Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-dnltx Total loading time: 0 Render date: 2024-04-23T11:00:21.627Z Has data issue: false hasContentIssue false

Part II - Understanding between-host processes

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. 223 - 426
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

Abbott, K.C. & Dwyer, G. (2008) Using mechanistic models to understand synchrony in forest insect populations: the North American Gypsy Moth as a case study. American Naturalist, 172, 613624.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., Harvell, D. & Friedle, E. (2003) Rapid evolutionary dynamics and disease threats to biodiversity. Trends in Ecology & Evolution, 18, 589596.Google Scholar
Anderson, R.M. & May, R.M. (1978) Regulation and stability of host–parasite population interactions: I. Regulatory processes. The Journal of Animal Ecology, 47, 219247.CrossRefGoogle 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 Disease of Humans: Dynamics and Control. Oxford: Oxford University Press.Google Scholar
Baldwin, I.T. (1998) Jasmonate-induced responses are costly but benefit plants under attack in native populations. Proceedings of the National Academy of Sciences of the United States of America, 95, 81138118.Google Scholar
Barbosa, P. & Krischik, V.A. (1987) Influence of alkaloids on feeding preference of eastern deciduous forest trees by the gypsy moth Lymantria dispar. American Naturalist, 130, 5369.Google Scholar
Bjørnstad, O.N., Robinet, C. & Liebhold, A.M. (2010) Geographic variation in North American gypsy moth cycles: subharmonics, generalist predators, and spatial coupling. Ecology, 91, 106118.Google Scholar
Bonsall, M.B., Sait, S.M. & Hails, R.S. (2005) Invasion and dynamics of covert infection strategies in structured insect–pathogen populations. Journal of Animal Ecology, 74, 464474.Google Scholar
Boots, M., Greenman, J., Ross, D., et al. (2003) The population dynamical implications of covert infections in host–microparasite interactions. Journal of Animal Ecology, 72, 10641072.Google Scholar
Burden, J.P., Nixon, C.P., Hodgkinson, A.E., et al. (2003) Covert infections as a mechanism for long-term persistence of baculoviruses. Ecology Letters, 6, 524531.Google Scholar
Burnham, K. & Anderson, D. (2002) Model Selection and Multimodal Inference: A Practical Information-theoretic Approach. New York, NY: Springer.Google Scholar
Chamberlin, T.C. (1897) Studies for students: the method of multiple working hypotheses. The Journal of Geology, 5, 837848.Google Scholar
Clark, J.S. (2007) Models for Ecological Data: An Introduction. Princeton, NJ: Princeton University Press.Google Scholar
Cory, J.S. & Hoover, K. (2006) Plant-mediated effects in insect–pathogen interactions. Trends in Ecology & Evolution, 21, 278286.Google Scholar
D’Amico, V., Elkinton, J.S., Podgwaite, J.D., Buonaccorsi, J. & Dwyer, G. (2005) Pathogen clumping: an explanation for non-linear transmission of an insect virus. Ecological Entomology, 30, 383390.Google Scholar
Dieckmann, U. (2002) Adapative dynamics of pathogen-host interactions. In: Dieckmann, U., Sigmund, K. & Metz, H. (eds.), Adaptive Dynamics of Infectious Diseases: In Pursuit of Virulence Management (pp. 3959). Cambridge: Cambridge University Press.Google Scholar
Doane, C.C. & McManus, M.L. (eds.). (1981) The Gypsy Moth: Research Toward Integrated Pest Management. Washington, DC: USDA.Google Scholar
Duffy, M.A., Ochs, J.H., Penczykowski, R.M., et al. (2012) Ecological context influences epidemic size and parasite-driven evolution. Science, 335, 16361638.Google Scholar
Dwyer, G. (1991) The roles of density, stage, and patchiness in the transmission of an insect virus. Ecology, 72, 559574.Google Scholar
Dwyer, G., Dushoff, J., Elkinton, J.S., Burand, J.P. & Levin, S.A. (2002) Variation in susceptibility: lessons from an insect virus. In: Dieckmann, U., Sigmund, K. & Metz, H. (eds.), Adaptive Dynamics of Infectious Diseases: In Pursuit of Virulence Management (pp. 7484). Cambridge: Cambridge University Press.Google Scholar
Dwyer, G., Dushoff, J., Elkinton, J.S. & Levin, S.A. (2000) Pathogen-driven outbreaks in forest defoliators revisited: building models from experimental data. American Naturalist, 156, 105120.Google Scholar
Dwyer, G., Dushoff, J. & Yee, S.H. (2004) The combined effects of pathogens and predators on insect outbreaks. Nature, 430, 341345.CrossRefGoogle ScholarPubMed
Dwyer, G., Elkinton, J.S. & Buonaccorsi, J.P. (1997) Host heterogeneity in susceptibility and disease dynamics: tests of a mathematical model. American Naturalist, 150, 685707.Google Scholar
Dwyer, G., Firestone, J. & Stevens, T.E. (2005) Should models of disease dynamics in herbivorous insects include the effects of variability in host-plant foliage quality? American Naturalist, 165, 1631.Google Scholar
Eakin, L., Wang, M. & Dwyer, G. (2015) The effects of the avoidance of infectious hosts on infection risk in an insect–pathogen interaction. American Naturalist, 185, 100112.Google Scholar
Elderd, B.D. (2013) Developing models of disease transmission: insights from ecological studies of insects and their baculoviruses. PLoS Pathogens, 9, e1003372.CrossRefGoogle ScholarPubMed
Elderd, B.D. (2019) Bottom-up trait-mediated indirect effects decrease pathogen transmission in a tritrophic system. Ecology, 100, e02551.Google Scholar
Elderd, B.D., Dushoff, J. & Dwyer, G. (2008) Host–pathogen interactions, insect outbreaks, and natural selection for disease resistance. American Naturalist, 172, 829842.Google Scholar
Elderd, B.D., Dwyer, G. & Dukic, V. (2013) Population-level differences in disease transmission: a Bayesian analysis of multiple smallpox epidemics. Epidemics, 5, 146156.Google Scholar
Elderd, B.D. & Miller, T.E.X. (2016) Quantifying uncertainty in demographic models: Bayesian methods for Integral Projection Models (IPMs). Ecological Monographs, 86, 125144.Google Scholar
Elderd, B.D., Rehill, B.J., Haynes, K.J. & Dwyer, G. (2013) Induced plant defenses, host–pathogen interactions, and forest insect outbreaks. Proceedings of the National Academy of Sciences of the United States of America, 110, 14,97814,983.Google Scholar
Elderd, B.D. & Reilly, J. (2014) Warmer temperatures increase disease transmission and outbreak intensity in a host–pathogen system. Journal of Animal Ecology, 83, 838849.Google Scholar
Elkinton, J.S., Healy, W.M., Buonaccorsi, J.P., et al. (1996) Interactions among gypsy moths, white-footed mice, and acorns. Ecology, 77, 23322342.Google Scholar
Elkinton, J. & Liebhold, A. (1990) Population dynamics of gypsy moth in North America. Annual Review of Entomology, 35, 571596.Google Scholar
Ferrari, M.J., Grais, R.F., Bharti, N., et al. (2008) The dynamics of measles in sub-Saharan Africa. Nature, 451, 679684.Google Scholar
Fleming-Davies, A.E., Dukic, V., Andreasen, V. & Dwyer, G. (2015) Effects of host heterogeneity on pathogen diversity and evolution. Ecology Letters, 18, 12521261.Google Scholar
Fleming-Davies, A.E. & Dwyer, G. (2015) Phenotypic variation in overwinter environmental transmission of a baculovirus and the cost of virulence. American Naturalist, 186, 797806.Google Scholar
Fox, L.R. (1975) Cannibalism in natural-populations. Annual Review of Ecology and Systematics, 6, 87106.Google Scholar
Fuller, E., Elderd, B.D. & Dwyer, G. (2012) Pathogen persistence in the environment and insect–baculovirus interactions: disease-density thresholds, epidemic burnout, and insect outbreaks. American Naturalist, 179, E70E96.Google Scholar
Fuxa, J.R. (1982) Prevalence of viral infections in populations of fall armyworm, Spodoptera frugiperda, in Southeastern Louisiana. Environmental Entomology, 11, 239242.Google Scholar
Fuxa, J.R. (1987) Spodoptera frugiperda susceptibility to nuclear polyhedrosis-virus isolates with reference to insect migration. Environmental Entomology, 16, 218223.Google Scholar
Fuxa, J.R. & Geaghan, J.P. (1983) Multiple-regression analysis of factors affecting prevalence of nuclear polyhedrosis virus in Spodoptera frugiperda (Lepidoptera, Noctuidae) populations. Environmental Entomology, 12, 311316.Google Scholar
Fuxa, J., Mitchell, F. & Richter, A. (1988) Resistance of Spodoptera frugiperda (Lepidoptera: Noctuidae) to a nuclear polyhedrosis virus in the field and laboratory. BioControl, 33, 5563.Google Scholar
Gelman, A., Carlin, J.B., Stern, H.S. & Rubin, D.B. (2014) Bayesian Data Analysis. Boca Raton, FL: Taylor & Francis.Google Scholar
Godfree, R., Robertson, B., Bolger, T., Carnegie, M. & Young, A. (2011) An improved hexagon open-top chamber system for stable diurnal and nocturnal warming and atmospheric carbon dioxide enrichment. Global Change Biology, 17, 439451.Google Scholar
Grzywacz, D., Mushobozi, W.L., Parnell, M., Jolliffe, F. & Wilson, K. (2008) Evaluation of Spodoptera exempta nucleopolyhedrovirus (SpexNPV) for the field control of African armyworm (Spodoptera exempta) in Tanzania. Crop Protection, 27, 1724.Google Scholar
Hajek, A.E. (1999) Pathology and epizootiology of Entomophaga maimaiga infections in forest Lepidoptera. Microbiology and Molecular Biology Reviews, 63, 814835.Google Scholar
Hajek, A.E., Tobin, P.C. & Haynes, K.J. (2015) Replacement of a dominant viral pathogen by a fungal pathogen does not alter the collapse of a regional forest insect outbreak. Oecologia, 177, 785797.Google Scholar
Hall, S.R., Simonis, J.L., Nisbet, R.M., Tessier, A.J. & Caceres, C.E. (2009) Resource ecology of virulence in a planktonic host–parasite system: an explanation using dynamic energy budgets. American Naturalist, 174, 149162.Google Scholar
Harvell, C.D., Mitchell, C.E., Ward, J.R., et al. (2002) Ecology – Climate warming and disease risks for terrestrial and marine biota. Science, 296, 21582162.Google Scholar
Hassell, M.P., May, R.M., Pacala, S.W. & Chesson, P.L. (1991) The persistence of host–parasitoid associations in patchy environments. I. A general criterion. The American Naturalist, 138, 568583.Google Scholar
Hethcote, H.W. (1976) Qualitative analyses of communicable disease models. Mathematical Biosciences, 28, 335356.Google Scholar
Hilborn, R. & Mangel, M. (1997) The Ecological Detective: Confronting Models with Data. Princeton, NJ: Princeton University Press.Google Scholar
Hinds, W. & Dew, J. (1915) The grass worm or fall army worm. Bulletin no. 186, Alabama Agricultural Experiment Station.Google Scholar
Hobbs, N.T., Geremia, C., Treanor, J., et al. (2015) State-space modeling to support management of brucellosis in the Yellowstone bison population. Ecological Monographs, 85, 525556.Google Scholar
Hobbs, N.T. & Hooten, M.B. (2015) Bayesian Models: A Statistical Primer for Ecologists. Princeton, NJ: Princeton University Press.Google Scholar
Hooten, M.B. & Hobbs, N.T. (2015) A guide to Bayesian model selection for ecologists. Ecological Monographs, 85, 328.Google Scholar
Hoover, K., Grove, M., Gardner, M., et al. (2011) A gene for an extended phenotype. Science, 333, 14011401.Google Scholar
Hudson, A.I., Fleming-Davies, A.E., Páez, D.J. & Dwyer, G. (2016) Genotype-by-genotype interactions between an insect and its pathogen. Journal of Evolutionary Biology, 29, 24802490.Google Scholar
Hudson, P.J., Dobson, A.P. & Newborn, D. (1998) Prevention of population cycles by parasite removal. Science, 282, 22562258.Google Scholar
Hudson, P.J., Rizzoli, A., Grenfell, B.T., Heesterbeek, H. & Dobson, A.P. (2002) The Ecology of Wildlife Diseases. Oxford: Oxford University Press.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
Hunter, M. (2016) The Phytochemical Landscape: Linking Trophic Interactions and Nutrient Dynamics. Princeton, NJ:: Princeton University Press.Google Scholar
Hunter-Fujita, F.R., Entwistle, P., Evans, H., et al. (1998) Insect Viruses and Pest Management. Chichester: John Wiley & Sons Ltd.Google Scholar
Ionides, E., Bretó, C. & King, A. (2006) Inference for nonlinear dynamical systems. Proceedings of the National Academy of Sciences of the United States of America, 103, 18,43818,443.Google Scholar
Johnson, D.M., Liebhold, A.M. & Bjørnstad, O.N. (2006) Geographical variation in the periodicity of gypsy moth outbreaks. Ecography, 29, 367374.Google Scholar
Johnson, D.M., Liebhold, A.M., Bjørnstad, O.N. & McManus, M.L. (2005) Circumpolar variation in periodicity and synchrony among gypsy moth populations. Journal of Animal Ecology, 74, 882892.Google Scholar
Johnson, P.T., De Roode, J.C. & Fenton, A. (2015) Why infectious disease research needs community ecology. Science, 349, 1259504.Google Scholar
Karban, R. & Baldwin, I.T. (1997) Induced Responses to Herbivory. Chicago, IL: University of Chicago Press.Google Scholar
Keeling, M.J. & Rohani, P. (2008) Modeling Infectious Diseases in Humans and Animals. Princeton, NJ: Princeton University Press.Google Scholar
Kendall, B.E., Briggs, C.J., Murdoch, W.W., et al. (1999) Why do populations cycle? A synthesis of statistical and mechanistic modeling approaches. Ecology, 80, 17891805.Google Scholar
Kennedy, D.A., Dukic, V. & Dwyer, G. (2014) Pathogen growth in insect hosts: inferring the importance of different mechanisms using stochastic models and response-time data. American Naturalist, 184, 407423.Google Scholar
Kéry, M. (2010) Introduction to WinBUGS for Ecologists. Boston, MA: Academic Press.Google Scholar
King, A.A., Ionides, E.L., Pascual, M. & Bouma, M.J. (2008) Inapparent infections and cholera dynamics. Nature, 454, 877880.Google Scholar
Konishi, S. & Kitagawa, G. (2008) Information Criteria and Statistical Modeling. New York, NY: Springer Science & Business Media.Google Scholar
Kot, M. (2001) Elements of Mathematical Ecology. Cambridge: Cambridge University Press.Google Scholar
LaDeau, S.L., Glass, G.E., Hobbs, N.T., Latimer, A. & Ostfeld, R.S. (2011) Data–model fusion to better understand emerging pathogens and improve infectious disease forecasting. Ecological Applications, 21, 14431460.Google Scholar
Lee, K.P., Cory, J.S., Wilson, K., Raubenheimer, D. & Simpson, S.J. (2006) Flexible diet choice offsets protein costs of pathogen resistance in a caterpillar. Proceedings of the Royal Society of London B, 273, 823829.Google Scholar
Levin, S.A. (1992) The problem of pattern and scale in ecology: the Robert H. MacArthur award lecture. Ecology, 73, 19431967.Google Scholar
Link, W.A. & Barker, R.J. (2010) Bayesian Inference with Ecological Applications. London: Academic Press.Google Scholar
May, R.M. & Anderson, R.M. (1978) Regulation and stability of host–parasite population interactions: II. Destabilizing processes. Journal of Animal Ecology, 47, 249267.Google Scholar
May, R.M. & Anderson, R.M. (1979) Population biology of infectious diseases: Part II. Nature, 280, 455461.Google Scholar
McCallum, H. (2016) Models for managing wildlife disease. Parasitology, 143, 805820.Google Scholar
McCallum, H., Fenton, A., Hudson, P., et al. (2017) Breaking beta: deconstructing the parasite transmission function. Philosophical Transactions of the Royal Society of London B, 372, 20160084.Google Scholar
Miller, L.K. (2013) The Baculoviruses. New York, NY: Springer Science & Business Media.Google Scholar
Mitchell, E., McNeil, J., Westbrook, J., et al. (1991) Seasonal periodicity of fall armyworm (Lepidoptera: Noctuidae) in the Caribbean basin and northward to Canada. Journal of Entomological Science, 26, 3950.Google Scholar
Moreau, G. & Lucarotti, C.J. (2007) A brief review of the past use of baculoviruses for the management of eruptive forest defoliators and recent developments on a sawfly virus in Canada. The Forestry Chronicle, 83, 105112.Google Scholar
Myers, J.H. (1988) Can a general hypothesis explain population cycles of forest Lepidoptera? Advances in Ecological Research, 18, 179242.Google Scholar
Myers, J.H. & Cory, J.S. (2016) Ecology and evolution of pathogens in natural populations of Lepidoptera. Evolutionary Applications, 9, 231247.Google Scholar
New, L.F., Matthiopoulos, J., Redpath, S. & Buckland, S.T. (2009) Fitting models of multiple hypotheses to partial population data: investigating the causes of cycles in red grouse. American Naturalist, 174, 399412.Google Scholar
Nykänen, H. & Koricheva, J. (2004) Damage-induced changes in woody plants and their effects on insect herbivore performance: a meta-analysis. Oikos, 104, 247268.Google Scholar
O’Neill, P.D. (2002) A tutorial introduction to Bayesian inference for stochastic epidemic models using Markov chain Monte Carlo method. Mathematical Biosciences, 180, 103114.Google Scholar
Páez, D.J., Dukic, V., Dushoff, J., Fleming-Davies, A.E. & Dwyer, G. (2017) Eco-evolutionary theory and insect outbreaks. American Naturalist, 189, 616629.Google Scholar
Parker, B.J., Elderd, B.D. & Dwyer, G. (2010) Host behaviour and exposure risk in an insect–pathogen interaction. Journal of Animal Ecology, 79, 863870.Google Scholar
Pascual, M. & Bouma, M.J. (2009) Do rising temperatures matter? Ecology, 90, 906912.Google Scholar
Pepin, K.M., Kay, S.L., Golas, B.D., et al. (2017) Inferring infection hazard in wildlife populations by linking data across individual and population scales. Ecology Letters, 20, 275292.CrossRefGoogle ScholarPubMed
Pfennig, D.W. (2000) Effect of predator–prey phylogenetic similarity on the fitness consequences of predation: a trade-off between nutrition and disease? American Naturalist, 155, 335345.Google Scholar
Pitre, H.N. & Hogg, D.B. (1983) Development of the fall armyworm (Lepidoptera, Noctuidae) on cotton, soybean and corn. Journal of the Georgia Entomological Society, 18, 182187.Google Scholar
Polis, G.A. (1981) The evolution and dynamics of intraspecific predation. Annual Review of Ecology and Systematics, 12, 225251.Google Scholar
Redman, E.M., Wilson, K. & Cory, J.S. (2016) Trade-offs and mixed infections in an obligate-killing insect pathogen. Journal of Animal Ecology, 85, 12001209.Google Scholar
Reeson, A.F., Wilson, K., Cory, J.S., et al. (2000) Effects of phenotypic plasticity on pathogen transmission in the field in a Lepidoptera–NPV system. Oecologia, 124, 373380.Google Scholar
Reeson, A.F., Wilson, K., Gunn, A., Hails, R.S. & Goulson, D. (1998) Baculovirus resistance in the noctuid Spodoptera exempta is phenotypically plastic and responds to population density. Proceedings of the Royal Society of London B, 265, 17871791.Google Scholar
Reilly, J.R. & Elderd, B.D. (2014) Effects of biological control on long-term population dynamics: identifying unexpected outcomes. Journal of Applied Ecology, 51, 90101.Google Scholar
Richardson, M.L., Mitchell, R.F., Reagel, P.F. & Hanks, L.M. (2010) Causes and consequences of cannibalism in noncarnivorous insects. Annual Review of Entomology, 55, 3953.Google Scholar
Richter, A.R., Fuxa, J.R. & Abdelfattah, M. (1987) Effect of host plant on the susceptibility of Spodoptera frugiperda (Lepidoptera, Noctuidae) to a nuclear polyhedrosis virus. Environmental Entomology, 16, 10041006.Google Scholar
Rohrmann, G.F. (2014) Baculovirus nucleocapsid aggregation (MNPV vs SNPV): an evolutionary strategy, or a product of replication conditions? Virus Genes, 49, 351357.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
Rudolf, V.H. (2007) Consequences of stage-structured predators: cannibalism, behavioral effects, and trophic cascades. Ecology, 88, 29913003.Google Scholar
Rudolf, V.H.W., Kamo, M. & Boots, M. (2010) Cannibals in space: the coevolution of cannibalism and dispersal in spatially structured populations. American Naturalist, 175, 513524.Google Scholar
Schultz, J.C. & Baldwin, I.T. (1982) Oak leaf quality declines in response to defoliation by gypsy moth larvae. Science, 217, 149151.Google Scholar
Shikano, I., Shumaker, K.L., Peiffer, M., Felton, G.W. & Hoover, K. (2017) Plant-mediated effects on an insect–pathogen interaction vary with intraspecific genetic variation in plant defences. Oecologia, 183, 11211134.Google Scholar
Spiegelhalter, D.J., Best, N.G., Carlin, B.R. & van der Linde, A. (2002) Bayesian measures of model complexity and fit. Journal of the Royal Statistical Society Series B Statistical Methodology, 64, 583616.Google Scholar
Thompson, J. (2005) The Geographic Mosaic of Coevolution. Chicago, IL: University of Chicago Press.Google Scholar
Underwood, N., Rausher, M. & Cook, W. (2002) Bioassay versus chemical assay: measuring the impact of induced and constitutive resistance on herbivores in the field. Oecologia, 131, 211219.Google Scholar
Valicente, F.H., Tuelher, E.S., Pena, R.C., Andreazza, R. & Guimaraes, M.R.F. (2013) Cannibalism and virus production in Spodoptera frugiperda (JE Smith) (Lepidoptera: Noctuidae) larvae fed with two leaf substrates inoculated with Baculovirus spodoptera. Neotropical Entomology, 42, 191199.Google Scholar
Van Allen, B.G., Dillemuth, F.P., Flick, A.J., et al. (2017) Cannibalism and infectious disease: friend or foe? American Naturalist, 120, 299312.Google Scholar
Verity, R., Stevenson, M.D., Rossmo, D.K., Nichols, R.A. & Le Comber, S.C. (2014) Spatial targeting of infectious disease control: identifying multiple, unknown sources. Methods in Ecology and Evolution, 5, 647655.Google Scholar
Vilaplana, L., Wilson, K., Redman, E.M. & Cory, J.S. (2010) Pathogen persistence in migratory insects: high levels of vertically-transmitted virus infection in field populations of the African armyworm. Evolutionary Ecology, 24, 147160.Google Scholar
Wang, L., Hui, C., Sandhu, H.S., Li, Z. & Zhao, Z. (2015) Population dynamics and associated factors of cereal aphids and armyworms under global change. Scientific Reports, 5, 18801.Google Scholar
Watanabe, S. (2013) A widely applicable Bayesian information criterion. Journal of Machine Learning Research, 14, 867897.Google Scholar
Werner, E. & Peacor, S. (2003) A review of trait-mediated indirect interactions in ecological communities. Ecology, 84, 10831100.Google Scholar
Wild, S. (2017) African countries mobilize to battle invasive caterpillar. Nature, 543, 13.Google Scholar
Williams, D.W. & Liebhold, A.M. (1995) Herbivorous insects and global change: potential changes in the spatial distribution of forest defoliator outbreaks. Journal of Biogeography, 22, 665671.Google Scholar
Williams, T. & Hernádez, O. (2006) Costs of cannibalism in the presence of an iridovirus pathogen of Spodoptera frugiperda. Ecological Entomology, 31, 106113.Google Scholar
Wilson, K., Bjørnstad, O., Dobson, A., et al. (2002) Heterogeneities in macroparasite infections: Patterns and processes. In: Hudson, P., Rizzoli, A., Grenfell, B., Heesterbeek, H. & Dobson, A.P. (eds.), The Ecology of Wildlife Diseases (pp. 644). Oxford: Oxford University Press.Google Scholar
Wold, E.N. & Marquis, R.J. (1997) Induced defense in white oak: effects on herbivores and consequences for the plant. Ecology, 78, 13561369.Google Scholar
Woods, S. & Elkinton, J. (1987) Biomodal patterns of mortality from nuclear polyhedrosis virus in gypsy moth (Lymantria dispar) populations. Journal of Invertebrate Pathology,50,151157.Google Scholar

