Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-5g6vh Total loading time: 0 Render date: 2024-04-26T11:32:01.386Z Has data issue: false hasContentIssue false

Part III - Understanding wildlife disease ecology at the community and landscape level

Published online by Cambridge University Press:  28 October 2019

Kenneth Wilson
Affiliation:
Lancaster University
Andy Fenton
Affiliation:
University of Liverpool
Dan Tompkins
Affiliation:
Predator Free 2050 Ltd
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Wildlife Disease Ecology
Linking Theory to Data and Application
, pp. 427 - 643
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

References

Aguayo, J., Elegbede, F., Husson, C., Saintonge, F.X. & Marçais, B. (2014) Modeling climate impact on an emerging disease, the Phytophthora alni induced alder decline. Global Change Biology, 20, 32093221.Google Scholar
Alizon, S., de Roode, J. C. & Michalakis, Y. (2013) Multiple infections and the evolution of virulence. Ecology Letters, 16, 556567.CrossRefGoogle ScholarPubMed
Alizon, S., Hurford, A., Mideo, N. & Van Baalen, M. (2009) Virulence evolution and the trade‐off hypothesis: history, current state of affairs and the future. Journal of Evolutionary Biology, 22, 245259.Google Scholar
Alizon, S. & Michalakis, Y. (2015) Adaptive virulence evolution: the good old fitness-based approach. Trends in Ecology & Evolution, 30, 248254.Google Scholar
Altizer, S., Ostfeld, R.S., Johnson, P.T., Kutz, S. & Harvell, C.D. (2013) Climate change and infectious diseases: from evidence to a predictive framework. Science, 341(6145), 514519.Google Scholar
Amarasekare, P. (2003) Competitive coexistence in spatially structured environments: a synthesis. Ecology Letters, 6, 11091122.Google Scholar
Anderson, R.M. & May, R.M. (1982) Coevolution of hosts and parasites. Parasitology, 85, 411426.Google Scholar
Armstrong, R.A. & McGhee, R. (1980) Competitive exclusion. The American Naturalist, 115, 151170.Google Scholar
Baucom, R.S. & de Roode, J.C. (2011) Ecological immunology and tolerance in plants and animals. Functional Ecology, 25, 1828.Google Scholar
Bearchell, S.J., Fraaije, B.A., Shaw, M.W. & Fitt, B.D. (2005) Wheat archive links long-term fungal pathogen population dynamics to air pollution. Proceedings of the National Academy of Sciences of the United States of America, 102, 54385442.Google Scholar
Berendsen, R.L., Pieterse, C.M. & Bakker, P.A. (2012) The rhizosphere microbiome and plant health. Trends in Plant Science, 17, 478486.Google Scholar
Berngruber, T.W., Froissart, R., Choisy, M. & Gandon, S. (2013) Evolution of virulence in emerging epidemics. PLoS Pathogens, 9(3), e1003209.Google Scholar
Bert, D., Lasnier, J.-B., Capdevielle, X., Dugravot, A. & Desprez-Loustau, M.L. (2016) Powdery mildew decreases the radial growth of oak trees with cumulative and delayed effects over years. PLoS ONE, 11(5), e0155344.Google Scholar
Best, A., White, A. & Boots, M. (2014) The coevolutionary implications of host tolerance. Evolution, 68, 14261435.Google Scholar
Bever, J.D., Mangan, S.A. & Alexander, H.M. (2015) Maintenance of plant species diversity by pathogens. Annual Review of Ecology, Evolution, and Systematics, 46, 305325.Google Scholar
Brown, J.H., Gillooly, J.F., Allen, A.P., Savage, V.M. & West, G.B. (2004) Toward a metabolic theory of ecology. Ecology, 85, 17711789.Google Scholar
Brown, J.K. (2003) A cost of disease resistance: paradigm or peculiarity? Trends in Genetics, 19, 667671.CrossRefGoogle ScholarPubMed
Budde, K.B., Nielsen, L.R., Ravn, H.P. & Kjær, E.D. (2016) The natural evolutionary potential of tree populations to cope with newly introduced pests and pathogens – lessons learned from forest health catastrophes in recent decades. Current Forestry Reports, 2, 1829.Google Scholar
Bull, J.J. (1994) Perspective: virulence. Evolution, 48, 14231437.Google Scholar
Bull, J.J. & Ebert, D. (2008) Invasion thresholds and the evolution of nonequilibrium virulence. Evolutionary Applications, 1, 172182.Google Scholar
Burdon, J.J., Thrall, P.H. & Ericson, L. (2013) Genes, communities & invasive species: understanding the ecological and evolutionary dynamics of host–pathogen interactions. Current Opinion in Plant Biology, 16(4), 400405.Google Scholar
Busby, P.E., Ridout, M. & Newcombe, G. (2016) Fungal endophytes: modifiers of plant disease. Plant Molecular Biology, 90, 645655.CrossRefGoogle ScholarPubMed
Chesson, P. (2000) Mechanisms of maintenance of species diversity. Annual Review of Ecology and Systematics, 31, 343366.Google Scholar
Chuine, I., de Cortazar-Atauri, I.G., Kramer, K. & Hänninen, H. (2013) Plant development models. In: Schwarz, M.D. (ed.), Phenology: An Integrative Environmental Science (pp. 275293). Dordrecht: Springer.Google Scholar
Combes, C. (2001) Parasitism: The Ecology and Evolution of Intimate Interactions. Chicago, IL: University of Chicago Press.Google Scholar
Cronin, J.P., Rúa, M.A. & Mitchell, C.E. (2014) Why is living fast dangerous? Disentangling the roles of resistance and tolerance of disease. The American Naturalist, 184, 172187.CrossRefGoogle ScholarPubMed
Crous, P.W. & Groenewald, J.Z. (2005) Hosts, species and genotypes: opinions versus data. Australasian Plant Pathology, 34, 463470.Google Scholar
Cunniffe, N.J., Koskella, B., Metcalf, C.J.E., et al. (2015) Thirteen challenges in modelling plant diseases. Epidemics, 10, 610.Google Scholar
Dantec, C.F., Ducasse, H., Capdevielle, X., et al. (2015) Escape of spring frost and disease through phenological variations in oak populations along elevation gradients. Journal of Ecology, 103, 10441056.Google Scholar
Desprez-Loustau, M.L., Feau, N., Mougou-Hamdane, A. & Dutech, C.C. (2011) Interspecific and intraspecific diversity in oak powdery mildews in Europe: coevolution history and adaptation to their hosts. Mycoscience, 52, 165173.Google Scholar
Desprez-Loustau, M.L., Robin, C., Buee, M., et al. (2007) The fungal dimension of biological invasions. Trends in Ecology & Evolution, 22, 472480.Google Scholar
Desprez-Loustau, M.L., Saint-Jean, G., Barres, B., Dantec, C. & Dutech, C.C. (2014) Oak powdery mildew changes growth patterns in its host tree: host tolerance response and potential manipulation of host physiology by the parasite. Annals of Forest Science, 71, 563573.Google Scholar
Desprez-Loustau, M.L., Vitasse, Y., Delzon, S., et al. (2010) Are plant pathogen populations adapted for encounter with their host? A case study of phenological synchrony between oak and an obligate fungal parasite along an altitudinal gradient. Journal of Evolutionary Biology, 23, 8797.Google Scholar
Doumayrou, J., Avellan, A., Froissart, R. & Michalakis, Y. (2013) An experimental test of the transmission–virulence trade-off hypothesis in a plant virus. Evolution, 67, 477486.Google Scholar
Ducousso, A., Guyon, J.P. & Kremer, A. (1996) Latitudinal and altitudinal variation of bud burst in western populations of sessile oak (Quercus petraea (Matt) Liebl). Annals of Forest Science, 53, 775782.Google Scholar
Edwards, M.C. & Ayres, P.G. (1982) Seasonal changes in resistance of Quercus petraea (sessile oak) leaves to Microsphaera alphitoides. Transactions of the British Mycological Society, 78, 569571.Google Scholar
Emmons, C.W. (1930) Cicinnobolus cesatii, a study in host–parasite relationships. Bulletin of the Torrey Botanical Club, 57, 421441.Google Scholar
Ennos, R.A. (2015) Resilience of forests to pathogens: an evolutionary ecology perspective. Forestry, 88, 4152.Google Scholar
Escriu, F., Fraile, A. & García-Arenal, F. (2003) The evolution of virulence in a plant virus. Evolution, 57, 755765.Google Scholar
Feau, N., Decourcelle, T., Husson, C., Desprez Loustau, M.L. & Dutech, C.C. (2011) Finding single copy genes out of sequenced genomes for multilocus phylogenetics in non-model fungi. PLoS ONE, 6(4), e18803.Google Scholar
Feau, N., Lauron-Moreau, A., Piou, D., et al. (2012) Niche partitioning of the genetic lineages of the oak powdery mildew complex. Fungal Ecology, 5, 154162.Google Scholar
Fisher, M.C., Henk, D.A., Briggs, C.J., et al. (2012) Emerging fungal threats to animal, plant and ecosystem health. Nature, 484(7393), 186194.Google Scholar
Fitt, B.D., Huang, Y., van den Bosch, F. & West, J.S. (2006) Coexistence of related pathogen species on arable crops in space and time. Annual Review of Phytopathology, 44, 163–82.Google Scholar
Flory, S.L. & Clay, K. (2013) Pathogen accumulation and long‐term dynamics of plant invasions. Journal of Ecology, 101, 607613.Google Scholar
Francl, L.J. (2001) The disease triangle: a plant pathological paradigm revisited. Plant Health Instructor, DOI:10.1094/PHI-T-2001-0517-01Google Scholar
Gilchrist, M.A., Sulsky, D.L. & Pringle, A. (2006). Identifying fitness and optimal life-history strategies for an asexual filamentous fungus. Evolution, 60, 970979.Google Scholar
Glawe, D.A. (2008) The powdery mildews: a review of the world’s most familiar (yet poorly known) plant pathogens. Annual Review of Phytopathology, 46, 2751.Google Scholar
Guillaume, F. & Rougemont, J. (2006) Nemo: an evolutionary and population genetics programming framework. Bioinformatics, 22, 25562557.Google Scholar
Hajji, M., Dreyer, E. & Marçais, B. (2009) Impact of Erysiphe alphitoides on transpiration and photosynthesis in Quercus robur leaves. European Journal of Plant Pathology, 125, 6372.Google Scholar
Halkett, F., Harrington, R., Hullé, M., et al. (2004) Dynamics of production of sexual forms in aphids: theoretical and experimental evidence for adaptive ‘coin-flipping’ plasticity. The American Naturalist, 163, E112E125.CrossRefGoogle ScholarPubMed
Hamelin, F.M., Bisson, A., Desprez-Loustau, M.L., Fabre, F. & Mailleret, L. (2016) Temporal niche differentiation of parasites sharing the same plant host: oak powdery mildew as a case study. Ecosphere, 7, e01517.Google Scholar
Hamelin, F.M., Castel, M., Poggi, S., Andrivon, D. & Mailleret, L. (2011) Seasonality and the evolutionary divergence of plant parasites. Ecology, 92, 21592166.Google Scholar
Huot, B., Yao, J., Montgomery, B.L. & He, S.Y. (2014) Growth–defense tradeoffs in plants: a balancing act to optimize fitness. Molecular Plant, 7, 12671287.Google Scholar
Jakuschkin, B., Fievet, V., Schwaller, L., et al. (2016) Deciphering the pathobiome: intra- and interkingdom interactions involving the pathogen Erysiphe alphitoides. Microbial Ecology, 72, 870880.Google Scholar
Jarosz, A.M. & Davelos, A.L. (1995) Effects of disease in wild plant populations and the evolution of pathogen aggressiveness. New Phytologist, 129, 371387.Google Scholar
Jeger, M.J. (2000) Theory and plant epidemiology. Plant Pathology, 49, 651658.Google Scholar
Jousimo, J., Tack, A.J., Ovaskainen, O., et al. (2014) Ecological and evolutionary effects of fragmentation on infectious disease dynamics. Science, 344(6189), 12891293.CrossRefGoogle ScholarPubMed
Keeling, M.J. & Rohani, P. (2008) Modeling Infectious Diseases in Humans and Animals. Princeton, NJ: Princeton University Press.Google Scholar
Kerling, L.C.P. (1966) The hibernation of the oak mildew. Plant Biology, 15, 7683.Google Scholar
Kermack, W.O. & McKendrick, A.G. (1927) A contribution to the mathematical theory of epidemics. Proceedings of the Royal Society of London A, 115, 700721.Google Scholar
Kisdi, E. (2012) F1000 Prime Recommendation of Hamelin FM et al., Ecology 2011, 92(12),2159–66. F1000 Prime.Google Scholar
Kiss, L., Russell, J.C., Szentiványi, O., Xu, X. & Jeffries, P. (2004) Biology and biocontrol potential of Ampelomyces mycoparasites, natural antagonists of powdery mildew fungi. Biocontrol Science and Technology, 14, 635651.CrossRefGoogle Scholar
Lenski, R.E. & May, R.M. (1994) The evolution of virulence in parasites and pathogens: reconciliation between two competing hypotheses. Journal of Theoretical Biology, 169, 253265.Google Scholar
Limkaisang, S., Cunnington, J.H, Wui, L.K., et al. (2006) Molecular phylogenetic analyses reveal a close relationship between powdery mildew fungi on some tropical trees and Erysiphe alphitoides, an oak powdery mildew. Mycoscience, 47, 327335.Google Scholar
Lively, C.M., de Roode, J.C., Duffy, M.A., Graham, A.L. & Koskella, B. (2014) Interesting open questions in disease ecology and evolution. The American Naturalist, 184(S1), S1S8.Google Scholar
Liyanage, A.D.S. & Royle, D.J. (1976) Overwintering of Sphaerotheca humuli, the cause of hop powdery mildew. Annals of Applied Biology, 83, 381394.Google Scholar
Loreau, M. (1992) Time scale of resource dynamics and coexistence through time partitioning. Theoretical Population Biology, 41, 401412.Google Scholar
Loreau, M. & Hector, A. (2001) Partitioning selection and complementarity in biodiversity experiments. Nature, 412(6842), 7276.Google Scholar
Madden, L.V., Hughes, G. & Bosch, F. (2007) The Study of Plant Disease Epidemics. St Paul, MN: American Phytopathological Society (APS Press).Google Scholar
Mailleret, L., Castel, M., Montarry, J. & Hamelin, F.M. (2012) From elaborate to compact seasonal plant epidemic models and back: is competitive exclusion in the details? Theoretical Ecology, 5, 311324.Google Scholar
Mailleret, L. & Lemesle, V. (2009) A note on semi-discrete modelling in the life sciences. Philosophical Transactions of the Royal Society of London A, 367, 47794799.Google ScholarPubMed
Marcais, B. & Desprez-Loustau, M.L. (2014) European oak powdery mildew: impact on trees, effects of environmental factors, and potential effects of climate change. Annals of Forest Science, 71, 633642.Google Scholar
Marcais, B., Kavkova, M. & Desprez-Loustau, M.L. (2009) Phenotypic variation in the phenology of ascospore production between European populations of oak powdery mildew. Annals of Forest Science, 66, 814.CrossRefGoogle Scholar
Marçais, B., Piou, D., Dezette, D. & Desprez-Loustau, M.L. (2017) Can oak powdery mildew severity be explained by indirect effects of climate on the composition of the Erysiphe pathogenic complex? Phytopathology, 107, 570579.Google Scholar
Menzel, A. (2000). Trends in phenological phases in Europe between 1951 and 1996. International Journal of Biometeorology, 44(2), 7681.Google Scholar
Montarry, J., Cartolaro, P., Delmotte, F., Jolivet, J. & Willocquet, L. (2008) Genetic structure and aggressiveness of Erysiphe necator populations during grapevine powdery mildew epidemics. Applied and Environmental Microbiology, 74, 63276332.Google Scholar
Mordecai, E.A. (2011) Pathogen impacts on plant communities: unifying theory, concepts, and empirical work. Ecological Monographs, 81, 429441.Google Scholar
Mougou, A., Dutech, C.C. & Desprez-Loustau, M.L. (2008) New insights into the identity and origin of the causal agent of oak powdery mildew in Europe. Forest Pathology, 38, 275287.Google Scholar
Mougou-Hamdane, A., Giresse, X., Dutech, C.C. & Desprez Loustau, M.L. (2010) Spatial distribution of lineages of oak powdery mildew fungi in France, using quick molecular detection methods. Annals of Forest Science, 67, 212.Google Scholar
Newcombe, G. (1998) A review of exapted resistance to diseases of Populus. European Journal of Forest Pathology, 28, 209216.CrossRefGoogle Scholar
Pasco, C., Montarry, J., Marquer, B. & Andrivon, D. (2016) And the nasty ones lose in the end: foliar pathogenicity trades off with asexual transmission in the Irish famine pathogen Phytophthora infestans. New Phytologist, 209, 334342.Google Scholar
Pautasso, M., Aas, G., Queloz, V. & Holdenrieder, O. (2013) European ash (Fraxinus excelsior) dieback – a conservation biology challenge. Biological Conservation, 158, 3749.Google Scholar
Pautasso, M., Holdenrieder, O. & Stenlid, J. (2005) Susceptibility to fungal pathogens of forests differing in tree diversity. In: Scherer-Lorenzen, M., Körner, C. & Schulze, E.-D. (eds.), Forest Diversity and Function (pp. 263289). Berlin: Springer.CrossRefGoogle Scholar
Pearson, R.C. & Gadoury, D.M. (1987) Cleistothecia, the source of primary inoculum for grape powdery mildew in New York. Phytopathology, 77, 15091514.Google Scholar
Penczykowski, R.M., Walker, E., Soubeyrand, S. & Laine, A.L. (2015) Linking winter conditions to regional disease dynamics in a wild plant–pathogen metapopulation. New Phytologist, 205, 11421152.CrossRefGoogle Scholar
Piepenbring, M., Hofmann, T.A., Kirschner, R., et al. (2011) Diversity patterns of Neotropical plant parasitic microfungi. Ecotropica, 17, 2740.Google Scholar
Plomion, C., Aury, J.M., Amselem, J., et al. (2018) Oak genome reveals facets of long lifespan. Nature Plants, 4, 440.CrossRefGoogle ScholarPubMed
Robinson, R.A. (1976) Plant Pathosystems. Berlin: Springer.Google Scholar
Roslin, T., Laine, A.-L. & Gripenberg, S. (2007) Spatial population structure in an obligate plant pathogen colonizing oak Quercus robur. Functional Ecology, 21, 11681177.Google Scholar
Roy, B.A. & Kirchner, J.W. (2000) Evolutionary dynamics of pathogen resistance and tolerance. Evolution, 54, 5163.Google Scholar
Sacristan, S. & Garcia-Arenal, F. (2008) The evolution of virulence and pathogenicity in plant pathogen populations. Molecular Plant Pathology, 9, 369384.Google Scholar
Schoch, C.L., Seifert, K.A., Huhndorf, S., et al. (2012) Nuclear ribosomal internal transcribed spacer (ITS) region as a universal DNA barcode marker for Fungi. Proceedings of the National Academy of Sciences of the United States of America, 109, 62416246.Google Scholar
Segarra, J., Jeger, M.J. & Van den Bosch, F. (2001) Epidemic dynamics and patterns of plant diseases. Phytopathology, 91, 10011010.CrossRefGoogle ScholarPubMed
Soularue, J.P. & Kremer, A. (2012) Assortative mating and gene flow generate clinal phenological variation in trees. BMC Evolutionary Biology, 12, 79.Google Scholar
Sparks, T.H. & Carey, P.D. (1995) The responses of species to climate over two centuries: an analysis of the Marsham phenological record, 1736–1947. Journal of Ecology, 83, 321.Google Scholar
Sparks, T.H., Carey, P.D. & Combes, J. (1997) First leafing dates of trees in Surrey between 1947 and 1996. The London Naturalist, 76, 1520.Google Scholar
Spotts, R.A. & Chen, P.M. (1984) Cold hardiness and temperature responses of healthy and mildew-infected terminal buds of apple during dormancy. Phytopathology, 74, 542544.Google Scholar
Stukenbrock, E.H. & McDonald, B.A. (2008) The origins of plant pathogens in agro-ecosystems. Annual Review of Phytopathology, 46, 75100.Google Scholar
Susi, H., Barrès, B., Vale, P.F. & Laine, A.L. (2015) Co-infection alters population dynamics of infectious disease. Nature Communications, 6, 5975.Google Scholar
Tack, A.J. & Laine, A.L. (2014) Ecological and evolutionary implications of spatial heterogeneity during the off‐season for a wild plant pathogen. New Phytologist, 202, 297308.Google Scholar
Takamatsu, S. (2013) Origin and evolution of the powdery mildews (Ascomycota, Erysiphales). Mycoscience, 54, 7586.CrossRefGoogle Scholar
Takamatsu, S., Braun, U., Limkaisang, S., et al. (2007) Phylogeny and taxonomy of the oak powdery mildew Erysiphe alphitoides sensu lato. Mycological Research, 111, 809826.Google Scholar
Takamatsu, S., Ito, H., Shiroya, Y., Kiss, L. & Heluta, V. (2015) First comprehensive phylogenetic analysis of the genus Erysiphe (Erysiphales, Erysiphaceae) I. The Microsphaera lineage. Mycologia, 107, 475489.Google Scholar
Tedersoo, L., Bahram, M., Põlme, S., et al. (2014) Global diversity and geography of soil fungi. Science, 346(6213), 1256688.Google Scholar
Tian, D., Traw, M.B., Chen, J. Q., Kreitman, M. & Bergelson, J. (2003) Fitness costs of R-gene-mediated resistance in Arabidopsis thaliana. Nature, 423(6935), 7477.Google Scholar
Tollenaere, C., Susi, H. & Laine, A.-L. (2016) Evolutionary and epidemiological implications of multiple infection in plants. Trends in Plant Science, 21, 8090.CrossRefGoogle ScholarPubMed
van den Berg, F., Bacaer, N., Metz, J.A.J., Lannou, C. & van den Bosch, F. (2011) Periodic host absence can select for both higher or lower parasite transmission rates. Evolutionary Ecology, 25, 121137.Google Scholar
Verdú, M. & Climent, J. (2007) Evolutionary correlations of polycyclic shoot growth in Acer (Sapindaceae). American Journal of Botany, 94, 13161320.Google Scholar
Viennot-Bourgin, G. (1968) Note sur des Erysiphacees. Bulletin Trimestriel de la Societe Mycologique de France, 84, 117118.Google Scholar
Viney, R. (1970) L’oïdium du Chêne: incident léger ou désastre. Revue Forestière Française, 22, 365369.CrossRefGoogle Scholar
Vitasse, Y. (2013) Ontogenic changes rather than difference in temperature cause understory trees to leaf out earlier. New Phytologist, 198, 149155.CrossRefGoogle ScholarPubMed
Vitasse, Y., François, C., Delpierre, N., et al. (2011) Assessing the effects of climate change on the phenology of European temperate trees. Agricultural and Forest Meteorology, 151, 969980.Google Scholar
Vuillemin, P. (1910a) Le déclin de la maladie du blanc du chêne. Bulletin de l’Office forestier du Centre et de l’Ouest, 347350.Google Scholar
Vuillemin, P. (1910b) Un ennemi naturel de l’Oïdium du Chêne. Bulletin de la Société Mycologique de France, 26.Google Scholar
Weis, A.E., Simms, E.L. & Hochberg, M.E. (2000) Will plant vigor and tolerance be genetically correlated? Effects of intrinsic growth rate and self-limitation on regrowth. Evolutionary Ecology, 14, 331352.Google Scholar
Woodward, R.C., Waldie, J.S.L. & Steven, H.M. (1929) Oak mildew and its control in forest nurseries. Forestry, 3, 3856.CrossRefGoogle Scholar
Zandt, P.A.V. & Mopper, S. (1998) A meta-analysis of adaptive deme formation in phytophagous insect populations. The American Naturalist, 152, 595604.Google Scholar

