke, B. B. 2005. Compendium of turfgrass diseases. 3rd ed. Am. Phyto. Soc., St. Paul, MN. Stier, J. 2006. Shorter mowing heights are hazardous to summer health. The Grass Roots 35: 4-5,7,9. Throssell, C. 1986. Management practices affecting putting green speed. The Bull Sheet 39: 4,7,9. USGA Green Section Staff. 1993. USGA recommendations for a method of putting green construction. USGA Green Section Record. 31: 1-3. Vargas, J. M. Jr. 2005. Management of turfgrass diseases. 3rd ed. CRC Press, Inc., Boca Raton, F L. pp. 19-22. Williams, D. W. and Powell, A. J. Jr. 1995. Dew removal and dollar spot on creeping bentgrass. Golf Course Management. 63: 49-52. Youngner, VB. 1969. Physiology of growth and development. In Hanson, A. A. and Juska, F.V. (ed.). Turfgrass science. ASA, Madison, WI. pp.187-216. 54 CHAPTER TWO TURFGRASS ROOTZONE MICROBIAL RESPONSES TO ROLLING ABSTRACT Relative abundance of microorganism community populations is useful information for estimating the effects of particular management strategies on rhizosphere soil (REF). A two-year rolling study was conducted during the growing seasons of 2008 and 2009 and relative abundance of soil microorganisms under a creeping bentgrass (Agrostis palustris) putting green was elucidated via phospholipid fatty acid (PLF A) analysis (REF). Treatments rolled twice in the morning (2x a.m.) and once in the morning (1x a.m.) both showed significant increases in particular PLFA abundance when compared to the control (P < 0.05) all of which are bacterial biomarkers. When combined to form specific PLF A taxonomic biomarker groups, relative abundance was found to be higher in all rolled treatments compared to the non-rolled control, while fimgal and other PLF A relative abundance remained relatively constant. A trend towards higher bacterial abundance in rolled treatments suggest a potential biological effect of lightweight rolling possibly attributed to the aforementioned increase in volumetric water content observed in rolled treatments (Chapter 1). Increases in bacterial relative abundance may play a role in dollar spot (Rutstroemia floccosum) reductions due to various mechanisms of fungal disease suppression often observed in many soil dwelling bacterial organisms. 55 INTRODUCTION The rhizosphere is a complex community of soil microflora and microfauna. Along with facilitating plant infection and disease, many microorganisms play beneficial roles in the root zone of a turfgrass stand. Microbial communities have been implicated in suppressing soilbome plant diseases, promoting plant growth, and in changing vegetation (Doran et al., 1996). Several studies have elucidated the significance of microbial activity in soil related to and responsible for the suppression of plant pathogens. For example, relationships have been found between microbial diversity and root disease suppression (Nitta, 1991; Workneh and van Bruggen, 1994). Rovira & Wildermuth (1981) indicated that the microbiota in a “rich” soil tends to reduce the severity of attack by many plant pathogens or, in other words, soils higher in microbiota content and diversity show tendencies towards general disease suppression. Bacterial populations in the soil and rhizosphere have long been of interest to agronomists and plant scientists, partially due to the observation that antagonism and competition between bacteria and many important plant pathogens can be exploited as possible control mechanisms. An active microbial population performs and plays a role in many beneficial activities and processes such as organic matter decomposition, assisting in plant nutrient acquisition, availability and recycling, and plant pathogen suppression (Sylvia et al., 1997). Thus, turfgrasses grown in rootzones containing lower microbial populations may be less healthy and possibly more easily affected by turfgrass pathogens, resulting in an overall lower quality turfgrass than those grown in the presence of higher microbial populations (Hodges, 1990; Couch, 1995). In turfgrass systems, successful suppression of fungal diseases such as dollar spot (Rutstroemiafloccosum) and 56 anthracnose (Colletotrichum cereale) has been achieved both in vitro and with field applications of an antibiotic producing Pseudomonas spp. (Powell, 1993; Uddin and Viji, 2002). These bacterial metabolites are thought to contribute to the antagonistic effects on plant pathogenic fungi. Bolton et al., (1985), Fraser et al., (1988), Kirchner et al., (1993), and Powlson et al., (1987) have documented associations between agricultural management practices and microbial community make-up and activities. Particular management practices in different systems may alter the composition and structure of soil microbial communities. This has been documented in agricultural soils by examining either whole soil microbial communities or specific functional groups (Ka et al., 1995; Lundquist et al., 1999; Webster et al., 2002; Clegg et al., 2003). Microbial populations fluctuate, and their community structure and composition is highly sensitive to change with intense management (Donnison et al., 2000). Disease suppressiveness has been achieved through crop rotation (Cook et al., 2002), intercropping (Schneider et al., 2003), residue destruction (Baird et al., 2003), organic amendments (Tilston et al., 2002), and tillage management practices (Sturz et al., 1997; Pankhurst et al., 2002). All of these cultureal practices contribute to altering the microbial community in a way that favors plant health and hinders disease development. A useful approach to estimating the fungal and bacterial abundance in soil is to measure chemical components that are specific for microbial groups. One such approach relies on bacteria and fungi having different fatty acid compositions in their phospholipids (Hardwood and Russell, 1984; Tunlid and White, 1992). Phospholipid fatty acids (PLFAs) are widely accepted as biomarkers that indicate viable microbial 57 biomass and provide a microbial community ‘fingerprint’ (Vestal and White, 1989; Zelles, 1999). Changes in soil microbial communities under many different experimental applications have been successfully measured using PLF A analysis. Fatty acid patterns have been used to distinguish microbial fluxes between different types of cultivation (Zelles et al. 1995), fertilizer treatments (Guo and Wang, 2009), pH gradients (Baath and Anderson 2003), agricultural management systems, seasons, soil types (Bossio et al. 1998), and changes in the soil microbial community due to pollution (Baath et al., 1992). A quantitative measure of microbial communities can be obtained using PLFA analysis as opposed to most current DNA-based fingerprinting methods (Muyzer et al., 1993; Kowalchuk et al., 1997). While many studies have shown the direct effect that high populations of microbes have on plant pathogens in the rhizosphere, much remains unknown with regard to their disease suppression mechanisms. Numerous species within several genera of bacteria have been reported to be effective biological control agents for root and foliar diseases on turfgrass. The mechanisms of biological control are thought to occur mainly through the process of antagonism. Antagonism can be defined as active opposition that results from the production of substances by one organism that are toxic to other organisms. Such substances can be classified as antibiotics that cause lysis or death of microbial competitors (Rovira and Wildermuth, 1981). Additionally, competition for food, oxygen, and space is often considered a form of antagonism or a distinct mechanism of biological inhibition all together (Baker and Cook, 1974). The purpose of this study was to determine the effects of lightweight rolling of creeping bentgrass putting greens in relation to soil microbial populations. Results of 58 dramatic decreases in disease incidence caused by the fungal pathogen R. floccosum, have been outlined in the previous chapter. Microbial community assessment was explored in order to test the hypothesis that rolling is having an effect on the community dynamics, possibly related to disease suppression. Microbial community composition was estimated using PLFA analysis of root zone samples from treatment plots in order to make comparisons among rolled and non-rolled turf stands. MATERIALS AND METHODS Field research was conducted at the Hancock Turfgrass Research Center on the campus of Michigan State University, East Lansing, Michigan, on an experimental putting green constructed in 2005 and seeded with creeping bentgrass (Agrostis stolonifera cv ‘Independence ). Research plots were 7 ft x 12 ft (2.13 m x 3.65 m) arranged in a randomized block design with three replications for each rolling treatment. Prior to initiating the study, each plot was randomly assigned a rolling treatment. Mowing took place between 6:00 and 7:00 a.m. and was implemented on all plots prior to any rolling. Each rolling treatment was carried out 5 days week.l (Monday-Friday) and were as follows: 1) Control (no rolling), 2) rolled once (1x) in the a.m. immediately after mowing, 3) rolled once (1x) in the p.m. when dew and guttation water had dissipated, and 4) rolled twice (2x) in the a.m. immediately after mowing. The morning rolling treatments were implemented between 7:00 and 8:30 a.m., and the afiemoon rolling between 1:00 and 2:00 p.m. Rolling of plots was conducted with a Tru-Turf ride-on greens roller, model RS48-11B, with a 39 inch (99cm) roll swath weighing 562 pounds (255kg) without an 59 operator. To ensure study uniformity, rolling was the only treatment that differed in each