References

Adlard, R.D., Miller, T.L. & Smit, N.J. (2015) The butterfly effect: parasite diversity, environment, and emerging disease in aquatic wildlife. Trends in Parasitology, 31, 160166.Google Scholar
Anderson, L.G., Dunn, A.M., Rosewarne, P.J. & Stebbing, P.D. (2015) Invaders in hot water: a simple decontamination method to prevent the accidental spread of aquatic invasive non-native species. Biological Invasions, 17, 22872297.Google Scholar
Anderson, L.G., White, P.C.L., Stebbing, P.D., Stentiford, G.D. & Dunn, A.M. (2014) Biosecurity and vector behaviour: evaluating the potential threat posed by anglers and canoeists as pathways for the spread of invasive non-native species and pathogens. PLoS ONE, 9, e92788.Google Scholar
Arundell, K., Dunn, A., Alexander, J., et al. (2015) Enemy release and genetic founder effects in invasive killer shrimp populations of Great Britain. Biological Invasions, 17, 14391451.CrossRefGoogle Scholar
Bandi, C., Dunn, A.M., Hurst, G.D.D. & Rigaud, T. (2001) Inherited microorganisms, sex-specific virulence and reproductive parasitism. Trends in Parasitology, 17, 8894.Google Scholar
Bass, D., Stentiford, G.D., Littlewood, D.T.J. & Hartikainen, H. (2015) Diverse applications of environmental DNA methods in parasitology. Trends in Parasitology, 31, 499513.Google Scholar
Blackburn, T.M., Pysek, P., Bacher, S., et al. (2011) A proposed unified framework for biological invasions. Trends in Ecology & Evolution, 26, 333339.Google Scholar
Bojko, J. (2017) Parasites of invasive crustacea: risks and opportunities for control. PhD thesis, University of Leeds, UK.Google Scholar
Bojko, J., Bacela-Spychalska, K., Stebbing, P.D., et al. (2017) Parasites, pathogens and commensals in the ‘low-impact’ non-native amphipod host Gammarus roeselii. Parasites & Vectors, 10, 193.Google Scholar
Bojko, J., Dunn, A.M., Stebbing, P.D., et al. (2015) Cucumispora ornata n. sp (Fungi: Microsporidia) infecting invasive ‘demon shrimp’ (Dikerogammarus haemobaphes) in the United Kingdom. Journal of Invertebrate Pathology, 128, 2230.Google Scholar
Bojko, J., Stebbing, P.D., Bateman, K.S., et al. (2013) Baseline histopathological survey of a recently invading island population of ‘killer shrimp’, Dikerogammarus villosus. Diseases of Aquatic Organisms, 106, 241253.Google Scholar
Bovy, H.C., Barrios-O’Neill, D., Emmerson, M.C., Aldridge, D.C. & Dick, J.T.A. (2015) Predicting the predatory impacts of the ‘demon shrimp’ Dikerogammarus haemobaphes, on native and previously introduced species. Biological Invasions, 17, 597607.Google Scholar
Bowers, R.G. & Turner, J. (1997) Community structure and the interplay between interspecific infection and competition. Journal of Theoretical Biology, 187, 95109.CrossRefGoogle ScholarPubMed
Bunke, M., Alexander, M.E., Dick, J.T.A., et al. (2015) Eaten alive: cannibalism is enhanced by parasites. Royal Society Open Science, 2, 140369.Google Scholar
Claessen, D., de Roos, A.M. & Persson, L. (2004) Population dynamic theory of size-dependent cannibalism. Proceedings of the Royal Society B-Biological Sciences, 271, 333340.Google Scholar
Colautti, R.I., Ricciardi, A., Grigorovich, I.A. & MacIsaac, H.J. (2004) Is invasion success explained by the enemy release hypothesis? Ecology Letters, 7, 721733.Google Scholar
DAISIE (2017) DAISIE European Invasive Alien Species Gateway, accessed 2017.Google Scholar
Dick, J.T.A. (1996) Post-invasion amphipod communities of Lough Neagh, Northern Ireland: influences of habitat selection and mutual predation. Journal of Animal Ecology, 65, 756767.Google Scholar
Dick, J.T.A., Alexander, M.E., Jeschke, J.M., et al. (2014) Advancing impact prediction and hypothesis testing in invasion ecology using a comparative functional response approach. Biological Invasions, 16, 735753.Google Scholar
Dick, J.T.A., Armstrong, M., Clarke, H.C., et al. (2010) Parasitism may enhance rather than reduce the predatory impact of an invader. Biology Letters, 6, 636638.Google Scholar
Dick, J.T.A., Laverty, C., Lennon, J.J., et al. (2017) Invader Relative Impact Potential: a new metric to understand and predict the ecological impacts of existing, emerging and future invasive alien species. Journal of Applied Ecology, 54, 12591267.Google Scholar
Dick, J.T.A., Montgomery, I. & Elwood, R.W. (1993) Replacement of the indigenous amphipod Gammarus duebeni celticus by the introduced Gammarus pulex – differential cannibalism and mutual predation. Journal of Animal Ecology, 62, 7988.Google Scholar
Dick, J.T.A., Montgomery, W.I. & Elwood, R.W. (1999) Intraguild predation may explain an amphipod replacement: evidence from laboratory populations. Journal of Zoology, 249, 463468.Google Scholar
Dick, J.T.A. & Platvoet, D. (1996) Intraguild predation and species exclusions in amphipods: the interaction of behaviour, physiology and environment. Freshwater Biology, 36, 375383.Google Scholar
Dobson, A., Lafferty, K.D., Kuris, A.M., Hechinger, R.F. & Jetz, W. (2008) Homage to Linnaeus: how many parasites? How many hosts? Proceedings of the National Academy of Sciences of the United States of America, 105, 11,48211,489.Google Scholar
Dudgeon, D., Arthington, A.H., Gessner, M.O., et al. (2006) Freshwater biodiversity: importance, threats, status and conservation challenges. Biological Reviews, 81, 163182.Google Scholar
Dunn, A.M. & Dick, J.T.A. (1998) Parasitism and epibiosis in native and non-native gammarids in freshwater in Ireland. Ecography, 21, 593598.Google Scholar
Dunn, A.M. & Hatcher, M.J. (2015) Parasites and biological invasions: parallels, interactions, and control. Trends in Parasitology, 31, 189199.Google Scholar
Dunn, A.M., Torchin, M.E., Hatcher, M.J., et al. (2012) Indirect effects of parasites in invasions. Functional Ecology, 26, 12621274.Google Scholar
EU. (2014). Regulation (EU) No. 1143/2014 of the European Parliament and of the Council of 22 October 2014 on the prevention and management of the introduction and spread of invasive alien species. https://eur-lex.europa.eu/legal-content/EN/TXT/?uri=celex%3A32014R1143Google 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
Fielding, N.J., MacNeil, C., Robinson, N., et al. (2005) Ecological impacts of the microsporidian parasite Pleistophora mulleri on its freshwater amphipod host Gammarus duebeni celticus. Parasitology, 131, 331336.CrossRefGoogle ScholarPubMed
Filipova, L., Petrusek, A., Matasova, K., Delaunay, C. & Grandjean, F. (2013) Prevalence of the crayfish plague pathogen Aphanomyces astaci in populations of the signal crayfish Pacifastacus leniusculus in France: evaluating the threat to native crayfish. PLoS ONE, 8, e70157.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
GBNNSS (2015) The Great Britain Invasive Non Native Species Strategy.Google Scholar
Grabner, D.S., Weigand, A.M., Leese, F., et al. (2015) Invaders, natives and their enemies: distribution patterns of amphipods and their microsporidian parasites in the Ruhr Metropolis, Germany. Parasites & Vectors, 8, 419.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. (2008) A keystone effect for parasites in intraguild predation? Biology Letters, 4, 534537.Google Scholar
Hatcher, M.J., Dick, J.T.A. & Dunn, A.M. (2012a) Disease emergence and invasions. Functional Ecology, 26, 12751287.Google Scholar
Hatcher, M.J., Dick, J.T.A. & Dunn, A.M. (2012b) Diverse effects of parasites in ecosystems: linking interdependent processes. Frontiers in Ecology and the Environment, 10, 186194.Google Scholar
Hatcher, M.J., Dick, J.T.A. & Dunn, A.M. (2014) Parasites that change predator or prey behaviour can have keystone effects on community composition. Biology Letters, 10.Google Scholar
Hatcher, M.J., Dick, J.T.A., Paterson, R.A., et al. (2015) Trait-mediated effects of parasites on invader–native interactions. In: Mehlhorn, H. (ed.), Host Manipulations by Parasites and Viruses (pp. 2947). Cham: Springer.Google Scholar
Hatcher, M.J. & Dunn, A.M. (2011) Parasites in Ecological Communities; From Interactions to Ecosystems. Cambridge: Cambridge University Press.Google Scholar
Holt, R.D. & Dobson, A.P. (2006) Extending the principles of community ecology to address the epidemiology of host–pathogen systems. In: Collinge, S.K.R. (ed.), Disease Ecology: Community Structure and Pathogen Dynamics (pp. 627). Oxford: Oxford University Press.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
Holt, R.D. & Pickering, J. (1985) Infectious-disease and species coexistence – a model of Lotka–Volterra form. American Naturalist, 126, 196211.Google Scholar
Holt, R.D. & Polis, G.A. (1997) A theoretical framework for intraguild predation. American Naturalist, 149, 745764.Google Scholar
Hudson, P. & Greenman, J. (1998) Competition mediated by parasites: biological and theoretical progress. Trends in Ecology & Evolution, 13, 387390.Google Scholar
Hudson, P.J., Dobson, A.P. & Lafferty, K.D. (2006) Is a healthy ecosystem one that is rich in parasites? Trends in Ecology & Evolution, 21, 381385.Google Scholar
IUCN (2017) Global invasive species database. www.iucngisd.org/gisd/, accessed November 2017.Google Scholar
Keane, R.M. & Crawley, M.J. (2002) Exotic plant invasions and the enemy release hypothesis. Trends in Ecology & Evolution, 17, 164170.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
Kelly, D.W. & Dick, J.T.A. (2005) Introduction of the non-indigenous amphipod Gammarus pulex alters population dynamics and diet of juvenile trout Salmo trutta. Freshwater Biology, 50, 127140.Google Scholar
Kelly, D.W., Dick, J.T.A. & Montgomery, W.I. (2002) The functional role of Gammarus (Crustacea, Amphipoda): shredders, predators, or both? Hydrobiologia, 485, 199203.Google Scholar
Kenna, D., Fincham, W.N., Dunn, A.M., Brown, L.E. & Hassall, C. (2017) Antagonistic effects of biological invasion and environmental warming on detritus processing in freshwater ecosystems. Oecologia, 183, 875886.Google Scholar
Kumschick, S., Gaertner, M., Vila, M., et al. (2015) Ecological impacts of alien species: quantification, scope, caveats, and recommendations. Bioscience, 65, 5563.Google Scholar
Laverty, C., Brenner, D., McIlwaine, C., et al. (2017) Temperature rise and parasitic infection interact to increase the impact of an invasive species. International Journal for Parasitology, 47, 291296.Google Scholar
Ledger, M.E. & Milner, A.M. (2015) Extreme events in running waters. Freshwater Biology, 60, 24552460.Google Scholar
Lefevre, T., Lebarbenchon, C., Gauthier-Clerc, M., et al. (2009) The ecological significance of manipulative parasites. Trends in Ecology & Evolution, 24, 4148.Google Scholar
MacLeod, C.J., Paterson, A.M., Tompkins, D.M. & Duncan, R.P. (2010) Parasites lost – do invaders miss the boat or drown on arrival? Ecology Letters, 13, 516527.Google Scholar
MacNeil, C., Dick, J.T. & Elwood, R.W. (1997) The trophic ecology of freshwater Gammarus spp. (Crustacea: Amphipoda): problems and perspectives concerning the functional feeding group concept. Biological Reviews of the Cambridge Philosophical Society, 72, 349364.Google Scholar
MacNeil, C., Dick, J.T.A., Hatcher, M.J., et al. (2003a) Parasite-mediated predation between native and invasive amphipods. Proceedings of the Royal Society of London Series B, 270, 13091314.Google Scholar
MacNeil, C., Dick, J.T.A., Platvoet, D. & Briffa, M. (2011) Direct and indirect effects of species displacements: an invading freshwater amphipod can disrupt leaf-litter processing and shredder efficiency. Journal of the North American Benthological Society, 30, 3848.Google Scholar
MacNeil, C., Elwood, R.W. & Dick, J.T.A. (1999) Predator–prey interactions between brown trout Salmo trutta and native and introduced amphipods; their implications for fish diets. Ecography, 22, 686696.Google Scholar
MacNeil, C., Fielding, N.J., Dick, J.T.A., et al. (2003b) An acanthocephalan parasite mediates intraguild predation between invasive and native freshwater amphipods (Crustacea). Freshwater Biology, 48, 20852093.Google Scholar
MacNeil, C., Fielding, N.J., Hume, K.D., et al. (2003c) Parasite altered micro-distribution of Gammarus pulex (Crustacea: Amphipoda). International Journal for Parasitology, 33, 5764.Google Scholar
MacNeil, C., Platvoet, D., Dick, J.T.A., et al. (2010) The Ponto-Caspian ‘killer shrimp’, Dikerogammarus villosus (Sowinsky, 1894), invades the British Isles. Aquatic Invasions, 5, 441445.Google Scholar
Mitchell, C.E. & Power, A.G. (2003) Release of invasive plants from fungal and viral pathogens. Nature, 421, 625627.Google Scholar
Mouritsen, K.N. & Poulin, R. (2010) Parasitism as a determinant of community structure on intertidal flats. Marine Biology, 157, 201213.Google Scholar
Ohgushi, T., Schmitz, O. & Holt, R.D. (2012) Trait-mediated Indirect Interactions: Ecological and Evolutionary Perspectives. Cambridge: Cambridge University Press.Google Scholar
Okamura, B. & Feist, S.W. (2011) Emerging diseases in freshwater systems. Freshwater Biology, 56, 627637.Google Scholar
Ovcharenko, M.O., Bacela, K., Wilkinson, T., et al. (2010) Cucumispora dikerogammarz n. gen. (Fungi: Microsporidia) infecting the invasive amphipod Dikerogammarus villosus: a potential emerging disease in European rivers. Parasitology, 137, 191204.Google Scholar
Parker, J.D., Torchin, M.E., Hufbauer, R.A., et al. (2013) Do invasive species perform better in their new ranges? Ecology, 94, 985994.Google Scholar
Paterson, R.A., Dick, J.T.A., Pritchard, D.W., et al. (2015) Predicting invasive species impacts: a community module functional response approach reveals context dependencies. Journal of Animal Ecology, 84, 453463.Google Scholar
Pimentel, D., Zuniga, R. & Morrison, D. (2005) Update on the environmental and economic costs associated with alien-invasive species in the United States. Ecological Economics, 52, 273288.Google Scholar
Polis, G.A., Myers, C.A. & Holt, R.D. (1989) The ecology and evolution of intraguild predation – potential competitors that eat each other. Annual Review of Ecology and Systematics, 20, 297330.Google Scholar
Poulin, R. (2010) Network analysis shining light on parasite ecology and diversity. Trends in Parasitology, 26, 492498.Google Scholar
Price, P.W., Westoby, M., Rice, B., et al. (1986) Parasite mediation in ecological interactions. Annual Review of Ecology and Systematics, 17, 487505.Google Scholar
Rewicz, T., Grabowski, M., MacNeil, C. & Bacela-Spychalska, K. (2014) The profile of a ‘perfect’ invader – the case of killer shrimp, Dikerogammarus villosus. Aquatic Invasions, 9, 267288.Google Scholar
Rewicz, T., Wattier, R., Grabowski, M., Rigaud, T. & Bacela-Spychalska, K. (2015) Out of the Black Sea: phylogeography of the invasive killer shrimp Dikerogammarus villosus across Europe. PLoS ONE, 10, e0118121.Google Scholar
Ricciardi, A. & MacIsaac, H.J. (2011) Impacts of Biological Invasions on Freshwater Ecosystems. Malden, MA: Wiley-Blackwell.Google Scholar
Roy, H. (2016) Invasive species: control wildlife pathogens too. Nature, 530, 281281.Google Scholar
Roy, H.E., Hesketh, H., Purse, B.V., et al. (2016) Alien pathogens on the horizon: opportunities for predicting their threat to wildlife. Conservation Letters, 10, 477484.Google Scholar
Rudolf, V.H.W. (2007) The interaction of cannibalism and omnivory: consequences for community dynamics. Ecology, 88, 26972705.Google Scholar
Sanchez, M.I., Rode, N.O., Flaven, E., et al. (2012) Differential susceptibility to parasites of invasive and native species of Artemia living in sympatry: consequences for the invasion of A. franciscana in the Mediterranean region. Biological Invasions, 14, 18191829.Google Scholar
Selakovic, S., de Ruiter, P.C. & Heesterbeek, H. (2014) Infectious disease agents mediate interaction in food webs and ecosystems. Proceedings of the Royal Society of London B, 281, 20132709.Google Scholar
Shannon, C., Quinn, C.H., Sutcliffe, C., et al. (2019) Exploring knowledge, perception of risk and biosecurity practices among researchers in the UK: a quantitative survey. Biological Invasions, 21, 303314.Google Scholar
Slothouber Galbreath, J.G.M., Smith, J.E., Becnel, J.J., Butlin, R.K. & Dunn, A.M. (2010) Reduction in post-invasion genetic diversity in Crangonyx pseudogracilis (Amphipoda: Crustacea): a genetic bottleneck or the work of hitchhiking vertically transmitted microparasites? Biological Invasions, 12, 191209.Google Scholar
Slothouber Galbreath, J.G.M., Smith, J.E., Terry, R.S., Becnel, J.J. & Dunn, A.M. (2004) Invasion success of Fibrillanosema crangonycis, n.sp., n.g.: a novel vertically transmitted microsporidian parasite from the invasive amphipod host Crangonyx pseudogracilis. International Journal for Parasitology, 34, 235244.Google Scholar
Smith, K.F., Sax, D.F. & Lafferty, K.D. (2006) Evidence for the role of infectious disease in species extinction and endangerment. Conservation Biology, 20, 13491357.Google Scholar
Sutcliffe, C., Quinn, C.H., Shannon, C., Glover, A. & Dunn, A.M. (2018) Exploring the attitudes to and uptake of biosecurity practices for invasive non-native species: views amongst stakeholder organisations working in UK natural environments. Biological Invasions, 20, 399411.Google Scholar
Terry, R.S., MacNeil, C., Dick, J.T.A., Smith, J.E. & Dunn, A.M. (2003) Resolution of a taxonomic conundrum: an ultrastructural and molecular description of the life cycle of Pleistophora mulleri (Pfeiffer 1895; Georgevitch 1929). Journal of Eukaryotic Microbiology, 50, 266273.Google Scholar
Terry, R.S., Smith, J.E., Sharpe, R.G., et al. (2004) Widespread vertical transmission and associated host sex-ratio distortion within the eukaryotic phylum Microspora. Proceedings of the Royal Society of London Series B, 271, 17831789.Google Scholar
Torchin, M.E., Lafferty, K.D., Dobson, A.P., McKenzie, V.J. & Kuris, A.M. (2003) Introduced species and their missing parasites. Nature, 421, 628630.Google Scholar
Torchin, M.E., Lafferty, K.D. & Kuris, A.M. (2002) Parasites and marine invasions. Parasitology, 124, S137S151.Google Scholar
Wattier, R.A., Haine, E.R., Beguet, J., et al. (2007) No genetic bottleneck or associated microparasite loss in invasive populations of a freshwater amphipod. Oikos, 116, 19411953.Google Scholar
Wilkinson, T.J., Rock, J., Whiteley, N.M., Ovcharenko, M.O. & Ironside, J.E. (2011) Genetic diversity of the feminising microsporidian parasite Dictyocoela: new insights into host-specificity, sex and phylogeography. International Journal for Parasitology, 41, 959966.Google Scholar