References

Auld, S.K., Hall, S.R., Ochs, J.H., Sebastian, M. & Duffy, M.A. (2014) Predators and patterns of within-host growth can mediate both among-host competition and evolution of transmission potential of parasites. American Naturalist, 18, S77S90.Google Scholar
Bertram, C.R., Pinkowski, M., Hall, S.R., Duffy, M.A. & Cáceres, C.E. (2013) Trait-mediated indirect effects, predators, and disease: test of a size-based model. Oecologia, 173, 10231032.Google Scholar
Bidegain, G., Powell, E.N., Klinck, J.M., Ben-Horin, T. & Hofmann, E.E. (2016) Marine infectious disease dynamics and outbreak thresholds: contact transmission, pandemic infection, and the potential role of filter feeders. Ecosphere, 7, e01286.CrossRefGoogle Scholar
Brooks, J.L. & Dodson, S.I. (1965) Predation, body size, and composition of plankton. Science, 150, 2835.Google Scholar
Buck, J., Truong, L. & Blaustein, A. (2011) Predation by zooplankton on Batrachochytrium dendrobatidis: biological control of the deadly amphibian chytrid fungus? Biodiversity and Conservation, 20, 35493553.Google Scholar
Byers, J.E., Malek, A.J., Quevillon, L.E., Altman, I. & Keogh, C.L. (2015) Opposing selective pressures decouple pattern and process of parasitic infection over small spatial scale. Oikos, 124, 15111519.Google Scholar
Cáceres, C.E., Hall, S.R., Duffy, M.A., Tessier, A.J., Helmle, C. & MacIntyre, S. (2006) Physical structure of lakes constrains epidemics in Daphnia populations. Ecology, 87, 14381444.Google Scholar
Cáceres, C.E., Knight, C.J. & Hall, S.R. (2009) Predator spreaders: predation can enhance parasite success in a planktonic host–parasite system. Ecology, 90, 28502858.Google Scholar
Cáceres, C.E., Tessier, A.J., Duffy, M.A. & Hall, S.R. (2014) Disease in freshwater zooplankton: what have we learned and where are we going? Journal of Plankton Research, 36, 326333.Google Scholar
Choisy, M. & Rohani, P. (2006) Harvesting can increase severity of wildlife disease epidemics. Proceedings of the Royal Society of London Series B, 273, 20252034.Google Scholar
Civitello, D.J., Pearsall, S., Duffy, M.A. & Hall, S.R. (2013) Parasite consumption and host interference can inhibit disease spread in dense populations. Ecology Letters, 16, 626634.Google Scholar
Civitello, D.J., Penczykowski, R.M., Smith, A.N., et al. (2015) Resources, key traits, and the size of fungal epidemics in Daphnia populations. Journal of Animal Ecology, 84, 10101017.Google Scholar
Coors, A. & De Meester, L. (2011) Fitness and virulence of a bacterial endoparasite in an environmentally stressed crustacean host. Parasitology, 138, 122131.Google Scholar
Cressler, C.E., Nelson, W.A., Day, T. & McCauley, E. (2014) Disentangling the interaction among host resources, the immune system and pathogens. Ecology Letters, 17, 284293.Google Scholar
de Roos, A.M. & Persson, L. (2013) Population and Community Ecology of Ontogenetic Development. Princeton, NJ: Princeton University Press.Google Scholar
Decaestecker, E., De Meester, L. & Ebert, D. (2002) In deep trouble: habitat selection constrained by multiple enemies in zooplankton. Proceedings of the National Academy of Science of the United States of America, 99, 54815485.Google Scholar
Department for Environment Food and Rural Affairs (2016) Summary of badger control monitoring during 2016. www.gov.uk/government/uploads/system/uploads/attachment_data/file/578436/summary-badger-control-monitoring-2016.pdfGoogle Scholar
Donnelly, C.A., Woodroffe, R., Cox, D.R., et al. (2003) Impact of localized badger culling on tuberculosis incidence in British cattle. Nature, 426, 834837.Google Scholar
Duffy, M.A. (2007) Selective predation, parasitism, and trophic cascades in a bluegill–Daphnia–parasite system. Oecologia, 153, 453460.Google Scholar
Duffy, M.A. (2009) Staying alive: the post-consumption fate of parasite spores and its implications for disease dynamics. Limnology and Oceanography, 54, 770773.Google Scholar
Duffy, M.A., Cáceres, C.E., Hall, S.R., Tessier, A.J. & Ives, A.R. (2010) Temporal, spatial, and between-host comparisons of patterns of parasitism in lake zooplankton. Ecology, 91, 33223331.Google Scholar
Duffy, M.A. & Hall, S.R. (2008) Selective predation and rapid evolution can jointly dampen effects of virulent parasites on Daphnia populations. American Naturalist, 171, 499510.Google Scholar
Duffy, M.A., Hall, S.R., Cáceres, C.E. & Ives, A.R. (2009) Rapid evolution, seasonality and the termination of parasite epidemics. Ecology, 90, 14411448.Google Scholar
Duffy, M.A., Hall, S.R., Tessier, A.J. & Huebner, M. (2005) Selective predators and their parasitized prey: are epidemics in zooplankton under top-down control? Limnology and Oceanography, 50, 412420.Google Scholar
Duffy, M.A., Housley, J.M., Penczykowski, R.M., Cáceres, C.E. & Hall, S.R. (2011) Unhealthy herds: indirect effects of predators enhance two drivers of disease spread. Functional Ecology, 25, 945953.Google Scholar
Duffy, M.A., James, T.Y. & Longworth, A. (2015) Ecology, virulence, and phylogeny of Blastulidium paedophthorum, a widespread brood parasite of Daphnia spp. Applied & Environmental Microbiology, 81, 54865496.Google Scholar
Duffy, M.A., Ochs, J.H., Penczykowski, R.M., et al. (2012) Ecological context influences epidemic size and parasite-mediated selection. Science, 335, 16361638.Google Scholar
Elser, M.M., Vonende, C.N., Sorrano, P. & Carpenter, S.R. (1987) Chaoborus populations: response to food web manipulation and potential effects on zooplankton communities. Canadian Journal of Zoology, 65, 28462852.Google Scholar
González, M.J. & Tessier, A.J. (1997) Habitat segregation and interactive effects of multiple predators on a prey assemblage. Freshwater Biology, 38, 179191.Google Scholar
Goren, L. & Ben-Ami, F. (2017) To eat or not to eat infected food: a bug’s dilemma.Hydrobiologia, 798, 2532.Google Scholar
Groner, M.L. & Relyea, R.A. (2015) Predators reduce Batrachochytrium dendrobatidis infection loads in their prey. Freshwater Biology, 60, 16991704.Google Scholar
Hall, S.R., Becker, C.R., Simonis, J.L., et al. (2009) Friendly competition: evidence for a dilution effect among competitors in a planktonic host–parasite system. Ecology, 90, 791801.Google Scholar
Hall, S.R., Duffy, M.A. & Cáceres, C.E. (2005) Selective predation and productivity jointly drive complex behavior in host–parasite systems. American Naturalist, 165, 7081.Google Scholar
Hall, S.R., Sivars-Becker, L., Becker, C., et al. (2007) Eating yourself sick: transmission of disease as a function of feeding biology of hosts. Ecology Letters, 10, 207218.Google Scholar
Hall, S.R., Smyth, R., Becker, C.R., et al. (2010) Why are Daphnia in some lakes sicker? Disease ecology, habitat structure, and the plankton. BioScience, 60, 363375.CrossRefGoogle Scholar
Hall, S.R., Tessier, A.J., Duffy, M.A., Huebner, M. & Cáceres, C.E. (2006) Warmer does not have to mean sicker: temperature and predators can jointly drive timing of epidemics. Ecology, 87, 16841695.CrossRefGoogle Scholar
Harvell, D., Aronson, R., Baron, N., et al. (2004) The rising tide of ocean diseases: unsolved problems and research priorities. Frontiers in Ecology and the Environment, 2, 375382.Google Scholar
Hesse, O., Engelbrecht, W., Laforsch, C. & Wolinska, J. (2012) Fighting parasites and predators: how to deal with multiple threats? BMC Ecology, 12, 12.Google Scholar
Hite, J.L., Bosch, J., Fernández-Beaskoetxea, S., Medina, D. & Hall, S.R. (2016) Joint effects of habitat, zooplankton, host stage structure and diversity on amphibian chytrid. Proceedings of the Royal Society of London B, 283, 20160832.Google Scholar
Holt, R.D. & Roy, M. (2007) Predation can increase the prevalence of infectious disease. American Naturalist, 169, 690699.Google Scholar
Hudson, P.J. (1986) The effect of a parasitic nematode on the breeding production of red grouse. Journal of Animal Ecology, 55, 8592.Google Scholar
Hudson, P.J., Dobson, A.P. & Newborn, D. (1992) Do parasites make prey vulnerable to predation? Red grouse and parasites. Journal of Animal Ecology, 61, 681692.Google Scholar
Johnson, A. & Brunner, J. (2014) Persistence of an amphibian ranavirus in aquatic communities. Diseases of Aquatic Organisms, 111, 129138.Google Scholar
Johnson, P.T.J., Dobson, A., Lafferty, K.D., et al. (2010) When parasites become prey: ecological and epidemiological significance of eating parasites. Trends in Ecology & Evolution, 25, 362371.Google Scholar
Johnson, P.T.J., Stanton, D.E., Preu, E.R., Forshay, K.J. & Carpenter, S.R. (2006) Dining on disease: how interactions between parasite infection and environmental conditions affect host predation risk. Ecology, 87, 19731980.Google Scholar
Kagami, M., Van Donk, E., de Bruin, A., Rijkeboer, M. & Ibelings, B.W. (2004) Daphnia can protect diatoms from fungal parasitism. Limnology and Oceanography, 49, 680685.Google Scholar
Keeling, M.J. & Rohani, P. (2008) Modeling Infectious Diseases in Humans and Animals. Princeton, NJ: Princeton University Press.Google Scholar
Keesing, F., Belden, L.K., Daszak, P., et al. (2010) Impacts of biodiversity on the emergence and transmission of infectious diseases. Nature, 468, 647652.Google Scholar
Keymer, A., Crompton, D.W.T. & Walters, D.E. (1983) Parasite population biology and host nutrition – dietary fructose and Moniliformis (Acanthocephala). Parasitology, 87, 265278.Google Scholar
Kistler, R.A. (1985) Host-age structure and parasitism in a laboratory system of two hymenopterous parasitoids and larvae of Zabrotes subfasciatus (Coleoptera, Bruchidae). Environmental Entomology, 14, 507511.Google Scholar
Lafferty, K.D. (2004) Fishing for lobsters indirectly increases epidemics in sea urchins. Ecological Applications, 14, 15661573.Google Scholar
Lafferty, K.D., Harvell, C.D., Conrad, J.M., et al. (2015) Infectious diseases affect marine fisheries and aquaculture economics. Annual Review of Marine Science, 7, 471496.Google Scholar
Lass, S. & Bittner, K. (2002) Facing multiple enemies: parasitised hosts respond to predator kairomones. Oecologia, 132, 344349.Google Scholar
Levi, T., Kilpatrick, A.M., Mangel, M. & Wilmers, C.C. (2012) Deer, predators, and the emergence of Lyme disease. Proceedings of the National Academy of Sciences of the United States of America, 109, 10,94210,947.CrossRefGoogle ScholarPubMed
Li, J., Kolivras, K.N., Hong, Y., et al. (2014) Spatial and temporal emergence pattern of Lyme disease in Virginia. The American Journal of Tropical Medicine and Hygiene, 91, 11661172.Google Scholar
Lindeque, P.M. & Turnbull, P.C.B. (1994) Ecology and epidemiology of anthrax in the Etosha National Park, Namibia. Onderstepoort Journal of Veterinary Research, 61, 7183.Google Scholar
Michael, E. & Bundy, D.A.P. (1992) Nutrition, immunity and helminth infection: effect of dietary protein on the dynamics of the primary antibody response to Trichuris muris (Nematoda) in CBA/Ca mice. Parasite Immunology, 14, 169183.Google Scholar
Mittelbach, G.G. (1981) Patterns of invertebrate size and abundance in aquatic habitats. Canadian Journal of Fisheries and Aquatic Sciences, 38, 896904.Google Scholar
Morters, M.K., Restif, O., Hampson, K., et al. (2013) Evidence-based control of canine rabies: a critical review of population density reduction. Journal of Animal Ecology, 82, 614.Google Scholar
Navarro, C., de Lope, F., Marzal, A. & Møller, A.P. (2004) Predation risk, host immune response, and parasitism. Behavioral Ecology, 15, 629635.Google Scholar
Orlofske, S.A., Jadin, R.C., Preston, D.L. & Johnson, P.T.J. (2012) Parasite transmission in complex communities: predators and alternative hosts alter pathogenic infections in amphibians. Ecology, 93, 12471253.Google Scholar
Ostfeld, R.S. & Holt, R.D. (2004) Are predators good for your health? Evaluating evidence for top-down regulation of zoonotic disease reservoirs. Frontiers in Ecology and the Environment, 2, 1320.Google Scholar
Ostfeld, R.S. & Keesing, F. (2000) Biodiversity and disease risk: the case of Lyme disease [Biodiversidad y Riesgo de Enfermedades: El Caso de la Enfermedad de Lyme]. Conservation Biology, 14, 722728.Google Scholar
Pace, M.L., Cole, J.J., Carpenter, S.R. & Kitchell, J.F. (1999) Trophic cascades revealed in diverse ecosystems. Trends in Ecology and Evolution, 14, 483488.CrossRefGoogle ScholarPubMed
Packer, C., Holt, R.D., Hudson, P.J., Lafferty, K.D. & Dobson, A.P. (2003) Keeping the herds healthy and alert: implications of predator control for infectious disease. Ecology Letters, 6, 797802.Google Scholar
Pastorok, R.A. (1981) Prey vulnerability and size selection by Chaoborus larvae. Ecology, 62, 13111324.Google Scholar
Penczykowski, R.M., Hall, S.R., Civitello, D.J. & Duffy, M.A. (2014) Habitat structure and ecological drivers of disease. Limnology and Oceanography, 59, 340348.Google Scholar
Ramirez, R.A. & Snyder, W.E. (2009) Scared sick? Predator–pathogen facilitation enhances exploitation of a shared resource. Ecology, 90, 28322839.Google Scholar
Rapti, Z. & Cáceres, C.E. (2016) Effects of intrinsic and extrinsic host mortality on disease spread. Bulletin of Mathematical Biology, 78, 235253.Google Scholar
Rohr, J.R., Civitello, D.J., Crumrine, P.W., et al. (2015) Predator diversity, intraguild predation, and indirect effects drive parasite transmission. Proceedings of the National Academy of Sciences of the United States of America, 112, 30083013.Google Scholar
Salkeld, D.J., Padgett, K.A. & Jones, J.H. (2013) A meta-analysis suggesting that the relationship between biodiversity and risk of zoonotic pathogen transmission is idiosyncratic. Ecology Letters, 16, 679686.Google Scholar
Searle, C.L., Mendelson, J.R., Green, L.E. & Duffy, M.A. (2013) Daphnia predation on the amphibian chytrid fungus and its impacts on disease risk in tadpoles. Ecology and Evolution, 3, 41294138.Google Scholar
Smith, V. (2007) Host resource supplies influence the dynamics and outcome of infectious disease. Integrative and Comparative Biology, 47, 310316.Google Scholar
Snyder, W.E. & Ives, A.R. (2001) Generalist predators disrupt biological control by a specialist parasitoid. Ecology, 82, 705716.Google Scholar
Spitze, K. (1985) Functional response of an ambush predator: Chaoborus americanus predation on Daphnia pulex. Ecology, 66, 938949.Google Scholar
Strauss, A.T., Civitello, D.J., Cáceres, C.E. & Hall, S.R. (2015) Success, failure and ambiguity of the dilution effect among competitors. Ecology Letters, 18, 916926.Google Scholar
Strauss, A.T., Shocket, M.S., Civitello, D.J., et al. (2016) Habitat, predators, and hosts regulate disease in Daphnia through direct and indirect pathways. Ecological Monographs, 86, 393411.Google Scholar
Tessier, A.J. & Woodruff, P. (2002) Cryptic trophic cascade along a gradient of lake size. Ecology, 83, 12631270.Google Scholar
Thomas, S.H., Bertram, C., van Rensburg, K., Caceres, C.E. & Duffy, M.A. (2011) Spatiotemporal dynamics of free-living stages of a bacterial parasite of zooplankton. Aquatic Microbial Ecology, 63, 265272.Google Scholar
Turney, S., Gonzalez, A. & Millien, V. (2014) The negative relationship between mammal host diversity and Lyme disease incidence strengthens through time. Ecology, 95, 32443250.Google Scholar
Valois, A.E. & Burns, C.W. (2016) Parasites as prey: Daphnia reduce transmission success of an oomycete brood parasite in the calanoid copepod Boeckella. Journal of Plankton Research, 38, 12811288.Google Scholar
Williamson, C.E., Overholt, E.P., Pilla, R.M., et al. (2015) Ecological consequences of long-term browning in lakes. Scientific Reports, 5, 18666.Google Scholar
Wilson, K. & Cotter, S.C. (2009) Density-dependent prophylaxis in insects. In: Whitman, D.W. & Ananthakrishnan, T.N. (eds.), Phenotypic Plasticity of Insects: Mechanisms and Consequences (pp. 137176). Boca Raton, FL: CRC Press.Google Scholar
Wood, C.L. & Lafferty, K.D. (2013) Biodiversity and disease: a synthesis of ecological perspectives on Lyme disease transmission. Trends in Ecology & Evolution, 28, 239247.Google Scholar