plot. The root zone mix consisted of a sand based 80:20 (sandzpeat v/v) mix constructed to USGA recommendations (U SGA, 1993). Sand topdressing was applied to the entire research area on a light, fiequent (bi/tri-weekly) basis throughout the growing season in order to simulate typical golf course putting green maintenance practices. No vertical mowing or core cultivation occurred on the research plots during the course of the study in order to minimize turfgrass and soil disruption. Fungicides were not applied to any of the research plots during the duration of the study. Insecticides and herbicides were applied only on an as needed basis, and were applied uniformly over the entire study area when necessary. Soil samples from plots were taken at a depth of 1 in (2.54 cm) with a 1 in (2.54 cm) diameter soil probe. Ten cores were randomly taken from each plot on September 25, 2009. Top growth, including turfgrass leaves, stems, roots, and debris was removed from each core by cutting with a sterile razor blade just below the thatch/soil interface. Cut ends were gently shook so that loose soil that remained in thatch and roots would be included in the sample. Unused top growth was discarded. The remaining upper rootzone soil was homogenized by vigorous mixing in plastic bags. Thirty grams from each of the representative, homogenized soil samples was measured and placed into separate 3 1/8 X 5 1/2 inch paper coin envelopes before being lyophilized. Freeze dried soil samples from each plot were ground using a pestle and mortar until they passed through an ATM #40 (425 um) U.S. standard testing mesh sieve (Advantech Manufacturing Inc., New Berlin, WI). Ten grams of the ground material from each sample were stored in 2 oz. Nasco 60 Whirl-Pak bags at -20° C prior to their shipment to the University of Wisconsin, Madison, Department of Soil Science for PLFA analysis by Dr. Teri Balser’s laboratory staff. Lipids were extracted from 3 g (dry weight) subsamples using a modified Bligh and Dyer (1959) technique as described in Balser and Firestone (2005). Samples were analyzed using a Hewlett-Packard Agilent 6890A gas chromatograph (GC) (Agilent Tech. Co., Santa Clara, CA) equipped with an Agilent Ultra-2 (5% phenyl)- methylpolysiloxane capillary column (25m by 0.2m by 0.33pm) and flame ionization detector (F ID). Peaks were identified using a mix of known FAME standards and comparing retention times or estimated chain lengths (ECL) from each sample output to a naming table. These naming tables have ECL’s for hundreds of known lipids and a f‘naming window”, which is the accepted amount that an unknown can vary from the ECL and still be assigned a particular name. The fatty acid analyses were carried out by an MIDI Sherlock microbial identification system (Version 4.5, MIDI, Newark, NJ). Lipid data were converted from raw peak area to nmol/g soil in order to facilitate statistical analysis. Peak area was converted to ug carbon (C) using conversion factors from internal standards run on the GC. Conversion factors are equal to pg C of the standard divided by the area of the standard on the GC and are typically the averages of two internal standard lipids, 9:0 and 19:0 in order to capture the response of both short- and long-chain lipids. This number was then divided by the unit weight soil in each sample (approximately 3g), and finally converted to umol C g soil’I using the molecular weight of each lipid. 61 Processed lipid data are expressed as abundance (nmollipid gsoil-l), mole fraction (nmollipidx nmoltotampid-IQ (O-l)), or mole percent (mole fraction*100, (0-100%)). Mole fraction and mole percent are normalized by the total biomass in a sample and are thus measures of the relative abundance of any given lipid. Mole fraction is appropriate for use in ordination analyses (after transformation, e.g. arcsine, square-root), while mole percent (mole fraction multiplied by 100) is an easy-to-interpret value. Lipid abundances are the absolute amount of a given lipid extracted per gram of soil. Because the quantity of lipid per cell is reasonably constant, and the lipid extraction is highly quantitative (i.e. close to 100% extraction efficiency), abundance is, in effect, an estimate of microbial biomass. Total abundance is total biomass, and the abundance of key indicators reflects the biomass of the group it represents (T. Balser, pers. comm). Fatty acid nomenclature used in this study is as follows: total number of carbon atomsznumber of double bonds, followed by the position of the double bond ((0) from the methyl end of the molecule. Cis and trans geometries are indicated by the suffixes c and t, respectively. The prefixes a and 1' refer to anteiso- and iso-branching, respectively. Methyl groups on the tenth carbon atom from the carboxyl end of the molecule are indicated by. The positions of the hydroxyl (OH) groups are noted when necessary, while ey indicates cyclopropane fatty acids (Bossio et al., 2006). Each PLFA value is represented by the mean of three soil extraction replications. Particular individual fatty acids extracted from samples have been used as signature indicators for various taxonomic groups of microorganisms. Combinations of particular PLFAs were considered to be representative of groups 0f organisms of interest in the soil. The PLFAs 18:1(o9c, 18:2(o6c, and 18:3(o6c were 62 used to represent soil fungi (Myers et al., 2001 and Vestal and White, 1989). Gram- positive (gm+) bacteria were represented by i15:0, 015:0, 15:0, i16:O, 17:0, i17:0 and a17:0, while 16:1m7c, cyl7:0, cyl9:0 and 18:1co9t were used to indicate gram-negative (gm-) bacteria (Ratledge and Wilkinson, 1988 and Zogg et al., 1997). Total bacteria were represented by the combined gm- and gm+ bacterial indicators (Ratledge and Wilkinson, 1988 and Zogg et al., 1997). Treatment effects on PLFA relative abundance were analyzed using the PROC Mixed procedure in the SAS software (SAS Institute Inc. 2009, Cary, NC, USA). Analysis of variance (ANOVA) followed by Dunnett’s test, for the comparison of treatment means to the control for individual fatty acid mole percentages, were carried out. Combined fatty acids representing groups of microorganisms were also subjected to ANOVA as well as mean comparisons based on Dunnett’s test. A principal component analysis (PCA) of the arcsine transformed mole fraction PLFA data was carried out with the correlation matrix. RESULTS AND DISCUSSION Mole percent is considered an indication of relative abundance in relation to total PLF A abundance. As FID response from gas chromatography is proportional to molecular mass, results were expressed as molar percentages. Thirty fatty acids were identified for use in data analysis. Individual fatty acid relative abundances were compared in order to investigate possible differences due to season long lightweight rolling treatments. All PLFAs used in analyses were evaluated for normal distribution before being subjected to significance testing. Significant differences in particular fatty 63 acid mole percentages existed between rolled and control treatments in rootzone soils of the creeping bentgrass putting green. These PLFA indicators, mole percentages, and references, are reported in Table 2.01. Of the thirty microorganism indicator PLFAs extracted, eight were significantly higher in mole percentage in the 2x a.m. rolled treatment than in the untreated control (P S 0.10). These included the straight chain saturated fatty acids 12:0, 14:0, 15:0, 18:0, 19:0cy, 20:0, 9:0 (bacterial indicators) and the monounsaturated fatty acid 18:1w9t (actinomycete indicator). Other significant differences in individual PLFAs existed between the 1x a.m. treatment and the control (P S 0.10). These included cyclopropane fatty acids 17:0cy and 19:0cy, straight chain saturated fatty acids 18:0, 20:0, and 9:0, and the monounsaturated fatty acid 18:1 w7c (all bacterial indicators); the polyunsaturated fatty acid 18:2(06c (saprophytic fungi indicator); and the methyl branched fatty acid 117:] and monounsaturated fatty acid 18:1(o9t (both actinomycete indicators). 64 Table 2.01. Relative abundance (mol %) of PLFA indicators in turfgrass root zones. PLF A 12:0 14:0 114:0 30H 14:1(05e 15:0 a15:0 115:0 16:0 [16:0 116:1 I-I 16:1m5c 16:1(o7c 16:1(o9c 17:0 (117:0 17:0cy 117:0 17:1m8c 18:0 18:10)5c 18:10)9c 18:1(n9t 18:3cu6c 19:0 19:0cy 20:0 9:0 117:1 18:206c 18:1037c Rolling Treatment Control 0.7446 0.9251 0.3279 0.2097 0.4878 1.1085 2.4031 15.949 0.5792 0.5080 36.410 2.6360 0.4481 0.1997 0.3455 0.5650 0.3455 0.4281 0.6418 0.4178 7.5882 0.2344 0.5733 0.1188 0.5445 0.2255 0.2318 1.2652 3.6361 4.9095 1x a.m. 0.7797 1.0461 0.4073 0.2225 0.5458 1.4009 2.9951 15.599 0.7681 0.5875 31.143 3.0052 0.5341 0.2736 0.4717 0.77704 "' 0.4717 0.5226 0.9294 * 1.2252 7.4477 0.29724 * 0.6938 0.1500 0.8611 ** 0.33765 ** 0.28886 * 1.6838 * 4.0150 6.0582 * 1x p.m. 0.8962 1.0315 0.3889 0.2189 0.5477 1.3757 2.8863 15.540 0.7288 0.6872 32.753 2.9218 0.5392 0.2506 0.4321 0.7137 0.4321 0.6033 0.7805 0.8376 7.5651 0.2694 0.6054 0.1235 0.7305 0.2714 0.2576 1.5674 3.1897 5.2525 2): a.m 1.0465 ** 1.1277 ** 0.4011 0.1428 0.5591 "' 1.3671 2.9857 15.714 0.7313 0.4279 31.896 2.9093 0.5150 0.1780 0.4319 0.7324 0.4319 0.5518 0.8558 ** 0.4048 8.2416 0.3214 *"‘ 0.6865 0.1483 0.7731 * 0.3068 * 0.2904 * 1.5392 3.6111 5.8397 PLFA Biomarker't Bacteria Bacteria Unknown (Gram -) bacteria Bacteria (Gram +) bacteria (Gram +) bacteria Bacteria and fungi (Gram +) bacteria (Gram -) bacteria Arbuscular mycorrhizae Gram- bacteria Unknown Bacteria (Gram +) bacteria Gram - /anaerobes (Gram +) bacteria (Gram -) bacteria Bacteria Unknown Saprophytic or ectotrophic fungi Actinomycetes Saprophytic/ ectotrophic fungi Bacteria Gram - / anaerobic bacteria Bacteria Bacteria Actinomycetes Saprophytic fungi (Gram -) bacteria *, " Means within the same row are significantly higher than the control according to Dunnett’s test at the P S 0.10 and 0.05 probability levels respectively. 