Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-22dnz Total loading time: 0 Render date: 2024-04-26T10:43:51.111Z Has data issue: false hasContentIssue false

16 - The Science Required to Deliver Epichloë Endophytes to Commerce

from Part V - Application and Commercialisation of Endophytes in Crop Production

Published online by Cambridge University Press:  01 April 2019

Trevor R. Hodkinson
Affiliation:
Trinity College Dublin
Fiona M. Doohan
Affiliation:
University College Dublin
Matthew J. Saunders
Affiliation:
Trinity College Dublin
Brian R. Murphy
Affiliation:
Trinity College Dublin
Get access

Summary

In some environments, the survival and production of ryegrass and fescue is heavily reliant on its mutualistic association with Epichloë endophytes. Epichloë endophytes produce a range of bioactive alkaloids, or secondary metabolites that can be effective in deterring insect pests, although some have also been shown to be toxic to grazing animals. These endophytes are being used in grassland farming systems in Australia, New Zealand, USA and some parts of South America. However, to achieve this outcome there has been considerable investment into developing a research pipeline for delivery of animal-safe endophyte strains that are still capable of deterring insect pests and providing protection against abiotic stresses. The pipeline starts with the discovery and isolation of endophytes from wild populations of ryegrass and fescue, characterisation of the known alkaloids they produce, use of genetic markers to determine the relationship between known well-characterised strains and new strains entering the collection, determination of their bioactivity against insect pests of economic significance, understanding issues of compatibility of a strain of interest with the elite germplasm into which it has been inoculated, determining ease of transmission to subsequent seed generations, and ensuring there will be no or minimal animal health and welfare issues associated with using the strain in grazing systems.

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2019

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

ACIL Allen Consulting (2017). New Zealand’s Science System: Case Studies, Report to the Ministry of Business, Innovation and Employment. Sydney, Australia: ACIL Allen Consulting.Google Scholar
Aiken, G. E. and Strickland, J. R. (2013). Forages and Pastures Symposium: Managing the tall fescue–fungal endophyte symbiosis for optimum forage-animal production. Journal of Animal Science, 91, 23692378.CrossRefGoogle ScholarPubMed
Aldrich, C. G., Rhodes, M. T., Miner, J. L., Kerley, M. S. and Paterson, J. A. (1993). The effects of endophyte-infected tall fescue consumption and use of a dopamine antagonist on intake, digestibility, body temperature, and blood constituents in sheep. Journal of Animal Science, 71, 158163.CrossRefGoogle ScholarPubMed
Arachevaleta, M., Bacon, C. W., Hoveland, C. S. and Radcliffe, D. E. (1989). Effect of the tall fescue endophyte on plant response to environmental stress. Agronomy Journal, 81, 8390.CrossRefGoogle Scholar
Bacetty, A. A., Snook, M. E., Glenn, A. E. et al. (2007). Nematotoxic effects of endophyte-infected tall fescue toxins and extracts in an in vitro bioassay using the nematode Pratylenchus scribneri. In Research and Practice New Zealand Grassland Association 13: 6th International Symposium on Fungal Endophytes of Grasses, ed. Popay, A. J. and Thom, E. R.. Palmerston, New Zealand: New Zealand Grassland Association, pp. 357361.Google Scholar
Bacon, C. W. (1993). Abiotic stress tolerances (moisture, nutrients) and photosynthesis in endophyte infected tall fescue. Agriculture Ecosystems and Environment, 44, 123141.CrossRefGoogle Scholar
Bacon, C. W. (1995). Toxic endophyte-infected tall fescue and range grasses: historic perspectives. Journal of Animal Science, 73, 861870.CrossRefGoogle ScholarPubMed
Bacon, C. W. and Siegel, M. R. (1988). Endophyte parasitism of tall fescue. Journal of Production Agriculture, 1, 4555.CrossRefGoogle Scholar
Bacon, C. W. and White, J. F. Jr. (1994). Stains, media, and procedures for analyzing endophytes. In Biotechnology of Endophytic Fungi of Grasses, ed. Bacon, C. W. and White, J. F.. Boca Raton, FL: CRC Press, pp. 4756.Google Scholar
Bacon, C. W., Porter, J. K., Robbins, J. D. and Luttrell, E. S. (1977). Epichloë typhina from toxic tall fescue grasses. Applied Environmental Microbiology, 34, 576581.CrossRefGoogle ScholarPubMed
Baldauf, M. W., Mace, W. J. and Richmond, D. S. (2011). Endophyte-mediated resistance to black cutworm as a function of plant cultivar and endophyte strain in tall fescue. Environmental Entomology, 40, 639647.CrossRefGoogle ScholarPubMed
Balfourier, F., Charmet, G. and Ravel, C. (1998). Genetic differentiation within and between natural populations of perennial and annual ryegrass (Lolium perenne and L. rigidum). Heredity, 81, 100110.Google Scholar
Ball, D. M., Schmidt, S. P., Lacefield, G. D., Hoveland, C. S. and Young, W. C. III. (2003). Tall fescue/endophyte concepts. Oregon Tall Fescue Commission Special Publication, 103.Google Scholar
Ball, D. M., Lacefield, G. D., Agee, C. S. and Hoveland, C. S. (2007). Introduction and acceptance of novel endophyte tall fescue in USA. In Research and Practice New Zealand Grassland Association 13: 6th International Symposium on Fungal Endophytes of Grasses, ed. Popay, A. J. and Thom, E. R.. Palmerston, New Zealand: New Zealand Grassland Association, pp. 249251.Google Scholar
Ball, O. J. P., Prestidge, R. A. and Sprosen, J. M. (1993). Effect of plant age and endophyte viability on peramine and lolitrem B concentration in perennial ryegrass seedlings. In Proceedings of the Second International Symposium on Acremonium/Grass Interactions, ed. Hume, D. E., Latch, G. C. M. and H. S. Easton. Palmerston, New Zealand: Grasslands Research Centre, pp. 6366.Google Scholar
