Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-75dct Total loading time: 0 Render date: 2024-05-17T20:28:44.373Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 August 2012

Melanie J. Hatcher
Affiliation:
University of Bristol
Alison M. Dunn
Affiliation:
University of Leeds
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
Parasites in Ecological Communities
From Interactions to Ecosystems
, pp. 393 - 438
Publisher: Cambridge University Press
Print publication year: 2011

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

Abrams, P. A. (1992). Predators that benefit prey and prey that harm predators: unusual effects of interacting foraging adaptations. American Naturalist, 140, 573–600.CrossRefGoogle Scholar
Abrams, P. A. & Matsuda, H. (1996). Positive indirect effects between prey species that share predators. Ecology, 77, 610–616.CrossRefGoogle Scholar
Abrams, P. A. & Roth, J. D. (1994). The effects of enrichment of three-species food-chains with nonlinear functional-responses. Ecology, 75, 1118–1130.CrossRefGoogle Scholar
Agrawal, A. A. & Kotanen, P. M. (2003). Herbivores and the success of exotic plants: a phylogenetically controlled experiment. Ecology Letters, 6, 712–715.CrossRefGoogle Scholar
Agrawal, A. A., Kotanen, P. M., Mitchell, C. E., Power, A. G., Godsoe, W. & Klironomos, J. (2005). Enemy release? An experiment with congeneric plant pairs and diverse above- and belowground enemies. Ecology, 86, 2979–2989.CrossRefGoogle Scholar
Aguirre, A. A., Ostfeld, R. S., Tabor, G. M., House, C. A. & Pearl, M. C. (eds) (2002). Conservation Medicine: Ecological Health in Practice. New York: Oxford University Press.
Alexander, H. M. & Holt, R. D. (1998). The interaction between plant competition and disease. Perspectives in Plant Ecology Evolution and Systematics, 1, 206–220.CrossRefGoogle Scholar
Aliabadi, B. W. & Juliano, S. A. (2002). Escape from gregarine parasites affects the competitive interactions of an invasive mosquito. Biological Invasions, 4, 283–297.CrossRefGoogle ScholarPubMed
Allan, B. F., Keesing, F. & Ostfeld, R. S. (2003). Effect of forest fragmentation on Lyme disease risk. Conservation Biology, 17, 267–272.CrossRefGoogle Scholar
Allan, B. F., Langerhans, R. B., Ryberg, W. A., Landesman, W. J., Griffin, N. W., Katz, R. S., Oberle, B. J., Schutzenhofer, M. R., Smyth, K. N., St Maurice, A., Clark, L., Crooks, K. R., Hernandez, D. E., McLean, R. G., Ostfeld, R. S. & Chase, J. M. (2009). Ecological correlates of risk and incidence of West Nile virus in the United States. Oecologia, 158, 699–708.CrossRefGoogle ScholarPubMed
Altizer, S., Dobson, A., Hosseini, P., Hudson, P., Pascual, M. & Rohani, P. (2006). Seasonality and the dynamics of infectious diseases. Ecology Letters, 9, 467–484.CrossRefGoogle ScholarPubMed
Ameloot, E., Verlinden, G., Boeckx, P., Verheyen, K. & Hermy, M. (2008). Impact of hemiparasitic Rhinanthus angustifolius and R. minor on nitrogen availability in grasslands. Plant and Soil, 311, 255–268.CrossRefGoogle Scholar
Amundsen, P. A., Lafferty, K. D., Knudsen, R., Primicerio, R., Klemetsen, A. & Kuris, A. M. (2009). Food web topology and parasites in the pelagic zone of a subarctic lake. Journal of Animal Ecology, 78, 563–572.CrossRefGoogle ScholarPubMed
Anderson, A., McCormack, S., Helden, A. J., Sheridan, H., Kinsella, A. & Purvis, G. (2011). The potential of parasitoid Hymenoptera as bioindicators of arthropod diversity in agricultural grasslands. Journal of Applied Ecology. In press.CrossRef