References

Abrams, P.A. (1995) Implications of dynamically variable traits for identifying, classifying, and measuring direct and indirect effects in ecological communities. American Naturalist, 146, 112134.Google Scholar
Ackery, P.R. & Vane-Wright, R.I. (1984) Milkweed Butterflies: Their Cladistics and Biology. Ithaca, NY: Cornell University Press.Google Scholar
Adamo, S. & Parsons, N. (2006) The emergency life-history stage and immunity in the cricket, Gryllus texensis. Animal Behaviour, 72, 235244.Google Scholar
Adamo, S., Roberts, J., Easy, R. & Ross, N. (2008) Competition between immune function and lipid transport for the protein apolipophorin III leads to stress-induced immunosuppression in crickets. Journal of Experimental Biology, 211, 531538.Google Scholar
Agrawal, A.A., Petschenka, G., Bingham, R.A., Weber, M.G. & Rasmann, S. (2012) Toxic cardenolides: chemical ecology and coevolution of specialized plant–herbivore interactions. New Phytologist, 194, 2845.Google Scholar
Alizon, S., Hurford, A., Mideo, N. & Van Baalen, M. (2009) Virulence evolution and the trade-off hypothesis: history, current state of affairs and the future. Journal of Evolutionary Biology, 22, 245259.Google Scholar
Alizon, S. & Michalakis, Y. (2015) Adaptive virulence evolution: the good old fitness-based approach. Trends in Ecology & Evolution, 30, 248254.Google Scholar
Altizer, S., Bartel, R. & Han, B.A. (2011) Animal migration and infectious disease risk. Science, 331, 296302.Google Scholar
Altizer, S. & de Roode, J.C. (2015) Monarchs and their debilitating parasites: immunity, migration, and medicinal plant use. In: Oberhauser, K.O., Altizer, S. & Nail, K. (eds.), Monarchs in a Changing World: Biology and Conservation of an Iconic Insect (pp. 8393). Ithaca, NY:Cornell University Press.Google Scholar
Altizer, S., Hobson, K., Davis, A., de Roode, J. & Wassenaar, L. (2015) Do healthy monarchs migrate farther? Tracking natal origins of parasitized vs. uninfected monarch butterflies overwintering in Mexico. PLoS ONE, 10, e0141371.Google Scholar
Altizer, S., Ostfeld, R.S., Johnson, P.T.J., Kutz, S. & Harvell, C.D. (2013) Climate change and infectious diseases: from evidence to a predictive framework. Science, 341, 514519.CrossRefGoogle ScholarPubMed
Altizer, S.M. (2001) Migratory behaviour and host–parasite co-evolution in natural populations of monarch butterflies infected with a protozoan parasite. Evolutionary Ecology Research, 3, 611632.Google Scholar
Altizer, S.M. & Oberhauser, K.S. (1999) Effects of the protozoan parasite Ophryocystis elektroscirrha on the fitness of monarch butterflies (Danaus plexippus). Journal of Invertebrate Pathology, 74, 7688.Google Scholar
Altizer, S.M., Oberhauser, K.S. & Brower, L.P. (2000) Associations between host migration and the prevalence of a protozoan parasite in natural populations of adult monarch butterflies. Ecological Entomology, 25, 125139.Google Scholar
Altizer, S.M., Oberhauser, K.S. & Geurts, K.A. (2004) Transmission of the protozoan parasite, Ophryocystis elektroscirrha, in monarch butterfly populations: implications for prevalence and population-level impacts. In: Oberhauser, K.S. & Solensky, M. (eds.), The Monarch Butterfly: Biology and Conservation (pp. 203218). Ithaca, NY: Cornell University Press.Google Scholar
Anderson, R.M. & May, R.M. (1982) Coevolution of hosts and parasites. Parasitology, 85, 411426.Google Scholar
Anderson, R.M. & May, R.M. (1991) Infectious Diseases of Humans – Dynamics and Control. Oxford: Oxford University Press.Google Scholar
Andrews, H. (2015) Changes in water availability and variability affect plant defenses and herbivore responses in grassland forbs. Master’s thesis, University of Michigan.Google Scholar
Antia, R., Levin, B.R. & May, R.M. (1994) Within-host population dynamics and the evolution and maintenance of microparasite virulence. American Naturalist, 144, 457472.Google Scholar
Barriga, P.A., Sternberg, E.D., Lefèvre, T., de Roode, J.C. & Altizer, S. (2016) Occurrence and host specificity of a neogregarine protozoan in four milkweed butterfly hosts (Danaus spp.). Journal of Invertebrate Pathology, 140, 7582.Google Scholar
Bartel, R.A., Oberhauser, K.S., de Roode, J.C. & Altizer, S. (2011) Monarch butterfly migration and parasite transmission in eastern North America. Ecology, 92, 342351.Google Scholar
Batalden, R.V. & Oberhauser, K.S. (2015) Potential changes in eastern North American monarch migration in response to an introduced milkweed, Asclepias curassavica. In: Oberhauser, K.S., Nail, K.R. & Altizer, S. (eds.), Monarchs in a Changing World: Biology and Conservation of an Iconic Butterfly. Ithaca, NY: Cornell University Press.Google Scholar
Baucom, R.S. & de Roode, J.C. (2011) Ecological immunology and tolerance in plants and animals. Functional Ecology, 25, 1828.Google Scholar
Bauer, S. & Hoye, B.J. (2014) Migratory animals couple biodiversity and ecosystem functioning worldwide. Science, 344, 1242552.Google Scholar
Bowlin, M.S., Bisson, I.A., Shamoun-Baranes, J., et al. (2010) Grand challenges in migration biology. Integrative and Comparative Biology, 50, 261279.Google Scholar
Bradley, C.A. & Altizer, S. (2005) Parasites hinder monarch butterfly flight: implications for disease spread in migratory hosts. Ecology Letters, 8, 290300.Google Scholar
Bremermann, H.J. & Pickering, J. (1983) A game-theoretical model of parasite virulence. Journal of Theoretical Biology, 100, 411426.Google Scholar
Bremermann, H.J. & Thieme, H.R. (1989) A competitive exclusion principle for pathogen virulence. Journal of Mathematical Biology, 27, 179190.Google Scholar
Brower, L.P. (1995) Understanding and misunderstanding the migration of the monarch butterfly (Nymphalidae) in North America: 1857–1995. Journal of the Lepidopterists’ Society, 49, 304385.Google Scholar
Brower, L.P. & Fink, L.S. (1985) A natural toxic defense system – cardenolides in butterflies versus birds. Annals of the New York Academy of Sciences, 443, 171188.Google Scholar
Brower, L.P., Ryerson, W.N., Coppinger, L. & Glazier, S.C. (1968) Ecological chemistry and the palatability spectrum. Science, 161, 13491351.Google Scholar
Brower, L.P., Taylor, O.R., Williams, E.H., et al. (2012) Decline of monarch butterflies overwintering in Mexico: is the migratory phenomenon at risk? Insect Conservation and Diversity, 5, 95100.Google Scholar
Buehler, D.M., Tieleman, B.I. & Piersma, T. (2010) How do migratory species stay healthy over the annual cycle? A conceptual model for immune function and for resistance to disease. Integrative and Comparative Biology, 50, 346357.Google Scholar
Choisy, M. & de Roode, J.C. (2014) The ecology and evolution of animal medication: genetically fixed response versus phenotypic plasticity. American Naturalist, 184, S31S46.Google Scholar
Civitello, D.J., Penczykowski, R.M., Hite, J.L., Duffy, M.A. & Hall, S.R. (2013) Potassium stimulates fungal epidemics in Daphnia by increasing host and parasite reproduction. Ecology, 94, 380388.Google Scholar
Clough, D., Prykhodko, O. & Råberg, L. (2016) Effects of protein malnutrition on tolerance to helminth infection. Biology Letters, 12.Google Scholar
Cory, J.S. & Hoover, K. (2006) Plant-mediated effects in insect–pathogen interactions. Trends in Ecology and Evolution, 21, 278286.Google Scholar
Costello, M.J. (2009) How sea lice from salmon farms may cause wild salmonid declines in Europe and North America and be a threat to fishes elsewhere. Proceedings of the Royal Society of London B, 276, 33853394.Google Scholar
Cousineau, S.V. & Alizon, S. (2014) Parasite evolution in response to sex-based host heterogeneity in resistance and tolerance. Journal of Evolutionary Biology, 27, 27532766.Google Scholar
Couture, J.J., Serbin, S.P. & Townsend, P.A. (2015) Elevated temperature and periodic water stress alter growth and quality of common milkweed (Asclepias syriaca) and monarch (Danaus plexippus) larval performance. Arthropod–Plant Interactions, 9, 149161.Google Scholar
de Roode, J.C. & Altizer, S. (2010) Host–parasite genetic interactions and virulence–transmission relationships in natural populations of monarch butterflies. Evolution, 64, 502514.Google Scholar
de Roode, J.C., Chi, J., Rarick, R.M. & Altizer, S. (2009) Strength in numbers: high parasite burdens increase transmission of a protozoan parasite of monarch butterflies (Danaus plexippus). Oecologia, 161, 6775.Google Scholar
de Roode, J.C., Gold, L.R. & Altizer, S. (2007) Virulence determinants in a natural butterfly–parasite system. Parasitology, 134, 657668.Google Scholar
de Roode, J.C., Lefèvre, T. & Hunter, M.D. (2013) Self-medication in animals. Science, 340, 150151.Google Scholar
de Roode, J.C., Lopez Fernandez de Castillejo, C., Faits, T. & Alizon, S. (2011a) Virulence evolution in response to anti-infection resistance: toxic food plants can select for virulent parasites of monarch butterflies. Journal of Evolutionary Biology, 24, 712722.Google Scholar
de Roode, J.C., Pedersen, A.B., Hunter, M.D. & Altizer, S. (2008) Host plant species affects virulence in monarch butterfly parasites. Journal of Animal Ecology, 77, 120126.Google Scholar
de Roode, J.C., Rarick, R.M., Mongue, A.J., Gerardo, N.M. & Hunter, M.D. (2011b) Aphids indirectly increase virulence and transmission potential of a monarch butterfly parasite by reducing defensive chemistry of a shared food plant. Ecology Letters, 14, 453461.Google Scholar
de Roode, J.C., Yates, A.J. & Altizer, S. (2008) Virulence–transmission trade-offs and population divergence in virulence in a naturally occurring butterfly parasite. Proceedings of the National Academy of Sciences of the United States of America, 105, 74897494.Google Scholar
Dingle, H. (1996) Migration: The Biology of Life on the Move. Oxford: Oxford University Press.Google Scholar
Dobler, S., Dalla, S., Wagschal, V. & Agrawal, A.A. (2012) Community-wide convergent evolution in insect adaptation to toxic cardenolides by substitutions in the Na,K-ATPase. Proceedings of the National Academy of Sciences of the United States of America, 109, 13,04013,045.Google Scholar
Dwyer, G. & Elkinton, J.S. (1995) Host dispersal and the spatial spread of insect pathogens. Ecology, 76, 12621275.Google Scholar
Epstein, J.H., McKee, J., Shaw, P., et al. (2006) The Australian white ibis (Threskiornis molucca) as a reservoir of zoonotic and livestock pathogens. EcoHealth, 3, 290298.Google Scholar
Evans, K.L., Newton, J., Gaston, K.J., et al. (2012) Colonisation of urban environments is associated with reduced migratory behaviour, facilitating divergence from ancestral populations. Oikos, 121, 634640.Google Scholar
Felton, G.W., Duffey, S.S., Vail, P.V., Kaya, H.K. & Manning, J. (1987) Interaction of nuclear polyhedrosis virus with catechols: potential incompatability for host-plant resistence against noctuid larvae. Journal of Chemical Ecology, 13, 947957.Google Scholar
Flack, A., Fiedler, W., Blas, J., et al. (2016) Costs of migratory decisions: a comparison across eight white stork populations. Science Advances, 2, e1500931.Google Scholar
Folstad, I., Nilssen, A.C., Halvorsen, O. & Andersen, J. (1991) Parasite avoidance: the cause of post-calving migrations in Rangifer? Canadian Journal of Zoology, 69, 24232429.Google Scholar
Forbey, J.S. & Hunter, M.D. (2012) The herbivore’s prescription: a pharm-ecological perspective on host plant use by vertebrate and invertebrate herbivores. In: Iason, G.R., Dicke, M. & Hartley, S.E. (eds.),The Ecology of Plant Secondary Matabolites: From Genes to Global Processes (pp. 78100). Cambridge: Cambridge University Press.Google Scholar
Frank, S.A. (1996) Models of parasite virulence. Quarterly Review of Biology, 71, 3778.Google Scholar
Gandon, S., Mackinnon, M.J., Nee, S. & Read, A.F. (2001) Imperfect vaccines and the evolution of pathogen virulence. Nature, 414, 751756.Google Scholar
Gandon, S. & Michalakis, Y. (2000) Evolution of parasite virulence against qualitative or quantitative host resistance. Proceedings of the Royal Society of London B, 267, 985990.Google Scholar
Gilbert, N.I., Correia, R.A., Silva, J.P., et al. (2016) Are white storks addicted to junk food? Impacts of landfill use on the movement and behaviour of resident white storks (Ciconia ciconia) from a partially migratory population. Movement Ecology, 4, 7.Google Scholar
Gowler, C.D., Leon, K.E., Hunter, M.D. & de Roode, J.C. (2015) Secondary defense chemicals in milkweed reduce parasite infection in monarch butterflies, Danaus plexippus. Journal of Chemical Ecology, 41, 520523.Google Scholar
Graham, R.I., Grzywacz, D., Mushobozi, W.L. & Wilson, K. (2012) Wolbachia in a major African crop pest increases susceptibility to viral disease rather than protects. Ecology Letters, 15, 9931000.Google Scholar
Gustafsson, K.M., Agrawal, A.A., Lewenstein, B.V. & Wolf, S.A. (2015) The monarch butterfly through time and space: the social construction of an icon. Bioscience, 65, 612622.Google Scholar
Hall, R.J., Altizer, S. & Bartel, R.A. (2014) Greater migratory propensity in hosts lowers pathogen transmission and impacts. Journal of Animal Ecology, 83, 10681077.Google Scholar
Hegemann, A., Matson, K.D., Both, C. & Tieleman, B.I. (2012) Immune function in a free-living bird varies over the annual cycle, but seasonal patterns differ between years. Oecologia, 170, 605618.Google Scholar
Hoang, K.M., Tao, L., Hunter, M.D. & de Roode, J.C. (2017) Host diet affects the morphology of a butterfly parasite. Journal of Parasitology, 103, 228236.Google Scholar
Howard, E., Aschen, H. & Davis, A.K. (2010) Citizen science observations of monarch butterfly overwintering in the southern United States. Psyche: A Journal of Entomology, 2010, 689301.Google Scholar
Hsieh, H.Y., Liere, H., Soto, E.J. & Perfecto, I. (2012) Cascading trait-mediated interactions induced by ant pheromones. Ecology and Evolution, 2, 21812191.Google Scholar
Hunter, M.D. (2016) The Phytochemical Landscape. Linking Trophic Interactions and Nutrient Dynamics. Princeton, NJ: Princeton University Press.Google Scholar
Hunter, M.D., Malcolm, S.B. & Hartley, S.E. (1996) Population-level variation in plant secondary chemistry and the population biology of herbivores. Chemoecology, 7, 4556.Google Scholar
Hunter, M.D. & Schultz, J.C. (1993) Induced plant defenses breached? Phytochemical induction protects an herbivore from disease. Oecologia, 94, 195203.Google Scholar
Johns, S. & Shaw, A.K. (2016) Theoretical insight into three disease-related benefits of migration. Population Ecology, 58, 213221.Google Scholar
Johnson, P.T.J., de Roode, J.C. & Fenton, A. (2015) Why infectious disease research needs community ecology. Science, 349, 1259504.Google Scholar
Johnson, P.T.J., Preston, D.L., Hoverman, J.T. & Richgels, K.L.D. (2013) Biodiversity decreases disease through predictable changes in host community competence. Nature, 494, 230233.Google Scholar
Keating, S.T. & Yendol, W.G. (1987) Influence of selected host plants on gypsy moth (Lepidoptera, Lymantriidae) larval mortality caused by a baculovirus. Environmental Entomology, 16, 459462.Google Scholar
Krkošek, M., Ford, J.S., Morton, A., et al. (2007a) Declining wild salmon populations in relation to parasites from farm salmon. Science, 318, 17721775.Google Scholar
Krkošek, M., Gottesfeld, A., Proctor, B., et al. (2007b) Effects of host migration, diversity and aquaculture on sea lice threats to Pacific salmon populations. Proceedings of the Royal Society of London B, 274, 31413149.Google Scholar
Krkošek, M., Lewis, M.A. & Volpe, J.P. (2005) Transmission dynamics of parasitic sea lice from farm to wild salmon. Proceedings of the Royal Society of London B, 272, 689696.Google Scholar
Lank, D.B., Butler, R.W., Ireland, J. & Ydenberg, R.C. (2003) Effects of predation danger on migration strategies of sandpipers. Oikos, 103, 303319.Google Scholar
Lefèvre, T., Chiang, A., Kelavkar, M., et al. (2012) Behavioural resistance against a protozoan parasite in the monarch butterfly. Journal of Animal Ecology, 81, 7079.Google Scholar
Lefèvre, T., Oliver, L., Hunter, M.D. & de Roode, J.C. (2010) Evidence for trans-generational medication in nature. Ecology Letters, 13, 14851493.Google Scholar
Lefèvre, T., Williams, A.J. & de Roode, J.C. (2011) Genetic variation for resistance, but not tolerance, to a protozoan parasite in the monarch butterfly. Proceedings of the Royal Society of London B, 278, 751759.Google Scholar
Leong, K.L.H., Kaya, H.K., Yoshimura, M.A. & Frey, D.F. (1992) The occurrence and effect of a protozoan parasite, Ophryocystis elektroscirrha (Neogregarinida, Ophryocystidae) on overwintering monarch butterflies, Danaus plexippus (Lepidoptera, Danaidae) from two California winter sites. Ecological Entomology, 17, 338342.Google Scholar
Levin, S. & Pimentel, D. (1981) Selection of intermediate rates of increase in parasite–host systems. American Naturalist, 117, 308315.Google Scholar
Liere, H. & Larsen, A. (2010) Cascading trait-mediation: disruption of a trait-mediated mutualism by parasite-induced behavioral modification. Oikos, 119, 13941400.Google Scholar
Mackinnon, M.J., Gandon, S. & Read, A.F. (2008) Virulence evolution in response to vaccination: the case of malaria. Vaccine, 26, C42C52.Google Scholar
Malcolm, S.B. (1994) Milkweeds, monarch butterflies and the ecological significance of cardenolides. Chemoecology, 5, 101117.Google Scholar
Malcolm, S.B. & Brower, L.P. (1989) Evolutionary and ecological implications of cardenolide sequestration in the monarch butterfly. Experientia, 45, 284295.Google Scholar
Malcolm, S.B. & Zalucki, M.P. (1996) Milkweed latex and cardenolide induction may resolve the lethal plant defence paradox. Entomologia Experimentalis et Applicata, 80, 193196.Google Scholar
Matson, K.D., Horrocks, N.P., Tieleman, B.I. & Haase, E. (2012) Intense flight and endotoxin injection elicit similar effects on leukocyte distributions but dissimilar effects on plasma-based immunological indices in pigeons. Journal of Experimental Biology, 215, 37343741.Google Scholar
May, R.M. & Anderson, R.M. (1983) Epidemiology and genetics in the coevolution of parasites and hosts. Proceedings of the Royal Society of London B, 219, 281313.Google Scholar
McKay, A.F., Ezenwa, V.O. & Altizer, S. (2016a) Consequences of food restriction for immune defense, parasite infection, and fitness in monarch butterflies. Physiological and Biochemical Zoology, 89, 389401.Google Scholar
McKay, A.F., Ezenwa, V.O. & Altizer, S. (2016b) Unravelling the costs of flight for immune defenses in the migratory monarch butterfly. Integrative and Comparative Biology, 56, 278289.Google Scholar
McKinnon, L., Smith, P.A., Nol, E., et al. (2010) Lower predation risk for migratory birds at high latitudes. Science, 327, 326327.Google Scholar
McLaughlin, R.E. & Myers, J. (1970) Ophryocystis elektroscirrha sp. n., a neogregarine pathogen of monarch butterfly Danaus plexippus (L.) and the Florida queen butterfly D. gilippus berenice Cramer. Journal of Protozoology, 17, 300305.Google Scholar
Møller, A.P. & Erritzøe, J. (1998) Host immune defence and migration in birds. Evolutionary Ecology, 12, 945953.Google Scholar
Nagano, C.D., Sakai, W.H., Malcolm, S.B., et al. (1993) Spring migration of monarch butterflies in California. In: Zalucki, M.P. (ed.), Biology and Conservation of the Monarch Butterfly (pp. 217232). Los Angeles, CA: Natural History Museum of Los Angeles County.Google Scholar
Nebel, S., Buehler, D.M., MacMillan, A. & Guglielmo, C.G. (2013) Flight performance of western sandpipers, Calidris mauri, remains uncompromised when mounting an acute phase immune response. Journal of Experimental Biology, 216, 27522759.Google Scholar
Owen, J., Moore, F., Panella, N., et al. (2006) Migrating birds as dispersal vehicles for West Nile virus. EcoHealth, 3, 79.Google Scholar
Owen, J. & Moore, F.R. (2008a) Relationship between energetic condition and indicators of immune function in thrushes during spring migration. Canadian Journal of Zoology, 86, 638647.Google Scholar
Owen, J.C. & Moore, F.R. (2006) Seasonal differences in immunological condition of three species of thrushes. The Condor, 108, 389398.Google Scholar
Owen, J.C. & Moore, F.R. (2008b) Swainson’s thrushes in migratory disposition exhibit reduced immune function. Journal of Ethology, 26, 383388.Google Scholar
Penczykowski, R.M., Lemanski, B.C., Sieg, R.D., et al. (2014) Poor resource quality lowers transmission potential by changing foraging behaviour. Functional Ecology, 28, 12451255.Google Scholar
Petschenka, G., Fandrich, S., Sander, N., et al. (2013) Stepwise evolution of resistance to toxic cardenolides via genetic substitutions in the NA+/K+-ATPase of milkweed butterflies (Lepidoptera, Danaini). Evolution, 67, 27532761.Google Scholar
Pierce, A.A., de Roode, J.C. & Tao, L. (2016) Comparative genetics of Na+/K+-ATPase in monarch butterfly populations with varying host plant toxicity. Biological Journal of the Linnean Society, 119, 194200.Google Scholar
Pierce, A.A., Zalucki, M.P., Bangura, M., et al. (2014) Serial founder effects and genetic differentiation during worldwide range expansion of monarch butterflies. Proceedings of the Royal Society of London B, 281, 20142230.Google Scholar
Piersma, T. (1997) Do global patterns of habitat use and migration strategies co-evolve with relative investments in immunocompetence due to spatial variation in parasite pressure? Oikos, 80, 623631.Google Scholar
Pleasants, J.M. & Oberhauser, K.S. (2013) Milkweed loss in agricultural fields because of herbicide use: effect on the monarch butterfly population. Insect Conservation and Diversity, 6, 135144.Google Scholar
Plowright, R.K., Foley, P., Field, H.E., et al. (2011) Urban habituation, ecological connectivity and epidemic dampening: the emergence of Hendra virus from flying foxes (Pteropus spp.). Proceedings of the Royal Society of London B, 278, 37033712.Google Scholar
Price, P.W., Bouton, C.E., Gross, P., et al. (1980) Interactions among three tropic levels: influence of plants on interactions between insect herbivores and natural enemies. Annual Review of Ecology and Systematics, 11, 4165.Google Scholar
Råberg, L., Graham, A.L. & Read, A.F. (2009) Decomposing health: tolerance and resistance to parasites in animals. Philosophical Transactions of the Royal Society of London, Series B: Biological Sciences, 364, 3749.Google Scholar
Råberg, L., Sim, D. & Read, A.F. (2007) Disentangling genetic variation for resistance and tolerance to infectious disease in animals. Science, 318, 318320.Google Scholar
Rappole, J.H., Derrickson, S.R. & Hubálek, Z. (2000) Migratory birds and spread of West Nile virus in the Western Hemisphere. Emerging Infectious Diseases, 6, 319.Google Scholar
Rasmann, S. & Agrawal, A.A. (2011) Latitudinal patterns in plant defense: evolution of cardenolides, their toxicity and induction following herbivory. Ecology Letters, 14, 476483.Google Scholar
Read, A.F., Baigent, S.J., Powers, C., et al. (2015) Imperfect vaccination can enhance the transmission of highly virulent pathogens. PLoS Biology, 13, e1002198.Google Scholar
Sasaki, A. & Iwasa, Y. (1991) Optimal growth schedule of pathogens within a host: switching between lytic and latent cycles. Theoretical Population Biology, 39, 201239.Google Scholar
Satterfield, D.A., Altizer, S., Williams, M.-K. & Hall, R.J. (2017) Environmental persistence influences infection dynamics for a butterfly pathogen. PLoS ONE, 12, e0169982.Google Scholar
Satterfield, D.A., Maerz, J.C. & Altizer, S. (2015) Loss of migratory behaviour increases infection risk for a butterfly host. Proceedings of the Royal Society of London B, 282, 20141734.Google Scholar
Satterfield, D.A., Maerz, J.C., Hunter, M.D., et al. (2018) Migratory monarchs that encounter resident monarchs show life-history differences and higher rates of parasite infection. Ecology Letters, 21, 1670–1680.Google Scholar
Satterfield, D.A., Villablanca, F.X., Maerz, J.C. & Altizer, S. (2016) Migratory monarchs wintering in California experience low infection risk compared to monarchs breeding year-round on non-native milkweed. Integrative and Comparative Biology, 56, 343352.Google Scholar
Satterfield, D.A., Wright, A.E. & Altizer, S. (2013) Lipid reserves and immune defense in healthy and diseased migrating monarchs Danaus plexippus. Current Zoology, 59, 393402.Google Scholar
Shaw, A.K., Binning, S.A., Hall, S.R. & Michalakis, Y. (2016) Migratory recovery from infection as a selective pressure for the evolution of migration. The American Naturalist, 187, 491501.Google Scholar
Simmons, A.M. & Rogers, C.E. (1991) Dispersal and seasonal occurrence of Noctuidonema guyanense, an ectoparasitic nematode of adult fall armyworm (Lepidoptera: Noctuidae), in the United States 2. Journal of Entomological Science, 26, 136148.Google Scholar
Speight, M.R., Hunter, M.D. & Watt, A.D. (2008) The Ecology of Insects: Concepts and Applications, 2nd edn. Oxford: Wiley-Blackwell.Google Scholar
Sternberg, E.D., Lefèvre, T., Li, J., et al. (2012) Food plant derived disease tolerance and resistance in a natural butterly–plant–parasite interaction. Evolution, 66, 33673376.Google Scholar
Sternberg, E.D., Li, H., Wang, R., Gowler, C. & de Roode, J.C. (2013) Patterns of host–parasite adaptation in three populations of monarch butterflies infected with a naturally occurring protozoan disease: virulence, resistance, and tolerance. American Naturalist, 182, E235E248.Google Scholar
Sternberg, E.D., de Roode, J.C. & Hunter, M.D. (2015) Trans‐generational parasite protection associated with paternal diet. Journal of Animal Ecology, 84, 310321.Google Scholar
Tao, L., Gowler, C.D., Ahmad, A., Hunter, M.D. & de Roode, J.C. (2015) Disease ecology across soil boundaries: effects of below-ground fungi on above-ground host–parasite interactions. Proceedings of the Royal Society of London B, 282, 20151993.Google Scholar
Tao, L., Hoang, K.M., Hunter, M.D. & de Roode, J.C. (2016) Fitness costs of animal medication: anti‐parasitic plant chemicals reduce fitness of monarch butterfly hosts. Journal of Animal Ecology, 85, 12461254.Google Scholar
Taylor, C.M., Laughlin, A.J. & Hall, R.J. (2016) The response of migratory populations to phenological change: a migratory flow network modelling approach. Journal of Animal Ecology, 85, 648659.Google Scholar
Taylor, C.M. & Norris, D.R. (2010) Population dynamics in migratory networks. Theoretical Ecology, 3, 6573.Google Scholar
Urquhart, F.A. (1976) Found at last: the monarch’s winter home. National Geographic, 161173.Google Scholar
Urquhart, F.A. & Urquhart, N.R. (1978) Autumnal migration routes of the eastern population of the monarch butterfly (Danaus p. plexippus L.; Danaidae; Lepidoptera) in North America to the overwintering site in the Neovolcanic Plateau of Mexico. Canadian Journal of Zoology, 56, 17591764.Google Scholar
Van Baalen, M. & Sabelis, M.W. (1995) The dynamics of multiple infection and the evolution of virulence. American Naturalist, 146, 881910.Google Scholar
Van der Ree, R., McDonnell, M., Temby, I., Nelson, J. & Whittingham, E. (2006) The establishment and dynamics of a recently established urban camp of flying foxes (Pteropus poliocephalus) outside their geographic range. Journal of Zoology, 268, 177185.Google Scholar
Van Gils, J.A., Munster, V.J., Radersma, R., et al. (2007) Hampered foraging and migratory performance in swans infected with low-pathogenic avian influenza A virus. PLoS ONE, 2, e184.Google Scholar
Van Zandt, P.A. & Agrawal, A.A. (2004) Specificity of induced plant responses to specialist herbivores of the common milkweed Asclepias syriaca. Oikos, 104, 401409.Google Scholar
Vane-Wright, R.I. (1993) The Columbus hypothesis: an explanation for the dramatic 19th century range expansion of the monarch butterfly. In: Malcolm, S.B. & Zalucki, M.P. (eds.), Biology and Conservation of the Monarch Butterfly (pp. 179187). Los Angeles, CA: Natural History Museum of Los Angeles County.Google Scholar
Vannette, R.L. & Hunter, M.D. (2011) Plant defence theory re-examined: nonlinear expectations based on the costs and benefits of resource mutualisms. Journal of Ecology, 99, 6676.Google Scholar
Vidal, O. & Rendón-Salinas, E. (2014) Dynamics and trends of overwintering colonies of the monarch butterfly in Mexico. Biological Conservation, 180, 165175.Google Scholar
Visser, M.E., Perdeck, A.C., van Balen, J. & Both, C. (2009) Climate change leads to decreasing bird migration distances. Global Change Biology, 15, 18591865.Google Scholar
Werner, E.E. & Peacor, S.D. (2003) A review of trait-mediated indirect interactions in ecological communities. Ecology, 84, 10831100.Google Scholar
Wilcove, D.S. (2008) No Way Home: The Decline of the World’s Great Animal Migrations. Washington, DC: Island Press.Google Scholar
Woods, E.C., Hastings, A.P., Turley, N.E., Heard, S.B. & Agrawal, A.A. (2012) Adaptive geographical clines in the growth and defense of a native plant. Ecological Monographs, 82, 149168.Google Scholar
Woodson, R.E. (1954) The North American species of Asclepias L. Annals of the Missouri Botanical Garden, 41, 1211.Google Scholar
Zalucki, M.P., Brower, L.P. & Malcolm, S.B. (1990) Oviposition by Danaus plexippus in relation to cardenolide content of 3 Asclepias species in the southeastern USA. Ecological Entomology, 15, 231240.Google Scholar
Zalucki, M.P. & Clarke, A.R. (2004) Monarchs across the Pacific: the Columbus hypothesis revisited. Biological Journal of the Linnean Society, 82, 111121.Google Scholar
Zalucki, M.P., Malcolm, S.B., Paine, T.D., et al. (2001) It’s the first bites that count: survival of first-instar monarchs on milkweeds. Austral Ecology, 26, 547555.Google Scholar
Zehnder, C.B. & Hunter, M.D. (2007) Interspecific variation within the genus Asclepias in response to herbivory by a phloem-feeding insect herbivore. Journal of Chemical Ecology, 33, 20442053.Google Scholar
Zehnder, C.B. & Hunter, M.D. (2008) Effects of nitrogen deposition on the interaction between an aphid and its host plant. Ecological Entomology, 33, 2430.Google Scholar
Zehnder, C.B. & Hunter, M.D. (2009) More is not necessarily better: the impact of limiting and excessive nutrients on herbivore population growth rates. Ecological Entomology, 34, 535543.Google Scholar
Zhan, S., Zhang, W., Niitepõld, K., et al. (2014) The genetics of monarch butterfly migration and warning colouration. Nature, 514, 317321.Google Scholar
Zhen, Y., Aardema, M.L., Medina, E.M., Schumer, M. & Andolfatto, P. (2012) Parallel molecular evolution in a herbivore community. Science, 337, 16341637.Google Scholar
Zhu, H., Casselman, A. & Reppert, S.M. (2008) Chasing migration genes: a brain expressed sequence tag resource for summer and migratory monarch butterflies (Danaus plexippus). PLoS ONE, 3, e1345.Google Scholar