References

Allen, P.C. (1987) Physiological response of chicken gut tissue to coccidial infection: comparative effects of Eimeria acervulina and Eimeria mitis on mucosal mass, carotenoid content, and brush border enzyme activity. Poultry Science, 66, 13061315.Google Scholar
Allen, P.C. (1997) Production of free radical species during Eimeria maxima infections in chickens. Poultry Science, 76, 814821.Google Scholar
Alonso-Alvarez, C., Bertrand, S., Faivre, B., Chastel, O. & Sorci, G. (2007) Testosterone and oxidative stress: the oxidation handicap hypothesis. Proceedings of the Royal Society of London B, 274, 819.Google Scholar
Amundsen, T. (2000) Why are female birds ornamented? Trends in Ecology & Evolution, 15, 149155.Google Scholar
Amundsen, T.P.H. & Pärn, H. (2006) Female coloration: review of functional and non functional hypotheses. In Hill, G.E. & McGraw, K.J. (eds.), Bird Coloration. Volume 2. Function and Evolution (pp. 280345). Cambridge, MA: Harvard University Press.Google Scholar
Andersson, M. (1994) Sexual Selection. Princeton, NJ: Princeton University Press.Google Scholar
Andersson, M. & Simmons, L.W. (2006) Sexual selection and mate choice. Trends in Ecology & Evolution, 21, 296302.Google Scholar
Biron, D.G. & Loxdale, H.D. (2012) Host–parasite molecular cross-talk during the manipulative process of a host by its parasite. The Journal of Experimental Biology, 216, 148.Google Scholar
Bortolotti, G.R., Marchant, T., Blas, J. & Cabezas, S. (2009a) Tracking stress: localisation, deposition and stability of corticosterone in feathers. Journal of Experimental Biology, 212, 14771482.Google Scholar
Bortolotti, G.R., Mougeot, F., Martínez-Padilla, J., Webster, L.M.I. & Piertney, S.B. (2009b) Physiological stress mediates the honesty of social signals. PLoS ONE, 4, e4983.Google Scholar
Chenoweth, S.F., Doughty, P. & Kokko, H. (2006) Can non-directional male mating preferences facilitate honest female ornamentation? Ecology Letters, 9, 179184.Google Scholar
Cobbold, T.S. (1873) Contributions to our knowledge of grouse disease, including the description of a new species of entozoon, with remarks on a case of rot in the hare. Veterinarian, 46, 163172.Google Scholar
Cornwallis, C.K. & Uller, T. (2009) Towards an evolutionary ecology of sexual traits. Trends in Ecology & Evolution, 25, 145152.Google Scholar
Costantini, D. (2008) Oxidative stress in ecology and evolution: lessons from avian studies. Ecology Letters, 11, 12381251.Google Scholar
Cotton, S., Fowler, K. & Pomiankowski, A. (2004) Do sexual ornaments demonstrate heightened condition-dependent expression as predicted by the handicap hypothesis? Proceedings of the Royal Society of London B, 271, 771783.Google Scholar
Darwin, C.R. (1871) Descent of Man, and Selection in Relation to Sex. London: John Murray.Google Scholar
Ferrari, N., Cattadori, I.M., Nespereira, J., Rizzoli, A. & Hudson, P.J. (2004) The role of host sex in parasite dynamics: field experiments on the yellow-necked mouse Apodemus flavicollis. Ecology Letters, 7, 8894.Google Scholar
Fisher, R.A. (1930) The Genetical Theory of Natural Selection. Oxford: Clarendon Press.Google Scholar
Folstad, I. & Karter, A.J. (1992) Parasites, bright males, and the immunocompetence handicap. The American Naturalist, 139, 603622.Google Scholar
Gómez, P., Ashby, B. & Buckling, A. (2015) Population mixing promotes arms race host–parasite coevolution. Proceedings of the Royal Society of London B, 282, 20142297.Google Scholar
Goodwin, T.W. (1984) The Biochemistry of Carotenoids. London: Springer.Google Scholar
Grafen, A. (1990) Biological signals as handicaps. Journal of Theoretical Biology, 144, 517546.Google Scholar
Gustafsson, L., Nordling, D., Andersson, M.S., Sheldon, B.C. & Qvarnström, A. (1994) Infectious diseases, reproductive effort and the cost of reproduction in birds. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 346, 323331.Google Scholar
Haines, J.A. (2010) Female ornamentation in red grouse and its potential role in sexual selection. MPhil, University of Aberdeen.Google Scholar
Halliwell, B. & Gutteridge, J. (2007) Free Radicals in Biology and Medicine. New York, NY: Oxford University Press.Google Scholar
Hamilton, W.D. & Zuk, M. (1982) Heritable true fitness and bright birds: a role for parasites? Science, 218, 384387.Google Scholar
Haydon, D.T., Shaw, D.J., Cattadori, I.M., Hudson, P.J. & Thirgood, S.J. (2002) Analysing noisy time-series: describing regional variation in the cyclic dynamics of red grouse. Proceedings of the Royal Society of London B, 269, 16091617.Google Scholar
Hill, G.E. (2011) Condition-dependent traits as signals of the functionality of vital cellular processes. Ecology Letters, 14, 625634.Google Scholar
Horak, P., Saks, L., Zilmer, M., Karu, U. & Zilmer, K. (2007) Do dietary antioxidants alleviate the cost of immune activation? An experiment with greenfinches. The American Naturalist, 170, 625635.Google Scholar
Hudson, P.J. (1986a) The effect of a parasitic nematode on the breeding production of red grouse. Journal of Animal Ecology, 55, 85–92.Google Scholar
Hudson, P.J. (1986b) The Red Grouse: The Biology and Management of a Wild Gamebird. Fordingbridge: The Game Conservancy Trust.Google Scholar
Hudson, P.J., Dobson, A.P., Cattadori, I.M., et al. (2002) Trophic interactions and population growth rates: describing patterns and identifying mechanisms. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 357, 12591271.Google Scholar
Hudson, P.J., Dobson, A.P. & Newborn, D. (1998) Prevention of population cycles by parasite removal. Science, 282, 14.Google Scholar
Hudson, P.J., Newborn, D. & Dobson, A.P. (1992) Regulation and stability of a free-living host–parasite system: Trichostrongylus tenuis in red grouse. I. Monitoring and parasite reduction experiments. Journal of Animal Ecology, 61, 477486.Google Scholar
Husak, J.F. & Moore, I.T. (2008) Stress hormones and mate choice. Trends in Ecology & Evolution, 23, 532534.Google Scholar
Jenkins, D., Watson, A. & Miller, G.R. (1963) Population studies on red grouse, Lagopus lagopus scoticus (Lath.) in north-east Scotland. Journal of Animal Ecology, 32, 317376.Google Scholar
Kodric-Brown, A. & Brown, J.H. (1984) Truth in advertising: the kinds of traits favored by sexual selection. The American Naturalist, 124, 305322.Google Scholar
Kotiaho, J.S. & Puurtinen, M. (2007) Mate choice for indirect genetic benefits: scrutiny of the current paradigm. Functional Ecology, 21, 638644.Google Scholar
Kraaijeveld, K., Kraaijeveld-Smit, F.J.L. & Komdeur, J. (2007) The evolution of mutual ornamentation. Animal Behaviour, 74, 657677.Google Scholar
Lambin, X., Krebs, C.J., Moss, R., Stenseth, N.C. & Yoccoz, N.G. (1999) Population cycles and parasitism. Science, 286, 24252425.Google Scholar
Lande, R. (1980) Sexual dimorphism, sexual selection, and adaptation in polygenic characters. Evolution, 34, 292305.Google Scholar
Larbi, A., Kempf, J. & Pawelec, G. (2007) Oxidative stress modulation and T cell activation. Experimental Gerontology, 42, 852858.Google Scholar
Lenz, T.L., Eizaguirre, C., Rotter, B., Kalbe, M. & Milinski, M. (2013) Exploring local immunological adaptation of two stickleback ecotypes by experimental infection and transcriptome-wide digital gene expression analysis. Molecular Ecology, 22, 774786.Google Scholar
Lochmiller, R.L. (1996) Immunocompetence and animal population regulation. Oikos, 76, 594602.Google Scholar
Lochmiller, R.L. & Deerenberg, C. (2000) Trade-offs in evolutionary immunology: just what is the cost of immunity? Oikos, 88, 8798.Google Scholar
Lovat, L. (1911) The Grouse in Health and in Disease. London: Smith & Elder.Google Scholar
Martínez-de la Puente, J., Merino, S., Tomás, G., et al. (2010) The blood parasite Haemoproteus reduces survival in a wild bird: a medication experiment. Biology Letters, 6, 663665.Google Scholar
Martínez-Padilla, J., Mougeot, F., Perez-Rodriguez, L. & Bortolotti, G.R. (2007) Nematode parasites reduce carotenoid-based signalling in male red grouse. Biology Letters, 3, 161164.Google Scholar
Martínez-Padilla, J., Mougeot, F., Webster, L.M.I., Perez-Rodriguez, L. & Piertney, S.B.B. (2010) Testing the interactive effects of testosterone and parasites on carotenoid-based ornamentation in a wild bird. Journal of Evolutionary Biology, 23, 902913.Google Scholar
Martínez-Padilla, J., Pérez-Rodríguez, L., Mougeot, F., Ludwig, S. & Redpath, S.M. (2014a) Intra-sexual competition alters the relationship between testosterone and ornament expression in a wild territorial bird. Hormones and Behavior, 65, 435444.Google Scholar
Martínez-Padilla, J., Pérez-Rodríguez, L., Mougeot, F., Ludwig, S.C. & Redpath, S.M. (2014b) Experimentally elevated levels of testosterone at independence reduce fitness in a territorial bird. Ecology, 95, 10331044.Google Scholar
Martínez-Padilla, J., Redpath, S.M., Zeineddine, M. & Mougeot, F. (2014c) Insights into population ecology from long-term studies of red grouse Lagopus lagopus scoticus. Journal of Animal Ecology, 83, 8598.Google Scholar
Martínez-Padilla, J., Vergara, P., Mougeot, F. & Redpath, S.M. (2012) Parasitized mates increase infection risk for partners. The American Naturalist, 179, 811820.Google Scholar
Martínez-Padilla, J., Vergara, P., Perez-Rodriguez, L., et al. (2011) Condition- and parasite-dependent expression of a male-like trait in a female bird. Biology Letters, 7, 364367.Google Scholar
McGraw, K.J. (2006) Mechanics of carotenoid-based coloration. In: Hill, G.E. & McGraw, K.J. (eds.),Bird Coloration (pp. 177242). Cambridge, MA: Harvard University Press.Google Scholar
Moss, R. & Watson, A. (2001) Population cycles in birds of the Grouse family (Tetranoidae). Advances in Ecological Research, 32, 53111.Google Scholar
Moss, R. Watson, A. & Parr, R. (1996) Experimental prevention of a population cycle in Red grouse. Ecology, 77, 15121530.Google Scholar
Moss, R., Watson, A., Trenholm, I.B. & Parr, R. (1993) Cecal threadworms Trichostrongylus tenuis in red grouse Lagopus lagopus scoticus – effects of weather and host density upon estimated worm burdens. Parasitology, 107, 199209.Google Scholar
Mougeot, F., Dawson, A., Redpath, S.M. & Leckie, F. (2005b) Testosterone and autumn territorial behavior in male red grouse Lagopus lagopus scoticus. Hormones and Behavior, 47, 576584.Google Scholar
Mougeot, F., Evans, S.A. & Redpath, S.M. (2005a) Interactions between population processes in a cyclic species: parasites reduce autumn territorial behaviour of male red grouse. Oecologia, 144, 289298.Google Scholar
Mougeot, F., Irvine, J.R., Seivwright, L., Redpath, S.M. & Piertney, S. (2004) Honest sexual signaling in male red grouse. Behavioral Ecology, 15, 930937.Google Scholar
Mougeot, F., Martínez-Padilla, J., Blount, J.D., et al. (2010) Oxidative stress and the effect of parasites on a carotenoid-based ornament. Journal of Experimental Biology, 213, 400407.Google Scholar
Mougeot, F., Martínez-Padilla, J., Perez-Rodriguez, L. & Bortolotti, G.R. (2007a) Carotenoid-based colouration and ultraviolet reflectance of the sexual ornaments of grouse. Behavioral Ecology and Sociobiology, 61, 741751.Google Scholar
Mougeot, F., Martínez-Padilla, J., Webster, L.M.I., et al. (2009) Honest sexual signalling mediated by parasite and testosterone effects on oxidative balance. Proceedings of the Royal Society of London B, 276, 10931100.Google Scholar
Mougeot, F., Perez-Rodriguez, L., Martínez-Padilla, J., Leckie, F. & Redpath, S.M. (2007b) Parasites, testosterone and honest carotenoid-based signalling of health. Functional Ecology, 21, 886898.Google Scholar
Mougeot, F., Piertney, S.B.B., Leckie, F., et al. (2005c) Experimentally increased aggressiveness reduces population kin structure and subsequent recruitment in red grouse Lagopus lagopus scoticus. Journal of Animal Ecology, 74, 488497.Google Scholar
Mougeot, F. & Redpath, S.M. (2004) Sexual ornamentation relates to immune function in male red grouse Lagopus lagopus scoticus. Journal of Avian Biology, 35, 425433.Google Scholar