1' PLFA biomarker indicators according to: Bossio et al., 1998; Ratledge and Wilkinson, 1988; Zogg et al., 1997; Frostegérd and Baath, 1996; Myers etal., 2001; Vestal and White, 1989; Turpeinen et al., 2004; T. Balser, personal Communication, 2009. 65 The 1x p.m. rolling treatment resulted in no significant differences in any of the 30 indicator PLFAs when compared to the control, but remained higher in relative abundance than the non-rolled control in most of the bacterial indicators. The 1x p.m. treatment also did not stray statistically from the other rolling treatments in any of the individual PLF A relative abundances. Conversely, the untreated control had significantly higher proportions of the PLFA 16:1m5c (Arbuscular mycorrhizae indicator) compared to the 1x a.m. rolling treatment. Commonly used combinations of PLFAs were analyzed and used to represent taxonomic groups of interest in the root zone soil samples and make comparisons among rolling treatments, PLFA combinations are reported in Table 2.02. These combinations rendered significant differences among treatments as indicated in Figure 2.01. Gram- negative bacterial PLF A relative abundances were significantly higher (P < 0.10) in plots rolled 1x a.m. (10.70%) compared to the control (8.66%), with a standard error (SE) of 0.8941. Treatments 1x p.m. and 2x a.m. resulted in 0.963% and 1.599% higher gram negative bacteria PLFA abundances than the control respectively. Gram-positive bacterial PLFAs were not statistically different in any treatment but were slightly higher in all rolled treatments than the control. Total bacterial PLF A mole percentages were significantly higher (P < 0.10) in the 1x a.m. rolled treatment compared to the non-rolled control. A trend towards higher total bacterial PLF A indicators existed in all rolled treatments when compared to the control with a difference of 2.15% and 2.82% in the 1x p.m. and 2x a.m. treatments respectively. 66 935m .owo_ .223 one 38> :oom ..E 8 E32 .382: one 38me dosh“: 36323::on $55 .33 .8183: 933m SSE 3383 ...m 8 wwoN mama: £85653 can owcocmm one Como. .moo”: .ohsfio— one 33.5230 “63339880 .89 .onie 3.883 ..9 so wwom $2 .882; Ea 839m .3 ... .9: .98.. .92 92.192.~ 3.5.2 853m ...fimaéso .33 .5323 one Emmoaem dos—”w— MSE ..9 s wwom £2 685:? .592 .59: .0292 .33 mean 2332 «team one owoocmm ”woe ..B B Emmom our: .o”: 66: .onfl .926 6an mood—8:85 535““ 95..» 82.0.5.3: Eco—ha:— e5 wage :8. ...e .0 $50M #02 .._e 8 335: 32 ...e 5 23933 32 ...e .0 Eaepox KS ...e e :53 82 .5; es. 936 ”32 ...e .... £853. 62: 68395 2... £33.60. 000— ...e .0 coma—:05. ”N02 £ch use com—02 :00— .cEU use 00.202 N03 £qu use com—oz 32 coy—em use 953 $2 Exam 23 mac? ”:3— ...oxem 05 $.03 32 ...e 3 83o: 88 ...e .0 :.> 82 ...e .0 noswceoe ”32 ...e .... 9&8: :3. ...a a £328 KS. ...a a 523. 000— .mew._e> 5ch ..e .0 =03?— Emmcowefia Emmcowe0=< EmEoweE< EmEomeE< EmEoweE< 8:555 625 Simona—00m .8250 EmE0weE< Emmcoweucxx MED; coco—60.0 0:05:59 5502005 30033.2 00.. 000.. 8253 £030 5505 40% .30— w=_._% 40% 8:00 seen hoe—ham seen ..oEEzm gems e8. 82285 00% meo— w:_._% £23 eofiEsm 2030 .5883. 4:33 8358 40% 6:00 £93 8:25 £080 =e oxek 00% wee. 40% 3:09 £030 0305 40% nee. haw 40% 8:0D :88 :90er 40% meo— 40% ue=0fl :83 858% 40% «ee— 40% 8:03 saw neoafieimhm. d% 35.83% utmomeogefi Sunken. hmkshwmgu §QEQ~§§QR SNEQSBE matefiefituk NQUUQNU kNubQQQLNkKW «9%. metefiohzmmm 35:52:: 633% w0=050~§eek :Mxecfis nczesoheemk 3.05%?»0 eczefiehammk utmoeexoaiatethzmnk m§30\0m.50 gefiehemem 82.9.80”— Eat—2.002 02.0.3an 023me Seesaw—O .mcowofiea $83.50 008800 «memewe meooto oumEomeEe £5 mEmEewuo .etowoem .mo.m 29¢. 72 much research is still required in order to draw any substantial conclusions. PLF A profiles can be used to fingerprint the structure of soil microbial communities and measure their biomass (Bossio and Scow 1998; Bossio et al., 1998). These methods are free of the distortion associated with the requirements for quantitative removal of microbes from surfaces or the selectivity associated with growth on artificial media (White 1988; Bossio & Scow 1998). PLFA measurement can provide detailed information about the structure of the active microbial community because only lipids from living organisms are measured (Vestal & White 1989). As a result, PLFA profiles can be useful in predicting and manipulating physical and chemical factors in soils to sustain long-term pathogen suppression (Chen et al., 1988). It must be noted that microbial populations, particularly biomass and metabolic activities, in soil environments are extraordinarily dynamic, dependent on factors such as temperature, moisture, radiation, and atmosphere, and can change drastically throughout the growing season and even over the course of one day (F. Dazzo, personal communication, March 1, 2010). Population fluxes occur between areas within close proximity to one another as well. While the study area examined was maintained relatively consistently throughout (with exception to rolling treatments), many spatial discrepancies can result with regard to the abovementioned environmental factors. For this reason, relative abundance was chosen for comparative analysis in order to limit the misinterpretation of population estimates. By measuring the amount PLFAs belonging to particular taxonomic groups in relation to the total amount of PLF As extracted, conclusions could be drawn regarding population affects due to rolling with minimal concern to variations due to sampling parameters. 73 Perhaps, future rolling studies can implement multiple soil sampling dates for PLFA analysis during a growing season. Additionally, incorporating molecular-based techniques by which community diversity and qualification can be elucidated should be considered. Identifying key microorganisms associated with cultural practices such as rolling and microbial disease suppression is crucial in understanding the intricacies in disease control mechanisms. This initial study is meant to serve as a foundation for future projects aimed at elucidating soil microbial characteristics linked to turfgrass disease suppression. 74 LITERATURE CITED Austin, B., Dickinson, C. H., and Goodfellow, M. 1977. Antagonistic interactions of phylloplane bacteria with Drechslera dictyoides (Drechsler) Shoemaker. Can. J. of Microbio. 23: 710. Béath, E., Frostegard, A., and Fritze, H. 1992. Soil bacterial biomass, activity, phospholipid fatty acid pattern, and pH tolerance in an area polluted with alkaline dust deposition. App. and Env. Microbio. 58: 4026-4031. Baath, E. and Anderson, T. H. 2003. Comparison of soil fungal/bacterial ratios in a pH gradient using physiological and PLFA-based techniques. Soil Biol. & Biochem. 35: 955- 963. Baird, R.E., Watson, CE, and Scruggs, M. 2003. Relative longevity of Macrophomina phaseolina and associated mycobiota on residual soybean roots in soil. Plant Dis. 87: 563-566. Baker, K. F. and Cook, R. J. 1974. Biological Control of Plant Pathogens. San Francisco, CA: W. H. Freeman and Company. p. 431. Baldwin, N. A., Capper, A. L., and Yarham, D. J. 1991. Evaluation of biological agents for the control of take-all patch (Gaeumannomyces gramim’s) of fine turf. In Beemster, A. B. R. ed. Developments in Agricultural and Managed-Forest Ecology. Amsterdam: Elsevier Science Publishers, pp. 231-235. Balser, T. C. and Firestone, M. K. 2005. Linking microbial community composition and soil processes in two California ecosystems. Biogeochemisry. 73: 395-415. Bardgett, R. D., Lovell, R. D., Hobbs, P. J., and Jarvis, S. C. 1999. Seasonal changes in soil microbial communities along a fertility gradient of temperate grasslands. Soil Biol. Biochem. 31: 021-1030. Bligh, E. G. and Dyer, W. J. 1959. A rapid method of total lipid extraction and purification. Can. J. of Biochem. and Phys. 37: 911-917. Bolton, J., Elliot, L. F ., Papendickc, P. R., and Bezdiccek, D. F. 1985. Soil microbial biomass and selected soil enzyme activities; effect of fertilization and cropping practices. Siol Biol. Biochem. 17: 297-302. Bossio, D. A., Fleck, J. A., Scow, K. M., and Fujii, R. 2006. Alteration of soil microbial communities and water quality in restored wetlands. Soil Biol. Biochem. 38: 1223-1233. Bossio, D. A. and Scow, K. M. 1998. Impacts of carbon and flooding on soil microbial communities: phospholipid fatty acid profiles and substrate utilization patterns. Microbial Ecology. 35: 265-278. 75 Bossio, D. A., Scow, K. M., Gunapala, N., and Graham, K.J. 1998. Determinants of soil microbial communities: Effects of agricultural management, season, and soil type on phospholipid fatty acid profiles. Microbial Ecol. 36: 1-12. Chen, W., Hoitink, H. A. J ., and Madden, L. V. 1988. Microbial activity and biomass in container media for predicting suppressiveness to damping-off caused by Pythium ultimum. Phytopathology. 78: 1447- 1450. Clegg, C. D., Lovell, R. D. L., and Hobbs, P. J. 2003. The impact of grassland management regime on the community structure of selected bacterial groups in soils. FEMS Microbiol. Ecol. 43:263-270. Cook, R. J., Schillinger, W. F ., and Christensen, N. W. 2002. Rhizoctonia root rot and take-all of wheat in diverse direct-seed spring cropping systems. Can. J. Plant Pathol. 