References

Bielby, J., Fisher, M.C., Clare, F.C., Rosa, G.M. & Garner, T.W.J. (2015) Host species vary in infection probability, sub-lethal effects, and costs of immune response when exposed to an amphibian parasite. Scientific Reports, 5, 18.Google Scholar
Briggs, C.J., Vredenburg, V.T.,Knapp, R.A. & Rachowicz, L.J. (2005) Investigating the population-level effects of chytridiomycosis: an emerging infectious disease of amphibians. Ecology, 86, 31493159.Google Scholar
Hanlon, S.M., Lynch, K.J., Kerby, J. & Parris, M.J. (2015) Batrachochytrium dendrobatidis exposure effects on foraging efficiencies and body size in anuran tadpoles. Diseases of Aquatic Organisms, 112, 237242.Google Scholar
Jani, A.J., Knapp, R.A. & Briggs, C.J. (2017) Epidemic and endemic pathogen dynamics correspond to distinct host population microbiomes at a landscape scale. Proceedings of the Royal Society of London B, 284, 20170944.Google Scholar
Kilpatrick, A.M., Briggs, C.J. & Daszak, P. (2010) The ecology and impact of chytridiomycosis: an emerging disease of amphibians. Trends in Ecology and Evolution, 25, 109118.Google Scholar
Knapp, R.A. & Matthews, K.R. (2000) Non-native mountain fish introductions and the decline of the mountain yellow-legged frog from within protected areas. Conservation Biology, 14, 428438.Google Scholar
Knapp, R.A., Fellers, G.M., Kleeman, P.M., et al. (2016) Large-scale recovery of an endangered amphibian despite ongoing exposure to multiple stressors. Proceedings of the National Academy of Sciences of the United States of America, 113, 11,88911,894.Google Scholar
Longcore, J.E., Pessier, A.P. &Nichols, D.K. (1999) Batrachochytrium dendrobatidis gen. et sp. nov., a chytrid pathogenic to amphibians. Mycologia, 91, 219227.Google Scholar
Voyles, J., Berger, L., Young, S., et al. (2007) Electrolyte depletion and osmotic imbalance in amphibians with chytridiomycosis. Diseases of Aquatic Organisms, 77, 113118.Google Scholar
Voyles, J., Young, S., Berger, L., et al. (2009) Pathogenesis of chytridiomycosis, a cause of catastrophic amphibian declines. Science, 326, 58.Google Scholar
Vredenburg, V.T., Knapp, R.A., Tunstall, T.S. & Briggs, C.J. (2010) Dynamics of an emerging disease drive large-scale amphibian population extinctions. Proceedings of the National Academy of Sciences of the United States of America, 107, 96899694.Google Scholar

References

Boyle, D.G., Boyle, D.B., Olsen, V., Morgan, J.A.T. & Hyatt, A.D. (2004) Rapid quantitative detection of chytridiomycosis (Batrachochytrium dendrobatidis) in amphibian samples using real-time Taqman PCR assay. Diseases of Aquatic Organisms, 60, 141148.Google Scholar
Briggs, C.J., Knapp, R.A. & Vredenburg, V.T. (2010) Enzootic and epizootic dynamics of the chytrid fungal pathogen of amphibians. Proceedings of the National Academy of Sciences of the United States of America, 107, 96959700.Google Scholar
Easterling, M.R., Ellner, S.P. & Dixon, P.M. (2000) Size-specific sensitivity: applying a new structured population model. Ecology, 81, 694708.Google Scholar
Jani, A.J., Knapp, R.A. & Briggs, C.J. (2017) Epidemic and endemic pathogen dynamics correspond to distinct host population microbiomes at a landscape scale. Proceedings of the Royal Society of London B, 284, 20170944.Google Scholar
Wilber, M.Q., Knapp, R.A., Toothman, M. & Briggs, C.J. (2017) Resistance, tolerance and environmental transmission dynamics determine host extinction risk in a load-dependent amphibian disease. Ecology Letters, 30, 11691181.Google Scholar
Wilber, M.Q., Langwig, K.E., Kilpatrick, A.M., McCallum, H.I. & Briggs, C.J. (2016) Integral projection models for host–parasite systems with an application to amphibian chytrid fungus. Methods in Ecology and Evolution, 7, 11821194.Google Scholar
Woodhams, D.C., Alford, R.A., Briggs, C.J., Johnson, M. & Rollins-Smith, L.A. (2008) Life-history trade-offs influence disease in changing climates: strategies of an amphibian pathogen. Ecology, 89, 16271639.Google Scholar

References

Briggs, C.J., Knapp, R.A. & Vredenburg, V.T. (2010) Enzootic and epizootic dynamics of the chytrid fungal pathogen of amphibians. Proceedings of the National Academy of Sciences of the United States of America, 107, 96959700.Google Scholar
Wilber, M.Q., Knapp, R.A., Toothman, M. & Briggs, C.J. (2017) Resistance, tolerance and environmental transmission dynamics determine host extinction risk in a load-dependent amphibian disease. Ecology Letters, 30, 11691181.Google Scholar