Ball, O. J. P., Miles, C. O. and Prestidge, R. A. (1997). Ergopeptine alkaloids and Neotyphodium lolii-mediated resistance in perennial ryegrass against adult Heteronychus arator (Coleoptera: Scarabaeidae). Journal of Economic. Entomology, 90, 13821391.CrossRefGoogle Scholar
Barker, D. J., Hume, D. E. and Quigley, P. E. (1997). Negligible physiological responses to water deficit in endophyte-infected and uninfected perennial ryegrass. In Neotyphodium/Grass Interactions, ed. Bacon, C. W. and Hill, N. S.. New York: Plenus Press, pp. 137139.CrossRefGoogle Scholar
Barker, G. M., Patchett, B. J. and Cameron, N. E. (2015a). Epichloë uncinata infection and loline content afford Festulolium grasses protection from black beetle (Heteronychus arator). New Zealand Journal of Agricultural Research, 58, 3556.CrossRefGoogle Scholar
Barker, G. M., Patchett, B. J. and Cameron, N. E. (2015b). Epichloë uncinata infection and loline content protect Festulolium grasses from crickets (Orthoptera: Gryllidae). Journal of Economic Entomology, 108, 789–97.CrossRefGoogle ScholarPubMed
Beck, P. A., Gunter, S. A., Lusby, K. S. et al. (2008). Animal performance and economic comparison of novel and toxic endophyte tall fescues to cool-season annuals. Journal of Animal Science, 86, 20432055.CrossRefGoogle ScholarPubMed
Bell, N., Townsend, R., Popay, A., Mercer, C. and Jackson, T. (2011). Black beetle: lessons from the past and options for the future. New Zealand Grassland Association Pasture Persistence Symposium. Grassland Research and Practice Series, 15, 119124Google Scholar
Biermacher, J. and Beck, P. (2013). Study reveals legume-fescue mixture economics. The Samuel Roberts Noble Foundation Ag News and Views, 31, 4.Google Scholar
Bluett, S. J., Thom, E. R., Clark, D. A., Macdonald, K. A. and Minneé, E. M. K. (2003). Milksolids production from cows grazing perennial ryegrass containing AR1 or wild endophyte. Proceedings of the New Zealand Grassland Association, 65, 8390.CrossRefGoogle Scholar
Boling, J. A. (1985). Endophytic fungus and tall fescue utilization by ruminants. The Professional Animal Scientist, 1, 1922.CrossRefGoogle Scholar
Bond, J., Powell, J. B. and Weinland, B. T. (1984). Behaviour of steers grazing several varieties of tall fescue during summer conditions. Agronomy Journal, 76, 707709.CrossRefGoogle Scholar
Borrill, M. (1976). Temperate grasses. In Evolution of Crop Plants, ed. Simmonds, N.. London: Longman, pp. 137142.Google Scholar
Bouton, J. H. (2000). The use of endophytic fungi for pasture improvement in the USA. In Proceedings of the 4th International Neotyphodium/Grass Interactions Symposium. Soest, Germany, ed. Paul, V. H. and Dapprich, P. D.. Soest, Germany: Universitat Paderborn, Gesamthochschulepp, pp. 163168.Google Scholar
Bouton, J. H., Gates, R. N., Belesky, D. P. and Owsley, M. (1993). Yield and persistence of tall fescue in the southeastern coastal plain after removal of its endophyte. Agronomy Journal, 85, 5255.CrossRefGoogle Scholar
Bouton, J. H., Duncan, R. R., Gates, R. N., Hoveland, C. S. and Wood, D. T. (1997). Registration of ‘Jesup’ tall fescue. Crop Science, 37, 10111012.CrossRefGoogle Scholar
Bouton, J. H., Hill, N. S. and Hoveland, C. S. et al. (2000a). Performance of tall fescue cultivars infected with non-toxic endophytes. In Proceedings of the 4th International Neotyphodium/Grass Interactions Symposium. Soest, Germany, ed. Paul, V. H. and Dapprich, P. D.. Soest, Germany: Universitat Paderborn, Gesamthochschulepp, pp. 179185.Google Scholar
Bouton, J. H., Hill, N. S. and Hoveland, C. S. et al. (2000b). Infection of Tall Fescue Cultivars with Non-toxic Endophytes. In Proceedings 55th Southern Forage Crop Improvement Conference, Raleigh, NC, June 12–14, pp. 17.Google Scholar
Bouton, J. H., Hill, N. S. Hoveland, C. S. et al. (2001). Georgia State Report to SERAIEG Group 8, Chapel Hill TN, USA, November 5–6, 2001, pp. 920.Google Scholar
Bouton, J. H., Latch, G. C. M., Hill, N. S. et al. (2002). Reinfection of tall fescue cultivars with non-ergot alkaloid-producing endophytes. Agronomy Journal, 94, 567574.Google Scholar
Brendemuehl, J. P., Boosinger, T. R., Pugh, D. G. and Shelby, R. A. (1994). Influence of endophyte-infected tall fescue on cyclicity, pregnancy rate and early embryonic loss in the mare. Theriogenology 42, 489500.CrossRefGoogle ScholarPubMed
Breen, J. P. (1993). Enhanced resistance to three species of aphids (Homoptera: Aphididae) in Acremonium endophyte-infected turfgrasses. Journal of Economic Entomology, 86, 12791286.CrossRefGoogle Scholar
Briggs, L., Tapper, B., Sprosen, J., Mace, W. and Finch, S. (2017). Development of an enzyme-linked immunosorbent assay for the detection of lolines in pastures. Food and Agricultural Immunology, 28, 10581070.CrossRefGoogle Scholar
Briggs, L. R., Sprosen, J. M., Tapper, B. A. and Easton, H. S. (2007). AR1 quality assurance by ELISA. In Research and Practice New Zealand Grassland Association 13: 6th International Symposium on Fungal Endophytes of Grasses, ed. Popay, A. J. and Thom, E. R.. Palmerston, New Zealand: New Zealand Grassland Association, pp. 289291.Google Scholar
Brown, M. A., Brown, A. H. Jr., Jackson, W. G. and Miesner, J. R. (1996). Milk production in Angus, Brahman, and reciprocal-cross cows grazing common Bermuda grass or endophyte-infected tall fescue. Journal of Animal Science, 74, 20582066.CrossRefGoogle ScholarPubMed
Burke, J. M., Brauer, D. K. and Looper, M. L. (2004). Use of novel endophyte-infected tall fescue for cow–calf production in Arkansas. American Society of Animal Science Annual Meeting, 82 (S. 1), 91.Google Scholar
Bush, L. P., Wilkinson, H. H. and Schardl, C. L. (1997). Bioprotective alkaloids of grass-fungal endophyte symbioses. Plant Physiology, 114, 17.CrossRefGoogle ScholarPubMed
Caradus, J. R., Lovatt, S. and Belgrave, B. (2013). Adoption of forage technologies. Proceedings of New Zealand Grassland Association, 75, 3944.CrossRefGoogle Scholar