24: 349-358. Couch, H. B. 1995. Diseases of Turfgrass, 3rd ed. Krieger Publishing, Malabar, FL. pp. 376. Dazzo, F. Personal Communication. Michigan State University. March 1, 2010. Donnison, L. M., Griffith, G. S., Hedger, J., Hobbs, P. J., and Bardgett, R. D. 2000. Management influences on soil microbial communities and their function in botanically diverse haymedows of northern England and Wales. Soil Biol. Biochem. 32: 253-263. Doran, J. W., Sarrantonio, M., and Liebig, M. A. 1996. Soil health and sustainability. Adv. Agron. 56: 2-54. Fraser, D. G., Doran, J. W., Sahs, W. W., and Lesoing, G.W. 1988, Soil microbial populations and activities under conventional and organic management. J. Environ. Qual. 17: 585-590. Frostegard, A. and Baath E. 1996. The use of phospholipid fatty acid analysis to estimate bacterial and fimgal biomass in soil. Biol. Fert. Soils. 22: 59-65. Grayston, S. J., Griffith, G. S., Mawdley, J. L., Campbell, C. D., and Bardgett, R. D. 2001. Accounting for variability in soil microbial communities of temperate upland grassland ecosystems. Soil Biol. Biochem. 33: 533-551. Guo, H. C. and Wang, G. H. 2009. Phosphorus status and microbial community of paddy soil with the grth of annual ryegrass (Lolium multiflorum Lam.) under different phosphorus fertilizer treatments. J. of Zhejiang University-Sci. 10: 761-768. Hardwood, J. L. and Russell, N. J. 1984. Lipids in Plants and Microbes. George Allen & Unwin, London, UK. pp. 192. 76 Hodges, C. F. 1990. The microbiology of non-pathogens and minor root pathogens in high sand content greens. Golf Course Manage. 58: 60-75. Hodges, C. F ., Campbell, D. A., and Christians, N. 1994. Potential biocontrol of Sclerotinia homoeocarpa and Bipolaris sorokiniana on the phylloplane of Poa pratensis with strains of Pseudomonas spp. Plant Path. 43: 500-506. Ka, J. 0., Burauel, P., Bronson, J. A., Holben, W. E., and Tiedje, J. M. 1995. DNA probe analysis of microbial community selected in field by long-term 2,4-D applications. Soil Sci. Soc. Am. J. 59: 1581-1587. Killharn, K. 1994. Soil ecology. Cambridge Univ. Press, Cambridge, England. Kirchner, M. J., Wollum II, A. F ., and King, L. D. 1993. Soil microbial populations and activities in reduced chemical input agro-ecosystems. Soil Sci. Soc. Amer. J. 57: 1289- 1295. Kobayashi, D. Y. and El-Barrad, N. E. H. 1996. Selection of bacterial antagonists using enrichment cultures for the control of summer patch disease in Kentucky bluegrass. Microbiology. 32: 106-1 10. Kobayashi, D. Y., Gugliehnoni, M., and Clarke, B. B. 1995. Isolation of the chitinolytic bacteria Xanthomonas maltophilia and Serratia marcescens as biological control agents for summer patch disease of turfgrass. Soil Biol. Biochem. 27: 1479-1487. Kowalchuk, G. A., Stephen, J. R., De Boer, J. I., Prosser, J. 1., Embley, M. T., and Woldendorp, J. W. 1997. Analysis of proteobacteria ammoniaoxidising bacteria in coastal dunes using denaturing gradient gel electrophoresis and sequencing of PCR amplified 16S rDNA fiagments. App. Environ. Microbio. 63: 1489-1497. Liang, C., Fujinuma, R., and Balser, T. C. 2008. Comparing PLFA and amino sugars for microbial analysis in an Upper Michigan old grth forest. Soil Bio. and Biochem. 40: 2063-2065. Liu, L. X., Hsiang, T., Carey, K., and Eggens, J. L. 1995. Microbial populations and suppression of dollar spot disease in creeping bentgrass with inorganic and organic amendments. Plant Dis. 79: 144-147. Lundquist, E. J., Scow, K. M., Jackson, L. E., Uesugi, S. L., and Johnson, C. R. 1999. Rapid response of soil microbial communities from conventional, low input, and organic farming systems to a wet/dry cycle. Soil Biol. Biochem. 31: 1661-1675. Meyers, R. T., Zak, D. R., White, D. C., and Peacock, A. 2001. Landscape-level patterns of microbial community composition and substrate use in upland forest ecosystems. Soil Sci. Soc. of Amer. J. 65: 359-367. 77 Muyzer, G., De Waal, E. C., and Uitterlinden, A. G., 1993. Profiling of complex microbial populations by denaturing gradient gel electrophoresis analysis of polymerase chain reaction-amplified genes coding for 168 rRNA. Appl. and Envir. Microbio. 59: 695-700. Nelson, E. B. 1992. Biological control of turfgrass diseases. Information Bulletin 220. Cornell Cooperative Extension, Cornell University, Ithaca, NY. Nelson, E. B. Microbial mechanisms of biological disease control. 1998. In: Sticklen, M. B. and Kenna, M. P. Turfgrass Biotechnology: Cell and Molecular Genetic Approaches to Turfgrass Improvement. Chelsea, MI. Ann Arbor Press. pp. 55-92. Nelson, EB. and Craft, C. M. 1991. Introduction and establishment of strains of Enterobacter cloacae in golf course turf for the biological control of dollar spot. Plant Dis. 75: 510-514. Nelson, E. B. and Crafi, C. M. 1992. A miniaturized and rapid bioassay for the selection of soil bacteria suppressive to Pythium blight of turfgrasses. Phytopath. 82: 206—210. Nitta, T. 1991. Diversity of root fungal floras: its implications for soil-bome diseases and crop growth. Jpn. Agric. Res. 25: 6-11. Pankhurst, C.E., McDonald, H.J.B., Hawke, G., and Kirkby, CA. 2002. Effect of tillage and stubble management on chemical and microbiological properties and the development of suppression towards cereal root disease in soils from two sites in NSW Australia. Soil Biol. Biochem. 34: 833-840. Pankhurst, C. E., Pierret, A., Hawke, B. G., and Kirby, J. M. 2002. Microbiological and chemical properties of soil associated with macropores at different depths in a red-duplex soil in NSW Australia. Plant Soil. 238: 11-20. Powell, J. F., Vargas, J. M., and Nair, M. G. 2000. Management of dollar spot on creeping bentgrass with metabolites of Pseudomonas aureofaciens (TX-l). Plant Dis. 84: 19-24. Powlson, D.S., Brookes, RC, and Christensen, B.T. 1987. Measurement of soil microbial biomass provides an early indication of changes in total soil organic matter due to straw incorporation. Soil Biol. Biochem. 19: 159-164. Raaijmakers, J. M., Weller, D. M., and Thomashow, L. S. 1997. Frequency of antibiotic producing Pseudomonas spp. in natural environments. Appl. Environ. Microbiol. 63: 881-87 78 Raaijmakers, J. M. and Weller, D.M. 1998. Natural plant protection by 2,4- diacetylphloroglucinol-producing Pseudomonas spp. in take-all decline soils. Mol. Plant Microbe Interact. 11: 144-52 Ratledge, C. and Wilkinson, S. G. 1988. Microbial Lipids, Academic Press, London, England. Reuter, H. M., Schumann, G. L., Matheny, M. L., and Hatch, R. T. 1991. Suppression of dollar (Sclerotinia homoeocarpa) and brown patch (Rhizoctonia solam') on creeping bentgrass by an isolate of Streptomyces. Phytopathology. 81: 124. Rodriguez, R. and Pfender, W. F. 1997. Antibiosis and antagonism of Sclerotinia homeocarpa and Drechslera poae by Pseudomonas fluorescens Pf-S in vitro and in planta. Phytopathology 87: 614-621. Rovira, A. D. and Wildermuth. G. B. 1981. The nature and mechanisms of suppression. In Asher, M. J. C. and Shipton, P. ed., Biology and control of take-all. AcademicPress, Inc. Ltd., London, U.K. pp. 385-415. SAS Institute. 2009. SAS. STAT User’s Guide. Version 9.2. SAS Institute, Cary, North Carolina. Schneider, 0., Aubertot, J. N., Roger-Estrade, J., and Dore, T. 2003. Analysis and modeling of the amount of oilseed rape residues left at the soil surface after different soil tillage operations. 7‘h Int. Conf. on Plant Path. Tours France, 3-5 December 2003. Schumann, G. L. and Reuter, H. M. 1993. Suppression Of dollar spot with wheat bran topdressings. Biol. Cult. Tests Cont. Plant Dis. 7:113. Stark, J. M. and Firestone, M. K. 1995. Mechanisms for soil moisture effects on activity of nitrifying bacteria. App. Env. Microbio. 61 : 21 8-221 . Sturz, A.V., Carter, M.R., and Johnston, H.W. 1997. A review of plant disease pathogen interactions and microbial antagonism under conservation tillage in temperate humid agriculture. Soil Tillage Res. 41: 169-189. Sylvia, D. M., Fuhrrnan, J. J., Hartel, P. G., and Zuberer, D. A. 1997. Principles and Applications of Soil Microbiology. Prentice-Hall, Inc., Englewood Cliffs, NJ. pp. 640. Tilston, E., Pitt, L. D., and Groenhof, A. C. 2002. Composted recycled organic matter suppresses soil-bome diseases of field crops. New Phytol. 154: 731-740. Thompson, D. C., Clarke, B. B., and Kobayashi, D. Y. 1996. Evaluation of bacterial antagonists for reduction of summer patch symptoms in Kentucky bluegrass. Plant Dis. 80: 856-862. 79 Tunlid, A. and White, D. C., 1992. Biochemical analysis of biomass, community structure, nutritional status, and metabolic communities in soil. In Stozky, G., Bollag, J. M. ed., Soil Biochemistry. Vol. 7. Marcel Dekker, New York, NY. pp. 229-262. Turpeinen, R., Kairesalo, T., and Haggblom, M. M. 2004. Microbial community structure and activity in arsenic-, chromium-, and copper-contaminated soils. FEMS Microbiol. Ecol. 47(1):39—50. Uddin, W. and Viji, G. 2002. Biological control of turfgrass diseaes. In Gnanamanickam, S.S., Editor, 2002. Biological Control of Crop Diseases, Marcel Dekker, New York, NY. pp. 313—337. Vargas, J. M. Jr. 1999. Biological control: a work in progress. Golf Course Management 68: 55-58. Vestal, J. R. and White, D. C. 1989. Lipid analysis in microbial ecology: quantitative approches to the study of microbial community. Bioscience. 39: 535-541. Viji, G., Uddin, W., and Romaine, C. P. 2000. Evaluation of bacterial antagonists from spent mushroom substrate for suppression of turfgrass diseases. Phytopathology. 90: S81. Viji, G., Uddin, W., and Romaine, C. P. 2000. Efficacy and timing of application of Pseudomonas aeruginosa for control of ryegrass blast. Phytopathology. 