Anderson, R. A., Koella, J. C. & Hurd, H. (1999). The effect of Plasmodium yoelii nigeriensis infection on the feeding persistence of Anopheles stephensi Liston throughout the sporogonic cycle. Proceedings of the Royal Society of London Series B – Biological Sciences, 266, 1729–1733.CrossRefGoogle ScholarPubMed
Anderson, R. C. (1972). The ecological relationships of meningeal worm and native cervids in North America. Journal of Wildlife Diseases, 8, 304–310.CrossRefGoogle ScholarPubMed
Anderson, R. M. & May, R. M. (1981). The population dynamics of micro-parasites and their invertebrate hosts. Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 291, 451–524.CrossRefGoogle Scholar
Anderson, R. M. & May, R. M. (1982). Coevolution of hosts and parasites. Parasitology, 85, 411–426.CrossRefGoogle ScholarPubMed
Anderson, R. M. & May, R. M. (1985). Age-related changes in the rate of disease transmissions: implications for the design of vaccination programs. Journal of Hygiene, 94, 365–436.CrossRefGoogle Scholar
Anderson, R. M. & May, R. M. (1986). The invasion, persistence and spread of infectious-diseases within animal and plant communities. Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 314, 533–570.CrossRefGoogle ScholarPubMed
Anderson, R. M. & May, R. M. (1991). Infectious Diseases of Humans: Dynamics and Control. Oxford: Oxford University Press.Google Scholar
Andreadis, T. G. (1980). Nosema pyrausta infection in Macrocentrus grandii, a braconid parasite of the European corn-borer, Ostrinia nubilalis. Journal of Invertebrate Pathology, 35, 229–233.CrossRefGoogle Scholar
Antia, R., Regoes, R. R., Koella, J. C. & Bergstrom, C. T. (2003). The role of evolution in the emergence of infectious diseases. Nature, 426, 658–661.CrossRefGoogle ScholarPubMed
Arim, M. & Marquet, P. A. (2004). Intraguild predation: a widespread interaction related to species biology. Ecology Letters, 7, 557–564.CrossRefGoogle Scholar
Arinaminpathy, N. & McLean, A. R. (2009). Evolution and emergence of novel human infections. Proceedings of the Royal Society of London Series B – Biological Sciences, 276, 3937–3943.CrossRefGoogle ScholarPubMed
Arnold, A. E., Maynard, Z., Gilbert, G. S., Coley, P. D. & Kursar, T. A. (2000). Are tropical fungal endophytes hyperdiverse?Ecology Letters, 3, 267–274.CrossRefGoogle Scholar
Aron, J. L. & May, R. M. (1982). The population dynamics of malaria. In Anderson, R. M. (ed.), The Population Dynamics of Infectious Diseases: Theory and Applications. London: Chapman & Hall.Google Scholar
Askary, H. & Brodeur, J. (1999). Susceptibility of larval stages of the aphid parasitoid Aphidius nigripes to the entomopathogenic fungus Verticillium lecanii. Journal of Invertebrate Pathology, 73, 129–132.CrossRefGoogle ScholarPubMed
Auger, P., Mchich, R., Chowdhury, T., Sallet, G., Tchuente, M. & Chattopadhyay, J. (2009). Effects of a disease affecting a predator on the dynamics of a predator–prey system. Journal of Theoretical Biology, 258, 344–351.CrossRefGoogle ScholarPubMed
Augspurger, C. K. (1983). Seed dispersal of the tropical tree, Platypodium elegans, and the escape of its seedlings from fungal pathogens. Journal of Ecology, 71, 759–771.CrossRefGoogle Scholar
Augspurger, C. K. (1984). Seedling survival of tropical tree species – interactions of dispersal distance, light-gaps, and pathogens. Ecology, 65, 1705–1712.CrossRefGoogle Scholar
Bailes, E., Gao, F., Bibollet-Ruche, F., Courgnaud, V., Peeters, M., Marx, P. A., Hahn, B. H. & Sharp, P. M. (2003). Hybrid origin of SIV in chimpanzees. Science, 300, 1713.CrossRefGoogle ScholarPubMed