Mougeot, F., Redpath, S.M., Leckie, F. & Hudson, P.J. (2003a) The effect of aggressiveness on the population dynamics of a territorial bird. Nature, 421, 737739.Google Scholar
Mougeot, F., Redpath, S.M., Moss, R., et al. (2003b) Territorial behaviour and population dynamics in red grouse Lagopus lagopus scoticus. I. Population experiments. Journal of Animal Ecology, 72, 10731082.Google Scholar
Mougeot, F., Redpath, S.M. & Piertney, S.B. (2006) Elevated spring testosterone increases parasite intensity in male red grouse. Behavioral Ecology, 17, 117125.Google Scholar
Mougeot, F., Redpath, S.M., Piertney, S.B. & Hudson, P.J. (2005d) Separating behavioral and physiological mechanisms in testosterone-mediated trade-offs. The American Naturalist, 166, 158168.Google Scholar
Mougeot, F.L., Ádám, Z., Martínez-Padilla, J., et al. (2016) Parasites, mate attractiveness and female feather corticosterone levels in a socially monogamous bird. Behavioral Ecology and Sociobiology, 70, 277283.Google Scholar
Newborn, D. & Foster, R. (2002) Control of parasite burdens in wild red grouse Lagopus lagopus scoticus through the indirect application of anthelmintics. Journal of Applied Ecology, 39, 909914.Google Scholar
Owen-Ashley, N.T., Hasselquist, D. & Wingfield, J.C. (2004) Androgens and the immunocompetence handicap hypothesis: Unraveling direct and indirect pathways of immunosuppression in song sparrows. The American Naturalist, 164, 490505.Google Scholar
Paterson, S. & Piertney, S.B. (2011) Frontiers in host–parasite ecology and evolution. Molecular Ecology, 20, 869871.Google Scholar
Pérez-Rodríguez, L. (2009) Carotenoids in evolutionary ecology: re-evaluating the antioxidant role. BioEssays, 31, 11161126.Google Scholar
Pérez-Rodríguez, L., de Blas, E.G., Martínez-Padilla, J., Mougeot, F. & Mateo, R. (2016) Carotenoid profile and vitamins in the combs of the red grouse (Lagopus lagopus scoticus): implications for the honesty of a sexual signal. Journal of Ornithology, 157, 145153.Google Scholar
Piertney, S.B.B., Lambin, X., MacColl, A.D.C., et al. (2008) Temporal changes in kin structure through a population cycle in a territorial bird, the red grouse Lagopus lagopus scoticus. Molecular Ecology, 17, 25442551.Google Scholar
Poiani, A., Goldsmith, A.R. & Evans, M.R. (2000) Ectoparasites of house sparrows (Passer domesticus): an experimental test of the immunocompetence handicap hypothesis and a new model. Behavioral Ecology and Sociobiology, 47, 230242.Google Scholar
Potts, G.R., Tapper, S.C. & Hudson, P.J. (1984) Population fluctuations in Red grouse – analysis of bag records and a simulation-model. Journal of Animal Ecology, 53, 2136.Google Scholar
Poulin, R. & Thomas, F. (2008) Epigenetic effects of infection on the phenotype of host offspring: parasites reaching across host generations. Oikos, 117, 331335.Google Scholar
Prudic, K.L., Jeon, C., Cao, H. & Monteiro, A. (2011) Developmental plasticity in sexual roles of butterfly species drives mutual sexual ornamentation. Science, 331, 7375.Google Scholar
Quigley, B.J.Z., García López, D., Buckling, A., McKane, A.J. & Brown, S.P. (2012) The mode of host–parasite interaction shapes coevolutionary dynamics and the fate of host cooperation. Proceedings of the Royal Society of London B, 279, 3742.Google Scholar
Redpath, S.M., Mougeot, F., Leckie, F. & Evans, A.D. (2006a) The effects of autumn testosterone on survival and productivity in red grouse Lagopus lagopus scoticus. Animal Behaviour, 71, 12971305.Google Scholar
Redpath, S.M., Mougeot, F., Leckie, F.M., Elston, D.A. & Hudson, P.J. (2006b) Testing the role of parasites in driving the cyclic population dynamics of a gamebird. Ecology Letters, 9, 410418.Google Scholar
Richardson, W.S., Spivak, H., Hudson, J.E., Budacz, M.A. & Hunter, J.G. (1997) Teflon buttress inhibits recanalization of the ‘uncut’ Roux limb. Gastroenterology, 112, A1468A1468.Google Scholar
Roberts, M.L., Buchanan, K.L. & Evans, M.R. (2004) Testing the immunocompetence handicap hypothesis: a review of the evidence. Animal Behaviour, 68, 227239.Google Scholar
Romero, F.J., Bosch-Morell, F., Romero, M.J., et al. (1998) Lipid peroxidation products and antioxidants in human disease. Environmental Health Perspectives, 106, 12291234.Google Scholar
Romero, L.M. (2004) Physiological stress in ecology: lessons from biomedical research. Trends in Ecology & Evolution, 19, 249255.Google Scholar
Saino, N., Møller, A.P. & Bolzern, A.M. (1995) Testosterone effects on the immune system and parasite infestations in the barn swallow (Hirundo rustica): an experimental test of the immunocompetence hypothesis. Behavioral Ecology, 6, 397404.Google Scholar
Schmid-Hempel, P. (2011) Evolutionary Parasitology: The Integrated Study of Infections, Immunology, Ecology, and Genetics. Oxford: Oxford University Press.Google Scholar
Schmid-Hempel, P.E. & Ebert, D. (2003) On the evolutionary ecology of specific immune defence. Trends in Ecology & Evolution, 18, 2732.Google Scholar
Seivwright, L., Redpath, S.M., Mougeot, F., Watts, L. & Hudson, P.J. (2004) Faecal egg counts provide a reliable measure of Trichostrongylus tenuis intensities in free-living red grouse Lagopus lagopus scoticus. Journal of Helminthology, 78, 6976.Google Scholar
Seivwright, L.J., Redpath, S.M., Mougeot, F., Leckie, F. & Hudson, P.J. (2005) Interactions between intrinsic and extrinsic mechanisms in a cyclic species: testosterone increases parasite infection in red grouse. Proceedings of the Royal Society of London B, 272, 22992304.Google Scholar
Shaw, J.L. (1988) Arrested development of Trichostrongylus tenuis as 3rd stage larvae in red grouse. Research in Veterinary Science, 45, 256258.Google Scholar
Shaw, J.L. & Moss, R. (1989) The role of parasite fecundity and longevity in the success of Trichostrongylus tenuis in low-density red grouse populations. Parasitology, 99, 253258.Google Scholar
Shaw, J.L., Moss, R. & Pike, A.W. (1989) Development and survival of the free-living stages of Trichostrongylus tenuis, a cecal parasite of red grouse Lagopus lagopus scoticus. Parasitology, 99, 105113.Google Scholar
Sheldon, B.C. & Verhulst, S. (1996) Ecological immunology: costly parasite defences and trade-offs in evolutionary ecology. Trends in Ecology & Evolution, 11, 317321.Google Scholar
Splettstoesser, W.D. & Schuff-Werner, P. (2002) Oxidative stress in phagocytes – ‘the enemy within’. Microscopy Research and Technique, 57, 441455.Google Scholar
Stear, M.J., Fitton, L., Innocent, G.T., et al. (2007) The dynamic influence of genetic variation on the susceptibility of sheep to gastrointestinal nematode infection. Journal of the Royal Society Interface, 4, 767.Google Scholar
Tarjuelo, R., Vergara, P. & Martínez-Padilla, J. (2016) Intra-sexual competition modulates calling behavior and its association with secondary sexual traits. Behavioral Ecology and Sociobiology, 70, 16331641.Google Scholar
Thirgood, S.J., Redpath, S.M., Haydon, D.T., et al. (2000) Habitat loss and raptor predation: disentangling long- and short-term causes of red grouse declines. Proceedings of the Royal Society of London B, 267, 651656.Google Scholar
Thrall, P.H., Antonovics, J. & Bever, J.D. (1997) Sexual transmission of disease and host mating systems: within-season reproductive success. The American Naturalist, 149, 485506.Google Scholar
Turchin, P. (2003) Complex Population Dynamics: A Theoretical/Empirical Synthesis. Princeton, NJ: Princeton University Press.Google Scholar
Vergara, P. & Martínez-Padilla, J. (2012) Social context decouples the relationship between a sexual ornament and testosterone levels in a male wild bird. Hormones and Behavior, 62, 407412.Google Scholar
Vergara, P., Martínez-Padilla, J., Mougeot, F., Leckie, F. & Redpath, S.M. (2012a) Environmental heterogeneity influences the reliability of secondary sexual traits as condition indicators. Journal of Evolutionary Biology, 25, 2028.Google Scholar
Vergara, P., Martínez-Padilla, J., Redpath, S.M. & Mougeot, F. (2011) The ornament–condition relationship varies with parasite abundance at population level in a female bird. Naturwissenschaften, 98, 897902.Google Scholar
Vergara, P., Mougeot, F., Martínez-Padilla, J., Leckie, F. & Redpath, S.M. (2012b) The condition dependence of a secondary sexual trait is stronger under high parasite infection level. Behavioral Ecology, 23, 502511.Google Scholar
Vergara, P., Redpath, S.M., Martínez‐Padilla, J. & Mougeot, F. (2012c) Environmental conditions influence red grouse ornamentation at a population level. Biological Journal of the Linnean Society, 107, 788798.Google Scholar
von Schantz, T., Bensch, S., Grahn, M., Hasselquist, D. & Wittzell, H. (1999) Good genes, oxidative stress and condition-dependent sexual signals. Proceedings of the Royal Society of London B, 266, 112.Google Scholar
Watson, A. (1985) Social class, socially-induced loss, recruitment and breeding of red grouse. Oecologia, 67, 493498.Google Scholar
Watson, A. & Moss, R. (1988) Spacing behaviour and population limitation in red grouse. The Auk, 105, 207208.Google Scholar
Watson, A. & Moss, R. (2008) Grouse. London: Collins.Google Scholar
Watson, A., Moss, R. & Rae, S. (1998) Population dynamics of Scottish rock ptarmigan cycles. Ecology, 79, 11741192.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
Webster, L.M.I., Mello, L.V., Mougeot, F., et al. (2011) Identification of genes responding to nematode infection in red grouse. Molecular Ecology Resources, 11, 305313.Google Scholar
Webster, L.M.I.P., Paterson, S., Mougeot, F., Martínez-Padilla, J. & Piertney, S.B.B. (2011) Transcriptomic response of red grouse to gastro-intestinal nematode parasites and testosterone: implications for population dynamics. Molecular Ecology, 20, 920931.Google Scholar
Wenzel, M.A., Douglas, A., James, M.C., Redpath, S.M. & Piertney, S.B.B. (2016) The role of parasite-driven selection in shaping landscape genomic structure in red grouse (Lagopus lagopus scotica). Molecular Ecology, 25, 324341.Google Scholar
Wenzel, M.A., James, M.C., Douglas, A. & Piertney, S.B. (2015a) Genome-wide association and genome partitioning reveal novel genomic regions underlying variation in gastrointestinal nematode burden in a wild bird. Molecular Ecology, 24, 41754192.Google Scholar
Wenzel, M.A. & Piertney, S.B. (2014) Fine-scale population epigenetic structure in relation to gastro-intestinal parasite load in red grouse (Lagopus lagopus scotica). Molecular Ecology, 23, 42564273.Google Scholar
Wenzel, M.A. & Piertney, S.B.B. (2015) Digging for gold nuggets: uncovering novel candidate genes for variation in gastrointestinal nematode burden in a wild bird species. Journal of Evolutionary Biology, 28, 807825.Google Scholar
Wenzel, M.A., Webster, L.M.I., Paterson, S., et al. (2013) A transcriptomic investigation of handicap models in sexual selection. Behavioral Ecology and Sociobiology, 67, 221234.Google Scholar
Wenzel, M.A., Webster, L.M.I., Paterson, S. & Piertney, S.B. (2015b) Identification and characterisation of 17 polymorphic candidate genes for response to parasitic nematode (Trichostrongylus tenuis) infection in red grouse (Lagopus lagopus scotica). Conservation Genetics Resources, 7, 2328.Google Scholar
Wilfert, L. & Schmid-Hempel, P. (2008) The genetic architecture of susceptibility to parasites. BMC Evolutionary Biology, 8, 187.Google Scholar
Wilson, G.R. & Wilson, L.P. (1978) Haematology weight and condition of red grouse (Lagopus lagopus scoticus) infected with caecal threadworms (Trichostrongylus tenuis). Research in Veterinary Science, 25, 331336.Google Scholar
Wongrak, K., Daş, G., von Borstel, U.K. & Gauly, M. (2015) Genetic variation for worm burdens in laying hens naturally infected with gastro-intestinal nematodes. British Poultry Science, 56, 1521.Google Scholar
Zahavi, A. (1975) Mate selection – a selection for a handicap. Journal of Theoretical Biology, 53, 205214.Google Scholar
Zuk, M. & Stochr, A. (2002) Immune defense and host life history. The American Naturalist, 160, S9S22.Google Scholar