90: S80. Webster, G., Embley, T. M., and Prosser, J. I. 2002. Grassland management regimens reduced small-scale heterogeneity and species diversity of B-proteobacterial ammonia oxidizer populations. Appl. Environ. Microbiol. 68: 20-30. White, D. C. 1988. Validation of quantative analysis of microbial biomass, community structure, and metabolic activity. Arch. Hydrobiol. Beih. Ergeben Limnol. 31: 1—18. Wong, P. T. W. and Baker, R. 1981. Control of wheat take-all and Ophiobolus patch of Agrostis turfgrass by fluorescent pseudomonads from a F usarium-suppressive soil. Phytopath. 71: 1008. Wong, P. T. W. and Baker, R. 1984. Suppression of wheat take-all and Ophiobolus patch by fluorescent pseudomonads from a Fusarium-suppressive soil. Soil Biol. Biochem. 16: 397-403. Wong, P. T. W. and Baker, R. 1985. Control of wheat take-all and Ophiobolus patch of turfgrass by fluorescent pseudomanads. In Parker, C. A., Rovira, A. D., Moore, K. J ., Wong, P. T. W. and Kollmorgen, J. F. eds. Ecology and Management of Soilborne Plant Pathogens. pp. 151-153. 80 Workneh, F. and van Bruggen, A. H. C. 1994. Microbial density, composition, and diversity in organically and conventionally managed rhizosphere soil in relation to suppression of corky root of tomatoes. Appl. Soil Ecol. 1: 219-30. Yao, H., He, 2., Wilson, M. J., and Campbell, C. D. 2000. Microbial biomass and community structure in a sequence of soils with increasing fertility and changing land use. Microb. Ecol. 40: 223-237. Zelles, L. 1999. Fatty acid patterns of phospholipids and lipoplysaccharides in the characterization of microbial communities in soil: a review. Bio. and Fert. of Soils. 29: 111-129. Zelles, L., Rackwitz, R., Bai, Q.Y., Beck, T., and Beese, F., 1995. Discrimination of . microbial diversity by fatty acid profiles of phospholipids and lipopolysaccharides in differently cultivated soils. Plant and soil. 170: 115-122. Zhang, Z. and Yuen, G. Y. 1999. Biological control of Bipolaris sorokiniana on tall fescue by Stenotrophomonas maltophilia strain C3. Phytopath. 89: 817-822. Zogg, G. P., Zak, D. R., Ringelberg, D. B., MacDonald, N. W., Pregitzer, K. S., and White, D. C. 1997. Compositional and functional shifts in microbial communities due to soil warming. Soil Sci. Soc. of Amer. J. 61: 475-481. 81 CHAPTER THREE DOLLAR SPOT REDUCTION THROUGH LIGHTWEIGHT ROLLING ON CREEPING BENTGRASS PUTTING GREENS ABSTRACT Hypovirulent and virulent isolates of Rutstroemia floccosum were compared for mycelial growth, oxalic acid and cell wall degrading enzyme production. Virulence of isolates VCG-B, VCG-C, and CS was confirmed while a known hypovirulent strain, Sh12B, was unable to cause disease symptoms. Isolates were grown for 14 days on potato dextrose agar (PDA) amended with the acid indicator dye bromophenol blue. All isolates were capable of changing the media from purple to yellow indicating acid production, with VCG-B showing statistically less of a color change than the three other isolates. Oxalic acid (0A) concentrations were measured via HPLC in culture filtrates grown in potato dextrose broth (PDB) for 15 days. Virulent isolates produced significantly higher amounts of OA than isolate ShIZB. The cell wall degrading enzymes cellulase and pectin methylesterase were assayed in culture filtrates grown on isolated creeping bentgrass cell walls. All isolates exhibited pectin methylesterase activity, with isolate ShIZB displaying the highest activity relative to mycelia dry weight. Cellulase activity was highest in isolate Sh12B as well relative to mycelia dry weight. Mycelia] grth of the hypovirulent isolate was less than the virulent isolates in all experiments; however enzyme activity and oxalic acid production remained steady. These results suggest that oxalic acid and extracellular enzymes are produced by hypovirulent strains of R. floccosum, indicating that virulence may not be solely determined by the production of these particular compounds. 82 INTRODCTION Association between oxalic acid (0A) production and pathogenicity has been reported for a number of phytopathogenic fungi (Bateman and Beer, 1965; Kritzman et al., 1977; Magro et al., 1984; Marciano et al., 1983; Godoy et al., 1990). Extensive research has documented the importance of OA production, particularly by Sclerotinia spp. as a factor in pathogenesis. However, the role of additional factors such as extracellular enzymes remains debated (Callahan and Rowe, 1991). Mutants of Sclerotinia sclerotiorum that are deficient in the ability to synthesize 0A are non- pathogenic, whereas revertant strains that regain their OA biosynthetic capacity exhibit normal virulence (Godoy et al., 1990.) Mechanisms by which 0A secretion might enhance virulence of certain pathogens have long been investigated. Oxalic acid appears to be involved in pathogenesis by lowering pH in advance of infected tissues to enhance the activity of extracellular enzymes produced by the pathogen (Hodgkinson 1977; Dutton and Evans 1996). Other theories center around OA being directly toxic to the plant, thus weakening the plant, facilitating further invasion (Noyes and Hancock, 1981). This occurs by the chelation of cell wall Ca2+ by the oxalate anion which is thought to compromise the function of Ca2+- dependent defense responses (Bateman and Beer, 1965). Variation in production of 0A in liquid medium by hypovirulent and virulent strains of Cryphonectria parasitica (Murr.) Barr has been reported (McCarroll, 1978). Virulent strains produced twice as much calcium oxalate as hypovirulent strains. Hypovirulence is a phenotype associated with reduced virulence in fungal plant pathogens, and has been associated with several mechanisms of action (Bharathan and 83 Tavantzis, 1990; Boland, 1992; Elliston, 1982; Zhou and Boland, 1997). The dollar spot pathogen Rutstroemia floccosum (Syn. Sclerotinia homoeocarpa) produces oxalic acid in vitro, though to a much lesser extent than other Sclerotinia members such as S. sclerotiorum and S. trifoliorum (Beaulieu, 2008). Very little is known regarding the pathogenesis of R. floccosum, and whether or not 0A is an important factor related to turfgrass infection. Understanding the mechanisms of action by which hypovirulence influences the fungal pathogen with regard to 0A and extracellular enzyme production, may give valuable insight into the role that oxalic acid plays in the infection process. Elucidating the role that OA plays in R. floccosum pathogenicity, allows for ‘ investigations into novel methods of dollar spot control related to plant defense mechanisms. For instance, the expression of oxalate oxidase in transgenic plants has been shown to be successful in conveying resistance to a diverse array of pathogens particularly those that produce OA (Thompson et al. 1995; Liang et al., 2001; Hu et al., 2003). Since many pathogens produce CA as a toxin (as mentioned above), and it is thought to play a major role in pathogenesis, the overall breakdown of CA by oxalate oxidase may be detrimental to the pathogen, as well as beneficial to the plant with the generation of H202 possibly playing a role in defense signaling. Within the context of this thesis, different mechanisms related to dollar spot reduction as a result of lightweight rolling were investigated. One hypothesis not discussed in these chapters centered on the idea that rolling simulates a wounding effect in turfgrass plants, which subsequently induces an enzymatic response. Two enzymes of interest (oxalate oxidase, and peroxidase) were assayed for in grass tissues from various rolling treatments (data not shown). Unsuccessful attempts investigating plant oxalate 84 oxidase activity led to an assessment of pathgoenicity factors in the dollar spot pathogen R. floccosum. The purpose of this study, therefore, was to examine in vitro oxalate production by virulent and hypovirulent strains of the fimgus and to clarify the effect of hypovirulent agents on DA and enzyme production. The goal of this research was to examine whether oxalic acid and extracellular secreted cell wall degrading enzymes are important in the infection of creeping bentgrass (A grostis palustris) by R. floccosum. MATERIALS AND METHODS Fungal Isolates. Isolates of R. floccosum used in this study are summarized in Table 3.01. All isolates were grown on potato dextrose agar (PDA) (Becton Dickinson, Sparks, MD), and subcultured onto fresh PDA plates using colonized agar plugs (5 mm diam.) 7-10 days prior to each test. The hypovirulent strain ShIZB, was obtained from Dr. Greg Boland at the University of Guelph, ON, Canada. The presence of dsRNA and hypovirulent characteristics in strain ShIZB is discussed in Zhou and Boland (1996). Table 3.01. Rutstroemiafloccosum isolates studied and their source of obtainment. Isolate Source Vegetative compatibility strain B (VCG-B) Dr. Joseph M. Vargas Jr., Michigan State University. East Lansing, MI Vegetative compatibility strain C (V CG-C) Dr. Joseph M. Vargas Jr., Michigan State University. East Lansing, MI Common Strain (CS) Dr. Joseph M. Vargas Jr., Michigan State University. East Lansing, MI Hypovirulent Sclerotinia homoeocarpa Dr. Greg J. Boland, University of Guelph. (Sh12B) Guelph, Ontario 85 Isolate Virulence Testing Virulence for each isolate was tested under greenhouse conditions by inoculating 9 cm diameter cups of 10 wk old creeping bentgrass (Agrostis palustris cv. Crenshaw) with 6 mm colonized agar plugs of actively growing R. floccosum isolates. Each virulence trial was replicated four times for each of the four isolates tested. Inoculated plants were kept in enclosed chambers surrounded by plastic in a greenhouse maintained at approximately 26-28 C with an approximate 