Bally, M. & Garrabou, J. (2007). Thermodependent bacterial pathogens and mass mortalities in temperate benthic communities: a new case of emerging disease linked to climate change. Global Change Biology, 13, 2078–2088.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, 88–94.CrossRefGoogle ScholarPubMed
Baranowski, E., Ruiz-Jarabo, C. M. & Domingo, E. (2001). Evolution of cell recognition by viruses. Science, 292, 1102–1105.CrossRefGoogle ScholarPubMed
Bardgett, R. D., Smith, R. S., Shiel, R. S., Peacock, S., Simkin, J. M., Quirk, H. & Hobbs, P. J. (2006). Parasitic plants indirectly regulate below-ground properties in grassland ecosystems. Nature, 439, 969–972.CrossRefGoogle ScholarPubMed
Barrett, L. G., Kniskern, J. M., Bodenhausen, N., Zhang, W. & Bergelson, J. (2009). Continua of specificity and virulence in plant host–pathogen interactions: causes and consequences. New Phytologist, 183, 513–529.CrossRefGoogle ScholarPubMed
Bartha, I., Simon, P. & Muller, V. (2008). Has HIV evolved to induce immune pathogenesis?Trends in Immunology, 29, 322–328.CrossRefGoogle ScholarPubMed
Bascompte, J., Jordano, P., Melián, C. J. & Olesen, J. M. (2003). The nested assembly of plant–animal mutualistic networks. Proceedings of the National Academy of Sciences of the United States of America, 100, 9383–9387.CrossRefGoogle ScholarPubMed
Bascompte, J. & Melian, C. J. (2005). Simple trophic modules for complex food webs. Ecology, 86, 2868–2873.CrossRefGoogle Scholar
Bataille, A., Cunningham, A. A., Cedeno, V., Cruz, M., Eastwood, G., Fonseca, D. M., Causton, C. E., Azuero, R., Loayza, J., Martinez, J. D. C. & Goodman, S. J. (2009a). Evidence for regular ongoing introductions of mosquito disease vectors into the Galapagos Islands. Proceedings of the Royal Society of London Series B – Biological Sciences, 276, 3769–3775.CrossRefGoogle ScholarPubMed
Bataille, A., Cunningham, A. A., Cedeno, V., Patino, L., Constantinou, A., Kramer, L. D. & Goodman, S. J. (2009b). Natural colonization and adaptation of a mosquito species in Galapagos and its implications for disease threats to endemic wildlife. Proceedings of the National Academy of Sciences of the United States of America, 106, 10230–10235.CrossRefGoogle ScholarPubMed
Bauer, A., Trouve, S., Gregoire, A., Bollache, L. & Cezilly, F. (2000). Differential influence of Pomphorhynchus laevis (Acanthocephala) on the behaviour of native and invader gammarid species. International Journal for Parasitology, 30, 1453–1457.CrossRefGoogle ScholarPubMed
Bauer, L. S., Miller, D. L., Maddox, J. V. & McManus, M. L. (1998). Interactions between a Nosema sp. (Microspora: Nosematidae) and nuclear polyhedrosis virus infecting the gypsy moth, Lymantria dispar (Lepidoptera: Lymantriidae). Journal of Invertebrate Pathology, 72, 147–153.CrossRefGoogle Scholar
Beckstead, J. & Parker, I. M. (2003). Invasiveness of Ammophila arenaria: release from soil-borne pathogens?Ecology, 84, 2824–2831.CrossRefGoogle Scholar
Bedhomme, S., Agnew, P., Vital, Y., Sidobre, C. & Michalakis, Y. (2005). Prevalence-dependent costs of parasite virulence. PLoS Biology, 3, 1403–1408.CrossRefGoogle ScholarPubMed
Begon, M. & Bowers, R. G. (1994). Host–host–pathogen models and microbial pest-control: the effect of host self-regulation. Journal of Theoretical Biology, 169, 275–287.CrossRefGoogle ScholarPubMed
Begon, M. & Bowers, R. G. (1995). Beyond host–pathogen dynamics. In Grenfell, B. T. & Dobson, A. P. (eds), Ecology of Infectious Diseases in Natural Populations. Cambridge: Cambridge University Press, pp. 478–509.CrossRefGoogle Scholar