References

Altizer, S., Nunn, C.L., Thrall, P.H., et al. (2003) Social organization and parasite risk in mammals: integrating theory and empirical studies. Annual Review of Ecology Evolution and Systematics, 34, 517547.Google Scholar
Anderson, R.M., Fraser, C., Ghani, A.C., et al. (2004). Epidemiology, transmission dynamics and control of SARS: the 2002–2003 epidemic. Philosophical Transactions of the Royal Society of London Series B: Biological Sciences, 359, 10911105.Google Scholar
Anderson, R.M. & May, R.M. (1979) Population biology of infectious diseases. Part I. Nature, 280, 361367.Google Scholar
AusVet. (2005) Tasmanian devil facial tumour disease response. Technical Workshop: 29–31 August 2005. Hobart (Tasmania): Department of Primary Industries, Water, and Environment.Google Scholar
Beeton, N. (2011) Population and disease modelling in the Tasmanian devil. PhD thesis, University of Tasmania.Google Scholar
Beeton, N. & McCallum, H. (2011) Models predict that culling is not a feasible strategy to prevent extinction of Tasmanian devils from facial tumour disease. Journal of Applied Ecology, 48, 13151323.Google Scholar
Brown, G.K., Tovar, C., Cooray, A.A., et al. (2016) Mitogen-activated Tasmanian devil blood mononuclear cells kill devil facial tumour disease cells. Immunology and Cell Biology, 94, 673679.Google Scholar
Bruniche-Olsen, A., Jones, M.E., Austin, J.J., Burridge, C.P. & Holland, B.R. (2014) Extensive population decline in the Tasmanian devil predates European settlement and Devil Facial Tumor Disease. Biology Letters, 10, 20140619.Google Scholar
Burgman, M.A. (2005) Risks and Decisions for Conservation and Environmental Management. New York, NY: Cambridge University Press.Google Scholar
Coulson, T. (2012) Integral projections models, their construction and use in posing hypotheses in ecology. Oikos, 121, 13371350.Google Scholar
Daszak, P., Cunningham, A.A. & Hyatt, A.D. (2000) Emerging infectious diseases of wildlife – threats to biodiversity and human health. Science, 287, 443449.Google Scholar
De Castro, F. & Bolker, B. (2005) Mechanisms of disease-induced extinction. Ecology Letters, 8, 117126.Google Scholar
Dewar, E. (2013) Understanding behaviour, stress and disease in Tasmanian devils: implications for selective adaptations. Honours thesis, University of Tasmania.Google Scholar
Doherty, T.S., Dickman, C.R., Johnson, C.N., et al. (2017) Impacts and management of feral cats Felis catus in Australia. Mammal Review, 47(2), 8397. https://doi.org/10.1111/mam.12080Google Scholar
Donnelly, C.A., Ghani, A.C., Leung, G.M., et al. (2003) Epidemiological determinants of spread of causal agent of severe acute respiratory syndrome in Hong Kong. The Lancet, 361, 17611766.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
Ebert, D. & Bull, J.J. (2003) Challenging the trade-off model for the evolution of virulence: is virulence management feasible? Trends in Microbiology, 11, 1520.Google Scholar
Epstein, B., Jones, M., Hamede, R., et al. (2016) Rapid evolutionary response to a transmissible cancer in Tasmanian devils. Nature Communications, 7, 12684.Google Scholar
Frankham, R. (2008) Genetic adaptation to captivity in species conservation programs. Molecular Ecology, 17, 325333.Google Scholar
Frankham, R., Ballou, J.D., Eldridge, M.D.B., et al. (2011) Predicting the probability of outbreeding depression. Conservation Biology, 25, 465475.Google Scholar
Frankham, R., Ballou, J.D., Ralls, K., et al. (2017) Genetic Management of Fragmented Animal and Plant Populations. Oxford: Oxford University Press.Google Scholar
Galvani, A.P. & May, R.M. (2005) Dimensions of superspreading. Nature, 438, 293.Google Scholar
Hamede, R., Bashford, J., Jones, M. & McCallum, H. (2012) Simulating devil facial tumour disease outbreaks across empirically derived contact networks. Journal of Applied Ecology, 49, 447456.Google Scholar
Hamede, R.K., Bashford, J., McCallum, H. & Jones, M. (2009) Contact networks in a wild Tasmanian devil (Sarcophilus harrisii) population: using social network analysis to reveal seasonal variability in social behaviour and its implications for transmission of devil facial tumour disease. Ecology Letters, 12, 11471157.Google Scholar
Hamede, R.K., Beeton, N.J., Carver, S. & Jones, M.E. (2017) Untangling the model muddle: empirical tumour growth in Tasmanian devil facial tumour disease. Scientific Reports, 7(1), 6217.Google Scholar
Hamede, R.K., McCallum, H. & Jones, M. (2008) Seasonal, demographic and density-related patterns of contact between Tasmanian devils (Sarcophilus harrisii): implications for transmission of devil facial tumour disease. Austral Ecology, 33, 614614.Google Scholar
Hamede, R.K., McCallum, H. & Jones, M. (2013) Biting injuries and transmission of Tasmanian devil facial tumour disease. Journal of Animal Ecology, 82, 182190.Google Scholar
Hamede, R.K., Pearse, A.M., Swift, K., Barmuta, L.A., Murchison, E.P. & Jones, M.E. (2015) Transmissible cancer in Tasmanian devils: localized lineage replacement and host population response. Proceedings of the Royal Society of London B, 282, 20151468.Google Scholar
Hawkins, C.E., Baars, C., Hesterman, H., et al. (2006) Emerging disease and population decline of an island endemic, the Tasmanian devil Sarcophilus harrisii. Biological Conservation, 131, 307324.Google Scholar
Hawkins, C.E., McCallum, H., Mooney, N., Jones, M. & Holdsworth, M. (2009) Sarcophilus harrisii. IUCN Red List of threatened species. Version 2009. 1.Google Scholar
Hollings, T., Jones, M., Mooney, N. & McCallum, H. (2014) Trophic cascades following the disease-induced decline of an apex predator, the Tasmanian devil. Conservation Biology, 28, 3675.Google Scholar
Hollings, T., Jones, M., Mooney, N. & McCallum, H. (2016) Disease-induced decline of an apex predator drives invasive dominated states and threatens biodiversity. Ecology, 97, 394405.Google Scholar
Hollings, T., McCallum, H., Kreger, K., Mooney, N. & Jones, M. (2015) Relaxation of risk-sensitive behaviour of prey following disease-induced decline of an apex predator, the Tasmanian devil. Proceedings of the Royal Society of London B, 282, 20150124.Google Scholar
Huxtable, S.J., Lee, D.V., Wise, P. & Save the Tasmanian Devil Program. (2015) Metapopulation management of an extreme disease scenario. In: Armstrong, D.P., Hayward, M.W., Moro, D. & Seddon, B.P. (eds.), Advances in Reintroduction Biology of Australian and New Zealand Fauna. Clayton, Victoria, Australia: CSIRO.Google Scholar
Jones, M.E. (2003) Convergence in ecomorphology and guild structure among marsupial and placental carnivores. In: Jones, M.E., Dickman, C.R. & Archer, M. (eds.), Predators With Pouches: The Biology of Carnivorous Marsupials. Melbourne, Australia: CSIRO Publishing.Google Scholar
Jones, M.E., Cockburn, A., Hamede, R., et al. (2008) Life-history change in disease-ravaged Tasmanian devil populations. Proceedings of the National Academy of Sciences of the United States of America, 105, 10,02310,027.Google Scholar
Jones, M.E., Jarman, P.J., Lees, C.M., et al. (2007) Conservation management of Tasmanian devils in the context of an emerging, extinction-threatening disease: devil facial tumor disease. EcoHealth, 4, 326337.Google Scholar
Jones, M.E., Paetkau, D., Geffen, E.L.I. & Moritz, C. (2004) Genetic diversity and population structure of Tasmanian devils, the largest marsupial carnivore. Molecular Ecology, 13, 21972209.Google Scholar
Karu, N., Wilson, R., Hamede, R., et al. (2016) Discovery of biomarkers for Tasmanian devil cancer (DFTD) by metabolic profiling of serum. Journal of Proteome Research, 15, 38273840.Google Scholar
Kerr, P. J., Liu, J., Cattadori, I., et al. (2015) Myxoma virus and the leporipoxviruses: an evolutionary paradigm. Viruses, 7, 10201061.Google Scholar
Kreiss, A., Brown, G.K., Tovara, C., Lyons, A.B. & Woods, G.M. (2015). Evidence for induction of humoral and cytotoxic immune responses against devil facial tumor disease cells in Tasmanian devils Sarcophilus harrisii immunized with killed cell preparations. Vaccine, 33, 30163025.Google Scholar
Lachish, S., Jones, M. & McCallum, H. (2007) The impact of disease on the survival and population growth rate of the Tasmanian devil. Journal of Animal Ecology, 76, 926936.Google Scholar
Lachish, S., McCallum, H. & Jones, M. (2009) Demography, disease and the devil: life-history changes in a disease-affected population of Tasmanian devils (Sarcophilus harrisii). Journal of Animal Ecology, 78, 427436.Google Scholar
Lachish, S., McCallum, H., Mann, D., Pukk, C. & Jones, M.E. (2010) Evaluation of selective culling of infected individuals to control Tasmanian devil facial tumor disease. Conservation Biology, 24, 841851.Google Scholar
Lafferty, K.D. & Kuris, A.M. (2002) Trophic strategies, animal diversity and body size. Trends in Ecology & Evolution, 17, 507513.Google Scholar
Lazenby, B. (2009) Habitat identification and hair tube surveys for the Endangered New Holland Mouse in Tasmania with a focus on the St Helens area. www.northeastbioregionalnetwork.org.au/docs/NEBN%20final%20report.pdf (downloaded 08/06/2017): North East Bioregional Network.Google Scholar
Lazenby, B.T., Tobler, M.W., Brown, W.E., et al. (2018) Density trends and demographic signals uncover the long‐term impact of transmissible cancer in Tasmanian devils. Journal of Applied Ecology, 55, 13681379. https://doi.org/10.1111/1365–2664.13088Google Scholar
Le Roex, N., Berrington, C.M., Hoal, E.G. & Van Helden, P.D. (2015) Selective breeding: the future of TB management in African buffalo? Acta Tropica, 149, 3844.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
May, R.M. & Anderson, R. M. (1979) Population biology of infectious diseases. Part II. Nature, 280, 455461.Google Scholar
McCallum, H. (2005) Inconclusiveness of chytridiomycosis as the agent in widespread frog declines. Conservation Biology, 19, 14211430.Google Scholar
McCallum, H. & Jones, M. (2006) To lose both would look like carelessness: Tasmanian devil facial tumour disease. PLoS Biology, 4, e3421674.Google Scholar
McCallum, H. & Jones, M. (2010) Sins of omission and sins of commission: St Thomas Aquinas and the devil. Australian Zoologist, 35, 307314.Google Scholar
McCallum, H. & Jones, M. (2012) Infectious cancer in wildlife. In: Aguirre, A., Daszak, P. & Ostfeld, R. (eds.), Conservation Medicine: Applied Cases of Ecological Health (pp. 270283). Oxford: Oxford University Press.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
Metcalf, C.J.E., Graham, A.L., Martinez-Bakker, M. & Childs, D.Z. (2016) Opportunities and challenges of Integral Projection Models for modelling host–parasite dynamics. Journal of Animal Ecology, 85, 343355.Google Scholar
Metzger, M.J., Reinisch, C., Sherry, J. & Goff, S.P. (2015) Horizontal transmission of clonal cancer cells causes leukemia in soft-shell clams. Cell, 161, 255263.Google Scholar
Metzger, M.J., Villalba, A., Carballal, M.J., et al. (2016) Widespread transmission of independent cancer lineages within multiple bivalve species. Nature, 534, 705709.Google Scholar
Mooney, N. (2004) The devil’s new hell. Nature Australia, 28, 34.Google Scholar
Murchison, E.P. (2008) Clonally transmissible cancers in dogs and Tasmanian devils. Oncogene, 27, S19S30.Google Scholar
Murchison, E.P., Schulz-Trieglaff, O.B., Ning, Z., et al. (2012) Genome sequencing and analysis of the Tasmanian devil and its transmissible cancer. Cell, 148, 780791.Google Scholar
Murchison, E.P., Wedge, D.C., Alexandrov, L.B., et al. (2014) Transmissable dog cancer genome reveals the origin and history of an ancient cell lineage. Science, 343, 437440.Google Scholar
Murgia, C., Pritchard, J.K., Kim, S.Y., Fassati, A. & Weiss, R.A. (2006) Clonal origin and evolution of a transmissible cancer. Cell, 126, 477487.Google Scholar
Nak, D., Nak, Y., Cangul, I.T. & Tuna, B. (2005) A clinico-pathological study on the effect of vincristine on transmissible venereal tumour in dogs. Journal of Veterinary Medicine Series A, 52, 366370.Google Scholar
Pearse, A.M. & Swift, K. (2006) Transmission of devil facial-tumour disease – an uncanny similarity in the karyotype of these malignant tumours means that they could be infective. Nature, 439, 549549.Google Scholar
Pearse, A.M., Swift, K., Hodson, P., et al. (2012) Evolution in a transmissible cancer: a study of the chromosomal changes in devil facial tumor (DFT) as it spreads through the wild Tasmanian devil population. Cancer Genetics, 205, 101–112.Google Scholar
Phalen, D.N., Frimberger, A., Pyecroft, S., et al. (2013) Vincristine chemotherapy trials and pharmacokinetics in Tasmanian devils with Tasmanian devil facial tumor disease. PLoS ONE, 8, e65133.Google Scholar
Phalen, D.N., Frimberger, A.E., Peck, S., et al. (2015) Doxorubicin and carboplatin trials in Tasmanian devils (Sarcophilus harrisii) with Tasmanian devil facial tumor disease. The Veterinary Journal, 206, 312316.Google Scholar
Pimm, S.L., Dollar, L. & Bass, O.L. Jr (2006) The genetic rescue of the Florida panther. Animal Conservation, 9, 115122.Google Scholar
Pye, R., Hamede, R., Siddle, H., et al. (2016a) Demonstration of immune responses against devil facial tumour disease in wild Tasmanian devils. Biology Letters, 12, 20160553.Google Scholar
Pye, R.J., Pemberton, D., Tovar, C., et al. (2016b) A second transmissible cancer in Tasmanian devils. Proceedings of the National Academy of Sciences of the United States of America, 113, 374379.Google Scholar
Pyecroft, S.B., Pearse, A.M., Loh, R., et al. (2007) Towards a case definition for devil facial tumour disease: what is it? EcoHealth, 4, 346351.Google Scholar
Raberg, 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, 364, 3749.Google Scholar
Robinson, A.C., Lawson, B., Toms, M.P., et al. (2010) Emerging infectious disease leads to rapid population declines of common British birds. PLoS ONE, 5, e12215.Google Scholar
Roelke-Parker, M.E., Munson, L., Packer, C., et al. (1996) A canine distemper virus epidemic in Serengeti lions (Panthera leo). Nature, 379, 441445.Google Scholar
Rogers, T., Fox, S., Pemberton, D. & Wise, P. (2016) Sympathy for the devil: captive-management style did not influence survival, body-mass change or diet of Tasmanian devils 1 year after wild release. Wildlife Research, 43, 544552.Google Scholar
Ruiz-Aravena, M. (2019). The Tasmanian devil and its transmissible cancer: physiology of the devil-DFTD interaction. PhD thesis, University of Tasmania.Google Scholar
Ruiz-Aravena, M., Jones, M.E., Carver, S., et al. (2018). Sex bias in ability to cope with cancer: Tasmanian devils and facial tumour disease. Proceedings of the Royal Society of London B, 285, 20182239.Google Scholar
Sakai, A.K., Allendorf, F.W., Holt, J.S., et al. (2001) The population biology of invasive species. Annual Review of Ecology and Systematics, 32, 305332.Google Scholar
Scudamore, J.M. & Harris, D.M. (2002) Control of foot and mouth disease: lessons from the experience of the outbreak in Great Britain in 2001. Revue scientifique et technique (International Office of Epizootics), 21, 699710.Google Scholar
Siddle, H.V., Kreiss, A., Tovar, C., et al. (2013) Reversible epigenetic down-regulation of MHC molecules by devil facial tumour disease illustrates immune escape by a contagious cancer. Proceedings of the National Academy of Sciences of the United States of America, 110, 51035108.Google Scholar
Tompkins, D.M., Carver, S., Jones, M.E., Krkosek, M. & Skerratt, L.F. (2015) Emerging infectious diseases of wildlife: a critical perspective. Trends in Parasitology, 31, 149159.Google Scholar
Tovar, C., Obendorf, D., Murchison, E.P., et al. (2011) Tumor-specific diagnostic marker for transmissible facial tumors of Tasmanian devils: immunohistochemistry studies. Veterinary Pathology, 48, 11951203.Google Scholar
Wayne, R.K., Geffen, E., Girman, D.J., et al. (1997) Molecular systematics of the Canidae. Systematic Biology, 46, 622653.Google Scholar
Wells, K., Hamede, R.K., Kerlin, D.H., et al. (2017a) Infection of the fittest: devil facial tumour disease has greatest effect on individuals with highest reproductive output. Ecology Letters, 20, 770778.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
Woinarski, J.C., Burbidge, A.A. & Harrison, P.L. (2015) Ongoing unraveling of a continental fauna: decline and extinction of Australian mammals since European settlement. Proceedings of the National Academy of Sciences of the United States of America, 112, 45314540.Google Scholar
Woinarski, J.C.Z., Burbidge, A.A. & Harrison, P.L. (2014) The Action Plan for Australian Mammals 2012. Melbourne, Australia: CSIRO Publishing.Google Scholar
Woods, G.M., Howson, L.J., Brown, G.K., et al. (2015) Immunology of a transmissible cancer spreading among Tasmanian devils. The Journal of Immunology, 195, 2329.Google Scholar