13 hr photoperiod in order to encourage humidity; cups were arranged in a completely randomized design. Disease ratings were taken by measuring the advancing fungal mycelium and subsequent infected area on each plant on a daily basis. Advancing mycelium radiated from the center inoculation point, causing subsequent infection of turfgrass in a circular pattern. Infected turfgrass was rated by measuring the area of bleached, water soaked patches on the inoculated cups. Mean disease measurements were used for statistical comparisons among dollar spot isolates. Oxalic Acid production on Bromophenol Blue amended Potato Dextrose Agar. In order to compare different strains of R. fiocossum and their ability to produce acid when grown in vitro, a slightly revised method from Steadman et al. (1994) was used. Fungal isolates growing on potato dextrose (PDA) agar for 10 days were transferred to PDA amended with 0.05g/l bromophenol blue (BB) adjusted to pH 6.0 with 1M NaOH. Bromophenol blue is an indicator of acid production due to its rapid and visible response indicated by a characteristic yellow color change when pH is below 3. Approximately 20 ml of BB amended PDA (BBPDA) was aseptically pipetted into 100 x 15 mm polystyrene Petri-plates. One plug (5mm) of agar containing each isolate was 86 inoculated onto the center of a BBPDA plate and placed under fluorescent light (photoperiod of 12 h) at room temperature (20-22 C), and observed for oxalic acid production for 14 days. Each strain grown on BBPDA was replicated 4 times and the experiment was repeated twice. Measurement of fungal radial growth was taken daily, along with color change measurements taken visually against a 5000K fluorescent light Porta-Trace (Gagne Inc.). Color change ratings were given on a scale of 0-5 with 0 indicating no color change and 5 indicating a complete yellowing of the plate. Mean color change and fungal radial growth measurements were used for statistical comparisons. Growth and acid production in potato dextrose broth medium Mycelial growth and production of oxalic acid were evaluated by culturing individual isolates in potato-dextrose broth (PDB). One 5 mm diameter agar plug from actively growing margins of 7-day-old colonies was transferred to 50 ml liquid medium in 125-ml Erlenmeyer flasks. Inoculated flasks were incubated on a rotary shaker at 100 rpm at 20-22 C. Mycelial dry weights were determined by vacuum filtration of culture contents through a layer of pre-weighed Whatrnan No.1 filter paper in order to separate mycelium from culture filtrates. Dry weight of mycelium was determined after filter paper with mycelium was dried in a 80 C oven for 72 h. Approximately 3 m1 of culture filtrate from each treatment was collected and measured for pH on five different sampling days, collected samples were maintained at - 20 C until analysis. Concentrations of oxalic acid in PDB filtrates were determined after 15 days of grth in PDB. Extracts were analyzed on a Waters XBridge C18 3.5 um 87 column (Waters Corp., Milford, MA) high pressure liquid chromatograph (HPLC)/ mass selective detector (MSD) for oxalic acid. The column diameter and length were 3/0x50 mm, and the system was run at ambient temperature. The mobile phase was run at 0.25 ml/min which started at 80 % 0.1 % formic acid (A) and 20 % 0.1 % formic acid in acetonitrile (B) and was held there for one minute. A gradient began moving to 10 % A at four minutes. This was held for 5 minutes and then back to original flow. The MSD was monitoring ions 46 and 45 daltons in electrospray positive mode. 10p] from each sample was used in analysis, and values are reported as pg oxalic acid ml sample". Growth and production of extracellular enzymes on isolated creeping bentgrass cell walls.~ In order to investigate the fungal secretion of cell wall degrading enzymes by R. floccosum, cell walls from creeping bentgrass (Agrostis palustris cv. ‘Crenshaw’) were isolated and used as the sole carbon source in a minimal growing medium for R. floccosum isolate inoculation. Cell walls were prepared according to English et al., (1971) by grinding the frozen tissue to a fine powder in liquid nitrogen with the aid of a mortar and pestle. This frozen powder was then ground in approximately 2.5 volumes (v/w) of cold 100 mM postassiurn phosphate buffer, pH 7.0. The insoluble material was collected on Whatman No. 1 filter paper by suction filtration. The residue was resuspended in 1 volume of cold buffer, before re-filtration. This washing procedure was repeated four times, using 1 volume of buffer each time. Buffer washes were followed by one wash with 1 volume of cold distilled water to remove salts. The residue was then suspended in 2.5 volumes of a cold mixture of chloroform and methanol (1:1 v/v) and ground with a mortar and pestle. 88 The insoluble material was collected on Whatman No. 1 filter paper, washed three times with 1 volume of the chloroform-methanol mixture at room temperature, then washed three times with 1 volume of acetone at room temperature. The residue remaining after the acetone extraction, which constitutes the cell walls used in this study, was air dried and stored at room temperature. Immediately preceding use, the cell walls were placed in a dessicator to remove residual water. The medium for liquid culture was prepared according to a modified protocol from Anderson (1978). In one liter the medium contained 0.25g K2HPO4, 0.25 g NH4NO3, 0.1 g MgSO4'7H20, 0.15 ml 2% FeCl. In 125 ml Erlenmeyer flasks 50 ml of medium was added along with 0.5 g of cell wall material, flasks were then autoclaved at 121 C for 20 min. Cell wall amended media flasks were inoculated with one colonized agar plug (5 mm diam.) fi'om 7-10 day old isolates growing on PDA. Inoculated flasks were incubated on a rotary shaker at 100 rpm at 20-22 C for 14 days. Filtrates were obtained from the liquid cultures by passage of the suspensions through Whatman No. 1 filter paper via suction filtration. Mycelia was weighed after drying at 80 C for 72 h. Isolates were replicated four times, and liquid filtrates were stored at -20 C until enzyme assays were performed. Enzyme assays were performed on culture filtrates for cellulase and pectin methylesterase activities following modified procedures described in Downie et al. (1998) Hagerrnan et al. (1985) and Taylor and Secor (1988). Pectin-containing medium was prepared by mixing 0.1% (w/v) pectin, and 1% (w/v) Type II agarose (Sigma Chemical Co., St. Louis, MO) in 0.2 M phosphate buffer adjusted to pH 5.3. The mixture 89 was dissolved by heating to a boil while stirring. 18 m1 of the hot mixture was aseptically dispensed into 100 X ISO-mm sterile petri dishes using a sterile 25 ml pipette. Carboxymethylcellulose (CMC) containing medium was prepared using the same procedure but substituting 0.1% CMC for the pectin. Once agar plates cooled, a cork borer was used to punch 3 holes, 5 mm in diameter and approximately 1.7 cm apart, in the solidified medium. The holes were arranged in three rows, in a 3 x 3 x 3 pattern and the wells were filled with (35 ul) with standard, control, or filtrate solutions. The assay was incubated at 28 C for 17 hours. Pectin containing gels were developed after incubation by flooding the assay plate with 10 ml of 0.05% (w/v) ruthenium red (Sigma Chemical Co.) in water for 30 min at 25 C. Excess dye was removed by washing the plate several dimes with deionized water. The diameters of the resulting rings or halos of activity were measured against a 5000K fluorescent light Porta-Trace (Gagne Inc., Johnson City, NY). The cellulose containing plates were developed in a similar fashion using 1% (w/v) Congo red (Sigma Chemical Co.) in water to stain the plates. The stain was removed after 15 min and plates were de-stained for 15 min with 1 M NaCl in 0.2 M phosphate buffer. Enzyme activity was assessed by measuring ring diameter similar to the pectin containing plates, and values were standardized by dividing halo diameter by mycelia dry weight from each isolate. Assays for each isolate were replicated three times, and experiments were repeated twice. Mean ring diameters were used for statistical comparisons among isolates using the PROC GLM procedure in SAS. RESULTS AND DISCUSSION 90 Virulence. Isolates VCG-B, VCG-C, and CS initiated dollar spot symptoms on cups of creeping bentrgrass ranging from 7.10 to 8.00 cm in diameter after 16 days of incubation (Table 3.02). Some differences existed among these isolates with VCG-C displaying significantly higher virulence ratings than the other isolates at 6, 8, 11, 13, and 16 days after inoculation (DAI). Isolate Sh12B failed to cause disease symptoms, and resulted in no significant difference from the control at any DAI. These results confirmed the If hypovirulence phenotype of the strain Sh12B. Mean virulence measurements resulted in no statistical differences among isolates VCG-B, VCG-C, and CS. However, all three of the aforementioned isolates resulted in statistically higher virulence ratings (P < 0.05) p than the control and isolate Sh12B; neither isolate resulted in any disease observation. Table 3.02. Disease ratings of R. floccosum isolates on swards of creeping bentgrass (Agrostis palustris cv. Crenshaw)*. Days After Inoculation Disease Rating (cm) IsoLrte 5T 6 8 11 13 16 Mean VCG-B 1.07 a 1.17 b 2.90 b 4.03 b 5.02 b - 7.10 b 3.55 a VCG-C 1.27 a 1.43 a 4.70 a 5.37 a 6.37 a 8.00 a 4.52 a CS 1.07 a 1.17 b 3.40 b 4.17 b 4.70 b 7.43 b 3.67 a ShIZB 0.00 b 0.00 c 0.00 c 0.00 c 0.00 c 0.00 c 0.00 b Control 0.00 b 0.00 c 0.00 c 0.00 c 0.00 c 0.00 c 0.00 b * Means are average disease ratings from four replications. ‘1' Means followed by different letters are significantly different according to Fisher’s LSD (P < 0.05). Oxalic acid production on bromophenol blue amended PDA. Acid production in vitro was measured by observing the characteristic color change of PDA amended with the acid indicator dye bromophenol blue. All isolates 91 displayed some degree of acid production, with evident color changing of the media from purple to yellow. Figure 3.01 reports mean color change ratings for the four dollar spot isolates tested after 15 days of growth on BBPDA. Isolate VCG-B resulted in significantly lower color change ratings after 15 days of growth compared to all other isolates tested (P < 0.05). t Figure 3.01. Mean media color change ratingsf from acid production of four R. floccosum isolates grown on bromophenol blue-amended potato dextrose agar for 15 days. 5 G? é4 '00 ~83 a 8. g2 .12: O .531 O U 0 VCG-B VCG-C Isolate "‘ Values are means of 8 replications fi'om two separate assays. 