Card, S. D., Rolston, M. P., Park, Z., Cox, N. and Hume, D. E. (2011). Fungal endophyte detection in pasture grass seed utilising the infection layer and comparison to other detection techniques. Seed Science & Technology, 39, 581592.CrossRefGoogle Scholar
Cheplick, G. P. (2004). Recovery from drought stress in Lolium perenne (Poaceae): are fungal endophytes detrimental? American Journal of Botany, 91, 19601968.CrossRefGoogle Scholar
Cheplick, G. P., Perera, A. and Koulouris, K. (2000). Effect of drought on the growth of Lolium perenne genotypes with and without fungal endophytes. Functional Ecology, 14, 657667.CrossRefGoogle Scholar
Clay, K. (1989). Clavicipitaceous endophyte of grasses: their potential as biocontrol agents. Mycological Research, 92, 112.CrossRefGoogle Scholar
Clay, K. (1990). Fungal endophytes of grasses. Annual Review of Ecology and Systematics, 21, 275297.CrossRefGoogle Scholar
Clay, K., Marks, S. and Cheplick, G. P. (1993). Effects of insect herbivory and fungal endophyte infection on competitive interactions among grasses. Ecology, 74, 17671777.CrossRefGoogle Scholar
Clement, S. L., Elberson, L. R., Waldron, B. L. and Quisenberry, S. S. (2007). Variable performance of bird cherry-oat aphid on Neotyphodium-infected tall fescue from Tunisia. In Research and Practice New Zealand Grassland Association 13: 6th International Symposium on Fungal Endophytes of Grasses, ed. Popay, A. J. and Thom, E. R.. Palmerston, New Zealand: New Zealand Grassland Association, pp. 337340.Google Scholar
Christensen, M. J., Bennett, R. J., Ansari, H. A. et al. (2008). Epichloë endophytes grow by intercalary hyphal extension in elongating grass leaves. Fungal Genetics and Biology, 45, 8493.CrossRefGoogle ScholarPubMed
Coffey, K. P., Coblentz, W. K. and Caldwell, J. D. (2015). Pre- and postweaning performance by cows and calves that grazed toxic or nontoxic endophyte-infected tall fescue pastures. The Professional Animal Scientist, 31, 577587.CrossRefGoogle Scholar
Cox, R. J. (2007). Polyketides, proteins and genes in fungi: programmed nano-machines begin to reveal their secrets. Organic and Biomolecular Chemistry, 5, 20102026.CrossRefGoogle ScholarPubMed
Danielson, D. A., Schmidt, S. P., King, C. C., Smith, L. A. and Webster, W. B. (1986). Fescue toxicity and reproduction in beef heifers. Journal of Animal Science, 63 (S.1), 296.Google Scholar
Davies, E., Lane, G. A., Latch, G. C. M. et al. (1993). Alkaloid concentrations in field-grown synthetic perennial ryegrass endophyte associations. In Proceedings of the Second International Symposium on Acremonium/Grass Interactions, ed. Hume, D. E., Latch, G. C. M. and H. S. Easton. Palmerston, New Zealand: Grasslands Research Centre, pp. 7276.Google Scholar
Ditsch, D. C. and Aiken, G. E. (2009). Endophyte infected tall fescue and small ruminant production: do we have a problem? Online. Forage & Grazinglands. doi: 10.1094/FG2009-1104-02-RS.CrossRefGoogle Scholar
Duckett, S. K., Lacy, R. C., Andrae, J. G. et al. (2007). Animal performance, carcass quality and economics of cattle finished after grazing endophyte-infected, endophyte-free or nonergot alkaloid-producing endophyte-infected tall fescue. In Research and Practice New Zealand Grassland Association 13: 6th International Symposium on Fungal Endophytes of Grasses, ed. Popay, A. J. and Thom, E. R.. Palmerston, New Zealand: New Zealand Grassland Association, pp. 253255.Google Scholar
Easton, H. S. (1999). Endophyte in New Zealand ryegrass pastures: an overview. Ryegrass endophyte: an essential New Zealand symbiosis. In Ryegrass endophyte: an Essential New Zealand Symbiosis, ed. Woodfield, D. R. and Matthew, C.. Grassland Research and Practice Series 7, pp. 19.Google Scholar
Easton, H. S. (2007). Grasses and Neotyphodium endophytes: co-adaptation and adaptive breeding. Euphytica, 154, 295306.CrossRefGoogle Scholar
Easton, H. S., Lane, G. A., Tapper, B. A. et al. (1996). Ryegrass endophyte-related heat stress in cattle. Proceedings of the New Zealand Grassland Association, 57, 3741.Google Scholar
Easton, H. S., Latch, G. C. M., Tapper, B. A. and Ball, O. J.-P. (2002). Ryegrass host genetic control of concentrations of endophyte-derived alkaloids. Crop Science, 42, 5157.Google ScholarPubMed
Easton, H. S., Lyons, T. B., Mace, W. J., Simpson, W. R. et al. (2007). Differential expression of loline alkaloids in perennial ryegrass infected with endophyte isolated from tall fescue. In Research and Practice New Zealand Grassland Association 13: 6th International Symposium on Fungal Endophytes of Grasses, ed. Popay, A. J. and Thom, E. R.. Palmerston, New Zealand: New Zealand Grassland Association, pp. 163165.Google Scholar
Essig, H. W., Cantrell, C. E., Withers, F. T. Jr. et al. (1989). Performance and profitability of cow-calf systems grazing on EF and EI KY-31 fescue. (Preliminary report). Proceeding of Tall Fescue Toxicosis Workshop, November 13–14, Atlanta, GA, p. 61.Google Scholar
Essig, H. W., Aremu, B., Cantrell, C. E., Boyd, M. E. and Withers, F. T. Jr. (1993). Impacts of endophyte-infected fescue on cow/calf production. The Professional Animal Scientist, 9, 6469.CrossRefGoogle Scholar
Faville, M. J., Briggs, L., Cao, M. et al. (2015). A QTL analysis of host plant effects on fungal endophyte biomass and alkaloid expression in perennial ryegrass. Molecular Breeding, 35, 161.CrossRefGoogle ScholarPubMed
Fleetwood, D. J., Scott, B., Lane, G. A., Tanaka, A. and Johnson, R. D. (2007). A complex ergovaline gene cluster in Epichloë endophytes of grasses. Applied and Environmental Microbiology, 73, 25712579.CrossRefGoogle ScholarPubMed
Fletcher, L. R. (1999). ‘Non-toxic’ endophytes in ryegrass and their effect on livestock health and production. In Ryegrass endophyte: an Essential New Zealand Symbiosis, ed. Woodfield, D. R. and Matthew, C.. Grassland Research and Practice Series 7, pp. 133139.Google Scholar