References

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. (1981) The population dynamics of microparasites and their invertebrate hosts. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 291, 451524.Google Scholar
Aznar, I., Frankena, K., More, S. J., et al. (2018). Quantification of Mycobacterium bovis transmission in a badger vaccine field trial. Preventive Veterinary Medicine, 149, 2937.Google Scholar
Beirne, C., Delahay, R., Hares, M. & Young, A. (2014) Age-related declines and disease-associated variation in immune cell telomere length in a wild mammal. PLoS ONE, 9, e108964.Google Scholar
Beirne, C., Waring, L., McDonald, R.A., Delahay, R. & Young, A. (2016) Age-related declines in immune response in a wild mammal are unrelated to immune cell telomere length. Proceedings of the Royal Society of London B, 283, 20152949.Google Scholar
Benton, C.H., Delahay, R.J., Robertson, A., et al. (2016) Blood thicker than water: kinship, disease prevalence and group size drive divergent patterns of infection risk in a social mammal. Proceedings of the Royal Society of London B, 283, 20160798.Google Scholar
Blanchong, J.A., Scribner, K.T., Kravchenko, A.N. & Winterstein, S.R. (2007) TB-infected deer are more closely related than non-infected deer. Biology Letters, 3, 104106.Google Scholar
Buzdugan, S.N., Vergne, T., Grosbois, V., Delahay, R.J. & Drewe, J.A. (2017) Inference of the infection status of individuals using longitudinal testing data from cryptic populations: towards a probabilistic approach to diagnosis. Scientific Reports, 7, 1111.Google Scholar
Carpenter, P.J., Pope, L.C., Greig, C., et al. (2005) Mating system of the Eurasian badger, Meles meles, in a high density population. Molecular Ecology, 14, 273284.Google Scholar
Carter, S.P., Chambers, M.A., Rushton, S.P., et al. (2012) BCG vaccination reduces risk of tuberculosis infection in vaccinated badgers and unvaccinated badger cubs. PLoS ONE, 7, e49833.Google Scholar
Carter, S.P., Delahay, R.J., Smith, G.C., et al. (2007) Culling-induced social perturbation in Eurasian badgers Meles meles and the management of TB in cattle: an analysis of a critical problem in applied ecology. Proceedings of the Royal Society of London B, 274, 27692777.Google Scholar
Chambers, M.A., Aldwell, F., Williams, G.A., et al. (2017) The effect of oral vaccination with Mycobacterium bovis BCG on the development of tuberculosis in captive European badgers (Meles meles). Frontiers in Cellular and Infection Microbiology, 7(6).Google Scholar
Chambers, M.A., Rogers, F., Delahay, R.J., et al. (2011) Bacillus Calmette–Guérin vaccination reduces the severity and progression of tuberculosis in badgers. Proceedings of the Royal Society of London B, 278, 19131920.Google Scholar
Cheeseman, C., Wilesmith, J., Stuart, F. & Mallinson, P. (1988) Dynamics of tuberculosis in a naturally infected badger population. Mammal Review, 18, 6172.Google Scholar
Clifton-Hadley, R., Wilesmith, J. & Stuart, F. (1993) Mycobacterium bovis in the European badger (Meles meles): epidemiological findings in tuberculous badgers from a naturally infected population. Epidemiology and Infection, 111, 919.Google Scholar
Clutton-Brock, T. & Sheldon, B.C. (2010) Individuals and populations: the role of long-term, individual-based studies of animals in ecology and evolutionary biology. Trends in Ecology & Evolution, 25, 562573.Google Scholar
Colchero, F., Jones, O.R. & Rebke, M. (2012) BaSTA: an R package for Bayesian estimation of age‐specific survival from incomplete mark–recapture/recovery data with covariates. Methods in Ecology and Evolution, 3, 466470.Google Scholar
Corner, L.A., O’Meara, D., Costello, E., Lesellier, S. & Gormley, E. (2012) The distribution of Mycobacterium bovis infection in naturally infected badgers. The Veterinary Journal, 194, 166172.Google Scholar
Delahay, R., Langton, S., Smith, G., Clifton‐Hadley, R. & Cheeseman, C. (2000) The spatio‐temporal distribution of Mycobacterium bovis (bovine tuberculosis) infection in a high‐density badger population. Journal of Animal Ecology, 69, 428441.Google Scholar
Delahay, R., Walker, N., Forrester, G., et al. (2006) Demographic correlates of bite wounding in Eurasian badgers, Meles meles L., in stable and perturbed populations. Animal Behaviour, 71, 10471055.Google Scholar
Delahay, R., Walker, N., Smith, G., et al. (2013) Long-term temporal trends and estimated transmission rates for Mycobacterium bovis infection in an undisturbed high-density badger (Meles meles) population. Epidemiology and Infection, 141, 14451456.Google Scholar
Donnelly, C.A., Wei, G., Johnston, W.T., et al. (2007) Impacts of widespread badger culling on cattle tuberculosis: concluding analyses from a large-scale field trial. International Journal of Infectious Diseases, 11, 300308.Google Scholar
Donnelly, C.A., Woodroffe, R., Cox, D., et al. (2003) Impact of localized badger culling on tuberculosis incidence in British cattle. Nature, 426, 834837.Google Scholar
Donnelly, C.A., Woodroffe, R., Cox, D., et al. (2006) Positive and negative effects of widespread badger culling on tuberculosis in cattle. Nature, 439, 843846.Google Scholar
Drewe, J.A., Tomlinson, A.J., Walker, N.J. & Delahay, R.J. (2010) Diagnostic accuracy and optimal use of three tests for tuberculosis in live badgers. PLoS ONE, 5, e11196.Google Scholar
Gaillard, J.-M., Festa-Bianchet, M., Yoccoz, N., Loison, A. & Toigo, C. (2000) Temporal variation in fitness components and population dynamics of large herbivores. Annual Review of Ecology and Systematics, 31, 367393.Google Scholar
Gaillard, J., Lemaître, J., Berger, V., et al. (2016) Life history axes of variation. In: The Encyclopedia of Evolutionary Biology (pp. 312323). Oxford: Academic Press.Google Scholar
Gallagher, J. & Clifton-Hadley, R. (2000) Tuberculosis in badgers; a review of the disease and its significance for other animals. Research in Veterinary Science, 69, 203217.Google Scholar
Gallagher, J., Muirhead, R. & Burn, K. (1976) Tuberculosis in wild badgers (Meles meles) in Gloucestershire: pathology. Veterinary Record, 98, 914.Google Scholar
Gallagher, J. & Nelson, J. (1979) Cause of ill health and natural death in badgers in Gloucestershire. Tuberculosis, 10, 1416.Google Scholar
George, S.C., Smith, T.E., Mac Cana, P.S., Coleman, R. & Montgomery, W.I. (2014) Physiological stress in the Eurasian badger (Meles meles): effects of host, disease and environment. General and Comparative Endocrinology, 200, 5460.Google Scholar
Graham, A.L., Hayward, A.D., Watt, K.A., et al. (2010) Fitness correlates of heritable variation in antibody responsiveness in a wild mammal. Science, 330, 662665.Google Scholar
Graham, J., Smith, G., Delahay, R., et al. (2013) Multi-state modelling reveals sex-dependent transmission, progression and severity of tuberculosis in wild badgers. Epidemiology and Infection, 141, 14291436.Google Scholar
Jenkins, H.E., Cox, D. & Delahay, R.J. (2012) Direction of association between bite wounds and Mycobacterium bovis infection in badgers: implications for transmission. PLoS ONE, 7, e45584.Google Scholar
Jenkins, H.E., Morrison, W., Cox, D., et al. (2008) The prevalence, distribution and severity of detectable pathological lesions in badgers naturally infected with Mycobacterium bovis. Epidemiology and Infection, 136, 13501361.Google Scholar
Johnson, P.T., Rohr, J.R., Hoverman, J.T., et al. (2012) Living fast and dying of infection: host life history drives interspecific variation in infection and disease risk. Ecology Letters, 15, 235242.Google Scholar
Joly, D.O., Samuel, M.D., Langenberg, J.A., et al. (2006) Spatial epidemiology of chronic wasting disease in Wisconsin white-tailed deer. Journal of Wildlife Diseases, 42, 578588.Google Scholar
Keeling, M.J. & Danon, L. (2009) Mathematical modelling of infectious diseases. British Medical Bulletin, 92, 3342.Google Scholar
Kelly, G.E., McGrath, G. & More, S.J. (2010) Estimating the extent of spatial association of Mycobacterium bovis infection in badgers in Ireland. Epidemiology and Infection, 138, 270279.Google Scholar
Kéry, M. & Schaub, M. (2011) Bayesian Population Analysis Using WinBUGS: A Hierarchical Perspective. New York, NY: Academic Press.Google Scholar
Kramer‐Schadt, S., Fernández, N., Eisinger, D., Grimm, V. & Thulke, H.H. (2009) Individual variations in infectiousness explain long‐term disease persistence in wildlife populations. Oikos, 118, 199208.Google Scholar
Kruuk, L.E. (2004) Estimating genetic parameters in natural populations using the ‘animal model’. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 359, 873890.Google Scholar
Lebreton, J.D., Nichols, J.D., Barker, R.J., Pradel, R. & Spendelow, J.A. (2009) Modeling individual animal histories with multistate capture–recapture models. Advances in Ecological Research, 41, 87173.Google Scholar
Lesellier, S., Palmer, S., Gowtage-Sequiera, S., et al. (2011) Protection of Eurasian badgers (Meles meles) from tuberculosis after intra-muscular vaccination with different doses of BCG. Vaccine, 29, 37823790.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
Macdonald, D., Harmsen, B., Johnson, P. & Newman, C. (2004) Increasing frequency of bite wounds with increasing population density in Eurasian badgers, Meles meles.Animal Behaviour, 67, 745751.Google Scholar
Macdonald, D.W., Newman, C., Buesching, C.D. & Johnson, P.J. (2008) Male-biased movement in a high-density population of the Eurasian badger (Meles meles). Journal of Mammalogy, 89, 10771086.Google Scholar
May, R.M. & Anderson, R.M. (1979) Population biology of infectious diseases: Part II. Nature, 280, 455461.Google Scholar
McDonald, J.L., Bailey, T., Delahay, R.J., et al. (2016) Demographic buffering and compensatory recruitment promotes the persistence of disease in a wildlife population. Ecology Letters, 19, 443449.Google Scholar
McDonald, J.L., Robertson, A. & Silk, M.J. (2018) Wildlife disease ecology from the individual to the population: Insights from a long‐term study of a naturally infected European badger population. Journal of Animal Ecology, 87, 101112.Google Scholar
McDonald, J.L., Smith, G.C., McDonald, R.A., Delahay, R.J. & Hodgson, D. (2014) Mortality trajectory analysis reveals the drivers of sex-specific epidemiology in natural wildlife–disease interactions. Proceedings of the Royal Society of London B, 281, 20140526.Google Scholar
McDonald, R.A., Delahay, R.J., Carter, S.P., Smith, G.C. & Cheeseman, C.L. (2008) Perturbing implications of wildlife ecology for disease control. Trends in Ecology & Evolution, 23, 5356.Google Scholar
Murhead, R.M. & Burns, K.J. (1974) Tuberculosis in wild badgers in Gloucestershire: epidemiology. Veterinary Record, 95, 552555.Google Scholar
Ostfeld, R.S., Levi, T., Jolles, A.E., et al. (2014) Life history and demographic drivers of reservoir competence for three tick-borne zoonotic pathogens. PLoS ONE, 9, e107387.Google Scholar
Palphramand, K., Delahay, R., Robertson, A., et al. (2017) Field evaluation of candidate baits for oral delivery of BCG vaccine to European badgers, Meles meles. Vaccine, 35, 44024407.Google Scholar
Pfister, C.A. (1998) Patterns of variance in stage-structured populations: evolutionary predictions and ecological implications. Proceedings of the National Academy of Sciences of the United States of America, 95, 213218.Google Scholar
Pradel, R. (2005) Multievent: an extension of multistate capture–recapture models to uncertain states. Biometrics, 61, 442447.Google Scholar
Rogers, L., Cheeseman, C., Mallinson, P. & Clifton‐Hadley, R. (1997) The demography of a high‐density badger (Meles meles) population in the west of England. Journal of Zoology, 242, 705728.Google Scholar
Rogers, L., Delahay, R., Cheeseman, C., et al. (1998) Movement of badgers (Meles meles) in a high–density population: individual, population and disease effects. Proceedings of the Royal Society of London B, 265, 12691276.Google Scholar
Roper, T. (2010) Badger (Collins New Naturalist Library, Book 114). London: HarperCollins UK.Google Scholar
Rushmore, J., Caillaud, D., Matamba, L., et al. (2013) Social network analysis of wild chimpanzees provides insights for predicting infectious disease risk. Journal of Animal Ecology, 82, 976986.Google Scholar
Shirley, M.D., Rushton, S.P., Smith, G.C., South, A.B. & Lurz, P.W. (2003) Investigating the spatial dynamics of bovine tuberculosis in badger populations: evaluating an individual-based simulation model. Ecological Modelling, 167, 139157.Google Scholar
Silk, M.J., Croft, D.P., Delahay, R.J., et al. (2017a) Using social network measures in wildlife disease ecology, epidemiology, and management. BioScience, 67, 245257.Google Scholar
Silk, M.J., Croft, D.P., Delahay, R.J., et al. (2017b) The application of statistical network models in disease research. Methods in Ecology and Evolution, 8, 10261041.Google Scholar
Sin, Y.W., Annavi, G., Dugdale, H.L., et al. (2014) Pathogen burden, co‐infection and major histocompatibility complex variability in the European badger (Meles meles). Molecular Ecology, 23, 50725088.Google Scholar
Smith, G.C., Delahay, R.J., McDonald, R.A. & Budgey, R. (2016) Model of selective and non-selective management of badgers (Meles meles) to control bovine tuberculosis in badgers and cattle. PLoS ONE, 11, e0167206.Google Scholar
Smith, G.C., McDonald, R.A. & Wilkinson, D. (2012) Comparing badger (Meles meles) management strategies for reducing tuberculosis incidence in cattle. PLoS ONE, 7, e39250.Google Scholar
Stearns, S.C. (1983) The influence of size and phylogeny on patterns of covariation among life-history traits in the mammals. Oikos, 41, 173187.Google Scholar
Thrall, P.H., Antonovics, J. & Hall, D.W. (1993) Host and pathogen coexistence in sexually transmitted and vector-borne diseases characterized by frequency-dependent disease transmission. American Naturalist, 142, 543552.Google Scholar
Tomlinson, A., Chambers, M., Carter, S., et al. (2013a) Heterogeneity in the risk of Mycobacterium bovis infection in European badger (Meles meles) cubs. Epidemiology and Infection, 141, 14581466.Google Scholar
Tomlinson, A., Chambers, M. & Delahay, R. (2012) Mycobacterium bovis infection in badger cubs: re-assessing the evidence for maternally derived immunological protection from advanced disease. Veterinary Immunology and Immunopathology, 148, 326330.Google Scholar
Tomlinson, A., Chambers, M., Wilson, G., McDonald, R.A. & Delahay, R. (2013b) Sex‐related heterogeneity in the life‐history correlates of Mycobacterium bovis infection in European badgers (Meles meles). Transboundary and Emerging Diseases, 60, 3745.Google Scholar
Tomlinson, A.J., Chambers, M.A., McDonald, R.A. & Delahay, R.J. (2015) Association of quantitative interferon‐γ responses with the progression of naturally acquired Mycobacterium bovis infection in wild European badgers (Meles meles). Immunology, 144, 263270.Google Scholar
Vicente, J., Delahay, R., Walker, N. & Cheeseman, C. (2007) Social organization and movement influence the incidence of bovine tuberculosis in an undisturbed high‐density badger Meles meles population. Journal of Animal Ecology, 76, 348360.Google Scholar
Weber, N., Bearhop, S., Dall, S.R., et al. (2013a) Denning behaviour of the European badger (Meles meles) correlates with bovine tuberculosis infection status. Behavioral Ecology and Sociobiology, 67, 471479.Google Scholar
Weber, N., Carter, S.P., Dall, S.R., et al. (2013b) Badger social networks correlate with tuberculosis infection. Current Biology, 23, R915R916.Google Scholar
Wendland, L.D., Wooding, J., White, C.L., et al. (2010) Social behavior drives the dynamics of respiratory disease in threatened tortoises. Ecology, 91, 12571262.Google Scholar
White, P.C. & Harris, S. (1995) Bovine tuberculosis in badger (Meles meles) populations in southwest England: the use of a spatial stochastic simulation model to understand the dynamics of the disease. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 349, 391413.Google Scholar
Wilkinson, D., Smith, G., Delahay, R. & Cheeseman, C. (2004) A model of bovine tuberculosis in the badger Meles meles: an evaluation of different vaccination strategies. Journal of Applied Ecology, 41, 492501.Google Scholar
Wilkinson, D., Smith, G., Delahay, R., et al. (2000) The effects of bovine tuberculosis (Mycobacterium bovis) on mortality in a badger (Meles meles) population in England. Journal of Zoology, 250, 389395.Google Scholar
Woodroffe, R., Donnelly, C., Johnston, W., et al. (2005) Spatial association of Mycobacterium bovis infection in cattle and badgers Meles meles. Journal of Applied Ecology, 42, 852862.Google Scholar
Woodroffe, R., Donnelly, C.A., Wei, G., et al. (2009) Social group size affects Mycobacterium bovis infection in European badgers (Meles meles). Journal of Animal Ecology, 78, 818827.Google Scholar