Begon, M., Bowers, R. G., Kadianakis, N. & Hodgkinson, D. E. (1992). Disease and community structure: the importance of host self-regulation in a host–host–pathogen model. American Naturalist, 139, 1131–1150.CrossRefGoogle Scholar
Begon, M., Sait, S. M. & Thompson, D. J. (1999). Host–parasite–parasitoid systems. In Hawkins, B. A. & Cornell, H. V. (eds), Theoretical Approaches to Biological Control. Cambridge: Cambridge University Press, pp. 327–348.CrossRefGoogle Scholar
Bell, S. S., White, A., Sherratt, J. A. & Boots, M. (2009). Invading with biological weapons: the role of shared disease in ecological invasion. Theoretical Ecology, 2, 53–66.CrossRefGoogle Scholar
Belliure, B., Janssen, A., Maris, P. C., Peters, D. & Sabelis, M. W. (2005). Herbivore arthropods benefit from vectoring plant viruses. Ecology Letters, 8, 70–79.CrossRefGoogle Scholar
Belliure, B., Janssen, A. & Sabelis, M. W. (2008). Herbivore benefits from vectoring plant virus through reduction of period of vulnerability to predation. Oecologia, 156, 797–806.CrossRefGoogle ScholarPubMed
Ben-Ami, F., Mouton, L. & Ebert, D. (2008). The effects of multiple infections on the expression and evolution of virulence in a Daphnia–endoparasite system. Evolution, 62, 1700–1711.CrossRefGoogle Scholar
Bennetts, R. E., White, G. C., Hawksworth, F. G. & Severs, S. E. (1996). The influence of dwarf mistletoe on bird communities in Colorado ponderosa pine forests. Ecological Applications, 6, 899–909.CrossRefGoogle Scholar
Benson, J., Driesche, R. G., Pasquale, A. & Elkinton, J. (2003). Introduced braconid parasitoids and range reduction of a native butterfly in New England. Biological Control, 28, 197–213.CrossRefGoogle Scholar
Berger, L., Speare, R., Daszak, P., Green, D. E., Cunningham, A. A., Goggin, C. L., Slocombe, R., Ragan, M. A., Hyatt, A. D., McDonald, K. R., Hines, H. B., Lips, K. R., Marantelli, G. & Parkes, H. (1998). Chytridiomycosis causes amphibian mortality associated with population declines in the rain forests of Australia and Central America. Proceedings of the National Academy of Sciences of the United States of America, 95, 9031–9036.CrossRefGoogle ScholarPubMed
Bernot, R. J. & Lamberti, G. A. (2008). Indirect effects of a parasite on a benthic community: an experiment with trematodes, snails and periphyton. Freshwater Biology, 53, 322–329.Google Scholar
Berthe, F. C. J., Roux, F., Adlard, R. D. & Figueras, A. (2004). Marteiliosis in molluscs: a review. Aquatic Living Resources, 17, 433–448.CrossRefGoogle Scholar
Bertolino, S., Lurz, P. W. W., Sanderson, R. & Rushton, S. P. (2008). Predicting the spread of the American grey squirrel (Sciurus carolinensis) in Europe: a call for a co-ordinated European approach. Biological Conservation, 141, 2564–2575.CrossRefGoogle Scholar
Bever, J. D., Westover, K. M. & Antonovics, J. (1997). Incorporating the soil community into plant population dynamics: the utility of the feedback approach. Journal of Ecology, 85, 561–573.CrossRefGoogle Scholar
Bezemer, T. M. & Dam, N. M. (2005). Linking aboveground and belowground interactions via induced plant defenses. Trends in Ecology & Evolution, 20, 617–624.CrossRefGoogle ScholarPubMed
Bilu, E. & Coll, M. (2007). The importance of intraguild interactions to the combined effect of a parasitoid and a predator on aphid population suppression. Biocontrol, 52, 753–763.CrossRefGoogle Scholar