Fletcher, L. R. (2005). Managing ryegrass-endophyte toxicoses. In Neotyphodium in Cool-Season Grasses, ed. Roberts, C. A., West, C. P. and Spiers, D. E.. Ames, IA: Blackwell Publishing, pp. 229241.Google Scholar
Fletcher, L. R. (2012). Novel endophytes in New Zealand grazing systems: the perfect solution or a compromise? In Epichloae, Endophytes of Cool Season Grasses: Implications, Utilization and Biology. Proceedings of the 7th International Symposium on Fungal Endophytes of Grasses, Lexington, Kentucky, USA, 28 June to 1 July 2010, ed. Young, C. A., Aiken, G. E., McCulley, R. L., Strickland, J. R. and Schardl, C. L., pp. 513.Google Scholar
Fletcher, L. R. and Harvey, I. C. (1981). An association of a Lolium endophyte with ryegrass staggers. New Zealand Veterinary Journal, 29, 185186.CrossRefGoogle ScholarPubMed
Fletcher, L. R. and Sutherland, B. L. (2009). Sheep responses to grazing ryegrass with AR37 endophyte. Proceedings of New Zealand Grassland Association, 71, 127133.CrossRefGoogle Scholar
Fletcher, L. R., Sutherland, B. L. and Fletcher, C. G. (1999). The impact of endophyte on the health and productivity of sheep grazing ryegrass-based pasture. Grassland Research and Practice Series, 7, 1117.Google Scholar
Fletcher, L. R., Fletcher, C. G. and Sutherland, B. L. (2000). The health and performance of sheep grazing a non-toxic tall fescue endophyte association. In Proceedings of the 4th International Neotyphodium/Grass Interactions Symposium. Soest, Germany, ed. Paul, V. H. and Dapprich, P. D.. Soest, Germany: Universitat Paderborn, Gesamthochschulepp, pp. 459464.Google Scholar
Forester, N. T., Lane, G. A., Steringa, M., Lamont, I. L. and Johnson, L. J. (2017). Contrasting roles of fungal siderophores in maintaining iron homeostasis in Epichloë festucae. Fungal Genetics and Biology, 111, 6072.CrossRefGoogle ScholarPubMed
Franzluebbers, A. J., Seman, D. H. and Stuedemann, J. A. (2009). Tall fescue persists and cattle perform well on a novel-endophyte association in the southern Piedmont USA. Forage & Grasslands, 7, doi: 10.1094/FG-2009-0227-01-RS.Google Scholar
Gallagher, R. T., Keogh, R. G., Latch, G. C. M. and Reid, C. S. W. (1978). The role of fungal tremorgens in ryegrass staggers. New Zealand Journal of Agricultural Research, 20, 431440.CrossRefGoogle Scholar
Gallagher, R. T., White, E. P. and Mortimer, P. H. (1981). Ryegrass staggers: isolation of potent neurotoxins lolitrem A and lolitrem B from staggers producing pastures. New Zealand Veterinary Journal, 29, 189190.CrossRefGoogle ScholarPubMed
Gallagher, R. T., Campbell, A. G., Hawkes, A. D. et al. (1982). Ryegrass staggers: the presence of lolitrem neurotoxins in ryegrass seed. New Zealand Veterinary Journal, 30, 183184.CrossRefGoogle ScholarPubMed
Gallagher, R. T., Hawkes, A. D., Steyn, P. S. and Vleggaar, R. (1984). Tremorgenic neurotoxins from perennial ryegrass causing ryegrass staggers disorder of livestock: Structure elucidation of lolitrem B. Journal of Chemistry Society Chemistry Communications, 1984, 614616.CrossRefGoogle Scholar
Garrett, L. W., Heimann, E. D., Pfander, W. H. and Wilson, L. L. (1980). Reproductive problems of pregnant mares grazing fescue pastures. Journal of Animal Science 51(S. 1), 237.Google Scholar
Gay, N., Boling, J. A., Dew, R. and Miksch, D. E. (1988). Effects of endophyte-infected tall fescue on beef cow-calf performance. Applied Agricultural Research, 3, 182186.Google Scholar
Goldson, S. L., McNeill, M. R., Proffitt, J. R. et al. (1993). Systematic mass rearing and release of Microctonus hyperodae (Hym.: Braconidae, Euphorinae), a parasitoid of the Argentine stem weevil Listronotus bonariensis (Col.: Curculionidae) and records of its establishment in New Zealand. Entomophaga, 38, 527536.CrossRefGoogle Scholar
Groppe, K., Sanders, I., Wiemken, A. and Boller, T. (1995). A microsatellite marker for studying the ecology and diversity of fungal endophytes (Epichloë spp.) in grasses. Applied and Environmental Microbiology, 61, 39433949.CrossRefGoogle ScholarPubMed
Gwinn, K. D., Collinsshepard, M. H. and Reddick, B. B. (1991). Tissue print-immunoblot, an accurate method for the detection of Acremonium coenophialium in tall fescue. Phytopathology, 81, 747748.CrossRefGoogle Scholar
Harvey, I., Fletcher, L. and Emms, L. (1982). Effects of several fungicides on the Lolium endophyte in ryegrass plants, seeds, and in culture. New Zealand journal of Agricultural Research, 25, 601606.CrossRefGoogle Scholar
Hill, N. S., Thompson, F. N., Dawe, D. L. and Stuedemann, J. A. (1994). Antibody binding of circulating ergopeptine alkaloids in cattle grazing tall fescue. American Journal of Veterinary Research, 55, 419424.Google Scholar
Hopkins, A. A. and Bouton, J. H. (2007). A new tall fescue-novel endophyte combination for the south-central USA. In Research and Practice New Zealand Grassland Association 13: 6th International Symposium on Fungal Endophytes of Grasses, ed. Popay, A. J. and Thom, E. R.. Palmerston, New Zealand: New Zealand Grassland Association, pp. 257258.Google Scholar
Hoveland, C. S. (1993). Importance and economic significance of the Acremonium endophytes to performance of animals and grass plant. Agricultural Ecosystem & Environment, 44, 312.CrossRefGoogle Scholar
Hoveland, C. S., Schmidt, S. P., King, Jr, C.C. et al. (1983). Steer performance and association of Acremonium coenophialum fungal endophyte on tall fescue pasture. Agronomy Journal, 75, 821824.CrossRefGoogle Scholar
Hoveland, C. S., Schmidt, S. P., King, C. C. Jr. and Clark, E. M. (1984). Association of fungal endophyte with seasonal gains of beef steers grazing tall fescue pasture. In Proceedings of European Grassland Federation, ed. Riley, H. and Skelvag, A. D.. Zurich, Switzerland: European Grassland Federation, pp. 382386.Google Scholar
Howard, M. D., Muntifering, R. B., Bradley, N. W., Mitchell, G. E. Jr. and Lowry, S. R. (1992). Voluntary intake and ingestive behaviour of steers grazing Johnstone or endophyte-infected Kentucky-31 tall fescue. Journal of Animal Science, 70, 12271237.CrossRefGoogle ScholarPubMed