1' Color change ratings were given on a scale from 0-5 with 0 indicating no color change, and 5 indicating a uniform change of color in the media fi'om purple to yellow. 1 Means followed by different letters indicate a significant difference according to Fisher’s LSD (P < 0.05). Error bars represent the standard error (SE) of the mean 92 Growth and acid production in PDB Oxalic acid production was measured in vitro by growing dollar spot isolates in potato dextrose broth. Culture pH was measured throughout the duration of the experiment every three days. Table 3.03 gives mean culture pH measurements. All isolates displayed similar culture pH after 3 days of growth, however after 6 days, VCG- B had significantly lower pH than both the control and isolate Sh12B. After 15 days of growth, both VCG-B and VCG-C cultures were significantly lower in pH than the control; while VCG-C was significantly lower than isolate Sh12B as well (P < 0.05). * Table 3.03. Mean culture pH measurements of four R. floccosum isolates growing in potato dextrose broth for 15 days. Days After Inoculation Culture pH Isolate 3 6 T 9 12 15 VCG-B 6.22a 6.02 c 5.64 c 5.29 b 5.07 bc VCG-C 6.21 a 6.03 bc 5.67bc 5.17 b 4.92 c CS 6.22 a 6.09 abc 5.81 be 5.69 ab 5.76 abc Sh12B 6.17a 6.10 ab 5.93 ab 6.11 a 5.97 ab Control 6.20a 6.15 a 6.16 a 6.16 a 6.20 a "' Each pH measurement represents the mean of four replications. T Means followed by different letters indicate a significant difference in pH according to Fisher’s LSD (P < 0.05). Oxalic acid levels were measured in culture filtrates after 15 days of growth. All isolates displayed some level of OA production in liquid broth media. VCG-B displayed the highest levels followed by the other virulent isolates VCG-C and CS (Figure 3.02). The hypovirulent isolate Sh12B displayed variable levels of oxalic acid production 93 among replications, yet was able to produce CA at levels significantly lower than the other isolates (Figure 3.02). When OA production was standardized to the dry weight of fungal mycelia in each flask, hypovirulent isolate ShIZB produced significantly higher levels of OA than VCG-B (figure 3.03). While not statistically significant, ShIZB had higher OA production relative to mycelial grth than isolates VCG-C and CS as well (figure 3.03). Figure 3.02. Oxalic acid production* of R. floccosum isolates grown in potato dextrose broth for 15 days. 50.0 — t 45.0 ~‘ 41:9 a 400 ‘1 :::: 22:: 33:-5;; 28.8 b ... 35-0 “‘ :3:3:3:3: E 30.02, 222222222 2’6 b° DVCG-B :3 25.0 2‘ 335553535 IVCG-C 3 3.3.3.5.} 145° ICS g 15.0 1 §:§:§:§:§ lControl g“ 10.0 — 5:323:33 5.0 ~ zzszszszs 0" d 0.0 J ..... Isolate * Means represent the average of 4 replications '1' Means followed by different letters are significantly different according to Fisher’s LSD (P < 0.05). Error bars represent the standard error (SE) of the mean. 94 Figure 3.03. Oxalic acid production related to mycelial dry weight* of R. floccosum isolates grown in potato dextrose broth for 15 days. mg oxalic acid mycelia dry wt. 450 40* 35 " 30 ‘r 25 ___..-.. 20 ‘m 15‘ 10 ‘“' El VCG-B ‘22? VCG-C -— -~— cs i 2,-_ 8 ShIZB I Control 17,46 ab 15.60 ab ooooo Isolate * Means represent the average of 4 replications 1' Means followed by different letters are significantly different according to Fisher’s LSD (P < 0.05). Error bars represent the standard error (SE) of the mean. Growth and production of extracellular enzymes on isolated creeping bentgrass cell walls. Isolates were assayed for their ability to produce pectolytic and cellulolytic enzymes in vitro and were compared. In vitro activities of pectin methylesterases (PME) and cellulase from filtrates of 14 day old developing hyphae growing on creeping bentgrass cell walls are reported in figures 3.04 and 3.05. Levels of PME and cellulase activity were adjusted with respect to dry weight of mycelia from cultures. 95 up Figure 3.04. Pectin methylesterase activity of four R. floccosum isolates grown in liquid culture with creeping bentgrass (Agrostis palustris) cell walls. 80‘2"" "“' " r ' r E’ 78 ‘* Z‘ '6 g 76 ‘*' ** ** ‘53 O: _A 74 7 v '2 e __- 7' a 72 5 a Q 70 ‘" > ’5 g 68 ""“ E” m 66 r 64 _ . Isolate * Means are the average of 12 measurements from 4 replications. EJ VCG-B VCG-C CS 8 Sh12B 1' Means followed by different letters are significantly different according to Fisher’s LSD (P < 0.05). Error bars represent the standard error (SE) of the mean. When PME levels were adjusted with respect to mycelia dry weight, isolate Sh12B exhibited significantly greater activity than the other three isolates (P < 0.05). For cellulase activity, adjusted means showed Sh12B to have significantly higher enzyme activity than all other strains with respect to mycelium production (P < 0.05), while VCG-C had the lowest cellulase activity among all isolates. 96 :3: Figure 3.05. Cellulase activity of four R. floccosum isolates grown in liquid culture with creeping bentgrass (Agrostis palustris) cell walls. 140—2 3 r: 120» a a :5; 100— 3E .2 3 80“ ‘5»? IE; 60*" a $5 402- .2 g 20— 0 U 0.. 118.42 a ..... El VCG-B 323:3:313 VCG-C 5:13:52? , cs In ShIZB Isolate * Means are the average of 12 measurements from 4 replications. ‘1' Means followed by different letters are significantly different according to Fisher’s LSD (P < 0.05). Error bars represent the standard error (SE) of the mean. CONCLUSIONS In order to investigate whether oxalic acid plays a crucial role in pathogen virulence, four isolates of R. floccosum were subjected to multiple assays related to OA production and extracellular enzyme activity. The isolate Sh12B, a known hypovirulent strain of R. floccosum, was evaluated and compared to three other known virulent isolates. 97 Virulence testing confirmed prior findings by Zhou and Boland (1997) that isolate Sh12B is indeed unable to maintain virulence on swards of creeping bentgrass. All other isolates displayed typical dollar spot disease symptoms on inoculated creeping bentgrass with bleached, circular spots, growing outward from the inoculation point, containing evident fluffy mycelium near the infected margins. Other studies involving hypovirulence and the production of 0A have shown hypovirulent fimgal strains exhibiting reduced or delayed accumulation of 0A. They hypovirulence phenotype in isolate ShIZB of R. floccosum was previously associated with a reduced growth rate, atypical colony morphology, and reduced virulence, as well as the presence of double stranded RNA. If pathogenesis is associated with the production of OA and/or extracellular enzymes as proposed by many authors in the past (Hancock, 1967; Lumsden, 1969, 1976, 1979; Hancock, 1967; Marciano et al., 1983; Noyes and Hancock, 1981), one could hypothesize that the hypovirulent isolate ShIZB should be deficient in both acid and cell wall degrading enzyme production. Contrary to the initial hypothesis, isolate ShIZB displayed marked acid production when grown on BBPDA plates, displaying a high degree of color change in the indicator medium. Cultures of ShIZB tended to grow at about half the rate of the other virulent isolates, however, after 15 days of growth, the hypovirulent isolate Sh12B exhibited a higher degree of color change compared to VCG-B, and similar color change ratings to isolates VCG-C and CS. This initial examination of acid production in vitro indicated a possible delayed, yet sufficient production of acid by the hypovirulent strain when compared to the other R. floccosum isolates. The virulent isolate VCG-B displayed 98 significantly less color change ratings in the medium, indicating that overall acid production in vitro may not be directly associated with virulence on creeping bentgrass. In order to confirm the production of 0A, and make comparisons among fungal isolates culture filtrates 0A levels were measured via HPLC/MS. The results confirmed the increased production of CA by virulent isolates compared to the hypovirulent counterpart. These results coincide with Callahan and Rowe (1991) and Zhou and Boland (1999) in that reduced or delayed accumulation of oxalic acid in hypovirulent isolates appears to be associated with reduced disease severity. However, the production of OA way relative to mycelial production was significantly higher in the hypovirulent isolate, indicating marked acid production even when fungal growth is at a minium. With these results, oxalic acid does not appear to be a sole pathogenic determinant in R. floccosum. In vivo studies should be done as well as the investigation into oxalic acid deficient mutants in R. floccosum to truly identify the importance of 0A in