Blackmore, M. S., Scoles, G. A. & Craig, G. B. (1995). Parasitism of Aedes aegypti and Ae albopictus (Diptera, Culicidae) by Ascogregarina spp. (Apicomplexa, Lecudinidae) in Florida. Journal of Medical Entomology, 32, 847–852.CrossRefGoogle Scholar
Blossey, B. & Notzold, R. (1995). Evolution of increased competitive ability in invasive nonindigenous plants: a hypothesis. Journal of Ecology, 83, 887–889.CrossRefGoogle Scholar
Bolker, B., Holyoak, M., Krivan, V., Rowe, L. & Schmitz, O. (2003). Connecting theoretical and empirical studies of trait-mediated interactions. Ecology, 84, 1101–1114.CrossRefGoogle Scholar
Bolzoni, L., Leo, G. A., Gatto, M. & Dobson, A. P. (2008). Body-size scaling in an SEI model of wildlife diseases. Theoretical Population Biology, 73, 374–382.CrossRefGoogle Scholar
Bonsall, M. B. (2004). The impact of diseases and pathogens on insect population dynamics. Physiological Entomology, 29, 223–236.CrossRefGoogle Scholar
Bonsall, M. B. & Hassell, M. P. (1997). Apparent competition structures ecological assemblages. Nature, 388, 371–373.CrossRefGoogle Scholar
Bonsall, M. B. & Hassell, M. P. (2000). The effects of metapopulation structure on indirect interactions in host–parasitoid assemblages. Proceedings of the Royal Society of London Series B – Biological Sciences, 267, 2207–2212.CrossRefGoogle ScholarPubMed
Bordes, F. & Morand, S. (2008). Helminth species diversity of mammals: parasite species richness is a host species attribute. Parasitology, 135, 1701–1705.CrossRefGoogle ScholarPubMed
Borer, E. T., Briggs, C. J. & Holt, R. D. (2007a). Predators, parasitoids, and pathogens: a cross-cutting examination of intraguild predation theory. Ecology, 88, 2681–2688.CrossRefGoogle ScholarPubMed
Borer, E. T., Briggs, C. J., Murdoch, W. W. & Swarbrick, S. L. (2003). Testing intraguild predation theory in a field system: does numerical dominance shift along a gradient of productivity?Ecology Letters, 6, 929–935.CrossRefGoogle Scholar
Borer, E. T., Hosseini, P. R., Seabloom, E. W. & Dobson, A. P. (2007b). Pathogen-induced reversal of native dominance in a grassland community. Proceedings of the National Academy of Sciences of the United States of America, 104, 5473–5478.CrossRefGoogle Scholar
Bosch, J. & Rincon, P. A. (2008). Chytridiomycosis-mediated expansion of Bufo bufo in a montane area of Central Spain: an indirect effect of the disease. Diversity and Distributions, 14, 637–643.CrossRefGoogle Scholar
Bostock, R. M. (2005). Signal crosstalk and induced resistance: straddling the line between cost and benefit. Annual Review of Phytopathology, 43, 545–580.CrossRefGoogle ScholarPubMed
Bouwma, A. M., Ahrens, M. E., DeHeer, C. J. & DeWayne Shoemaker, D. (2006). Distribution and prevalence of Wolbachia in introduced populations of the fire ant Solenopsis invicta. Insect Molecular Biology, 15, 89–93.CrossRefGoogle ScholarPubMed
Bowers, R. G. & Begon, M. (1991). A host–host–pathogen model with free-living infective stages, applicable to microbial pest-control. Journal of Theoretical Biology, 148, 305–329.CrossRefGoogle ScholarPubMed
Bowers, R. G. & Turner, J. (1997). Community structure and the interplay between interspecific infection and competition. Journal of Theoretical Biology, 187, 95–109.CrossRefGoogle ScholarPubMed
Bradley, D. J., Gilbert, G. S. & Martiny, J. B. H. (2008). Pathogens promote plant diversity through a compensatory response. Ecology Letters, 11, 461–469.CrossRefGoogle ScholarPubMed
Brasier, C. (2000). Plant pathology: the rise of the hybrid fungi. Nature, 405, 134–135.CrossRefGoogle Scholar