Hume, D. E., Popay, A. J., Cooper, J. P. et al. (2004). Effect of a novel endophyte on the productivity of perennial ryegrass (Lolium perenne) in New Zealand. In Proceedings of 5th International Symposium on Neotyphodium/Grass Interactions, ed. Kallenbach, R., Rosekrans, C. Jr. and Lock, T. R., Poster #313.Google Scholar
Hume, D. E., Cooper, B. M. and Panckhurst, K. A. (2009). The role of endophyte in determining the persistence and productivity of ryegrass, tall fescue and meadow fescue in Northland. Proceedings of the New Zealand Grassland Association, 71, 145150.CrossRefGoogle Scholar
Hume, D. E., Card, S. D. and Rolston, M. P. (2013). Effects of storage conditions on endophyte and seed viability in pasture grasses. Proceedings of the 22nd International Grassland Congress, Sydney, pp. 405408.Google Scholar
Gundel, P., Rudgers, J. and Ghersa, C. (2011). Incorporating the process of vertical transmission into understanding of host–symbiont dynamics. Oikos, 120, 11211128.CrossRefGoogle Scholar
Jensen, J. G. and Popay, A. J. (2004). Perennial ryegrass infected with AR37 endophyte reduces survival of porina larvae. New Zealand Plant Protection, 57, 323328.CrossRefGoogle Scholar
Jensen, J. G. and Popay, A. J. (2007). Reductions in root aphid population by non-toxic endophyte strains in tall fescue. In Research and Practice New Zealand Grassland Association 13: 6th International Symposium on Fungal Endophytes of Grasses, ed. Popay, A. J. and Thom, E. R.. Palmerston, New Zealand: New Zealand Grassland Association, pp. 341344.Google Scholar
Jensen, J. G., Popay, A. J. and Tapper, B. A. (2009). Argentine stem weevil adults are affected by meadow fescue endophyte and its loline alkaloids. New Zealand Plant Protection, 62, 1218.CrossRefGoogle Scholar
Johnson, R. D. and Voisey, C. R. (2017). Method for Detection of Viable Endophyte in Plants. US Patent number 9631242.Google Scholar
Johnson, R. D., Lane, G. A., Koulman, A. et al. (2015). A novel family of cyclic oligopeptides derived from ribosomal peptide synthesis of an in planta-induced gene, gigA, in Epichloë endophytes of grasses. Fungal Genetics & Biology, 85, 1424.CrossRefGoogle Scholar
Johnson, J. M., Aiken, G. E., Phillips, T. D. et al. (2012). Steer and pasture responses for a novel endophyte tall fescue developed for the upper transition zone. Journal of Animal Science, 90, 24022409.CrossRefGoogle ScholarPubMed
Johnson, L. J., de Bonth, A. C. M., Briggs, L. R. et al. (2013a). The exploitation of epichloae endophytes for agricultural benefit. Fungal Diversity, 60, 171188.CrossRefGoogle Scholar
Johnson, L. J., Koulman, A., Christensen, M. et al. (2013b). An extracellular siderophore is required to maintain the mutualistic interaction of Epichloë festucae with Lolium Perenne. PLoS Pathogens, 9: e1003332.CrossRefGoogle ScholarPubMed
Johnson, T., ed. (1983). Proceedings of Tall Fescue Toxicosis Workshop. Cooperative Extension Service. Athens, Georgia: University of Georgia College of Agriculture, pp. 126.Google Scholar
Kane, K. H. (2011). Effects of endophyte infection on drought stress tolerance of Lolium perenne accessions from the Mediterranean region. Environmental and Experimental Botany, 71, 337344.Google Scholar
Karimi, S., Mirlohi, A., Sabzalian, M. R., Tabatabaei, B. E. S. and Sharifnabi, B. (2012). Molecular evidence for Neotyphodium fungal endophyte variation and specificity within host grass species. Mycologia, 104, 12811290.CrossRefGoogle ScholarPubMed
Koulman, A., Lee, T. V., Fraser, K. et al. (2012). Identification of extracellular siderophores and a related peptide from the endophytic fungus Epichloë festucae in culture and endophyte-infected Lolium perenne. Phytochemistry, 75, 128139.CrossRefGoogle Scholar
Lane, G. A., Christensen, M. J. and Miles, C. O. (2000). Coevolution of fungal endophytes with grasses: the significance of secondary metabolites. In Microbial Endophytes, ed. Bacon, C. W. and White, J. F.. New York: Marcel Dekker, pp. 341388.Google Scholar
Latch, G. C. M. (1994). Influence of Acremonium endophytes on perennial ryegrass improvement. New Zealand Journal of Agricultural Research, 37, 311318.CrossRefGoogle Scholar
Latch, G. C. M. (1997). An overview of Neotyphodium/Grass interactions. In Neotyphodium/Grass Interactions, ed. Bacon, C. W. and Hill, N. S.. New York: Plenus Press, New York, pp. 111.Google Scholar
Latch, G. C. M. and Christensen, M. J. (1985). Artificial infections of grasses with endophytes. Annals of Applied Biology, 107, 1724.CrossRefGoogle Scholar
Latch, G. C. M., Christensen, M. J. and Samuels, G. J. (1984). Five endophytes of Lolium and Festuca in New Zealand. Mycotaxon, 20, 535550.Google Scholar
Latch, G. C. M., Christensen, M. J., Tapper, B. A. et al. (2000). Tall fescue Endophytes. U.S. Patent 6 111 170. Date issued: 29 August 2000.Google Scholar
Leuchtmann, A., Bacon, C. W., Schardl, C. L., White, J. F. and Tadych, M. (2014). Nomenclatural realignment of Neotyphodium species with genus Epichloë. Mycologia, 106, 202215.CrossRefGoogle Scholar
Leury, B., Henry, M., Kemp, S. and Webb-Ware, J. (2014). Developing Increased Understanding, Awareness and Potential Mitigation Strategies for Perennial Ryegrass Toxicosis in Sheep Production Systems. Project B.AHE.0039. North Sydney, Australia: Meat and Livestock Australia Limited.Google Scholar
Looper, M. L., Rorie, R. W., Person, C. N. et al. (2009). Influence of toxic endophyte-infected fescue on sperm characteristics and endocrine factors of yearling Brahman-influenced bulls. Journal of Animal Science, 87, 11841191.CrossRefGoogle ScholarPubMed
Lyons, P. C., Plattner, R. D. and Bacon, C. W. (1986). Occurrence of peptide and clavine ergot alkaloids in tall fescue grass. Science, 232, 487489.CrossRefGoogle ScholarPubMed