References

Adams, A.J., Kupferberg, S.J., Wilber, M.Q., et al. (2017) Extreme drought, host density, sex, and bullfrogs influence fungal pathogen infection in a declining lotic amphibian. Ecosphere, 8(3), e01740.Google Scholar
Allen, L.J.S. (2015) Stochastic Population and Epidemic Models: Persistence and Extinction. London: Springer International Publishing.Google Scholar
Altwegg, R. & Reyer, H.-U. (2003) Patterns of natural selection on size at metamorphosis in water frogs. Evolution, 57, 872882.Google Scholar
Anderson, R.M. & May, R.M. (1979) Population biology of infectious diseases: Part I. Nature, 280, 361367.Google Scholar
Anderson, R.M. & May, R.M. (1981) The population dynamics of microparasites and their invertebrate hosts. Philosophical Transactions of the Royal Society of London B, 291, 451524.Google Scholar
Anderson, R.M. & May, R.M. (1991) Infectious Diseases of Humans: Dynamics and Control. Oxford: Oxford University Press.Google Scholar
Becker, C.G., Greenspan, S.E., Tracy, K.E., et al. (2017) Variation in phenotype and virulence among enzootic and panzootic amphibian chytrid lineages. Fungal Ecology, 26, 4550.Google Scholar
Begon, M., Bennett, M., Bowers, R.G., et al. (2002) A clarification of transmission terms in host–microparasite models: numbers, densities and areas. Epidemiology and Infection, 129, 147153.Google Scholar
Bletz, M.C., Rosa, G.M., Andreone, F., et al. (2015) Widespread presence of the pathogenic fungus Batrachochytrium dendrobatidis in wild amphibian communities in Madagascar. Scientific Reports, 5, 110.Google Scholar
Boots, M., Best, A., Miller, M.R. & White, A. (2009) The role of ecological feedbacks in the evolution of host defence: what does theory tell us? Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 364, 2736.Google Scholar
Bosch, J., Martínez-Solano, I. & García-París, M. (2001) Evidence of a chytrid fungus infection involved in the decline of the common midwife toad (Alytes obstetricans) in protected areas of central Spain. Biological Conservation, 97, 331337.Google Scholar
Briggs, C.J., Knapp, R.A. & Vredenburg, V.T. (2010) Enzootic and epizootic dynamics of the chytrid fungal pathogen of amphibians. Proceedings of the National Academy of Sciences of the United States of America, 107, 96959700.Google Scholar
Briggs, C.J., Vredenburg, V.T., Knapp, R.A. & Rachowicz, L.J. (2005) Investigating the population-level effects of chytridiomycosis: an emerging infectious disease of amphibians. Ecology, 86, 31493159.Google Scholar
Chestnut, T., Anderson, C., Popa, R., et al. (2014) Heterogeneous occupancy and density estimates of the pathogenic fungus Batrachochytrium dendrobatidis in waters of North America. PLoS ONE, 9, e106790.Google Scholar
Clare, F.C., Halder, J.B., Daniel, O., et al. (2016) Climate forcing of an emerging pathogenic fungus across a montane multi-host community. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 371, 20150454.Google Scholar
Cohen, J.M., Venesky, M.D., Sauer, E.L., et al. (2017) The thermal mismatch hypothesis explains host susceptibility to an emerging infectious disease. Ecology Letters, 20, 184193.Google Scholar
Combes, C. (2000) Parasitism: The Ecology and Evolution of Intimate Interactions. Chicago, IL: The University of Chicago Press.Google Scholar
Courtois, E.A., Loyau, A., Bourgoin, M. & Schmeller, D.S. (2017) Initiation of Batrachochytrium dendrobatidis infection in the absence of physical contact with infected hosts – a field study in a high altitude lake. Oikos, 126, 843851.Google Scholar
Daszak, P., Cunningham, A.A. & Hyatt, A.D. (2003) Infectious disease and amphibian population declines. Diversity and Distributions, 9, 141150.Google Scholar
De Castro, F. & Bolker, B. (2005) Mechanisms of disease-induced extinction. Ecology Letters, 8, 117126.Google Scholar
Diekmann, O. & Heesterbeek, J.A.P. (2000) Mathematical Epidemiology of Infectious Disease: Model Building, Interpretation, and Analysis. New York, NY: John Wiley & Sons.Google Scholar
DiRenzo, G.V., Langhammer, P.F., Zamudio, K.R. & Lips, K.R. (2014) Fungal infection intensity and zoospore output of Atelopus zeteki, a potential acute chytrid supershedder. PLoS ONE, 9, e93356.Google Scholar
Doddington, B.J., Bosch, J., Oliver, J.A., et al. (2013) Context-dependent amphibian host population response to an invading pathogen. Ecology, 94, 17951804.Google Scholar
Drawert, B., Griesemer, M., Petzold, L.R. & Briggs, C.J. (2017) Using stochastic epidemiological models to evaluate conservation strategies for endangered amphibians. Journal of The Royal Society Interface, 14, 20170480.Google Scholar
Ellison, A.R., Tunstall, T., Direnzo, G.V., et al. (2015) More than skin deep: functional genomic basis for resistance to amphibian chytridiomycosis. Genome Biology and Evolution, 7, 286298.Google Scholar
Farrer, R.A., Weinert, L.A., Bielby, J., et al. (2011) Multiple emergences of genetically diverse amphibian-infecting chytrids include a globalized hypervirulent recombinant lineage. Proceedings of the National Academy of Sciences of the United States of America, 108, 18,73218,736.Google Scholar
Fisher, M.C., Bosch, J., Yin, Z., et al. (2009) Proteomic and phenotypic profiling of the amphibian pathogen Batrachochytrium dendrobatidis shows that genotype is linked to virulence. Molecular Ecology, 18, 415429.Google Scholar
Fisher, M.C., Henk, D.A., Briggs, C.J., et al. (2012) Emerging fungal threats to animal, plant and ecosystem health. Nature, 484, 186194.Google Scholar
Garmyn, A.,Rooij, P. van, Pasmans, F., et al. (2012) Waterfowl: potential environmental reservoirs of the chytrid fungus Batrachochytrium dendrobatidis. PLoS ONE, 7, e35038.Google Scholar
Garner, T.W.J., Walker, S., Bosch, J., et al. (2009) Life history tradeoffs influence mortality associated with the amphibian pathogen Batrachochytrium dendrobatidis. Oikos, 118, 783791.Google Scholar
Godfray, H.C.J., Briggs, C.J., Barlow, N.D., et al. (1999) A model of insect–pathogen dynamics in which a pathogenic bacterium can also reproduce saprophytically. Proceedings of the Royal Society of London B, 266, 233240.Google Scholar
Hagman, M. & Alford, R.A. (2015) Patterns of Batrachochytrium dendrobatidis transmission between tadpoles in a high-elevation rainforest stream in tropical Australia. Diseases of Aquatic Organisms, 115, 213221.Google Scholar
Hanlon, S.M., Lynch, K.J., Kerby, J. & Parris, M.J. (2015) Batrachochytrium dendrobatidis exposure effects on foraging efficiencies and body size in anuran tadpoles. Diseases of Aquatic Organisms, 112, 237242.Google Scholar
Jani, A.J., Knapp, R.A. & Briggs, C.J. (2017) Epidemic and endemic pathogen dynamics correspond to distinct host population microbiomes at a landscape scale. Proceedings of the Royal Society of London B, 284, 20170944.Google Scholar
Johnson, M.L. & Speare, R. (2003) Survival of Batrachochytrium dendrobatidis in water: quarantine and disease control implications. Emerging Infectious Diseases, 9, 922925.Google Scholar
Johnson, M.L. & Speare, R. (2005) Possible modes of dissemination of the amphibian chytrid Batrachochytrium dendrobatidis in the environment. Diseases of Aquatic Organisms, 65, 181186.Google Scholar
Kilburn, V., Ibáñez, R. & Green, D. (2011) Reptiles as potential vectors and hosts of the amphibian pathogen Batrachochytrium dendrobatidis in Panama. Diseases of Aquatic Organisms, 97, 127134.Google Scholar
Kilpatrick, A.M., Briggs, C.J. & Daszak, P. (2010) The ecology and impact of chytridiomycosis: an emerging disease of amphibians. Trends in Ecology and Evolution, 25, 109118.Google Scholar
Knapp, R.A., Fellers, G.M., Kleeman, P.M., et al. (2016) Large-scale recovery of an endangered amphibian despite ongoing exposure to multiple stressors. Proceedings of the National Academy of Sciences of the United States of America, 113, 11,88911,894.Google Scholar
Kolby, J.E., Ramirez, S.D., Berger, L., et al. (2015) Terrestrial dispersal and potential environmental transmission of the amphibian chytrid fungus (Batrachochytrium dendrobatidis). PLoS ONE, 10, e0125386.Google Scholar
Lande, R., Engen, S. & Saether, B.-E. (2003) Stochastic Population Dynamics in Ecology and Conservation. Oxford: Oxford University Press.Google Scholar
Langhammer, P.F., Lips, K.R., Burrowes, P.A., et al. (2013) A fungal pathogen of amphibians, Batrachochytrium dendrobatidis, attenuates in pathogenicity with in vitro passages. PLoS ONE, 8, e77630.Google Scholar
Langwig, K.E., Voyles, J., Wilber, M.Q., et al. (2015) Context-dependent conservation responses to emerging wildlife diseases. Frontiers in Ecology and the Environment, 13, 195202.Google Scholar
Laurance, W.F., McDonald, K.R. & Speare, R. (1996) Epidemic disease and the catastrophic decline of Australian rain forest frogs. Conservation Biology, 10, 406413.Google Scholar
Liew, N., Mazon Moya, M.J., Wierzbicki, C.J., et al. (2017) Chytrid fungus infection in zebrafish demonstrates that the pathogen can parasitize non-amphibian vertebrate hosts. Nature Communications, 8, 15048.Google Scholar
Lloyd-Smith, J.O., Cross, P.C., Briggs, C.J., et al. (2005) Should we expect population thresholds for wildlife disease? Trends in Ecology and Evolution, 20, 511519.Google Scholar
Longcore, J.E., Pessier, A.P. & Nichols, D.K. (1999) Batrachochytrium dendrobatidis gen. et sp. nov., a chytrid pathogenic to amphibians. Mycologia, 91, 219227.Google Scholar
Maniero, G.D. & Carey, C. (1997) Changes in selected aspects of immune function in the leopard frog, Rana pipiens, associated with exposure to cold. Journal of Comparative Physiology B, 167, 256263.Google Scholar
Marca, E.L., Lips, K.R., Lötters, S., et al. (2005) Catastrophic population declines and extinctions in neotropical harlequin frogs (Bufonidae: Atelopus). Biotropica, 37, 190201.Google Scholar
McCallum, H. (2012) Disease and the dynamics of extinction. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 367, 28282839.Google Scholar
McCallum, H., Barlow, N. & Hone, J. (2001) How should pathogen transmission be modelled? Trends in Ecology and Evolution, 16, 295300.Google Scholar
McCallum, H., Jones, M., Hawkins, C., et al. (2009) Transmission dynamics of Tasmanian devil facial tumor disease may lead to disease-induced extinction. Ecology, 90, 33793392.Google Scholar
McMahon, T.A., Brannelly, L.A., Chatfield, M.W.H., et al. (2013) Chytrid fungus Batrachochytrium dendrobatidis has nonamphibian hosts and releases chemicals that cause pathology in the absence of infection. Proceedings of the National Academy of Sciences of the United States of America, 110, 210215.Google Scholar
McMahon, T.A., Sears, B.F., Venesky, M.D., et al. (2014) Amphibians acquire resistance to live and dead fungus overcoming fungal immunosuppression. Nature, 511, 224227.Google Scholar
Medzhitov, R., Schneider, D.S. & Soares, M.P. (2012) Disease tolerance as a defense strategy. Science, 335, 936941.Google Scholar
Perez, R., Richards-Zawacki, C.L., Krohn, A.R., et al. (2014) Field surveys in Western Panama indicate populations of Atelopus varius frogs are persisting in regions where Batrachochytrium dendrobatidis is now enzootic. Amphibian and Reptile Conservation, 8, 3035.Google Scholar
Piotrowski, J.S., Annis, S.L. & Longcore, J.E. (2004) Physiology of Batrachochytrium dendrobatidis, a chytrid pathogen of amphibians. Mycologia, 96, 915.Google Scholar
Rachowicz, L.J. & Briggs, C.J. (2007) Quantifying the disease transmission function: effects of density on Batrachochytrium dendrobatidis transmission in the mountain yellow-legged frog Rana muscosa. The Journal of Animal Ecology, 76, 711721.Google Scholar
Raffel, T.R., Halstead, N.T., McMahon, T.A., Davis, A.K. & Rohr, J.R. (2015) Temperature variability and moisture synergistically interact to exacerbate an epizootic disease. Proceedings of the Royal Society of London B, 282, 20142039.Google Scholar
Raffel, T.R., Rohr, J.R., Kiesecker, J.M. & Hudson, P.J. (2006) Negative effects of changing temperature on amphibian immunity under field conditions. Functional Ecology, 20, 819828.Google Scholar
Raffel, T.R., Romansic, J.M., Halstead, N.T., et al. (2012) Disease and thermal acclimation in a more variable and unpredictable climate. Nature Climate Change, 3, 146151.Google Scholar
Råberg, L., Graham, A.L. & Read, A.F. (2009) Decomposing health: tolerance and resistance to parasites in animals. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 364, 3749.Google Scholar
Reeder, N.M.M., Pessier, A.P. & Vredenburg, V.T. (2012) A reservoir species for the emerging amphibian pathogen Batrachochytrium dendrobatidis thrives in a landscape decimated by disease. PLoS ONE, 7, e33567.Google Scholar
Refsnider, J.M., Poorten, T.J., Langhammer, P.F., Burrowes, P.A. & Rosenblum, E.B. (2015) Genomic correlates of virulence attenuation in the deadly amphibian chytrid fungus, Batrachochytrium dendrobatidis. G3: Genes|Genomes|Genetics, 5, 22912298.Google Scholar
Retallick, R.W.R., McCallum, H. & Speare, R. (2004) Endemic infection of the amphibian chytrid fungus in a frog community post-decline. PLoS Biology, 2, e351.Google Scholar
Retallick, R.W.R. & Miera, V. (2007) Strain differences in the amphibian chytrid Batrachochytrium dendrobatidis and non-permanent, sub-lethal effects of infection. Diseases of Aquatic Organisms, 75, 201207.Google Scholar
Rosenblum, E.B., James, T.Y., Zamudio, K.R., et al. (2013) Complex history of the amphibian-killing chytrid fungus revealed with genome resequencing data. Proceedings of the National Academy of Sciences of the United States of America, 110, 93859390.Google Scholar
Savage, A.E., Becker, C.G. & Zamudio, K.R. (2015) Linking genetic and environmental factors in amphibian disease risk. Evolutionary Applications, 8, 560572.Google Scholar
Savage, A.E. & Zamudio, K.R. (2011) MHC genotypes associate with resistance to a frog-killing fungus. Proceedings of the National Academy of Sciences of the United States of America, 108, 16,70516,710.Google Scholar
Savage, A.E. & Zamudio, K.R. (2016) Adaptive tolerance to a pathogenic fungus drives major histocompatibility complex evolution in natural amphibian populations. Proceedings of the Royal Society of London B, 283, 20153115.Google Scholar
Scheele, B.C., Guarino, F., Osborne, W., et al. (2014) Decline and re-expansion of an amphibian with high prevalence of chytrid fungus. Biological Conservation, 170, 8691.Google Scholar
Scheele, B.C., Skerratt, L.F., Grogan, L.F., et al. (2017) After the epidemic: ongoing declines, stabilizations and recoveries in amphibians afflicted by chytridiomycosis. Biological Conservation, 206, 3746.Google Scholar
Schmeller, D.S., Blooi, M., Martel, A., et al. (2014) Microscopic aquatic predators strongly affect infection dynamics of a globally emerged pathogen. Current Biology, 24, 176180.Google Scholar
Semlitsch, R.D. (1990) Effects of body size, sibship, and tail injury on the susceptibility of tadpoles to dragonfly predation. Canadian Journal of Zoology, 68, 10271030.Google Scholar
Skerratt, L.F., Berger, L., Speare, R., et al. (2007) Spread of chytridiomycosis has caused the rapid global decline and extinction of frogs. EcoHealth, 4, 125134.Google Scholar
Smith, D.C. (1987) Adult recruitment in chorus frogs: effects of size and date at metamorphosis. Ecology, 68, 344350.Google Scholar
Stockwell, M.P., Bower, D.S., Clulow, J. & Mahony, M.J. (2016) The role of non-declining amphibian species as alternative hosts for Batrachochytrium dendrobatidis in an amphibian community. Wildlife Research, 43, 341347.Google Scholar
Stockwell, M.P., Clulow, J. & Mahony, M.J. (2010) Host species determines whether infection load increases beyond disease-causing thresholds following exposure to the amphibian chytrid fungus. Animal Conservation, 13, 6271.Google Scholar
Stuart, S.N., Chanson, J.S., Cox, N.A., et al. (2004) Status and trends of amphibian declines and extinctions worldwide. Science, 306, 17831786.Google Scholar
Venesky, M.D., Liu, X., Sauer, E.L. & Rohr, J.R. (2013) Linking manipulative experiments to field data to test the dilution effect. Journal of Animal Ecology, 83, 557565.Google Scholar
Voyles, J., Young, S., Berger, L., et al. (2009) Pathogenesis of chytridiomycosis, a cause of catastrophic amphibian declines. Science, 326, 58.Google Scholar
Vredenburg, V.T., Knapp, R.A., Tunstall, T.S. & Briggs, C.J. (2010) Dynamics of an emerging disease drive large-scale amphibian population extinctions. Proceedings of the National Academy of Sciences of the United States of America, 107, 96899694.Google Scholar
Wilber, M.Q., Knapp, R.A., Toothman, M. & Briggs, C.J. (2017) Resistance, tolerance and environmental transmission dynamics determine host extinction risk in a load-dependent amphibian disease. Ecology Letters, 30, 11691181.Google Scholar
Wilber, M.Q., Langwig, K.E., Kilpatrick, A.M., McCallum, H.I. & Briggs, C.J. (2016) Integral projection models for host–parasite systems with an application to amphibian chytrid fungus. Methods in Ecology and Evolution, 7, 11821194.Google Scholar
Woodhams, D.C., Alford, R.A., Briggs, C.J., Johnson, M. & Rollins-Smith, L.A. (2008) Life-history trade-offs influence disease in changing climates: strategies of an amphibian pathogen. Ecology, 89, 16271639.Google Scholar
Woodhams, D.C., Bosch, J., Briggs, C.J., et al. (2011) Mitigating amphibian disease: strategies to maintain wild populations and control chytridiomycosis. Frontiers in Zoology, 8, 8.Google Scholar