References

Almberg, E.S., Cross, P.C., Dobson, A.P., et al. (2012) Parasite invasion following host reintroduction: a case study of Yellowstone’s wolves. Philosophical Transactions of the Royal Society of London B, 367, 28402851.Google Scholar
Almberg, E.S., Manlove, K.R., Cassirer, E.F., et al. (2016) Modeling management strategies for the control of bighorn sheep respiratory disease. Biennial Symposium of the Northern Wild Sheep and Goat Council, 20, 18.Google Scholar
Ayling, R.D., Bashiruddin, S.E. & Nicholas, R.A.J. (2004) Mycoplasma species and related organisms isolated from ruminants in Britain between 1990 and 2000. Veterinary Record, 155, 413416.Google Scholar
Berube, C.H., Festa-Bianchet, M. & Jorgenson, J.T. (1999) Individual differences, longevity, and reproductive senescence in bighorn ewes. Ecology, 80(8), 25552565.Google Scholar
Besser, T.E., Cassirer, E.F., Highland, M.A., et al. (2013) Bighorn sheep pneumonia: sorting out the cause of a polymicrobial disease. Preventive Veterinary Medicine, 108(2), 8593.Google Scholar
Besser, T.E., Cassirer, E.F., Potter, K.A., et al. (2008) Association of Mycoplasma ovipneumoniae infection with population-limiting respiratory disease in free-ranging Rocky Mountain bighorn sheep (Ovis canadensis canadensis). Journal of Clinical Microbiology, 46(2), 423430.Google Scholar
Besser, T.E., Cassirer, E.F., Potter, K.A., et al. (2014) Epizootic pneumonia of bighorn sheep following experimental exposure to Mycoplasma ovipneumoniae. PLoS ONE, 9(10), e110039.Google Scholar
Besser, T.E., Cassirer, E.F., Yamada, C., et al. (2012) Survival of bighorn sheep (Ovis canadensis) commingled with domestic sheep (Ovis aries) in the absence of Mycoplasma ovipneumoniae. Journal of Wildlife Diseases, 48(1), 168172.Google Scholar
Besser, T.E., Highland, M.A., Baker, K., et al. (2012) Causes of pneumonia epizootics among bighorn sheep, western United States, 2008–2010. Emerging Infectious Diseases, 18(3), 406414.Google Scholar
Borg, N.J., Mitchell, M.S., Lukacs, P.M., et al. (2017) Behavioral connectivity among bighorn sheep suggests potential for disease spread. The Journal of Wildlife Management, 81(1), 3845.Google Scholar
Boyce, W.M. (1992). Population viability analysis. Annual Review of Ecology and Systematics, 23, 481506.Google Scholar
Briggs, C.J., Vredenburg, V.T., Knapp, R.A., et al. (2005) Investigating the population-level effects of chytridiomycosis: an emerging infectious disease of amphibians. Ecology, 86(12), 31493159.Google Scholar
Bunch, T., Wu, C., Zhang, Y.-P., et al. (2006) Phylogenetic analysis of snow sheep (Ovis nivicola) and closely related taxa. Journal of Heredity, 97, 2130.Google Scholar
Bureau of Land Management. (2016) Management of domestic sheep and goats to sustain wild sheep. Release No. 1–1771, 2 March 2016.Google Scholar
Cahn, M.L., Conner, M.M., Schmitz, O.J., et al. (2011) Disease, population viability, and recovery of endangered Sierra Nevada bighorn sheep. The Journal of Wildlife Management, 75(8), 17531766.Google Scholar
Carpenter, T.E., Coggins, V.L., McCarthy, C., et al. (2014) A spatial risk assessment of bighorn sheep extirpation by grazing domestic sheep on public lands. Preventive Veterinary Medicine, 114(1), 310.Google Scholar
Cassirer, E.F., Manlove, K.R., Almberg, E.S., et al. (2018) Pneumonia in bighorn sheep: risk and resilience. Journal of Wildlife Management, 82(1), 3245.Google Scholar
Cassirer, E.F., Manlove, K.R., Plowright, R.K., et al. (2017) Evidence for strain‐specific immunity to pneumonia in bighorn sheep. The Journal of Wildlife Management, 81(1), 133143.Google Scholar
Cassirer, E.F., Plowright, R.K., Manlove, K.R., et al. (2013) Spatio‐temporal dynamics of pneumonia in bighorn sheep. Journal of Animal Ecology, 82(3), 518528.Google Scholar
Cassirer, E.F., Rudolph, K.M., Fowler, P., et al. (2001) Evaluation of ewe vaccination as a tool for increasing bighorn lamb survival following pasteurellosis epizootics. Journal of Wildlife Diseases, 37, 4957.Google Scholar
Cassirer, E. & Sinclair, A.R.E. (2007) Dynamics of pneumonia in a bighorn sheep metapopulation. The Journal of Wildlife Management, 71(4), 10801088.Google Scholar
Clifford, D.L., Schumaker, B.A., Stephenson, T.R., et al. (2009) Assessing disease risk at the wildlife–livestock interface: a study of Sierra Nevada bighorn sheep. Biological Conservation, 142(11), 25592568.Google Scholar
Coggins, V.L. (2006) Selenium supplementation, parasite treatment, and management of bighorn sheep at Lostine River, Oregon. Biennial Symposium of the Northern Wild Sheep and Goat Council, 15, 98106.Google Scholar
Coltman, D.W., Festa-Bianchet, M., Jorgenson, J.T., et al. (2002) Age-dependent sexual selection in bighorn rams. Proceedings of the Royal Society of London B, 269(1487), 165172.Google Scholar
Craft, M.E., Hawthorne, P.L., Packer, C., et al. (2008) Dynamics of a multihost pathogen in a carnivore community. Journal of Animal Ecology, 77(6), 12571264.Google Scholar
Cross, P.C., Lloyd‐Smith, J.O., Johnson, P.L., et al. (2005) Duelling timescales of host movement and disease recovery determine invasion of disease in structured populations. Ecology Letters, 8(6), 587595.Google Scholar
Dassanayake, R.P., Shanthalingam, S., Herndon, C.N., et al. (2010) Mycoplasma ovipneumoniae can predispose bighorn sheep to fatal Mannheimia haemolytica pneumonia. Veterinary Microbiology, 145(3), 354359.Google Scholar
DeCesare, N.J. & Pletscher, D.H. (2006) Movements, connectivity, and resource selection of Rocky Mountain bighorn sheep. Journal of Mammalogy, 87(3), 531538.Google Scholar
Epps, C.W., Wehausen, J.D., Bleich, V.C., et al. (2007) Optimizing dispersal and corridor models using landscape genetics. Journal of Applied Ecology, 44(4), 714724.Google Scholar
Felts, B.L., Walsh, D.P., Cassirer, E.F., Besser, T.E. & Jenks, J. (2016) Mycoplasma ovipneumoniae cross-strain transmissions in captive bighorn sheep. Biennial Symposium of the Northern Wild Sheep and Goat Council, 20, 7778.Google Scholar
Festa-Bianchet, M. (1988) Nursing behaviour of bighorn sheep: correlates of ewe age, parasitism, lamb age, birthdate and sex. Animal Behaviour, 36(5), 14451454.Google Scholar
Festa-Bianchet, M. (2012) The cost of trying: weak interspecific correlations among life-history components in male ungulates. Canadian Journal of Zoology, 90(9), 10721085.Google Scholar
Geist, V. (1971) Mountain Sheep. A Study in Behaviour and Evolution. Chicago, IL: University of Chicago Press.Google Scholar
George, J.L., Martin, D.J., Lukacs, P.M., et al. (2008) Epidemic pasteurellosis in a bighorn sheep population coinciding with the appearance of a domestic sheep. Journal of Wildlife Diseases, 44(2), 388403.Google Scholar
Hass, C.C. (1997). Seasonality of births in bighorn sheep. Journal of Mammalogy, 78(4), 12511260.Google Scholar
Heinse, L.M.,Hardesty, L.M. & Harris, R.B. (2016) Risk of pathogen spillover from domestic sheep and goat flocks on private land. Wildlife Society Bulletin, 40(4), 625633.Google Scholar
Hiendleder, S., Kaupe, B., Wassmuth, R., et al. (2002) Molecular analysis of wild and domestic sheep questions current nomenclature and provides evidence for domestication from two different subspecies. Proceedings of the Royal Society of London B, 269(1494), 893904.Google Scholar
Hobbs, N.T. & Miller, M.W. (1992) Interactions between pathogens and hosts: simulation of pasteurellosis epizootics in bighorn sheep populations. In: Wildlife 2001: Populations (pp.9971007). Dordrecht: Springer.Google Scholar
Hogg, J.T. (1984) Mating in bighorn sheep: multiple creative male strategies. Science, 225, 526530.Google Scholar
Hogg, J.T., Hass, C.C. & Jenni, D.A. (1992) Sex-biased maternal expenditure in Rocky Mountain bighorn sheep. Behavioral Ecology and Sociobiology, 31(4), 243251.Google Scholar
Jansen, B.D., Krausman, P.R., Heffelfinger, J.R., et al. (2007) Population dynamics and behavior of bighorn sheep with infectious keratoconjunctivitis. Journal of Wildlife Management, 71(2), 571575.Google Scholar
Jones, G.E., Keir, W.A. & Gilmour, J.S. (1985). The pathogenicity of Mycoplasma ovipneumoniae and Mycoplasma arginini in ovine and caprine tracheal organ cultures. Journal of Comparative Pathology, 95(4), 477487.Google Scholar
Jorgenson, J.T., Festa-Bianchet, M., Lucherini, M., et al. (1993) Effects of body size, population density, and maternal characteristics on age at first reproduction in bighorn ewes. Canadian Journal of Zoology, 71(12), 25092517.Google Scholar
Joseph, M.B., Mihaljevic, J.R., Arellano, A.L., et al. (2013) Taming wildlife disease: bridging the gap between science and management. Journal of Applied Ecology, 50(3), 702712.Google Scholar
Justice-Allen, A.E., Butler, E., Pebworth, J., et al. (2016) Investigation of pneumonia mortalities in a Mycoplasma-positive desert bighorn sheep population and detection of a different strain of Mycoplasma ovipneumoniae. Biennial Symposium of the Northern Wild Sheep and Goat Council, 20, 6872.Google Scholar
Kamath, P.L., Cross, P.C., Cassirer, E.F., et al. (2016) Genetic linkages among Mycoplasma ovipneumoniae outbreaks in wild and domestic sheep and goats. Biennial Symposium of the Northern Wild Sheep and Goat Council, 20, 113.Google Scholar
Keeling, M.J. & Rohani, P. (2008) Modeling Infectious Diseases in Humans and Animals. Princeton, NJ: Princeton University Press.Google Scholar
Kilpatrick, A.M., Daszak, P., Jones, M.J., et al. (2006) Host heterogeneity dominates West Nile virus transmission. Proceedings of the Royal Society of London B, 273, 23272333.Google Scholar
Klepac, P., Pomeroy, L.W., Bjørnstad, O.N., et al. (2009) Stage-structured transmission of phocine distemper virus in the Dutch 2002 outbreak. Proceedings of the Royal Society of London B, 276(1666), 24692476.Google Scholar
Kuussaari, M., Bommarco, R., Heikkinen, R.K., et al. (2009) Extinction debt: a challenge for biodiversity conservation. Trends in Ecology & Evolution, 24(10), 564571.Google Scholar
Langwig, K.E., Frick, W.F., Bried, J.T., et al. (2012) Sociality, density-dependence and microclimates determine the persistence of populations suffering from a novel fungal disease, white-nose syndrome. Ecology Letters, 15(9), 10501057.Google Scholar
Lloyd-Smith, J.O., George, D., Pepin, K.M., et al. (2009) Epidemic dynamics at the human–animal interface. Science, 326, 13621367.Google Scholar
Loison, A., Festa-Bianchet, M., Gaillard, J.M., et al. (1999) Age‐specific survival in five populations of ungulates: evidence of senescence. Ecology, 80(8), 25392554.Google Scholar
Long, R.A., Rachlow, J.L. & Kie, J.G. (2008) Effects of season and scale on response of elk and mule deer to habitat manipulation. Journal of Wildlife Management, 72(5), 11331142.Google Scholar
Maksimović, Z., De la Fe, C., Amores, J., et al. (2016) Comparison of phenotypic and genotypic profiles among caprine and ovine Mycoplasma ovipneumoniae strains. Veterinary Record, 180(7), 180.Google Scholar
Manlove, K.R., Cassirer, E.F., Cross, P.C., et al. (2014) Costs and benefits of group living with disease: a case study of pneumonia in bighorn lambs (Ovis canadensis). Proceedings of the Royal Society of London B, 281(1797), 20142331.Google Scholar
Manlove, K., Cassirer, E.F., Cross, P.C., et al. (2016a) Disease introduction is associated with a phase transition in bighorn sheep demographics. Ecology, 97(10), 25932602.Google Scholar
Manlove, K.R., Cassirer, E.F., Plowright, R.K., et al. (2017) Contact and contagion: probability of transmission given contact varies with demographic state in bighorn sheep. Journal of Animal Ecology, 86(4), 908920.Google Scholar
Manlove, K.R., Walker, J.G., Craft, M.E., et al. (2016b) ‘One Health’ or three? Publication silos among the One Health disciplines. PLoS Biology, 14(4), e1002448.Google Scholar
McAdoo, C., Wolff, P. & Cox, M. (2010) Investigation of Nevada’s 2009–2010 East Humboldt Range and Ruby Mountain bighorn dieoff. Biennial Symposium of the Northern Wild Sheep and Goal Council, 17, 5152.Google Scholar
Meldrum, G.E. & Ruckstuhl, K.E. (2009) Mixed-sex group formation by bighorn sheep in winter: trading costs of synchrony for benefits of group living. Animal Behaviour, 77(4), 919929.Google Scholar
Miller, D.S., Hoberg, E., Weiser, G., et al. (2012) A review of hypothesized determinants associated with bighorn sheep (Ovis canadensis) die-offs. Veterinary Medicine International, 2012, 796527.Google Scholar
Miller, M.W., Vayhinger, J.E., Bowden, D.C., et al. (2000) Drug treatment for lungworm in bighorn sheep: reevaluation of a 20-year-old management prescription. Journal of Wildlife Management, 64(2), 505512.Google Scholar
Mincher, B.J., Ball, R.D., Houghton, T.P., et al. (2008) Some aspects of geophagia in Wyoming bighorn sheep (Ovis canadensis). European Journal of Wildlife Research, 54(2), 193198.Google Scholar
Monello, R.J., Murray, D.L. & Cassirer, E.F. (2001) Ecological correlates of pneumonia epizootics in bighorn sheep herds. Canadian Journal of Zoology, 79(8), 14231432.Google Scholar
Niang, M., Rosenbusch, R.F., Andrews, J.J., et al. (1998) Demonstration of a capsule on Mycoplasma ovipneumoniae. American Journal of Veterinary Research, 59(5), 557562.Google Scholar
Nicholas, R., Ayling, R. & McAuliffe, L. (2008) Mycoplasma Diseases of Ruminants: Disease, Diagnosis and Control. Cambridge, MA: CABI.Google Scholar
O’Brien, J.M., O’Brien, C.S., McCarthy, C., et al. (2014) Incorporating foray behavior into models estimating contact risk between bighorn sheep and areas occupied by domestic sheep. Wildlife Society Bulletin, 38(2), 321331.Google Scholar
Peel, A.J., Pulliam, J.R.C., Luis, A.D., et al. (2014) The effect of seasonal birth pulses on pathogen persistence in wild mammal populations. Proceedings of the Royal Society of London B, 281(1786), 20132962.Google Scholar
Plowright, R.K., Manlove, K.R., Besser, T.E., et al. (2017) Persistent carriers explain epidemiological features of pneumonia in bighorn sheep (Ovis canadensis). Ecology Letters, 20(10), 13251336.Google Scholar
Plowright, R.K., Manlove, K., Cassirer, E.F., et al. (2013) Use of exposure history to identify patterns of immunity to pneumonia in bighorn sheep (Ovis canadensis). PLoS ONE, 8(4), e61919.Google Scholar
Rezaei, H.R., Naderi, S., Chintauan-Marquier, I.C., et al. (2010) Evolution and taxonomy of the wild species of the genus Ovis (Mammalia, Artiodactyla, Bovidae). Molecular Phylogenetics and Evolution, 54, 315326.Google Scholar
Rifatbegović, M., Maksimović, Z. & Hulaj, B. (2011) Mycoplasma ovipneumoniae associated with severe respiratory disease in goats. Veterinary Record, 168, 565a.Google Scholar
Ruckstuhl, K.E. (1998) Foraging behaviour and sexual segregation in bighorn sheep. Animal Behaviour, 56(1), 99106.Google Scholar
Ruckstuhl, K.E. & Festa‐Bianchet, M. (2001) Group choice by subadult bighorn rams: trade‐offs between foraging efficiency and predator avoidance. Ethology, 107(2), 161172.Google Scholar
Scheele, B.C., Hunter, D.A., Brannelly, L.A., et al. (2017) Reservoir–host amplification of disease impact in an endangered amphibian. Conservation Biology, 31(3), 592600.Google Scholar
Sells, S.N., Mitchell, M.S., Nowak, J.J., et al. (2015) Modeling risk of pneumonia epizootics in bighorn sheep. The Journal of Wildlife Management, 79(2), 195210.Google Scholar
Shannon, J.M., Whiting, J.C., Larsen, R.T., et al. (2014) Population response of reintroduced bighorn sheep after observed commingling with domestic sheep. European Journal of Wildlife Research, 60(5), 737748.Google Scholar
Simmons, W.L., Daubenspeck, J.M., Osborne, J.D., et al. (2013) Type 1 and Type 2 strains of Mycoplasma pneumoniae form different biofilms. Microbiology, 159, 737747.Google Scholar
Smith, J.B., Jenks, J.A., Grovenburg, T.W., et al. (2014) Disease and predation: sorting out causes of a bighorn sheep (Ovis canadensis) decline. PLoS ONE, 9(2), e88271.Google Scholar
Subramaniam, R., Shanthalingam, S., Bavananthasivam, J., et al. (2014) Bighorn sheep × domestic sheep hybrids survive Mannheimia haemolytica challenge in the absence of vaccination. Veterinary Microbiology, 170, 278283.Google Scholar
Swinton, J., Harwood, J., Grenfell, B.T., et al. (1998) Persistence thresholds for phocine distemper virus infection in harbour seal Phoca vitulina metapopulations. Journal of Animal Ecology, 67, 5468.Google Scholar
USDA. (2015) Mycoplasma ovipneumoniae on US sheep operations. USDA-APHIS-VS-CEAH. Fort Collins, CO: USDA. 708.0615Google Scholar
Weiser, G.C., DeLong, W.J., Paz, J.L., et al. (2003) Characterization of Pasteurella multocida associated with pneumonia in bighorn sheep. Journal of Wildlife Diseases, 39(3), 536544.Google Scholar
Western Association of Fish and Wildlife Agencies Wild Sheep Working Group. (2012) Recommendations for domestic sheep and goat management in wild sheep habitat. Available from: www.wildsheepworkinggroup.com/resources/publications/, Western Association of Fish and Wildife Agencies Wild Sheep Working Group.Google Scholar
Whiting, J.C., Olson, D.D., Shannon, J.M., et al. (2012) Timing and synchrony of births in bighorn sheep: implications for reintroduction and conservation. Wildlife Research, 39(7), 565572.Google Scholar