Kuldau, G. and Bacon, C. (2008). Clavicipitaceous endophytes: Their ability to enhance resistance of grasses to multiple stresses. Biological Control, 46, 5771.CrossRefGoogle Scholar
Marahiel, M. A. (1997). Protein templates for the biosynthesis of peptide antibiotics. Chemistry and Biology, 4, 561567.CrossRefGoogle ScholarPubMed
Malinowski, D. P. and Belesky, D. P. (2000). Adaptations of endophyte-infected cool-season grasses to environmental stresses: mechanisms of drought and mineral stress tolerance. Crop Science, 40, 923940.CrossRefGoogle Scholar
Malinowski, D., Leuchtmann, A., Schmidt, D. and Nosberger, J. (1997). Growth and water status in meadow fescue (Festuca pratensis) is affected by Neotyphodium and Phialophora endophytes. Agronomy Journal, 89, 673678.CrossRefGoogle Scholar
Meister, B., Krauss, J., Härri, S. A., Schneider, M. V. and Müller, C. B. (2006). Fungal endosymbionts affect aphid population size by reduction of adult life span and fecundity. Basic and Applied Ecology, 7, 244252.CrossRefGoogle Scholar
Moate, P. J., Williams, S. R. O., Grainger, C. et al. (2012). Effects of wild-type, AR1 and AR37 endophyte-infected perennial ryegrass on dairy production in Victoria, Australia. Animal Production Science, 52, 11171130.CrossRefGoogle Scholar
Moon, C. D., Tapper, B. A. and Scott, B. (1999). Identification of Epichloë endophytes in planta by a microsatellite-based PCR fingerprinting assay with automated analysis. Applied and Environmental Microbiology, 65, 12681279.CrossRefGoogle ScholarPubMed
Panaccione, D. G., Johnson, R. D., Wang, J. et al. (2001). Elimination of ergovaline from a grass–Neotyphodium endophyte symbiosis by genetic modification of the endophyte. Proceedings of the National Academy of Sciences of the United States of America, 98, 1282012825.CrossRefGoogle ScholarPubMed
Parish, J. A., McCann, M. A., Watson, R. H. et al. (2003a). Use of non-ergot alkaloid-producing endophytes for alleviating tall fescue toxicosis in sheep. Journal of Animal Science, 81, 13161322.CrossRefGoogle Scholar
Parish, J. A., McCann, M. A., Watson, R. H. et al. (2003b). Use of non-ergot alkaloid-producing endophytes for alleviating tall fescue toxicosis in stocker cattle. Journal of Animal Science, 81, 28562868.CrossRefGoogle Scholar
Patchett, B. J., Chapman, R. B., Fletcher, L. R. and Gooneratne, S. R. (2008). Endophyte infected Festuca pratensis containing loline alkaloids deter feeding by Listronotus bonariensis. New Zealand Plant Protection, 61, 205209.CrossRefGoogle Scholar
Patchett, B. J., Gooneratne, S. R. and Chapman, R. B. (2011). Effects of loline-producing endophyte-infected meadow fescue ecotypes on New Zealand grass grub (Costelytra zealandica). New Zealand Journal of Agricultural Research, 54, 303313.CrossRefGoogle Scholar
Paterson, J., Forcherio, C., Larson, B., Samford, M. and Kerley, M. (1995). The effects of fescue toxicosis on beef cattle productivity. Journal of Animal Science, 73, 889–98.CrossRefGoogle ScholarPubMed
Pedersen, J. F., Lacefield, G. D. and Ball, D. M. (1990). A review of the agronomic characteristics of endophyte-free and endophyte infected tall fescue. Applied Agricultural Research, 5, 188194.Google Scholar
Pennell, C., Popay, A. J., Ball, O. J. P., Hume, D. E. and Baird, D. B. (2005). Occurrence and impact of pasture mealybug (Balanococcus poae) and root aphid (Aploneura lentisci) on ryegrass (Lolium spp.) with and without infection by Neotyphodium fungal endophytes. New Zealand Journal of Agricultural Research, 48, 329337.CrossRefGoogle Scholar
Peters, C. W., Grigsby, K. N., Aldrich, C. G. et al. (1992). Performance, forage utilization, and ergovaline consumption by beef cows grazing endophyte fungus-infected tall fescue, endophyte fungus-free tall fescue, or orchardgrass pastures. Journal of Animal Science, 70, 15501561.CrossRefGoogle ScholarPubMed
Philipson, M. N. and Christey, M. C. (1986). The relationship of host and endophyte during flowering, seed formation, and germination of Lolium perenne. New Zealand Journal of Botany, 24, 125134.CrossRefGoogle Scholar
Popay, A. J. and Hume, D. H. (2011). Endophytes improve ryegrass persistence by controlling insects. New Zealand Grassland Association Pasture Persistence Symposium. Grassland Research and Practice Series, 15, 149156.Google Scholar
Popay, A. J. and Rowan, D. D. (1994). Endophytic fungi as mediators of plant insect interactions. In Insect–Plant Interactions, ed. Bernays, E. A.. Boca Raton, FL: CRC Press, pp. 83103.Google Scholar
Popay, A. J. and Thom, E. R. (2009). Endophyte effects on major insect pests in Waikato dairy pasture. Proceedings of the New Zealand Grassland Association, 71, 121127.CrossRefGoogle Scholar
Popay, A. J., Hume, D. E., Baltus, J. G. et al. (1999). Field performance of perennial ryegrass (Lolium perenne) infected with toxin-free fungal endophytes (Neotyphodium spp.). In Ryegrass endophyte: an Essential New Zealand Symbiosis, ed. Woodfield, D. R. and Matthew, C.. Grassland Research and Practice Series 7, 113122.Google Scholar
Popay, A. J., Tapper, B. A. and Podmore, C. (2009). Endophyte-infected meadow fescue and loline alkaloids affect Argentine stem weevil larvae. New Zealand Plant Protection, 62, 1927.CrossRefGoogle Scholar
Porter, J. K. and Thompson, F. N. Jr. (1992). Effects of fescue toxicosis on reproduction in livestock. Journal of Animal Science, 70, 15941603.CrossRefGoogle ScholarPubMed
Potter, D. A., Patterson, C. G. and Redmond, C. T. (1992). Influence of turfgrass species and tall fescue endophyte on feeding ecology of Japanese beetle and southern masked chafer grubs (Coleoptera: Scarabaeidae). Journal of Economic Entomology, 85, 900909.CrossRefGoogle Scholar
Prestidge, R. A. and Gallagher, R. T. (1988). Endophyte fungus confers resistance to ryegrass: Argentine stem weevil larval studies. Ecological Entomology, 13, 429435.CrossRefGoogle Scholar