References

Aaen, S.M., Helgesen, K.O., Bakke, M.J., Kaur, K. & Horsberg, T.E. (2015) Drug resistance in sea lice: a threat to salmonid aquaculture. Trends in Parasitology, 31, 7281.Google Scholar
Altizer, S., Harvell, D. & Friedle, E. (2003) Rapid evolutionary dynamics and disease threats to biodiversity. Trends in Ecology and Evolution, 18, 589596.Google Scholar
Anderson, R. M. & May, R. M. (1978) Regulation and stability of host–parasite population interactions: I. Regulatory processes. Journal of Animal Ecology, 47, 219247.Google Scholar
Anderson, R. M. & May, R. M. (1979) Population biology of infectious diseases: Part I. Nature, 280, 361367.Google Scholar
Ashander, J. (2010) Effects of parasite exchange between wild and farmed salmon. MSc thesis, University of Alberta. DOI:10.6084/M9.FIGSHARE.1584651Google Scholar
Bateman, A.W., Peacock, S.J., Connors, B.M., et al. (2016) Recent failure in control of sea louse outbreaks on salmon in the Broughton Archipelago, British Columbia. Canadian Journal of Fisheries & Aquatic Sciences, 73, 11641172.Google Scholar
Beamish, R.J., Mahnken, C. & Neville, C.M. (2004) Evidence that reduced early marine growth is associated with lower marine survival of coho salmon. Transactions of the American Fisheries Society, 133, 2633.Google Scholar
Bjørn, P.A., Finstad, B. & Kristoffersen, R. (2001) Salmon lice infection of wild sea trout and Arctic char in marine and freshwaters: the effects of salmon farms. Aquaculture Research, 32, 947962.Google Scholar
Brauner, C.J., Sackville, M., Gallagher, Z., et al. (2012) Physiological consequences of the salmon louse (Lepeophtheirus salmonis) on juvenile pink salmon (Oncorhynchus gorbuscha): implications for wild salmon ecology and management, and for salmon aquaculture. Philosophical Transactions of the Royal Society of London: Series B, Biological Sciences, 367, 17701779.Google Scholar
Burridge, L., Weis, J.S., Cabello, F., Pizarro, J. & Bostick, K. (2010) Chemical use in salmon aquaculture: a review of current practices and possible environmental effects. Aquaculture, 306, 723.Google Scholar
Chittenden, C.M., Jensen, J.L.A., Ewart, D., et al. (2010) Recent salmon declines: a result of lost feeding opportunities due to bad timing? PLoS ONE, 5, e12423.Google Scholar
Comins, H.N. (1977) The development of insecticide resistance in the presence of migration. Journal of Theoretical Biology, 64, 177197.Google Scholar
Connors, B.M., Braun, D.C., Peterman, R.M.M., et al. (2012) Migration links ocean-scale competition and local ocean conditions with exposure to farmed salmon to shape wild salmon dynamics. Conservation Letters, 5, 304312.Google Scholar
Connors, B.M., Hargreaves, N.B., Jones, S.R.M. & Dill, L.M. (2010a) Predation intensifies parasite exposure in a salmonid food chain. Journal of Applied Ecology, 47, 13651371.Google Scholar
Connors, B.M., Krkošek, M. & Dill, L.M. (2008) Sea lice escape predation on their host. Biology Letters, 4, 455457.Google Scholar
Connors, B.M., Krkošek, M., Ford, J. & Dill, L.M. (2010b) Coho salmon productivity in relation to salmon lice from infected prey and salmon farms. Journal of Applied Ecology, 47, 13721377.Google Scholar
Connors, B.M., Lagasse, C. & Dill, L.M. (2011) What’s love got to do with it? Ontogenetic changes in drivers of dispersal in a marine ectoparasite. Behavioral Ecology, 22, 588593.Google Scholar
Costello, M.J. (2004) A checklist of best practice for sea lice control on salmon farms. Caligus, 8, 18.Google Scholar
Costello, M.J. (2006) Ecology of sea lice parasitic on farmed and wild fish. Trends in Parasitology, 22, 475483.Google Scholar
Costello, M.J. (2009) The global economic cost of sea lice to the salmonid farming industry. Journal of Fish Diseases, 32, 115.Google Scholar
Daszak, P., Cunningham, A. & Hyatt, A. (2000) Emerging infectious diseases of wildlife– threats to biodiversity and human health. Science, 287, 443.Google Scholar
De Castro, F. & Bolker, B. (2005) Mechanisms of disease-induced extinction. Ecology Letters, 8, 117126.Google Scholar
Dhondt, A., Dobson, A., Hochachka, W.M., et al. (2013) Multiple host transfers, but only one successful lineage in a continent-spanning emergent pathogen. Proceedings of the Royal Society of London B, 280, 20131068.Google Scholar
Eggers, D.M. (1978) Limnetic feeding behavior of juvenile sockeye salmon in Lake Washington and predator avoidance. Limnology and Oceanography, 23, 11141125.Google Scholar
FAO (2016) The State of the World Fisheries and Aquaculture (SOFIA) 2016. Rome: FAO.Google Scholar
Frazer, L.N., Morton, A. & Krkošek, M. (2012) Critical thresholds in sea lice epidemics: evidence, sensitivity and subcritical estimation. Proceedings of the Royal Society of London B, 279, 19501958.Google Scholar
Furey, N.B., Hinch, S.G., Bass, A.L., et al. (2016) Predator swamping reduces predation risk during nocturnal migration of juvenile salmon in a high-mortality landscape. Journal of Animal Ecology, 85, 948959.Google Scholar
Godwin, S.C., Dill, L.M., Krkošek, M. & Price, M.H.H. (2017) Reduced growth in wild juvenile sockeye salmon Oncorhynchus nerka infected with sea lice. Journal of Fish Biology, 91, 4157.Google Scholar
Godwin, S.C., Dill, L.M., Reynolds, J.D. & Krkošek, M. (2015) Sea lice, sockeye salmon, and foraging competition: lousy fish are lousy competitors. Canadian Journal of Fisheries & Aquatic Sciences, 72, 11131120.Google Scholar
Godwin, S.C., Krkošek, M., Reynolds, J.D., Rogers, L.A. & Dill, L.M. (2017) Heavy sea louse infection is associated with decreased stomach fullness in wild juvenile sockeye salmon. Canadian Journal of Fisheries and Aquatic Sciences, 75, 15871595.Google Scholar
Gould, F. (1998) Sustainability of transgenic insecticidal cultivars: integrating pest genetics and ecology. Annual Review of Entomology, 43, 701726.Google Scholar
Groner, M.L., Gettinby, G., Stormoen, M., Revie, C.W. & Cox, R. (2014) Modelling the impact of temperature-induced life history plasticity and mate limitation on the epidemic potential of a marine ectoparasite. PLoS ONE, 9, e88465.Google Scholar
Groot, C. & Margolis, L. (1991) Pacific Salmon Life Histories. Vancouver, B.C.: UBC Press.Google Scholar
Hamre, L.A., Eichner, C., Caipang, C.M.A., et al. (2013) The salmon louse Lepeophtheirus salmonis (Copepoda: Caligidae) life cycle has only two chalimus stages. PLoS ONE, 8, e73539.Google Scholar
Hargreaves, N. B. & LeBrasseur, R. J. (1985) Species selective predation on juvenile pink (Oncorhynchus gorbuscha) and chum salmon (O. keta) by coho salmon (O. kisutch). Canadian Journal of Fisheries and Aquatic Sciences, 42, 659668.Google Scholar
Hargreaves, N.B. & LeBrasseur, R.J. (1986) Size selectivity of coho (Oncorhynchus kisutch) preying on juvenile chum salmon (O. keta). Canadian Journal of Fisheries and Aquatic Sciences, 43, 581586.Google Scholar
Hatcher, M.J., Dick, J.T.A. & Dunn, A.M. (2006) How parasites affect interactions between competitors and predators. Ecology Letters, 9, 12531271.Google Scholar
Hatcher, M.J., Dick, J.T.A. & Dunn, A.M. (2012) Diverse effects of parasites in ecosystems: linking interdependent processes. Frontiers in Ecology and the Environment, 10, 186194.Google Scholar
Heuch, P.A., Nordhagen, J.R. & Schram, T.A. (2000) Egg production in the salmon louse [Lepeophtheirus salmonis (Krøyer)] in relation to origin and water temperature. Aquaculture Research, 31, 805814.Google Scholar
Holling, C.S. (1959) Some characteristics of simple types of predation and parasitism. The Canadian Entomologist, 91, 385398.Google Scholar
Hudson, P.J., Dobson, A.P. & Newborn, D. (1992) Do parasites make prey vulnerable to predation? Red grouse and parasites. Journal of Animal Ecology, 61, 681692.Google Scholar
Hudson, P. & Greenman, J. (1998) Competition mediated by parasites: biological and theoretical progress. Trends in Ecology and Evolution, 13, 387390.Google Scholar
Hudson, P.J., Rizzoli, A.P., Grenfell, B.T., Heesterbeek, J.A.P. & Dobson, A.P. (2002) Ecology of Wildlife Diseases. Oxford: Oxford University Press.Google Scholar
Ives, A.R. & Murray, D.L. (1997) Can sublethal parasitism destabilize predator–prey population dynamics? A model of snowshoe hares, predators and parasites. Journal of Animal Ecology, 66, 265278.Google Scholar
Jansen, P.A., Kristoffersen, A.B., Viljugrein, H., et al. (2012) Sea lice as a density-dependent constraint to salmonid farming. Proceedings of the Royal Society of London B, 279, 23302338.Google Scholar
Johnson, P.T.J., Stanton, D.E., Preu, E.R., Forshay, K.J. & Carpenter, S.R. (2006) Dining on disease: how interactions between infection and environment affect predation risk. Ecology, 87, 19731980.Google Scholar
Johnson, S.C. & Albright, L.J. (1991a) The developmental stages of Lepeophtheirus salmonis (Krøyer, 1837) (Copepoda: Caligidae). Canadian Journal of Zoology, 69, 929950.Google Scholar
Johnson, S.C. & Albright, L. J. (1991b) Development, growth, and survival of Lepeophtheirus salmonis (Copepoda: Caligidae) under laboratory conditions. Journal of the Marine Biological Association of the United Kingdom, 71, 425436.Google Scholar
Jones, K.E., Patel, N.G., Levy, M.A., et al. (2008) Global trends in emerging infectious diseases. Nature, 451, 990993.Google Scholar
Jones, S.R.M., Prosperi-Porta, G., Kim, E., Callow, P. & Hargreaves, N.B. (2006) The occurrence of Lepeophtheirus salmonis and Caligus clemensi (Copepoda: Caligidae) on three-spine stickleback Gasterosteus aculeatus in coastal British Columbia. The Journal of Parasitology, 92, 473480.Google Scholar
Kennedy, D.A., Kurath, G., Brito, I.L., et al. (2016) Potential drivers of virulence evolution in aquaculture. Evolutionary Applications, 9, 344354.Google Scholar
Kreitzman, M., Ashander, J., Driscoll, J., et al. (2018) An evolutionary ecosystem service: wild salmon sustain the effectiveness of parasite control on salmon farms. Conservation Letters, 11, e12395.Google Scholar
Krkošek, M. (2016) Population biology of infectious diseases shared by wild and farmed fish. Canadian Journal of Fisheries & Aquatic Sciences, 74, 620628.Google Scholar
Krkošek, M., Ashander, J., Frazer, L.N., & Lewis, M.A. (2013a) Allee effect from parasite spill-back. American Naturalist, 182, 640652.Google Scholar
Krkošek, M., Connors, B.M., Ford, H., et al. (2011a) Fish farms, parasites, and predators: implications for salmon population dynamics. Ecological Applications, 21, 897914.Google Scholar
Krkošek, M., Connors, B.M., Lewis, M.A. & Poulin, R. (2012) Allee effects may slow the spread of parasites in a coastal marine ecosystem. The American Naturalist, 179, 401412.Google Scholar
Krkošek, M., Connors, B.M., Morton, A., et al. (2011b) Effects of parasites from salmon farms on productivity of wild salmon. Proceedings of the National Academy of Sciences of the United States of America, 108, 14,70014,704.Google Scholar
Krkošek, M., Ford, J.S., Morton, A., et al. (2007a) Declining wild salmon populations in relation to parasites from farm salmon. Science, 318, 1772.Google Scholar
Krkošek, M., Gottesfeld, A., Proctor, B., et al. (2007b) Effects of host migration, diversity and aquaculture on sea lice threats to Pacific salmon populations. Proceedings of the Royal Society of London B, 274, 31413149.Google Scholar
Krkošek, M., Lewis, M.A., Morton, A., Frazer, L.N. & Volpe, J.P. (2006) Epizootics of wild fish induced by farm fish. Proceedings of the National Academy of Sciences of the United States of America, 103, 15,50615,510.Google Scholar
Krkošek, M., Lewis, M.A., Volpe, J.P. & Krkošek, M. (2005) Transmission dynamics of parasitic sea lice from farm to wild salmon. Proceedings of the Royal Society of London B, 272, 689696.Google Scholar
Krkošek, M., Morton, A., Volpe, J.P. & Lewis, M.A. (2009) Sea lice and salmon population dynamics: effects of exposure time for migratory fish. Proceedings of the Royal Society of London B, 276, 28192828.Google Scholar
Krkošek, M., Revie, C.W., Gargan, P.G., et al. (2013b) Impact of parasites on salmon recruitment in the Northeast Atlantic Ocean. Proceedings of the Royal Society of London B, 280, 20122359.Google Scholar
Kutz, S.J., Hoberg, E.P., Molnár, P.K., Dobson, A. & Verocai, G.G. (2014) A walk on the tundra: host–parasite interactions in an extreme environment. International Journal for Parasitology: Parasites and Wildlife, 3, 198208.Google Scholar
Lafferty, K.D. (1992) Foraging on prey that are modified by parasites. The American Naturalist, 140, 854867.Google Scholar
Lafferty, K.D. (1999) The evolution of trophic transmission. Parasitology Today, 15, 111115.Google Scholar
Lafferty, K.D. & Ben-Horin, T. (2013) Abalone farm discharges the withering syndrome pathogen into the wild. Frontiers in Microbiology, 4, 15.Google Scholar
Lenski, R. & May, R. (1994) The evolution of virulence in parasites and pathogens: reconciliation between two competing hypotheses. Journal of Theoretical Biology, 169, 253265.Google Scholar
Liu, Y., Rosten, T.W., Henriksen, K., et al. (2016) Comparative economic performance and carbon footprint of two farming models for producing Atlantic salmon (Salmo salar): land-based closed containment system in freshwater and open net pen in seawater. Aquacultural Engineering, 71, 112.Google Scholar
Losos, C.J.C., Reynolds, J.D. & Dill, L.M. (2010) Sex-selective predation by three spine sticklebacks on sea lice: a novel cleaning behaviour. Ethology, 116, 981989.Google Scholar
Marty, G.D., Saksida, S.M. & Quinn, T.J. (2010) Relationship of farm salmon, sea lice, and wild salmon populations. Proceedings of the National Academy of Sciences of the United States of America, 107, 22,59922,604.Google Scholar
May, R.M. & Anderson, R.M. (1991) Infectious Diseases of Humans. Oxford: Oxford University Press.Google Scholar
May, R.M. & Nowak, M.A. (1995) Coinfection and the evolution of parasite virulence. Proceedings of the Royal Society of London B, 261, 209215.Google Scholar
McEwan, G.F., Groner, M.L., Fast, M.D., Gettinby, G. & Revie, C.W. (2015) Using agent-based modelling to predict the role of wild refugia in the evolution of resistance of sea lice to chemotherapeutants. PLoS ONE, 10, 123.Google Scholar
McKinnell, S., Curchitser, E., Groot, K., Kaeriyama, M. & Trudel, M. (2014) Oceanic and atmospheric extremes motivate a new hypothesis for variable marine survival of Fraser River sockeye salmon. Fisheries Oceanography, 23, 322341.Google Scholar
Mennerat, A., Nilsen, F., Ebert, D., & Skorping, A. (2010) Intensive farming: evolutionary implications for parasites and pathogens. Evolutionary Biology, 37, 5967.Google Scholar
Messmer, A.M., Rondeau, E.B., Jantzen, S.G., et al. (2011) Assessment of population structure in Pacific Lepeophtheirus salmonis (Krøyer) using single nucleotide polymorphism and microsatellite genetic markers. Aquaculture, 320, 183192.Google Scholar
Morton, A. & Routledge, R. (2005) Mortality rates for juvenile pink (Oncorhynchus gorbuscha) and chum (O. keta) salmon infested with sea lice (Lepeophtheirus salmonis) in the Broughton Archipelago. Alaska Fishery Research Bulletin, 11, 146152.Google Scholar
Morton, A., Routledge, R.D. & Williams, R. (2005) Temporal patterns of sea louse infestation on wild Pacific salmon in relation to the fallowing of Atlantic salmon farms. North American Journal of Fisheries Management, 25, 811821.Google Scholar
Moss, J.H., Beauchamp, D.A., Cross, A.D., et al. (2005) Evidence for size-selective mortality after the first summer of ocean growth by pink salmon. Transactions of the American Fisheries Society, 134, 13131322.Google Scholar
Murray, A. (2011) A simple model to assess selection for treatment-resistant sea lice. Ecological Modelling, 222, 18541862.Google Scholar
Murray, A.G. & Salama, N.K.G. (2016) A simple model of the role of area management in the control of sea lice. Ecological Modelling, 337, 3947.Google Scholar
Nilsen, A., Nielsen, K.V., Biering, E. & Bergheim, A. (2017) Effective protection against sea lice during the production of Atlantic salmon in floating enclosures. Aquaculture, 466, 4150.Google Scholar
Orobko, M. (2016) Alternate stable states in coupled fishery–aquaculture systems. MSc thesis, University of Toronto. ProQuest Number: 10130697.Google Scholar
Packer, C., Holt, R.D., Hudson, P.J., Lafferty, K.D. & Dobson, A.P. (2003) Keeping the herds healthy and alert: implications of predator control for infectious disease. Ecology Letters, 6, 797802.Google Scholar
Parker, R.R. (1968) Marine mortality schedules of pink salmon of the Bella Coola River, central British Columbia. Journal of the Fisheries Research Board of Canada, 25, 757794.Google Scholar
Parker, R.R. (1969) Predator–prey relationship among pink and chum salmon fry and coho smolts in a central British Columbia inlet. Fisheries Research Board of Canada Manuscript Report Series, 1019.Google Scholar
Peacock, S.J., Bateman, A.W., Krkošek, M. & Lewis, M.A. (2016) The dynamics of coupled populations subject to control. Theoretical Ecology, 9, 365380.Google Scholar
Peacock, S.J., Connors, B.M., Krkošek, M., Irvine, J.R. & Lewis, M.A. (2014) Can reduced predation offset negative effects of sea louse parasites on chum salmon? Proceedings of the Royal Society of London B, 281, 20132913.Google Scholar
Peacock, S.J., Krkošek, M., Bateman, A. W. &. Lewis, M.A. (2015) Parasitism and food web dynamics of juvenile Pacific salmon. Ecosphere, 6, 116.Google Scholar
Peacock, S.J., Krkošek, M., Proboszcz, S., Orr, C. & Lewis, M.A. (2013) Cessation of a salmon decline with control of parasites. Ecological Applications, 23, 606620.Google Scholar
Pedersen, A.B., Jones, K.E., Nunn, C.L. & Altizer, S. (2007) Infectious diseases and extinction risk in wild mammals. Conservation Biology, 21, 12691279.Google Scholar
Pike, A.W. & Wadsworth, S.L. (2000), Sealice on salmonids: their biology and control. Advances in Parasitology, 44, 233337.Google Scholar
Price, M.H.H., Morton, A. & Reynolds, J. D. (2010) Evidence of farm-induced parasite infestations on wild juvenile salmon in multiple regions of coastal British Columbia, Canada. Canadian Journal of Fisheries and Aquatic Sciences, 67, 19251932.Google Scholar
Price, M.H.H., Proboszcz, S.L., Routledge, R.D., et al. (2011) Sea louse infection of juvenile sockeye salmon in relation to marine salmon farms on Canada’s west coast. PLoS ONE, 6, e16851.Google Scholar
Pruvot, M., Lejeune, M., Kutz, S., et al. (2016) Better alone or in ill company? the effect of migration and inter-species comingling on Fascioloides magna infection in elk. PLoS ONE, 11, e0159319.Google Scholar
Pruvot, M., Seidel, D., Boyce, M.S., et al. (2014) What attracts elk onto cattle pasture? Implications for inter-species disease transmission. Preventive Veterinary Medicine, 117, 326339.Google Scholar
Pulkkinen, K., Suomalainen, L.-R., Read, A.F., et al. (2010) Intensive fish farming and the evolution of pathogen virulence: the case of columnaris disease in Finland. Proceedings of the Royal Society of London B, 277, 593600.Google Scholar
Revie, C.W., Gettinby, G., Treasurer, J.W. & Wallace, C. (2003) Identifying epidemiological factors affecting sea lice Lepeophtheirus salmonis abundance on Scottish salmon farms using general linear models. Diseases of Aquatic Organisms, 57, 8595.Google Scholar
Ritchie, G., Mordue (Luntz), A.J., Pike, A.W. & Rae, G.H. (1996) Observations on mating and reproductive behaviour of Lepeophtheirus salmonis, Krøyer (Copepoda: Caligidae). Journal of Experimental Marine Biology and Ecology, 201, 285298.Google Scholar
Ruggerone, G.T. & Connors, B.M. (2015) Productivity and life history of sockeye salmon in relation to competition with pink and sockeye salmon in the North Pacific Ocean. Canadian Journal of Fisheries & Aquatic Sciences, 72, 818833.Google Scholar
Saksida, S. M., Morrison, D. & Revie, C.W. (2010) The efficacy of emamectin benzoate against infestations of sea lice, Lepeophtheirus salmonis, on farmed Atlantic salmon, Salmo salar L., in British Columbia. Journal of Fish Diseases, 33, 913917.Google Scholar
Schumaker, B. (2013) Risks of Brucella abortus spillover in the Greater Yellowstone Area. Revue scientifique et technique (International Office of Epizootics), 32, 7177.Google Scholar
Shaw, D.J., Grenfell, B.T. & Dobson, A.P. (1998) Patterns of macroparasite aggregation in wildlife host populations. Parasitology, 117, 597610.Google Scholar
Sundberg, L.-R., Ketola, T., Laanto, E., et al. (2016) Intensive aquaculture selects for increased virulence and interference competition in bacteria. Proceedings of the Royal Society of London B, 283, 20153069.Google Scholar
Tabashnik, B., Brévault, T. & Carrière, Y. (2013) Insect resistance to Bt crops: lessons from the first billion acres. Nature Biotechnology, 31, 510521.Google Scholar
Thorstad, E.B., Todd, C.D., Uglem, I., et al. (2015) Effects of salmon lice Lepeophtheirus salmonis on wild sea trout Salmo trutta – a literature review. Aquaculture Environment Interactions, 7, 91113.Google Scholar
Tian, H., Zhou, S., Dong, L., et al. (2015) Avian influenza H5N1 viral and bird migration networks in Asia. Proceedings of the National Academy of Sciences of the United States of America, 112, 172177.Google Scholar
Tompkins, D.M., Carver, S., Jones, M.E., Krkošek, M. & Skerratt, L.F. (2015) Emerging infectious diseases of wildlife: a critical perspective. Trends in Parasitology, 31, 149159.Google Scholar
van Baalen, M. & Sabelis, M. (1995) The dynamics of multiple infection and the evolution of virulence. The American Naturalist, 146, 881.Google Scholar
Van Boeckel, T.P., Brower, C., Gilbert, M., et al. (2015) Global trends in antimicrobial use in food animals. Proceedings of the National Academy of Sciences of the United States of America, 112, 56495654.Google Scholar
Viana, M., Cleaveland, S., Matthiopoulos, J., et al. (2015) Dynamics of a morbillivirus at the domestic–wildlife interface: canine distemper virus in domestic dogs and lions. Proceedings of the National Academy of Sciences of the United States of America, 112, 14641469.Google Scholar
Vollset, K.W., Dohoo, I., Karlsen, Ø., et al. (2018) Disentangling the role of sea lice on the marine survival of Atlantic salmon. ICES Journal of Marine Science, 75, 5060.Google Scholar
Vollset, K.W., Krontveit, R.I., Jansen, P.A., et al. (2015) Impacts of parasites on marine survival of Atlantic salmon: a meta-analysis. Fish and Fisheries, 17, 714730.Google Scholar
Watson, M.J. (2013) What drives population-level effects of parasites? Meta-analysis meets life-history. International Journal for Parasitology: Parasites and Wildlife, 2, 190196.Google Scholar