Briggs, C. J. (1993). Competition among parasitoid species on a stage-structured host and its effect on host suppression. American Naturalist, 141, 372–397.CrossRefGoogle Scholar
Briggs, C. J., Knapp, R. A. & Vredenburg, V. T. (2010). Enzootic and epizootic dynamics of the chytrid fungal pathogen of amphibians. Proceedings of the National Academy of Sciences of the United States of America, 107, 9695–9700.CrossRefGoogle ScholarPubMed
Brisson, D., Dykhuizen, D. E. & Ostfeld, R. S. (2008). Conspicuous impacts of inconspicuous hosts on the Lyme disease epidemic. Proceedings of the Royal Society of London Series B – Biological Sciences, 275, 227–235.CrossRefGoogle ScholarPubMed
Brodeur, J. & Rosenheim, J. A. (2000). Intraguild interactions in aphid parasitoids. Entomologia experimentalis et Applicata, 97, 93–108.CrossRefGoogle Scholar
Brook, B. W., Sodhi, N. S. & Bradshaw, C. J. A. (2008). Synergies among extinction drivers under global change. Trends in Ecology & Evolution, 23, 453–460.CrossRefGoogle ScholarPubMed
Brooks, D. R. & Hoberg, E. P. (2007). How will global climate change affect parasite–host assemblages?Trends in Parasitology, 23, 571–574.CrossRefGoogle ScholarPubMed
Brown, D. H. & Hastings, A. (2003). Resistance may be futile: dispersal scales and selection for disease resistance in competing plants. Journal of Theoretical Biology, 222, 373–388.CrossRefGoogle ScholarPubMed
Brown, S. P., Inglis, R. F. & Taddei, F. (2009). Evolutionary ecology of microbial wars: within-host competition and (incidental) virulence. Evolutionary Applications, 2, 32–39.CrossRefGoogle ScholarPubMed
Bruce, T. J. & Pickett, J. A. (2007). Plant defence signalling induced by biotic attacks. Current Opinion in Plant Biology, 10, 387–392.CrossRefGoogle ScholarPubMed
Bull, E. L., Heater, T. W. & Youngblood, A. (2004). Arboreal squirrel response to silvicultural treatments for dwarf mistletoe control in northeastern Oregon. Western Journal of Applied Forestry, 19, 133–141.Google Scholar
Bull, J. J. & Ebert, D. (2008). Invasion thresholds and the evolution of nonequilibrium virulence. Evolutionary Applications, 1, 172–182.CrossRefGoogle ScholarPubMed
Bullock, J. M. & Pywell, R. F. (2005). Rhinanthus: a tool for restoring diverse grassland?Folia Geobotanica, 40, 273–288.CrossRefGoogle Scholar
Burden, J. P., Nixon, C. P., Hodgkinson, A. E., Possee, R. D., Sait, S. M., King, L. A. & Hails, R. S. (2003). Covert infections as a mechanism for long-term persistence of baculoviruses. Ecology Letters, 6, 524–531.CrossRefGoogle Scholar
Bush, S. E. & Malenke, J. R. (2008). Host defence mediates interspecific competition in ectoparasites. Journal of Animal Ecology, 77, 558–564.CrossRefGoogle ScholarPubMed
Byers, J. E. (2009). Including parasites in food webs. Trends in Parasitology, 25, 55–57.CrossRefGoogle ScholarPubMed
Byrne, C. J., Holland, C. V., Kennedy, C. R. & Poole, W. R. (2003). Interspecific interactions between Acanthocephala in the intestine of brown trout: are they more frequent in Ireland?Parasitology, 127, 399–409.CrossRefGoogle ScholarPubMed
Caesar, A. J. & Caesar-Ton That, T. (2008). Rhizosphere bacterial communities associated with insect root herbivory of an invasive plant, Euphorbia esula/virgata. In Julien, M. H., Sforza, R., Bon, M. C., Evans, H. C., Hatcher, P. E., Hinz, H. L. & Rector, B. G. (eds), Proceedings of the XII Symposium on Biological Control of Weeds. Wallingford: CAB International, pp. 13–19.Google Scholar