Prestidge, R. A., Pottinger, R.P. and Barker, G. M. (1982). An association of Lolium endophyte with ryegrass resistance to Argentine stem weevil. Proceedings of the 35th New Zealand Weed and Pest Control Conference, 119–122.CrossRefGoogle Scholar
Prestidge, R. A., Barker, G. M. and Pottinger, R. P. (1991). The economic cost of Argentine stem weevil in pastures in New Zealand. Proceedings of the New Zealand Weed and Pest Control Conference, 44, 165170.CrossRefGoogle Scholar
Putnam, M. R., Bransby, D. I., Schumacher, J. et al. (1991). Effects of the fungal endophyte Acremonium coenophialum in fescue on pregnant mares and foal viability. American Journal of Veterinary Research 52, 20712074.Google ScholarPubMed
Realini, C. E., Duckett, S. K., Hill, N. S. et al. (2005). Effect of endophyte type on carcass traits, meat quality, and fatty acid composition of beef cattle grazing tall fescue. Journal of Animal Science, 8, 430439CrossRefGoogle Scholar
Reddick, B. and Collins, M. (1988). An improved method for detection of Acremonium coenophialum in tall fescue plants. Phytopathology, 78, 418420.CrossRefGoogle Scholar
Reidell, W., Kieckhefer, R., Petroski, R. and Powell, R. (1991). Naturally occurring and synthetic loline alkaloid derivatives: insect feeding behaviour modification and toxicity. Journal of Entomological Science, 26, 122129.CrossRefGoogle Scholar
Rhodes, M. T., Paterson, J. A., Kerley, M. S., Garner, H. E. and Laughlin, M. H. (1991). Reduced blood flow to peripheral and core body temperatures in sheep and cattle induced by endophyte infected tall fescue. Journal of Animal Science, 69, 20332043.CrossRefGoogle ScholarPubMed
Rolston, M. and Agee, C. (2007). Delivering quality seed to specification-the USA and NZ novel endophyte experience. In Research and Practice New Zealand Grassland Association 13: 6th International Symposium on Fungal Endophytes of Grasses, ed. Popay, A. J. and Thom, E. R.. Palmerston, New Zealand: New Zealand Grassland Association, pp. 229232.Google Scholar
Rolston, M., Hare, M., Moore, K. and Christensen, M. (1986). Viability of Lolium endophyte fungus in seed stored at different moisture contents and temperatures. New Zealand Journal of Experimental Agriculture, 14, 297300.CrossRefGoogle Scholar
Rolston, M. P., Archie, W. J. and Simpson, W. (2002). Tolerance of AR1 Neotyphodium endophyte to fungicides used in perennial ryegrass seed production. Proceedings New Zealand Plant Protection Conference, 55, 322325.CrossRefGoogle Scholar
Rolston, M. P., Mccloy, B. L., Harvey, I. C. and Chynoweth, R. W. (2009). Ryegrass (Lolium perenne) seed yield response to fungicides: a summary of 12 years of field research. New Zealand Plant Protection, 62, 343348.CrossRefGoogle Scholar
Rowan, D. D. and Gaynor, D. L. (1986). Isolation of feeding deterrents against Argentine stem weevil from ryegrass infected with the endophyte Acremonium loliae. Journal of Chemical Ecology, 12, 647658.CrossRefGoogle ScholarPubMed
Rowan, D. D., Hunt, M. B. and Gaynor, D. L. (1986). Peramine, a novel insect feeding deterrent from ryegrass infected with the endophyte Acremonium loliae. Journal of Chemistry Society Chemistry Communications, 1986, 935936.CrossRefGoogle Scholar
Rowan, D. D., Dymock, J. J. and Brimble, M. A. (1990). Effect of fungal metabolite peramine and analogs on feeding and development of Argentine stem weevil (Listronotus bonariensis). Journal of Chemical Ecology, 16, 16831695.CrossRefGoogle ScholarPubMed
Saikia, S., Takemoto, D., Tapper, B. A. et al. (2012). Functional analysis of an indole-diterpene gene cluster for lolitrem B biosynthesis in the grass endosymbiont Epichloë festucae. FEBS Lett, 586, 25632569.CrossRefGoogle ScholarPubMed
Saikkonen, K., Young, C. A., Helander, M. and. Schardl, C. L. (2016). Endophytic Epichloë species and their grass hosts: from evolution to applications. Plant Molecular Biology, 90, 665675.CrossRefGoogle ScholarPubMed
Schardl, C. L., Leuchtmann, A., Chung, K. R., Penny, D. and Siegel, M. R. (1997). Coevolution by common descent of fungal symbionts (Epichloë spp.) and grass hosts. Molecular Biology & Evolution, 14, 133143.CrossRefGoogle Scholar
Schardl, C. L., Leuchtmann, A. and Spiering, M. J. (2004). Symbioses of grasses with seedborne fungal endophytes. Annual Review of Plant Biology, 55, 315340.CrossRefGoogle ScholarPubMed
Schardl, C. L., Grossman, R. B., Nagabhyru, P., Faulkner, J. R. and Mallik, U. P. (2007). Loline alkaloids: currencies of mutualism. Phytochemistry, 68, 980996.CrossRefGoogle ScholarPubMed
Schardl, C. L., Craven, K. D., Speakman, S. et al. (2008). A novel test for host–symbiont codivergence indicates ancient origin of fungal endophytes in grasses. Systematic Biology, 57, 483498.CrossRefGoogle ScholarPubMed
Schardl, C. J., Young, C. A., Hesse, U. et al. (2013). Plant-symbiotic fungi as chemical engineers: multi-genome analysis of the Clavicipitaceae reveals dynamics of alkaloid loci. PLoS Genetics 9, e1003323.CrossRefGoogle ScholarPubMed
Schmidt, S. P. and Osborn, T. G. (1993). Effects of endophyte-infected tall fescue on animal performance. Agriculture Ecosystems and Environment, 44, 233262.CrossRefGoogle Scholar
Schmidt, S. P., Hoveland, C. S., Clark, E. M. et al. (1982). Association of an endophytic fungus with fescue toxicity in steers fed Kentucky 31 tall fescue seed or hay. Journal of Animal Science, 55, 1259.CrossRefGoogle ScholarPubMed
Schmidt, S. P., Danielson, A. D., Holliman, J. A., Grimes, H. W. and Webster, W. B. (1986). Fescue fungus suppresses growth and reproduction in replacement beef heifers. Highlights of Agricultural Research Alabama Agricultural Experiment Station, Auburn University, Auburn, 33, 15.Google Scholar
Schnitzius, J. M., Hill, N. S., Thompson, C. S. and Craig, A. M. (2001). Semiquantitative determination of ergot alkaloids in seed, straw, and digesta samples using a competitive enzyme-linked immunosorbent assay. Journal of Veterinary Diagnostic Investigation 13, 230237.CrossRefGoogle ScholarPubMed