References

Altizer, S., Davis, A.K., Cook, K.C. & Cherry, J.J. (2004a) Age, sex, and season affect the risk of mycoplasmal conjunctivitis in a southeastern house finch population. Canadian Journal of Zoology – Revue Canadienne De Zoologie, 82, 755763.Google Scholar
Altizer, S., Dobson, A., Hosseini, P., et al. (2006) Seasonality and the dynamics of infectious diseases. Ecology Letters, 9, 467484.Google Scholar
Altizer, S., Hochachka, W.M. & Dhondt, A.A. (2004b) Seasonal dynamics of mycoplasmal conjunctivitis in eastern North American house finches. Journal of Animal Ecology, 73, 309322.Google Scholar
Backstroem, N., Shipilina, D., Blom, M.P.K. & Edwards, S.V. (2013a) Cis-regulatory sequence variation and association with Mycoplasma load in natural populations of the house finch (Carpodacus mexicanus). Ecology and Evolution, 3, 655666.Google Scholar
Backstroem, N., Zhang, Q. & Edwards, S.V. (2013b) Evidence from a house finch (Haemorhous mexicanus) spleen transcriptome for adaptive evolution and biased gene conversion in passerine birds. Molecular Biology and Evolution, 30, 10461050.Google Scholar
Bonter, D.N. & Cooper, C.B. (2012) Data validation in citizen science: a case study from Project FeederWatch. Frontiers in Ecology and the Environment, 10, 305309.Google Scholar
Bouwman, K.M. & Hawley, D.M. (2010) Sickness behaviour acting as an evolutionary trap? Male house finches preferentially feed near diseased conspecifics. Biology Letters, 6, 462465.Google Scholar
Butcher, G.S., Fuller, M.R., Mcallister, L.S. & Geissler, P.H. (1990) An evaluation of the Christmas bird count for monitoring population trends of selected species. Wildlife Society Bulletin, 18, 129134.Google Scholar
Conn, P.B. & Cooch, E.G. (2009) Multistate capture–recapture analysis under imperfect state observation: an application to disease models. Journal of Applied Ecology, 46, 486492.Google Scholar
Cooper, C.B., Hochachka, W.M. & Dhondt, A.A. (2007) Contrasting natural experiments confirm competition between house finches and house sparrows. Ecology, 88, 864870.Google Scholar
Delaney, N.F., Balenger, S., Bonneaud, C., et al. (2012) Ultrafast evolution and loss of CRISPRs following a host shift in a novel wildlife pathogen, Mycoplasma gallisepticum. PLoS Genetics, 8(2), e1002511.Google Scholar
Dhondt, A.A., Altizer, S., Cooch, E.G., et al. (2005) Dynamics of a novel pathogen in an avian host: mycoplasmal conjunctivitis in house finches. Acta Tropica, 94, 7793.Google Scholar
Dhondt, A.A., Badyaev, A.V., Dobson, A.P., et al. (2006) Dynamics of mycoplasmal conjunctivitis in the native and introduced range of the host. Ecohealth, 3, 95102.Google Scholar
Dhondt, A.A., Decoste, J.C., Ley, D.H. & Hochachka, W.M. (2014) Diverse wild bird host range of Mycoplasma gallisepticum in eastern North America. PLoS ONE, 9(7), e103553.Google Scholar
Dhondt, A.A., Dhondt, K.V., Hawley, D.M. & Jennelle, C.S. (2007a) Experimental evidence for transmission of Mycoplasma gallisepticum in house finches by fomites. Avian Pathology, 36, 205208.Google Scholar
Dhondt, A.A., Dhondt, K.V. & Hochachka, W.M. (2015) Response of black-capped chickadees to house finch Mycoplasma gallisepticum. PLoS ONE, 10(4), e0124820.Google Scholar
Dhondt, A.A., Dhondt, K.V., Hochachka, W.M. & Schat, K.A. (2013) Can American goldfinches function as reservoirs for Mycoplasma gallisepticum? Journal of Wildlife Disease, 49, 4954.Google Scholar
Dhondt, A.A., Dhondt, K.V. & McCleery, B.V. (2008) Comparative infectiousness of three passerine bird species after experimental inoculation with Mycoplasma gallisepticum. Avian Pathology, 37, 635640.Google Scholar
Dhondt, A.A., Driscoll, M.J.L. & Swarthout, E.C.H. (2007b) House finch Carpodacus mexicanus roosting behaviour during the non-breeding season and possible effects of mycoplasmal conjunctivitis. Ibis, 149, 19.Google Scholar
Dhondt, A.A., States, S.L., Dhondt, K. . & Schat, K.A. (2012) Understanding the origin of seasonal epidemics of mycoplasmal conjunctivitis. Journal of Animal Ecology, 81, 9961003.Google Scholar
Dhondt, A.A., Tessaglia, D.L. & Slothower, R.L. (1998) Epidemic mycoplasmal conjunctivitis in house finches from Eastern North America. Journal of Wildlife Disease, 34, 265280.Google Scholar
Dhondt, K.V., Dhondt, A.A. & Ley, D.H. (2007c) Effects of route of inoculation on Mycoplasma gallisepticum infection in captive house finches. Avian Pathology, 36, 475479.Google Scholar
Duckworth, R.A., Badyaev, A.V., Farmer, K.L., Hill, G.E. & Roberts, S.R. (2003) First case of Mycoplasma gallisepticum infection in the western range of the house finch (Carpodacus mexicanus). Auk, 120, 528530.Google Scholar
Ebert, D. & Herre, E.A. (1996) The evolution of parasitic diseases. Parasitology Today, 12, 96101.Google Scholar
Elliott, J. & Arbib, R. Jr (1953) Origin and status of the house finch in the eastern United States. Auk, 70, 3137.Google Scholar
Faustino, C.R., Jennelle, C.S., Connolly, V., et al. (2004) Mycoplasma gallisepticum infection dynamics in a house finch population: seasonal variation in survival, encounter and transmission rate. Journal of Animal Ecology, 73, 651669.Google Scholar
Ferguson, N.A., Hermes, D., Leiting, V.A. & Kleven, S.H. (2003) Characterization of a naturally occurring infection of a Mycoplasma gallisepticum house finch-like strain in turkey breeders. Avian Diseases, 47, 523530.Google Scholar
Fischer, J.R., Stallknecht, D.E., Luttrell, M.P., Dhondt, A.A. & Converse, K.A. (1997) Mycoplasmal conjunctivitis in wild songbirds: the spread of a new contagious disease in a mobile host population. Emerging Infectious Diseases, 3, 6972.Google Scholar
Hartup, B.K., Bickal, J.M., Dhondt, A.A., Ley, D.H. & Kollias, G.V. (2001a) Dynamics of conjunctivitis and Mycoplasma gallisepticum infections in house finches. Auk, 118, 327333.Google Scholar
Hartup, B.K., Dhondt, A.A., Sydenstricker, K.V., Hochachka, W.M. & Kollias, G.V. (2001b) Host range and dynamics of mycoplasmal conjunctivitis among birds in North America. Journal of Wildlife Disease, 37, 7281.Google Scholar
Hartup, B.K. & Kollias, G.V. (1999) Field investigation of Mycoplasma gallisepticum infections in house finch (Carpodacus mexicanus) eggs and nestlings. Avian Diseases, 43, 572576.Google Scholar
Hawley, D.M., Briggs, J., Dhondt, A.A. & Lovette, I.J. (2008) Reconciling molecular signatures across markers: mitochondrial DNA confirms founder effect in invasive North American house finches (Carpodacus mexicanus). Conservation Genetics, 9, 637643.Google Scholar
Hawley, D.M., Davis, A.K. & Dhondt, A.A. (2007) Transmission-relevant behaviours shift with pathogen infection in wild house finches (Carpodacus mexicanus). Canadian Journal of Zoology – Revue Canadienne De Zoologie, 85, 752757.Google Scholar
Hawley, D.M., Dhondt, K.V., Dobson, A.P., et al. (2010) Common garden experiment reveals pathogen isolate but no host genetic diversity effect on the dynamics of an emerging wildlife disease. Journal of Evolutionary Biology, 23, 16801688.Google Scholar
Hawley, D.M., Hanley, D., Dhondt, A.A. & Lovette, I.J. (2006) Molecular evidence for a founder effect in invasive house finch (Carpodacus mexicanus) populations experiencing an emergent disease epidemic. Molecular Ecology, 15, 263275.Google Scholar
Hawley, D.M., Osnas, E.E., Dobson, A.P., et al. (2013) Parallel patterns of increased virulence in a recently emerged wildlife pathogen. PLoS Biology, 11(5), e1001570.Google Scholar
Hengeveld, H. 1989. Dynamics of Biological Invasions., London: Chapman and Hall.Google Scholar
Hill, G.E., Badyaev, A.V. & Belloni, V. (2012) House finch (Haemorhous mexicanus). In Rodewald, P.G. (ed.), The Birds of North America. Ithaca, NY:Cornell Laboratory of Ornithology.Google Scholar
Hochachka, W.M. & Dhondt, A.A. (2000) Density-dependent decline of host abundance resulting from a new infectious disease. Proceedings of the National Academy of Sciences of the United States of America, 97, 53035306.Google Scholar
Hochachka, W.M. & Dhondt, A.A. (2006) House finch (Carpodacus mexicanus) population- and group-level responses to a bacterial disease. Ornithological Monographs, 60(1), 3043.Google Scholar
Hochachka, W.M., Dhondt, A.A., Dobson, A., et al. (2013) Multiple host transfers, but only one successful lineage in a continent-spanning emergent pathogen. Proceedings of the Royal Society of London B, 280(1766).Google Scholar
Hochberg, M.E. & Holt, R.D. (1990) The coexistence of competing parasites. 1. The role of cross-species infection. American Naturalist, 136, 517541.Google Scholar
Holt, R.D., Dobson, A.P., Begon, M., Bowers, R.G. & Schauber, E.M. (2003) Parasite establishment in host communities. Ecology Letters, 6, 837842.Google Scholar
Hosseini, P.R., Dhondt, A.A. & Dobson, A. (2004) Seasonality and wildlife disease: how seasonal birth, aggregation and variation in immunity affect the dynamics of Mycoplasma gallisepticum in house finches. Proceedings of the Royal Society of London B, 271, 25692577.Google Scholar
Hosseini, P.R., Dhondt, A.A. & Dobson, A.P. (2006) Spatial spread of an emerging infectious disease: conjunctivitis in house finches. Ecology, 87, 30373046.Google Scholar
Jennelle, C.S., Cooch, E.G., Conroy, M.J. & Senar, J.C. (2007) State-specific detection probabilities and disease prevalence. Ecological Applications, 17, 154167.Google Scholar
Levisohn, S. & Kleven, S.H. (2000) Avian mycoplasmosis (Mycoplasma gallisepticum). Revue Scientifique et Technique de l’Office International des Epizooties, 19, 425442.Google Scholar
Ley, D.H., Berkhoff, J.E. & Mclaren, J.M. (1996) Mycoplasma gallisepticum isolated from house finches (Carpodacus mexicanus) with conjunctivitis. Avian Diseases, 40, 480483.Google Scholar
Ley, D.H., Hawley, D.M., Geary, S.J. & Dhondt, A.A. (2016) House finch (Haemorhous mexicanus) conjunctivitis, and Mycoplasma spp. isolated from North American wild birds, 1994–2015. Journal of Wildlife Diseases, 52, 669673.Google Scholar
Ley, D.H., Sheaffer, D.S. & Dhondt, A.A. 2006. Further western spread of Mycoplasma gallisepticum infection of house finches. Journal of Wildlife Diseases, 42, 429431.Google Scholar
Murray, J.D. (1993) Mathematical Biology. Second corrected edition. Berlin: Springer-Verlag.Google Scholar
Newton, I. (1972). Finches. London: Collins.Google Scholar
Osnas, E.E. & Dobson, A.P. (2010) Evolution of virulence when transmission occurs before disease. Biology Letters, 6, 505508.Google Scholar
Osnas, E.E. & Dobson, A.P. (2012) Evolution of virulence in heterogeneous host communities under multiple trade-offs. Evolution, 66, 391401.Google Scholar
Osnas, E.E., Hurtado, P.J. & Dobson, A.P. (2015) Evolution of pathogen virulence across space during an epidemic. American Naturalist, 185, 332342.Google Scholar
Pflaum, K., Tulman, E.R., Beaudet, J., Liao, X. & Geary, S.J. (2016) Global changes in Mycoplasma gallisepticum phase-variable lipoprotein gene vlhA expression during in vivo infection of the natural chicken host. Infection and Immunity, 84, 351355.Google Scholar
Potapov, A., Merrill, E., Pybus, M., Coltman, D. & Lewis, M.A. (2013) Chronic wasting disease: possible transmission mechanisms in deer. Ecological Modelling, 250, 244257.Google Scholar
Roberts, S.R., Nolan, P.M., Lauerman, L.H., Li, L.Q. & Hill, G.E. (2001) Characterization of the mycoplasmal conjunctivitis epizootic in a house finch population in the southeastern USA. Journal of Wildlife Diseases, 37, 8288.Google Scholar
Sauer, J.R., Fallon, J.E. & Johnson, R. (2003) Use of North American Breeding Bird Survey data to estimate population change for bird conservation regions. Journal of Wildlife Management, 67, 372389.Google Scholar
Shubber, Z., Mishra, S., Vesga, J.F. & Boily, M.C. (2014) The HIV Modes of Transmission model: a systematic review of its findings and adherence to guidelines. Journal of the International Aids Society, 17(1), 18928.Google Scholar
States, S.L., Hochachka, W.M. & Dhondt, A.A. (2009) Spatial variation in an avian host community: implications for disease dynamics. Ecohealth, 6, 540545.Google Scholar
Sydenstricker, K.V., Dhondt, A.A., Ley, D.H. & Kollias, G.V. (2005) Re-exposure of captive house finches that recovered from Mycoplasma gallisepticum infection. Journal of Wildlife Diseases, 41, 326333.Google Scholar
Tulman, E.R., Liao, X., Szczepanek, S.M., et al. (2012) Extensive variation in surface lipoprotein gene content and genomic changes associated with virulence during evolution of a novel North American house finch epizootic strain of Mycoplasma gallisepticum. Microbiology, 158, 20732088.Google Scholar
Williams, P.D., Dobson, A.P., Dhondt, K.V., Hawley, D.M. & Dhondt, A.A. (2014) Evidence of trade-offs shaping virulence evolution in an emerging wildlife pathogen. Journal of Evolutionary Biology, 27, 12711278.Google Scholar
Woolhouse, M.E.J., Haydon, D.T. & Antia, R. (2005) Emerging pathogens: the epidemiology and evolution of species jumps. Trends in Ecology & Evolution, 20, 238244.Google Scholar