Calisher, C. H., Childs, J. E., Field, H. E., Holmes, K. V. & Schountz, T. (2006). Bats: important reservoir hosts of emerging viruses. Clinical Microbiology Reviews, 19, 531–545.CrossRefGoogle ScholarPubMed
Callaway, R. M. & Pennings, S. C. (1998). Impact of a parasitic plant on the zonation of two salt marsh perennials. Oecologia, 114, 100–105.CrossRefGoogle ScholarPubMed
Callaway, R. M. & Ridenour, W. M. (2004). Novel weapons: invasive success and the evolution of increased competitive ability. Frontiers in Ecology and the Environment, 2, 436–443.CrossRefGoogle Scholar
Callaway, R. M., Thelen, G. C., Rodriguez, A. & Holben, W. E. (2004). Soil biota and exotic plant invasion. Nature, 427, 731–733.CrossRefGoogle ScholarPubMed
Cameron, T. C., Wearing, H. J., Rohani, P. & Sait, S. M. (2005). A koinobiont parasitoid mediates competition and generates additive mortality in healthy host populations. Oikos, 110, 620–628.CrossRefGoogle Scholar
Cardoza, Y. J., Lait, C. G., Schmelz, E. A., Huang, J. & Tumlinson, J. H. (2003). Fungus-induced biochemical changes in peanut plants and their effect on development of beet armyworm, Spodoptera exigua Hubner (Lepidoptera: Noctuidae) larvae. Environmental Entomology, 32, 220–228.CrossRefGoogle Scholar
Carroll, B., Russell, P., Gurnell, J., Nettleton, P. & Sainsbury, A. W. (2009). Epidemics of squirrelpox virus disease in red squirrels (Sciurus vulgaris): temporal and serological findings. Epidemiology and Infection, 137, 257–265.CrossRefGoogle ScholarPubMed
Chadha, M. S., Comer, J. A., Lowe, L., Rota, P. A., Rollin, P. E., Bellini, W. J., Ksiazek, T. G. & Mishra, A. C. (2006). Nipah virus-associated encephalitis outbreak, Siliguri, India. Emerging Infectious Diseases, 12, 235–240.CrossRefGoogle ScholarPubMed
Chapuis, J. L., Bousses, P. & Barnaud, G. (1994). Alien mammals, impact and management in the French sub-Antarctic islands. Biological Conservation, 67, 97–104.CrossRefGoogle Scholar
Chattopadhyay, J. & Bairagi, N. (2001). Pelicans at risk in Salton Sea: an eco-epidemiological model. Ecological Modelling, 136, 103–112.CrossRefGoogle Scholar
Chattopadhyay, J., Srinivasu, P. D. N. & Bairagi, N. (2003). Pelicans at risk in Salton Sea: an eco-epidemiological model-II. Ecological Modelling, 167, 199–211.CrossRefGoogle Scholar
Chen, H. W., Liu, W. C., Davis, A. J., Jordan, F., Hwang, M. J. & Shao, K. T. (2008). Network position of hosts in food webs and their parasite diversity. Oikos, 117, 1847–1855.CrossRefGoogle Scholar
Chilcutt, C. F. & Tabashnik, B. E. (1997). Host-mediated competition between the pathogen Bacillus thuringiensis and the parasitoid Cotesia plutellae of the diamondback moth (Lepidoptera: Plutellidae). Environmental Entomology, 26, 38–45.CrossRefGoogle Scholar
Childs, J. E., Mackenzie, J. S. & Richt, J. A. (eds) (2007a). Wildlife and Emerging Zoonotic Diseases: The Biology, Circumstances and Consequence of Cross-species Transmission. Berlin: Springer.CrossRef
Childs, J. E., Richt, J. A. & Mackenzie, J. S. (2007b). Introduction: conceptualizing and partitioning the emergence process of zoonotic viruses from wildlife to humans. In Childs, J. E., Mackenzie, J. S. & Richt, J. A. (eds) Wildlife and Emerging Zoonotic Diseases: The Biology, Circumstances and Consequences of Cross-species Transmission. Berlin: Springer, pp. 1–31.Google Scholar
Choisy, M. & Rohani, P. (2006). Harvesting can increase severity of wildlife disease epidemics. Proceedings of the Royal Society of London Series B – Biological Sciences, 273, 2025–2034.CrossRefGoogle ScholarPubMed