Stewart, A. (1986). Effect on the Lolium endophyte of nitrogen applied to perennial ryegrass seed crops. New Zealand Journal of Experimental Agriculture, 14, 393397.CrossRefGoogle Scholar
Strahan, S. R., Hemken, R. W., Jackson, J. A. Jr. et al. (1987). Performance of lactating dairy cows fed tall fescue forage. Journal of Dairy Science, 70, 12281234.CrossRefGoogle ScholarPubMed
Strickland, J. R., Looper, M. L., Matthews, J. C. et al. (2011). St. Anthony’s fire in livestock; causes, mechanisms and potential solutions. Journal of Animal Science, 89, 16031626.CrossRefGoogle ScholarPubMed
Stuedemann, J. A. and Thompson, F. N. (1993). Management strategies and potential opportunities to reduce the effects of endophyte-infested tall fescue on animal performance. In Proceedings of the Second International Symposium on Acremonium/Grass Interactions, ed. D. E. Hume, G. C. M. Latch and H. S. Easton. Palmerston, New Zealand: Grasslands Research Centre, pp. 103114.Google Scholar
Stuedemann, J. A., Breedlove, D. L., Pond, K. R. et al. (1989). Effect of endophyte (Acremonium coenophialum) infection of tall fescue and paddock exchange on intake and performance of grazing steers. Proceedings of the XVI International Grassland Congress, pp. 12431244.Google Scholar
Sullivan, T. J., Bultman, T. L. and Schoolcraft, J. (2016). Primers to amplify SNP markers in Epichloë canadensis (Clavicipitaceae). Applications in Plant Sciences, 4 (3), apps.1500078.CrossRefGoogle ScholarPubMed
Takach, J. E. and Young, C. A. (2014). Alkaloid genotype diversity of tall fescue endophytes. Crop Science, 54, 667678.CrossRefGoogle Scholar
Takach, J. E., Mittal, S., Swoboda, G. A. et al. (2012). Genotypic and chemotypic diversity of Neotyphodium endophytes in tall fescue from Greece. Applied and Environmental Microbiology, 78, 55015510.CrossRefGoogle ScholarPubMed
Tanaka, A., Tapper, B. A., Popay, A., Parker, E. J. and Scott, B. (2005). A symbiosis expressed non-ribosomal peptide synthetase from a mutualistic fungal endophyte of perennial ryegrass confers protection to the symbiotum from insect herbivory. Molecular Microbiology, 57, 10361050.CrossRefGoogle Scholar
Tapper, B. A. and Lane, G. A. (2004). Janthitrems found in a Neotyphodium endophyte of perennial ryegrass. In Proceedings of 5th International Symposium on Neotyphodium/Grass interactions, ed. Kallenbach, R., Rosekrans, C. and Lock, T. R., #301.Google Scholar
Tapper, B. A. and Latch, G. C. M. (1999). Selection against toxin production in endophyte-infected perennial ryegrass. In Ryegrass endophyte: an Essential New Zealand Symbiosis, ed. Woodfield, D. R. and Matthew, C.. Grassland Research and Practice Series 7, Napier, New Zealand, pp. 107111.Google Scholar
Taylor, M. C., Loch, W. E. and Ellersieck, M. (1985). Toxicity in pregnant pony mares grazing Kentucky-31 fescue pastures. Nutrition Report International 31, 787.Google Scholar
Thom, E. R., Popay, A. J., Waugh, C. D. and Minneé, E. M. (2014). Impact of novel endophytes in perennial ryegrass on herbage production and insect pests from pastures under dairy cow grazing in northern New Zealand. Grass Forage Science, 69, 191204.CrossRefGoogle Scholar
Timper, P., Gates, R. N. and Bouton, J. H. (2005). Response of Pratylenchus spp. In tall fescue infected with different strains of the fungal endophyte Neotyphodium coenophialum. Nematology, 7, 105110.CrossRefGoogle Scholar
van Zijll de Jong, E., Guthridge, K. M., Spangenberg, G. C. and Forster, J. W. (2003). Development and characterization of EST-derived simple sequence repeat (SSR) markers for pasture grass endophytes. Genome, 46, 277290.CrossRefGoogle ScholarPubMed
Watson, B. M. (2007). The effect of endophyte in perennial ryegrass and tall fescue on red and blackheaded pasture cockchafers. In Research and Practice New Zealand Grassland Association 13: 6th International Symposium on Fungal Endophytes of Grasses, ed. Popay, A. J. and Thom, E. R.. Palmerston, New Zealand: New Zealand Grassland Association, pp. 347352.Google Scholar
Watson, R. H., McCann, M. A., Parish, J. A. et al. (2004). Productivity of cow-calf pairs grazing tall fescue pastures infected with either the wild-type endophyte or a non-ergot alkaloid-producing endophyte strain, AR542. Journal of Animal Science, 82, 33883393.CrossRefGoogle ScholarPubMed
Welty, R., Azevedo, M. and Cooper, T. (1987). Influence of moisture content, temperature, and length of storage on seed germination and survival of endophytic fungi in seeds of tall fescue and perennial ryegrass. Phytopathology, 77, 893900.CrossRefGoogle Scholar
West, C. P., Izekor, E., Turner, K. E. and Elmi, A. A. (1993). Endophyte effects on growth and persistence of tall fescue along a water-supply gradient. Agronomy Journal, 85, 264270.CrossRefGoogle Scholar
Wilkinson, H. H., Siegel, M. R., Blankenship, J. D. et al. (2000). Contribution of fungal loline alkaloids to protection from aphids in a grass–endophyte mutualism. Molecular Plant–Microbe Interactions, 13, 10271033.CrossRefGoogle Scholar
Young, C. A., Bryant, M. K., Christensen, M. J. et al. (2005). Molecular cloning and genetic analysis of a symbiosis-expressed gene cluster for lolitrem biosynthesis from a mutualistic endophyte of perennial ryegrass. Molecular Genetics & Genomics, 274, 1329.CrossRefGoogle ScholarPubMed
Young, C. A., Felitti, S., Shields, K. et al. (2006). A complex gene cluster for indole-diterpene biosynthesis in the grass endophyte Neotyphodium lolii. Fungal Genetics & Biology, 43, 679693.CrossRefGoogle ScholarPubMed
Young, C. A., Charlton, N. D., Takach, J. E. et al. (2014). Characterization of Epichloë coenophiala within the US: Are all tall fescue endophytes created equal? Chemistry and Biology, 2, 95.Google ScholarPubMed
Zhuang, J., Marchant, M. A., Schardl, C. L. and Butler, C. M. (2005). Economic analysis of replacing endophyte-infected with endophyte-free tall fescue pastures. Agronomy Journal, 97, 711716.CrossRefGoogle Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×