References

Allen, J.E. & Maizels, R.M. (2011) Diversity and dialogue in immunity to helminths. Nature Reviews Immunology, 11, 375388.Google Scholar
Allen, J.E. & Sutherland, T.E. (2014) Host protective roles of type 2 immunity: parasite killing and tissue repair, flip sides of the same coin, Seminars in Immunology, 26, 329340.Google Scholar
Allotey, P. & Gyapong, M. (2008) Gender in tuberculosis research. International Journal of Tuberculosis and Lung Disease, 12, 831836.Google Scholar
Anderson, R.M. & Gordon, D.M. (1982) Processes influencing the distribution of parasite numbers within host populations with special emphasis on parasite-induced host mortalities. Parasitology, 85, 373398.Google Scholar
Anderson, R.M. & May, R.M. (1978) Regulation and stability of host–parasite population interactions I: regulatory processes. Journal of Animal Ecology, 47, 219247.Google Scholar
Anthony, R.M., Rutitzky, L.I., Urban, J.F., Stadecker, M.J. & Gause, W.C. (2007) Protective immune mechanisms in helminth infection. Nature Reviews Immunology, 7, 975987.Google Scholar
Audebert, F., Cassone, J., Hoste, H. & Durette-Desset, M.C. (2000) Morphogenesis and distribution of Trichostrongylus retortaeformis in the intestine of the rabbit. Journal of Helminthology, 74, 95107.Google Scholar
Audebert, F. & Durette-Desset, M.C. (2007) Do lagomorphs play a relay role in the evolution of the Trichostrongylina nematodes? Parasite, 14, 183197.Google Scholar
Audebert, F., Vuong, P.N. & Durette-Desset, M.C. (2003) Intestinal migrations of Trichostrongylus retortaeformis (Trichostrongylina, Trichostrongylidae) in the rabbit. Veterinary Parasitology, 112, 131146.Google Scholar
Blackwell, A.D., Gurven, M.D., Sugiyama, L.S., et al. (2011) Evidence for a peak shift in a humoral response to helminths: age profiles of IgE in the Shuar of Ecuador, the Tsimane of Bolivia, and the US NHANES. PLoS Neglected Tropical Diseases, 5, e1218.Google Scholar
Blackwell, A.D., Martin, M., Kaplan, H. & Gurven, M. (2013) Antagonism between two intestinal parasites in humans: the importance of co-infection for infection risk and recovery dynamics. Proceedings of the Royal Society of London B, 280, 20131671.Google Scholar
Bleay, C., Wilkes, C.P., Paterson, S. & Viney, M.E. (2007) Density-dependent immune responses against the gastrointestinal nematode Strongyloides ratti. International Journal for Parasitology, 37, 15011509.Google Scholar
Bowers, R.G. (1999) A baseline model for the apparent competition between many host strains: the evolution of host resistance. Journal of Theoretical Biology, 200, 6575.Google Scholar
Bourke, C.D., Maizels, R.M. & Mutapi, F. (2011) Acquired immune heterogeneity and its sources in human helminth infection. Parasitology, 138, 139159.Google Scholar
Brady, M.T., O’Neill, S.M., Dalton, J.P. & Mills, K.H. (1999) Fasciola hepatica suppresses a protective Th1 response against Bordetella pertussis. Infection and Immunity, 67, 53725378.Google Scholar
Brogden, K.A., Lehmkuhl, H.D. & Cutlip, R.C. (1998) Pasteurella haemolytica complicated respiratory infections in sheep and goats. Veterinary Research, 29, 233254.Google Scholar
Brooker, S., Akhwale, W., Pullan, R., et al. (2007) Epidemiology of Plasmodium–helminth co-infection in Africa: populations at risk, potential impact on anemia, and prospects for combining control. The American Journal of Tropical Medicine and Hygiene, 77(S6), 8898.Google Scholar
Cattadori, I.M., Boag, B., Bjørnstad, O.N., Cornell, S. & Hudson, P.J. (2005) Immuno-epidemiology and peak shift in a seasonal host-nematode system. Proceedings of the Royal Society of London B, 272, 11631169.Google Scholar
Cattadori, I.M., Boag, B. & Hudson, P.J. (2008) Parasite co-infection and interaction as drivers of host heterogeneity. International Journal for Parasitology, 38, 371380.Google Scholar
Cattadori, I.M., Pathak, A.K. & Ferrari, M.J. (2019) Changes in helminth–host interactions under external disturbances: dynamics of infection, parasite traits and host immune responses. Under review.Google Scholar
Cattadori, I.M., Wagner, B.R., Wodzinski, L.A., et al. (2014) Infections do not predict shedding in co-infections with two helminths from a natural system. Ecology, 95, 16841692.Google Scholar
Chase-Topping, M., Gally, D., Low, C., Matthews, L. & Woolhouse, M. (2008) Super-shedding and the link between human infection and livestock carriage of Escherichia coli O157. Nature Reviews Microbiology, 6, 904912.Google Scholar
Cizauskas, C.A., Turner, W.C., Pitts, N. & Getz, W.M. (2015) Seasonal patterns of hormones, macroparasites, and microparasites in wild African ungulates: the interplay among stress, reproduction, and disease. PLoS ONE, 10, 0120800.Google Scholar
Cornell, S., Bjørnstad, O.N., Cattadori, I.M., Boag, B. & Hudson, P.J. (2008) Seasonality, cohort-dependence and the development of immunity in a natural host-nematode system. Proceedings of the Royal Society of London B, 275, 473591.Google Scholar
Curtale, F., Wahab Hassanein, Y.A., Barduagni, P., et al. (2007) Human fascioliasis infection: gender differences within school-age children from endemic areas of the Nile Delta, Egypt. Transactions of the Royal Society of Tropical Medicine and Hygiene, 101, 155160.Google Scholar
Duerr, H.P., Dietz, K. & Eichner, M. (2003) On the interpretation of age–intensity profiles and dispersion patterns in parasitological surveys. Parasitology, 126, 87101.Google Scholar
Elias, D., Akuffo, H. & Britton, S. (2006) Helminths could influence the outcome of vaccines against TB in the tropics. Parasite Immunology, 28, 507513.Google Scholar
Gao, L., Zhou, F., Li, X. & Jin, Q. (2010) HIV/TB co-infection in mainland China: a meta-analysis. PLoS ONE, 5, e10736.Google Scholar
Garske, T. & Rhodes, C.J. (2008) The effect of superspreading on epidemic outbreak size distributions. Journal of Theoretical Biology, 253, 228237.Google Scholar
Geerts, S. & Gryseels, S. (2000) Drug resistance in human helminths: current situation and lessons from livestock. Clinical Microbiology Reviews, 13, 207222.Google Scholar
Ghosh, S., Ferrari, M.J., Pathak, A.K. & Cattadori, I.M. (2018). Changes in parasite traits, rather than intensity, affect the dynamics of infection under external perturbation. PLoS Computational Biology, 14(6), e1006167.Google Scholar
Girgis, N.M., Gundra, U.M. & Loke, P. (2013) immune regulation during helminth infections. PLoS Pathogens, 9, e1003250.Google Scholar
Graham, A., Cattadori, I.M., Lloyd-Smith, J., Ferrari, M. & Bjornstad, O.N. (2007) Transmission consequences of co-infection: cytokines writ large? Trends in Parasitology, 6, 284291.Google Scholar
Grenfell, B. T. & Anderson, R. M. (1989) Pertussis in England and Wales: an investigation of transmission dynamics and control by mass vaccination. Proceedings of the Royal Society of London B, 236, 213252.Google Scholar
Harvell, D., Altizer, S., Cattadori, I.M., Harrington, L. & Weil, E. (2009) Climate change and wildlife diseases: when does the host matter the most? Ecology, 90, 912920.Google Scholar
Hayes, K.S., Bancroft, A.J., Goldrick, M., et al. (2010) Exploitation of the intestinal microflora by the parasitic nematode Trichuris muris. Science, 328, 13911394.Google Scholar
Hernandez, A.D., Poole, A. & Cattadori, I.M. (2013) Climate changes influence free-living stages of soil-transmitted parasites of European rabbits. Global Change Biology, 19, 10281042.Google Scholar
Hoffman, R.S. & Smith, A.T. (2005) “Order Lagomorpha”. In: Wilson, D.E. & Reeder, D.M. (eds.), Mammal Species of the World: A Taxonomic and Geographic Reference (3rd ed.). Baltimore, MD: Johns Hopkins University Press.Google Scholar
Hudson, P.J. & Dobson, A.P. (1989) Population biology of Trichostrongylus tenuis, a parasite of economic importance for red grouse management. Parasitology Today, 5, 283291.Google Scholar
Hudson, P.J., Perkins, S.E. & Cattadori, I.M. (2008) The emergence of wildlife disease and the application of ecology. In: Ostfeld, R. (ed.), Infectious Disease Ecology: Effects of Ecosystems on Disease and of Disease on Ecosystems, (1st edn, pp. 347367). Princeton, NJ: Princeton University Press.Google Scholar
Izhar, R. & Ben‐Ami, F. (2015) Host age modulates parasite infectivity, virulence and reproduction. Journal of Animal Ecology, 84, 10181028.Google Scholar
Jackson, D.W. & Rohani, P. (2014) Perplexities of pertussis: recent global epidemiological trends and their potential causes. Epidemiology and Infection, 142, 672684.Google Scholar
Jackson, J.A., Friberg, I.M., Little, S. & Bradley, J.E. (2009) Review series on helminths, immune modulation and the hygiene hypothesis: immunity against helminths and immunological phenomena in modern human populations: coevolutionary legacies? Immunology, 126, 1827.Google Scholar
James, C.E., Hudson, A.L. & Davey, M.W. (2009) Drug resistance mechanisms in helminths: is it survival of the fittest? Trends in Parasitology, 25, 328335.Google Scholar
Jia, T.W., Melville, S., Utzinger, J., King, C.H. & Zhou, X.N. (2012) Soil-transmitted helminth reinfection after drug treatment: a systematic review and meta-analysis. PLoS Neglected Tropical Diseases, 6, e1621.Google Scholar
Kao, R.R., Gravenor, M.B., Charleston, B., et al. (2007) Mycobacterium bovis shedding patterns from experimentally infected calves and the effect of concurrent infection with bovine viral diarrhoea virus. Journal of The Royal Society Interface, 4, 545551.Google Scholar
Keiser, J. & Utzinger, J. (2008) Efficacy of current drugs against soil-transmitted helminth infections: systematic review and meta-analysis. Journal of the American Medical Association, 299, 19371948.Google Scholar
Keymer, A. (1982) Density-dependent mechanisms in the regulation of intestinal helminth populations. Parasitology, 84, 573587.Google Scholar
Lass, S., Hudson, P.J., Thakar, J., et al. (2013) Generating super-shedders: co-infection increases bacterial load and egg production of a gastrointestinal helminth. Journal of the Royal Society Interface, 10, 20120588.Google Scholar
Lloyd-Smith, J.O., Schreiber, S.J., Kopp, P.E. & Getz, W.M. (2005) Superspreading and the effect of individual variation on disease emergence. Nature, 438, 355359.Google Scholar
Luong, L.T., Vigliotti, B.A. & Hudson, P.J. (2011) Strong density-dependent competition and acquired immunity constrain parasite establishment: implications for parasite aggregation. International Journal for Parasitology, 41, 505511.Google Scholar
Maizels, R.M. (2009) Parasite immunomodulation and polymorphisms of the immune system. Journal of Biology, 8, 62.Google Scholar
Massoni, J., Cassone, J., Durette-Desset, M.C. & Audebert, F. (2011) Development of Graphidium strigosum (Nematoda, Haemonchidae) in its natural host, the rabbit (Oryctolagus cuniculus) and comparison with several Haemonchidae parasites of ruminants. Parasitology Research, 109, 2536.Google Scholar
McRae, K.M., Stear, M.J., Good, B. & Keane, O.M. (2015) The host immune response to gastrointestinal nematode infection in sheep. Parasite Immunology, 37, 605613.Google Scholar
Mignatti, A., Boag, B. & Cattadori, I.M. (2016) Host immunity shapes the impact of climate changes on the dynamics of parasite infections. Proceedings of the National Academy of Sciences of the United States of America, 113, 29702975.Google Scholar
Miller, M. R., White, A. & Boots, M. (2005) The evolution of host resistance: tolerance and control as distinct strategies. Journal of Theoretical Biology, 236, 198207.Google Scholar
Molnár, P.K., Kutz, S.J., Hoar, B.M. & Dobson, A.P. (2013) Metabolic approaches to understanding climate change impacts on seasonal host–macroparasite dynamics. Ecology Letters, 16, 921.Google Scholar
Murphy, L., Nalpas, N., Stear, M. & Cattadori, I.M. (2011) The role of immunity on the dynamics of chronic gastrointestinal nematode infections of rabbits. Parasite Immunology, 33, 287302.Google Scholar
Murphy, L., Pathak, A.K. & Cattadori, I.M. (2013) A co-infection with two gastrointestinal nematodes alters host immune responses and only partially parasite dynamics. Parasite Immunology, 35, 421432.Google Scholar
Pathak, A.K., Boag, B., Poss, M., Harvill, E. & Cattadori, I.M. (2011) Seasonal incidence of Bordetella bronchiseptica in an age-structured free-living rabbit population. Epidemiology and Infection, 14, 110.Google Scholar
Pathak, A.K., Creppage, K.E., Werner, J.R. & Cattadori, I.M. (2010) Immune regulation of a chronic bacteria infection and consequences for pathogen transmission. BMC Microbiology, 10, 226.Google Scholar
Pathak, A.K., Pelensky, C., Boag, B. & Cattadori, I.M. (2012) Immuno-epidemiology of chronic bacterial and helminth co-infections: observations from the field and evidence from the laboratory. International Journal for Parasitology, 42, 647655.Google Scholar
Paull, S.H. & Johnson, P.T.J. (2014) Experimental warming drives a seasonal shift in the timing of host–parasite dynamics with consequences for disease risk. Ecology Letters, 4, 445453.Google Scholar
Pedersen, A.B. & Antonovics, J. (2013) Anthelmintic treatment alters the parasite community in a wild mouse host. Biology Letters, 9, 20130205.Google Scholar
Poulin, R. (2007) Evolutionary Ecology of Parasites (2nd edn). Princeton, NJ: Princeton University Press.Google Scholar
Quinnell, R.J., Medley, G.F. & Keymer, A.E. (1990) The regulation of gastrointestinal helminth populations. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 330, 191201.Google Scholar
Raberg, L., Sim, D. & Read, A.F. (2007) Disentangling genetic variation for resistance and tolerance to infectious diseases in animals. Science, 318, 812814.Google Scholar
Raffel, T.R., Romansic, J.M., Halstead, N.T., et al. (2013) Disease and thermal acclimation in a more variable and unpredictable climate. Nature Climate Change, 3, 146151.Google Scholar
Redpath, S.A., Fonseca, N.M. & Perona-Wright, G. (2014) Protection and pathology during parasite infection: IL-10 strikes the balance. Parasite Immunology, 36, 233252.Google Scholar
Restif, O. & Koella, J.C. (2004) Concurrent evolution of resistance and tolerance to pathogens. The American Naturalist, 164, 90102.Google Scholar
Reynolds, A., Lindström, J., Johnson, P.C. & Mable, B.K. (2016) Evolution of drug-tolerant nematode populations in response to density reduction. Evolutionary Applications, 9, 726738.Google Scholar
Roy, B.A. & Kirchner, J.W. (2000) Evolutionary dynamics of pathogen resistance and tolerance. Evolution, 54, 5163.Google Scholar
Sabatelli, L., Ghani, A.C., Rodrigues, L.C., Hotez, P.J. & Brooker, S. (2008) Modelling heterogeneity and the impact of chemotherapy and vaccination against human hookworm parasite. Journal of the Royal Society Interface, 5, 13291341.Google Scholar
Stear, M.J. & Bishop, S.C. (1999) The curvilinear relationship between worm length and fecundity of Teladorsagia circumcincta. International Journal for Parasitology, 29, 777780.Google Scholar
Stear, M.J., Boag, B., Cattadori, I.M. & Murphy, L. (2009) Genetic variation in resistance to mixed, predominantly Teladorsagia circumcincta nematode infections of sheep: from heritabilities to gene identification. Parasite Immunology, 31, 274282.Google Scholar
Stear, M.J., Strain, S. & Bishop, S.C. (1999) Mechanisms underlying resistance to nematode infection. International Journal for Parasitology, 29, 5156.Google Scholar
Thakar, J., Pathak, A.K., Murphy, L., Albert, R. & Cattadori, I.M. (2012) Network model of immune responses reveals key effectors to single and co-infection kinetics by a respiratory bacterium and a gastrointestinal helminth. PLoS Computational Biology, 8, e1002345.Google Scholar
Thompson, H.V. & King, C.M. (1994) The European Rabbit. Oxford: Oxford University Press.Google Scholar
Tompkins, D.M. & Hudson, P.J. (1999) Regulation of nematode fecundity in the ring-necked pheasant (Phasianus colchicus): not just density dependence. Parasitology, 118, 417423.Google Scholar
Van Kuren, A., Boag, B., Hrubar, E. & Cattadori, I.M. (2013) Variability in the intensity of nematode larvae from gastrointestinal tissues of a natural herbivore. Parasitology, 140, 632640.Google Scholar
Woolhouse, M.E.J. (1992) A theoretical framework for the immunoepidemiology of helminth infection. Parasite Immunology, 14, 563578.Google Scholar
Woolhouse, M.E. (1998) Patterns in parasite epidemiology: the peak shift. Parasitology Today, 14, 428434.Google Scholar
Yazdanbakhsh, M. & Sacks, D.L. (2010) Why does immunity to parasites take so long to develop? Nature Reviews Immunology, 10, 8081.Google Scholar

References

Addink, E.A., De Jong, S.M., Davis, S.A., et al. (2010) The use of high-resolution remote sensing for plague surveillance in Kazakhstan. Remote Sensing of Environment, 114, 674681.Google Scholar
Bartlett, M.S. (1957) Measles periodicity and community size. Journal of the Royal Statistical Society, Series A, 120, 4871.Google Scholar
Brinkerhoff, R.J., Collinge, S.K., Ray, C. & Gage, K.L. (2010) Rodent and flea abundance fail to predict a plague epizootic in black-tailed prairie dogs. Vector-Borne and Zoonotic Diseases, 10, 4752.Google Scholar
Davis, S., Begon, M., De Bruyn, L., et al. (2004) Predictive thresholds for plague in Kazakhstan. Science, 304, 736738.Google Scholar
Davis, S., Klassovskiy, N., Ageyev, V., et al. (2007a) Plague metapopulation dynamics in a natural reservoir: the burrow system as the unit of study. Epidemiology and Infection, 135, 740748.Google Scholar
Davis, S., Leirs, H., Viljugrein, H., et al. (2007b) Empirical assessment of a threshold model for sylvatic plague. Journal of the Royal Society Interface, 4, 649657.Google Scholar
Davis, S., Trapman, P., Leirs, H., Begon, M. & Heesterbeek, J.A.P. (2008) The abundance threshold for plague as a critical percolation phenomenon. Nature, 454, 634637.Google Scholar
Grassberger, P. (1983) On the critical behaviour of the general epidemic process and dynamical percolation. Mathematical Biosciences, 63, 157172.Google Scholar
Heier, L., Storvik, G.O., Davis, S.A., et al. (2011) Emergence, spread, persistence and fade-out of sylvatic plague in Kazakhstan. Proceedings of the Royal Society of London B, 278, 29152923.Google Scholar
Holt, R.D., Dobson, A.P., Begon, M., Bowers, R.G. & Schauber, E.M. (2003) Parasite establishment in host communities. Ecology Letters, 6, 837842.Google Scholar
Kausrud, K.L., Viljugrein, H., Frigessi, A., et al. (2007) Climatically-driven synchrony of gerbil populations allows large-scale plague outbreaks. Proceedings of the Royal Society of London B, 274, 19631969.Google Scholar
Keeling, M.J. & Rohani, P.R. (2008) Modeling Infectious Diseases in Humans and Animals. Princeton, NJ: Princeton University Press.Google Scholar
Levick, B., Laudisoit, A., Wilschut, L., et al. (2015) The perfect burrow, but for what? Identifying local habitat conditions promoting the presence of the hosts and vectors of the Kazakh plague system. PLoS ONE, 10, e0136962.Google Scholar
Lloyd-Smith, J.O., Cross, P.C., Briggs, C.J., et al. (2005) Should we expect population thresholds for wildlife disease? Trends in Ecology and Evolution, 20, 511519.Google Scholar
Morse, SS. (1995) Factors in the emergence of infectious diseases. Emerging Infectious Diseases, 1, 715.Google Scholar
Onishchenko, G.G. & Kutyrev, V.V. (2004) Natural Plague Foci in the Caucasus, Caspian Sea Region, Middle Asia and Siberia. Moscow: Meditsina [in Russian].Google Scholar
Reijniers, J., Begon, M., Ageyev, V. & Leirs, H. (2014) Plague epizootic cycles in Central Asia. Biology Letters, 10, 20140302.Google Scholar
Reijniers, J., Davis, S., Begon, M., et al. (2012) A curve of thresholds governs plague epizootics in Central Asia. Ecology Letters, 15, 554560.Google Scholar
Samia, N.I., Chan, K.S. & Stenseth, N.C. (2007) A generalized threshold mixed model for analyzing nonnormal nonlinear time series, with application to plague in Kazakhstan. Biometrika, 94, 101118.Google Scholar
Samia, N.I., Kausrud, K.L., Heesterbeek, H., et al. (2011) Dynamics of the plague–wildlife–human system in Central Asia are controlled by two epidemiological thresholds. Proceedings of the National Academy of Sciences of the United States of America, 108, 14,52714,532.Google Scholar
Stenseth, N.C., Atshabar, B.B., Begon, M., et al. (2008) Plague: past, present and future. PLoS Medicine, 5, 913.Google Scholar
Webb, C.T., Brooks, C.P., Gage, K.L. & Antolin, M.F. (2006) Classic flea borne transmission does not drive plague epizootics in prairie dogs. Proceedings of the National Academy of Sciences of the United States of America, 103, 62366241.Google Scholar
Wilschut, L.I., Addink, E.A. Heesterbeek, J.A.P., et al. (2013a) Mapping the distribution of the main host for plague in a complex landscape in Kazakhstan: an object-based approach using SPOT-5 XS, Landsat 7 ETM+, SRTM and multiple Random Forests. International Journal of Applied Earth Observation and Geoinformation, 23, 8194.Google Scholar
Wilschut, L.I., Addink, E.A., Heesterbeek, H., et al. (2013b) Potential corridors and barriers for plague spread in central Asia. International Journal of Health Geographics, 12, 49.Google Scholar
Wilschut, L.I., Laudisoit, A., Hughes, N.K., et al. (2015) Spatial distribution patterns of plague hosts: point pattern analysis of the burrows of great gerbils in Kazakhstan. Journal of Biogeography, 42, 12811292.Google 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
×