pathogenesis. Isolate Sh12B produced degradative enzymes (PME and cellulase) at levels equal to, or higher than the virulent strains evaluated. These hydrolytic enzymes have been associated with infection of tissue by Sclerotinia spp. (Hancock, 1966; Lumsden, 1979). However, these results coincide with Godoy et al. (1990), who found that non-pathogenic mutants of Sclerotinia sclerotiorum produced equal or greater amounts of pectolytic and cellulolytic enzymes when compared to wild type isolates. This could be an indication that factors other than pectolytic or cellulolytic enzymes are factors in the virulence of R. floccosum. In vivo studies involving extracellular cell wall degrading enzymes should be conducted in order to draw further conclusions regarding their importance in R. floccosum virulence and pathogenesis. This initial study set out to investigate differences 99 among virulent and non-virulent (hypovirulent) strains, and whether oxalic acid and/or degradative hydrolytic enzymes were essential in host/pathogen interactions. Isolate Sh12B may retain mechanisms necessary for the production of these enzymes for saprophytic survival purposes. If the isolate is to survive in nature by the acquisition of nutrients via other organisms, these degradative enzymes most likely are needed in the breakdown of detritus. Oxalic acid is thought to play a major role in the pathogenesis of many plant pathogens, but the importance of 0A production by the dollar spot pathogen R. floccosum has yet to be defined. Investigations into the importance of 0A in pathogen virulence and infection could potentially lead to novel methods of disease control or reduction. The previously mentioned breakdown of OA via plant derived oxalate oxidase has been characterized as an option in transgenic plant production (Thompson et al. 1995; Liang et al., 2001; Hu et al., 2003). This same enzyme has been shown to be induced in certain grass species after wounding, and when subjected to heavy metal and temperature stress (Valentovicova et al., 2009; Le Deunff et al., 2004). Prior chapters in this thesis pertained to the reduction of dollar spot disease through lightweight rolling of creeping bentgrass. A proposed means in disease suppression could be related to oxalate oxidase activity increasing in response to the rolling treatment. Initial research investigating this hypothesis proved sporadic and inconclusive (Appendix). However, without a basic knowledge of the role that oxalic acid plays in R. floccosum pathogenesis, it is difficult to make presmnptions on whether 0A degradation by oxalate oxidase could be having part in dollar spot disease suppression. The hypovirulent strain ShIZB which was unable to produce disease 100 symptoms on swards of creeping bentgrass, still exhibited marked acid and cell wall degrading enzyme production in vitro. These findings suggest that while oxalic acid and extracellular enzyme production may be contributing factors in R. floccosum pathogenesis, alone or in tandem, their production does not necessarily translate to disease development. It is likely that oxalic acid plays a diminutive role in the pathogenicity and disease development of R. floccosum, however, in vivo studies on the production of 0A and potential cell wall degrading enzymes should be done. 101 LITERATURE CITED Anderson, A. J. 1978. Extracellular enzymes produced by Colletotrichum lindemuthianum and Helminthosporium maydis during grth on isolated bean and corn cell walls. Phytopathology. 68: 1585-1589. Bateman, D. F. and Beer, S. V. 1965. Simultaneous production and synergistic action of oxalic acid and polygalacturonase during pathogenesis by Sclerotium rolfsii. Phytopathology. 55: 204-21 1. ‘ Bharathan, N., and Tavantizis, S. M. 1990. Genetic diversity of double-stranded RNA from Rhizoctonia solani. Phytopathology. 80: 631—635. Boland, G. J. 1992. Hypovirulence and double-stranded RNA in Sclerotinia sclerotiorum. Can. J. Plant Pathol. 14: 10-17. Callahan, F. E., and Rowe, D. E. 1991. Use of a host-pathogeninteraction system to test wheter oxalic acid is the sole pathogenic determinant in the exudates of Sclerotinia trifoliorum. Phytopathology. 81 : 1546-1550. Downie, 3., Dirk, L. M. A., Hadfield, K. A., Wilkins, T. A., Bennett, A. B., and Bradford, K. J. 1998. A gel diffusion assay for quantification of pectin methylesterase activity. Analytical Biochem. 264: 149-157. Dutton, M. V., and Evans, C. S. 1996. Oxalate production by fungi: Its role in pathogenicity and ecology in the soil environment. Can. J. Microbiol. 42: 881—895. Elliston, J. E. 1982. Hypovirulence. Adv. Plant Pathol. 1: 1-33. English, P. D., Jurale, J. B., and Albersheim, P. 1971. Parameters affecting polysaccharide-degrading enzyme secretion by Colletotrichum lindemuthianum grown in culture. Plant Physiol. 47: 1-6. Godoy, G., Steadman, J. R., Dickman, M. B., and Dam, R. 1990. Use of mutants to demonstrate the role of oxalic acid in pathogenicity of Sclerotinia sclerotiorum on Phaseolus vulgaris. Physiol. and Molec. Plant Path. 37: 179-191. Hancock, J. G. 1966. Degradation of pectic substances associated with pathogenesis by Sclerotinia sclerotiorum in sunflower and tomato stems. Phytopathology. 56: 975-979. Hancock, J. G. 1967. Hemicellulose degradation in sunflower hypocotyls infected with Sclerotinia sclerotiorum. Phytopathology. 57: 203-206. 102 Hagerman, A. E., Blau, D. M., and McClure, A. L. 1985. Plate assay for determining the time of production of protease, cellulase, and pectinases by germinating fungal spores. Analytical Biochem. 151: 334-342. Hodgkinson, A. 1977. Oxalic acid in biology an dmedicine. Academic Press, New York. Hu, X., Bidney, D. L., Yalpani, N., Duvick, J. P., Crasta, 0., Folkerts, 0., and Lu, G. 2003. Overexpression of a gene encoding hydrogen peroxide-generating oxalate oxidase evokes defense responses in sunflower. Plant Physiol. 133: 170-181. Kritzman, G., Chet, 1., Henis, Y. 1977. The role of oxalic acid in the pathogenic behavior of Sclertium roljfsii Sacc. Exp. Mycol.. 1: 280-285. Kritzman, G., Chet, I. 1980. The role of Phenols in the pathogenicity of Botrytis cinerea. Phytoparasitica. 8:27-37. Le Deunff, E., Davoine, C., Le Dantec, C., Billard, J. P., and Huault, C. 2004. Oxadative burst and expression of germin/oxo genes during wounding of ryegrass leaf blades: comparison with senescence of leaf sheaths. The Plant Journal. 38: 421-431. Liang, H., Maynard, C. A., Allen, R. D., Powell, WA. 2001. Increased Septoria musiva resistance in transgenic hybrid poplar leaves expressing a wheat oxalate oxidase gene. Plant Mol. Biol. 45: 619-629 Lumsden R. D. 1969. Sclerotinia sclerotiorum infection of bean and the production of cellulase. Phytopathology. 59: 653-657 Lumsden, R. D. 1976. Pectolytic enzymes of Sclerotinia sclerotiorum and their localization in infected bean. Can. J. Bot. 54: 2630-2641. Lumsden, R. D. 1979. Histology and physiology of pathogenesis in plant diseases caused by Sclerotinia species. Phytopathology. 69: 890-896. Magro, P., Marciano, P., Di Lenna, P. 1984. Oxalic acid production and its role in pathogenesis of Sclerotinia sclerotiorum. FEMS Microbiology Letters. 24: 9-12. Marcino, P., DiLenna, P., Magro, P. 1983. Oxalic acid, cell wall-degrading enzymes and pH in pathogenesis and their significance in the virulence of two Sclerotinia sclerotiorum isolates on sunflower. Physiol. Plant Pathol. 22: 339-345. McCarroll, D. R. 1978. Pathogenesis of Endothia parasitica (Murr.) A. and A. Ph.D dissertation, University of Tennesee, Knoxville. Noyes, R. D., and Hancock, J. G. 1981. Role of oxalic acid in the Sclerotinia wilt of sunflower. Physiol. Plant Pathol. 18: 123-132. 103 Steadman, J. R., Marcinkowska, J ., and Rutledge, S. 1994. A semi-selective medium for isolation of Sclerotinia sclerotiorum. Can. J. of Plant Path. 16: 68-70. Thompson, C., Dunwell, J. M., Johnstone, C. E., Lay, V., Ray, J ., Schmitt, M., Watson, H., and Nisbet, G. 1995. Degradation of oxalic acid by transgenic oilseed rape plants expressing oxalate oxidase. Euphytica. 85: 169-172 Taylor, R. J. and Secor, G. A. 1988. An improved diffusion assay for quantifying the polygalacturonase content of Erwinia culture filtrates. Phytopathology. 78: 1101 -1 103. Tu, J. C. 1985. Tolerance of white bean (Phaseolus vulgaris) to white mold (Sclerotinia sclerotiorum) associated with tolerance to oxalic acid., Physiological Plant Pathology. 26: 111-117. Valentovicova, K., Haluskova, L., Huttova, J ., Mistrik, I., and Tamas, L. 2009. Effect of heavy metals and temperature on the oxalate oxidase activity and lignifications of metaxylem vessels in barley roots. Env. and Exp. Bot. 66: 457-462. Zhou, T. and Boland, G. J. 1997. Hypovirulence and double-strandedRNA in Sclerotinia homoeocarpa. Phytopathology. 87: 147-153. 104 APPENDIX from different eping bentgrass tissue A Figure A.1. Oxalate oxidase specific activity* in cre rolling treatments. we we 28% £28 case one? 0335 throughout the growing '1' Means followed by the same letter are statistically similar according to Fischer’s LSD. Bars ‘Means represent the average of 6 different assays performed on tissue represent the standard error of the mean. seasons of 2008 and 2009 105 from different eping bentgrass tissue \\\\\\\\\\\\\\\\ m\\\\\\\\ 1x a.m. 1x p. m. 2x a.m. Control m... m m. m m. m m 0 Wm. p.m.: 35381338 case 032:2 Rolling treatment sue throughout the growing * Means represent the average of 4 different assays performed on tis seasons of 2008 and 2009 ‘1' Means followed by the same letter are statistically similar according to Fischer’s LSD. Bars represent the standard error of the mean. 106