References

Albon, S.D., Irvine, R.J., Halvorsen, O., et al. (2017) Contrasting effects of summer and winter warming on body mass explain population dynamics in a food-limited Arctic herbivore. Global Change Biology, 23, 13741389. doi:10.1111/gcb.13435Google Scholar
Albon, S.D., Stien, A., Irvine, R.J., et al. (2002) The role of parasites in the dynamics of a reindeer population. Proceedings of the Royal Society of London B, 269, 16251632.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
Armour, J. (1989) The influence of host immunity on the epidemiology of trichostrongyle infections in cattle. Veterinary Parasitology, 32, 519.Google Scholar
Arneberg, P., Folstad, I. & Karter, A.J. (1996) Gastrointestinal nematodes depress food intake in naturally infected reindeer. Parasitology, 112, 213219.Google Scholar
Arneberg, P., Skorping, A., Grenfell, B. & Read, A.F. (1998) Host densities as determinants of abundance in parasite communities. Proceedings of the Royal Society of London B, 265, 12831289. DOI 10.1098/rspb.1998.0431Google Scholar
Begon, M., Townsend, C.R. & Harper, J. (2005) Ecology: From Individuals to Ecosystems, 4th edition. Oxford: Wiley-Blackwell.Google Scholar
Bjørkvoll, E., Lee, A.M., Grøtan, V., et al. (2016) Demographic buffering of life histories? Implications of the choice of measurement scale. Ecology, 97, 4047.Google Scholar
Boag, B. & Thomas, R.J. (1977) Epidemiological studies on gastro-intestinal nematode parasites of sheep: the seasonal number of generations and succession of species. Research in Veterinary Science, 22, 6277.Google Scholar
Bye, K. (1987). Abomasal nematodes from three Norwegian wild reindeer populations. Canadian Journal of Zoology, 65, 677680.Google Scholar
Bye, K. & Halvorsen, O. (1983). Abomasal nematodes of the Svalbard reindeer (Rangifer tarandus platyrhynchus Vrolik). Journal of Wildlife Diseases, 19, 101103.Google Scholar
Bye, K., Halvorsen, O. & Nilssen, K. (1987) Immigration and regional distribution of abomasal nematodes of Svalbard reindeer. Journal of Biogeography, 14, 451458.Google Scholar
Carlsson, A.M., Albon, S.D., Coulson, S.J., et al. (2018) Little impact of over-winter parasitism on a free-ranging ungulate in the high Arctic. Functional Ecology; 32, 10461056. https://doi.org/10.1111/1365–2435.13037Google Scholar
Carlsson, A.M., Irvine, R.J., Wilson, K., et al. (2012) Disease transmission in an extreme environment: nematode parasites infect reindeer during the Arctic winter. International Journal of Parasitology, 42,786795. doi:10.1016/j.ijpara.2012.05.007.Google Scholar
Carlsson, A.M., Irvine, R.J., Wilson, K. & Coulson, S.J. (2013) Adaptations to the Arctic: low-temperature development and cold tolerance in the free-living stages of a parasitic nematode from Svalbard. Polar Biology, 36, 9971005. doi:10.1007/s00300-013-1323-7.Google Scholar
Carlsson, A.M., Wilson, K. & Irvine, R.J. (2012) Development and application of a delayed-release anthelmintic intra-ruminal bolus system for experimental manipulation of nematode worm burdens. Parasitology, 139, 10861092.Google Scholar
Colditz, I.G., Watson, D.L., Gray, G.D. & Eady, S.J. (1996) Some relationships between age, immune responsiveness and resistance to parasites in ruminants. International Journal for Parasitology, 26, 869877. doi:10.1016/S0020-7519(96)80058-0.Google Scholar
Coulson, T., Catchpole, E.A., Albon, S.D., et al. (2001) Age, sex, density, winter weather, and population crashes in Soay sheep. Science, 292, 15281531.Google Scholar
Crofton, H.D. (1957) Nematode parasite populations in sheep on lowland farms. III. The seasonal incidence of species. Parasitology, 47, 304318.Google Scholar
Crofton, H.D. (1963) Nematode parasite populations in sheep and on pasture. Technical Communication No. 35 of the Commonwealth Bureau of Helminthology. St Albans, UK.Google Scholar
Dallas, J.F., Irvine, R.J., Halvorsen, O. & Albon, S.D. (2000) Identification by polymerase chain reaction (PCR) of Marshallagia marshalli and Ostertagia gruehneri from Svalbard reindeer. International Journal of Parasitology, 30, 863866.Google Scholar
Drózdz, J. (1965) Studies on helminths and helminthiases in Cervidae. I. Revision of the subfamily Ostertagiinae Sarwar, 1956 and an attempt to explain the phylogenesis of its representatives. Acta Parasitologica Polonica, 13, 445481.Google Scholar
Drózdz, J. (1995) Polymorphism in the Ostertagiinae Lopez-Neyra, 1947 and comments on the systematics of these nematodes. Systematic Parasitology, 32, 9199.Google Scholar
El-Azazy, O.M.E. (1995) Seasonal changes and inhibited development of the abomasal nematodes of sheep and goats in Saudi Arabia. Veterinary Parasitology, 58, 9198.Google Scholar
Forbes, A.B., Huckle, C.A., Gibb, M.J., Rook, A.J. & Nuthall, R. (2000) Evaluation of the effects of nematode parasitism on grazing behaviour, herbage intake and growth in young grazing cattle. Veterinary Parasitology, 90, 111118.Google Scholar
Fox, M.T. (1997) Pathophysiology of infection with gastrointestinal nematodes in domestic ruminants: recent developments. Veterinary Parasitology, 72, 285308.Google Scholar
Førland, E.J., Benestad, B., Hanssen-Bauer, I., Haugen, J.E. & Skaugen, T.E. (2011) Temperature and precipitation development at Svalbard 1900–2100. Advances in Meteorology, 2011, 893790. doi:10.1155/2012/893790.Google Scholar
Gjelten, H.J., Nordli, O., Isaksen, K., et al. (2016) Air temperature variations and gradients along the coast and fjords of western Spitsbergen. Polar Biology, 35, 29878. doi.org/10.3402/polar.v35.29878Google Scholar
Grenfell, B.T. (1988) Gastrointestinal nematode parasites and the stability and productivity of intensive ruminant grazing systems. Philosophical Transactions of the Royal Society of London B, 321, 541563.Google Scholar
Grenfell, B.T. (1992) Parasitism and the dynamics of ungulate grazing systems. American Naturalist, 139, 907929.Google Scholar
Grenfell, B.T., Wilson, K., Finkenstädt, B.F., et al. (1998) Noise and determinism in synchronized sheep dynamics. Nature, 394, 674677.Google Scholar
Gulland, F.M.D. (1992) The role of nematode parasites in Soay sheep (Ovis aries L.) mortality during a population crash. Parasitology, 105, 493503.Google Scholar
Halvorsen, O. & Bye, K. (1986). Parasitter i svalbardrein 1. Rundmark i lùpen [in Norwegian]. In: Øritsland, N.A. (ed.), Svalbardreinen og dens livsgrunnlag (pp. 120133). Oslo: Universitetsforlaget.Google Scholar
Halvorsen, O. & Bye, K. (1999) Parasites, biodiversity, and population dynamics in an ecosystem in the High Arctic. Veterinary Parasitology, 84, 205227.Google Scholar
Halvorsen, O., Stien, A., Irvine, J., Langvatn, R. & Albon, S. (1999) Evidence for continued transmission of parasitic nematodes in reindeer during the Arctic winter. International Journal of Parasitology, 29, 567579.Google Scholar
Hansen, B.B., Aanes, R., Herfindal, I., Kohler, J. & Sæther, B-E. (2011) Climate, icing, and wild arctic reindeer: past relationships and future prospects. Ecology, 92, 19171923.Google Scholar
Hansen, B.B., Aanes, R. & Sæther, B-E. (2010) Feeding-crater selection by High-arctic reindeer facing ice-blocked pastures. Canadian Journal of Zoology, 88, 170177.Google Scholar
Hansen, B.B., Isaksen, K., Benestad, R.E., et al. (2014) Warmer and wetter winters: characteristics and implications of an extreme event in the High Arctic. Environmental Research Letters, 9, 114021. doi:10.1088/1748-9326/9/11/114021Google Scholar
Heinzmann, D., Barbour, A.D. & Torgerson, P.R. (2009) Compound processes as models for clumped parasite data. Mathematical Biosciences, 222(1), 2735. DOI:10.1016/j.mbs.2009.08.007.Google Scholar
Hoar, B., Eberhardt, A. & Kutz, S. (2012a) Obligate larval inhibition of Ostertagia gruehneri in Rangifer tarandus? Causes and consequences in an Arctic system. Parasitology, 139, 13391345. doi:10.1017/S0031182012000601Google Scholar
Hoar, B.M., Ruckstuhl, K. & Kutz, S. (2012b) Development and availability of the free-living stages of Ostertagia gruehneri, an abomasal parasite of barrenground caribou (Rangifer tarandus groenlandicus) on the Canadian tundra. Parasitology, 139, 10931100.Google Scholar
Hoberg, E.P., Galbreath, K.E., Cook, J.A., Kutz, S.J. & Polley, L. (2012) Northern host–parasite assemblages: history and biogeography on the borderlands of episodic climate and environmental transition. Advances in Parasitology, 79, 197.Google Scholar
Hoberg, E.P., Kocan, A.A. & Richard, L.G. (2001) Gastrointestinal strongyles in wild ruminants. In: Samuel, W.M., Pybus, M.J. & Kocan, A.A. (eds.), Parasitic Diseases of Wild Mammals (pp. 193227). London: Manson Publishing/Veterinary Press.Google Scholar
Hudson, P.J., Cattadori, I.M., Boag, B. & Dobson, A.P. (2006) Climate disruption and parasite–host dynamics: patterns and processes associated with warming and the frequency of extreme climatic events. Journal of Helminthology, 80, 175182.Google Scholar
Hutchings, M.R., Kyriazakis, I., Papachristou, T.G., Gordon, I.J. & Jackson, F. (2000) The herbivores’ dilemma: trade-offs between nutrition and parasitism in foraging decisions. Oecologia, 124, 242251.Google Scholar
Igrashev, I.K. (1973) Helminths and Helminthoses of the Karakul Sheep. Tashkent, Uzbekistan: National Academy of Sciences of the Uzbekistan Soviet Socialist Republic, Institute of Zoology and Parasitology.Google Scholar
Irvine, R.J. (2000) Use of moxidectin treatment in the investigation of abomasal nematodiasis in wild reindeer (Rangifer tarandus platyrhynchus). Veterinary Record, 147, 570573.Google Scholar
Irvine, R.J. (2001) Contrasting life-history traits and population dynamics in two co-existing gastrointestinal nematodes of Svalbard reindeer. PhD Thesis. University of Stirling, Stirling.Google Scholar
Irvine, R.J., Stien, A., Dallas, J.F., et al. (2001) Contrasting regulation of fecundity in two abomasal nematodes of Svalbard reindeer (Rangifer tarandus platyrhynchus). Parasitology, 122, 673681.Google Scholar
Irvine, R.J., Stien, A., Halvorsen, O., Langvatn, R. & Albon, S.D. (2000) Life-history strategies and population dynamics of abomasal nematodes in Svalbard reindeer (Rangifer tarandus platyrhynchus). Parasitology, 120, 297311.Google Scholar
Jolles, A.E. & Ezenwa, V.O. (2015) Ungulates as model systems for the study of disease processes in natural populations. Journal of Mammalogy, 96, 415.Google Scholar
Kattsov, V.M., Källén, E., Symon, C., Arris, L. & Hill, B. (2005) Future climate change: modeling and scenarios for the Arctic. In: Arctic Climate Impact Assessment (pp. 100150). www.acia.uaf.eduGoogle Scholar
Kutz, S.J., Ducrocq, J., Verocai, G.G., et al. (2012) Parasites in ungulates of Arctic North America and Greenland: a view of contemporary diversity, ecology, and impact in a world under change. Advances in Parasitology, 79, 99252.Google Scholar
Kutz, S.J., Hoberg, E.P., Polley, L. & Jenkins, E.J. (2005) Global warming is changing the dynamics of Arctic host–parasite systems. Proceedings of the Royal Society of London B, 272, 25712576. doi:10.1098/rspb.2005.3285Google Scholar
Kutz, S.J., Jenkins, E.J., Veitch, A.M., et al. (2009) The Arctic as a model for anticipating, preventing, and mitigating climate change impacts on host–parasite interactions. Veterinary Parasitology, 3, 217228.Google Scholar
Lebreton, J.-D., Burnham, K. P., Clobert, J. & Anderson, D. R. (1992) Modeling survival and testing biological hypotheses using marked animals: a unified approach with case studies. Ecological Monographs, 62, 67118.Google Scholar
May, R.M. & Anderson, R.M. (1978) Regulation and stability of host–parasite population interactions. II. Destabilising processes. Journal of Animal Ecology, 47, 249267.Google Scholar
McCallum, H. (2000) Population Parameters. Oxford: Blackwell Science Ltd.Google Scholar
McCallum, H., Fenton, A., Hudson, P.J., et al. (2017) Breaking beta: deconstructing the parasite transmission function. Philosophical Transactions of the Royal Society of London Series B, 372, 20160084. doi:10.1098/rstb.2016.0084Google Scholar
Milner, J.M., Stien, A., Irvine, R.J., et al. (2003) Body condition in Svalbard reindeer and the use of blood parameters as indicators of condition and fitness. Canadian Journal of Zoology, 81, 15661578.Google Scholar
Morgan, E.R., Medley, G.F., Torgerson, P.R., Shaikenov, B.S. & Milner-Gulland, E.J. (2007) Parasite transmission in a migratory multiple host system. Ecological Modelling, 200, 511520. doi:10.1016/j.ecolmodel.2006.09.002.Google Scholar
Morgan, E.R., Shaikenov, B., Torgerson, P.R., Medley, G.F. & Milner-Gulland, E.J. (2005) Helminths of Saiga antelope in Kazakhstan: implications for conservation and livestock production. Journal of Wildlife Diseases, 41, 149162. https://doi.org/10.7589/0090-3558-41.1.149.Google Scholar
Murray, D.L., Cary, J.R. & Keith, L.B. (1997) Interactive effects of sublethal nematodes and nutritional status on snowshoe hare vulnerability to predation. Journal of Animal Ecology, 66, 250264.Google Scholar
Omsjoe, E.H., Stien, A., Irvine, R.J., et al. (2009) Evaluating capture stress and its effect on reproductive success of Svalbard reindeer. Canadian Journal Zoology, 87, 7385.Google Scholar
Pedersen, A.B. & Fenton, A. (2015) The role of antiparasite treatment experiments in assessing the impact of parasites on wildlife. Trends in Parasitology, 31, 200211. 10.1016/j.pt.2015.02.004.Google Scholar
Pedersen, A.B. & Grieves, T. J. (2008) The interaction of parasites and resource cause crashes in a wild mouse population. Journal of Animal Ecology, 77, 370377.Google Scholar
Redpath, S.M., Mougeot, F., Leckie, F.M., Elston, D.A. & Hudson, P. J. (2006) Testing the role of parasites in driving the cyclic population dynamics of a gamebird. Ecology Letters, 9, 410418.Google Scholar
Reimers, E. (1977) Population dynamics of two subpopulations of reindeer in Svalbard. Arctic and Alpine Research, 9, 369381.Google Scholar
Reimers, E. (1982) Winter mortality and population trends of reindeer on Svalbard, Norway. Arctic and Alpine Research, 14, 295300. doi:10.2307/1550792.Google Scholar
Reimers, E. & Ringberg, T. (1983) Seasonal changes in body weights of Svalbard reindeer from birth to maturity. Acta Zoologica Fennica, 175, 6972.Google Scholar
Ropstad, E., Johansen, O., King, C., et al. (1999) Comparison of plasma progesterone, transrectal ultrasound and pregnancy specific proteins (PSPB) used for pregnancy diagnosis in reindeer. Acta Veterinaria Scandinavia, 40, 151162.Google Scholar
Solberg, E.J., Jordhøy, P., Strand, O., et al. (2001) Effects of density-dependence and climate on the dynamics of a Svalbard reindeer population. Ecography, 24, 441451.Google Scholar
Stien, A., Ims, R.A., Albon, S.D., et al. (2012) Congruent responses to weather variability in high arctic herbivores. Biology Letters, 8, 10021005.Google Scholar
Stien, A., Irvine, R.J., Langvatn, R., et al. (2002a) The impact of gastrointestinal nematodes on wild reindeer: experimental and cross-sectional studies. Journal of Animal Ecology, 71, 937945.Google Scholar
Stien, A., Irvine, R.J., Langvatn, R., Albon, S.D. & Halvorsen, O. (2002b) The population dynamics of Ostertagia gruehneri in reindeer: a model for the seasonal and intensity dependent variation in nematode fecundity. International Journal of Parasitology, 32, 991996.Google Scholar
Stien, A., Loe, L.E., Mysterud, A., et al. (2010) Icing events trigger range displacement in a high-arctic ungulate. Ecology, 91, 915920.Google Scholar
Tompkins, D.M. & Begon, M. (1999) Parasites can regulate wildlife populations. Parasitology Today, 15, 311313.Google Scholar
Tompkins, D.M., Dobson, A.P., Arneberg, P., et al. (2002) Parasites and host population dynamics. In: Hudson, P., Rizzoli, A., Grenfell, B., Heesterbeek, H. & Dobson, A. (eds.), The Ecology of Wildlife Diseases (pp. 4562). New York, NY: Oxford University Press.Google Scholar
Tompkins, D.M., Dunn, A. M., Smith, M. J. & Telfer, S. (2011) Wildlife diseases: from individuals to ecosystems. Journal of Animal Ecology, 80, 1938.Google Scholar
Townsend, S.E., Newey, S., Thirgood, S.J., Matthews, L. & Haydon, D.T. (2009) Can parasites drive population cycles in mountain hares? Proceedings of the Royal Society of London B, 276, 16111617. doi:10.1098/rspb.2008.1669.Google Scholar
Tyler, N.J.C. (1987) Natural limitation of the abundance of the high Arctic Svalbard reindeer. PhD thesis, University of Cambridge.Google Scholar
Tyler, N.J.C., Forchhammer, M.C. & Øritsland, N.A. (2008) Nonlinear effects of climate and density in the dynamics of a fluctuating population of reindeer. Ecology, 98, 16751686.Google Scholar
Tyler, N. & Øritsland, N.A. (1989) Why don’t Svalbard reindeer migrate? Holarctic Ecology, 12, 369376.Google Scholar
Tyler, N. & Øritsland, N.A. (1999) Varig ustabilitet og bestandsregulering hos Svalbardrein (in Norwegian). In: Bengtson, S.A., Mehlum, F. & Severinsen, T. (eds.), Svalbardtundraens økologi (pp. 139147). Tromsø: Norsk Polarinstituttmdelelser nr. 150.Google Scholar
Vandergrift, K.J., Raffel, T.R. & Hudson, P.J. (2008) Parasites prevent summer breeding in white-footed mice, Peromyscus leucopus. Ecology, 89, 22512258.Google Scholar
Van der Wal, R. & Stien, A. (2014) High-arctic plants like it hot: a long-term investigation of between-year variability in plant biomass. Ecology, 95, 34143427.Google Scholar
Wegener, C. & Odasz-Albrigtsen, A.M. (1998) Do Svalbard reindeer regulate standing crop in the absence of predators? A test of the “exploitation ecosystems” model. Oecologia, 116, 202206.Google Scholar
White, G.C. & Burnham, K.P. (1999) Program Mark: survival estimation from populations of marked animals. Bird Study, 46, S120S139.Google Scholar
Wilson, K. & Grenfell, B.T. (1997) Generalised linear modelling for parasitologists. Parasitology Today, 13, 3338.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
×