Chomel, B. B. (2008). Control and prevention of emerging parasitic zoonoses. International Journal for Parasitology, 38, 1211–1217.CrossRefGoogle ScholarPubMed
Chong, J. H. & Oetting, R. D. (2007). Intraguild predation and interference by the mealybug predator Cryptolalemus montrouzieri on the parasitoid Leptomastix dactylopii. Biocontrol Science and Technology, 17, 933–944.CrossRefGoogle Scholar
Choo, K., Williams, P. D. & Day, T. (2003). Host mortality, predation and the evolution of parasite virulence. Ecology Letters, 6, 310–315.CrossRefGoogle Scholar
Chua, K. B., Bellini, W. J., Rota, P. A., Harcourt, B. H., Tamin, A., Lam, S. K., Ksiazek, T. G., Rollin, P. E., Zaki, S. R., Shieh, W. J., Goldsmith, C. S., Gubler, D. J., Roehrig, J. T., Eaton, B., Gould, A. R., Olson, J., Field, H., Daniels, P., Ling, A. E., Peters, C. J., Anderson, L. J. & Mahy, B. W. J. (2000). Nipah virus: a recently emergent deadly paramyxovirus. Science, 288, 1432–1435.CrossRefGoogle ScholarPubMed
Clay, K. (1996). Interactions among fungal endophytes, grasses and herbivores. Researches on Population Ecology, 38, 191–201.CrossRefGoogle Scholar
Clay, K. & Holah, J. (1999). Fungal endophyte symbiosis and plant diversity in successional fields. Science, 285, 1742–1744.CrossRefGoogle ScholarPubMed
Clay, K., Holah, J. & Rudgers, J. A. (2005). Herbivores cause a rapid increase in hereditary symbiosis and alter plant community composition. Proceedings of the National Academy of Sciences of the United States of America, 102, 12465–12470.CrossRefGoogle Scholar
Clay, K., Marks, S. & Cheplick, G. P. (1993). Effects of insect herbivory and fungal endophyte infection on competitive interactions among grasses. Ecology, 74, 1767–1777.CrossRefGoogle Scholar
Clay, K. & Schardl, C. (2002). Evolutionary origins and ecological consequences of endophyte symbiosis with grasses. American Naturalist, 160, S99–S127.CrossRefGoogle ScholarPubMed
Cleaveland, S., Haydon, D. T. & Taylor, L. (2007). Overviews of pathogen emergence: which pathogens emerge, when and why? In Childs, J. E., Mackenzie, J. S. & Richt, J. A. (eds) Wildlife and Emerging Zoonotic Diseases: The Biology, Circumstances and Consequences of Cross-species Transmission. Berlin: Springer, pp. 85–111.Google Scholar
Cleaveland, S., Kaare, M., Tiringa, P., Mlengeya, T. & Barrat, J. (2003). A dog rabies vaccination campaign in rural Africa: impact on the incidence of dog rabies and human dog-bite injuries. Vaccine, 21, 1965–1973.CrossRefGoogle ScholarPubMed
Cohen, J. E., Jonsson, T. & Carpenter, S. R. (2003). Ecological community description using the food web, species abundance, and body size. Proceedings of the National Academy of Sciences of the United States of America, 100, 1781–1786.CrossRefGoogle ScholarPubMed
Colautti, R. I., Muirhead, J. R., Biswas, R. N. & MacIsaac, H. J. (2005). Realized vs apparent reduction in enemies of the European starling. Biological Invasions, 7, 723–732.CrossRefGoogle 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, 721–733.CrossRefGoogle Scholar
Collinge, S. K., Johnson, W. C., Ray, C., Matchett, R., Grensten, J., Cully, J. F., Gage, K. L., Kosoy, M. Y., Loye, J. E. & Martin, A. P. (2005). Testing the generality of a trophic-cascade model for plague. Ecohealth, 2, 102–112.CrossRefGoogle Scholar
Collinge, S. K. & Ray, C. (2006). Disease Ecology: Community Structure and Pathogen Dynamics. Oxford: Oxford University Press.CrossRefGoogle Scholar
Connell, J. H. (1971). On the role of natural enemies in preventing competitive exclusion in some marine animals and in rain forest trees. In Boer, P. J. & Gradwell, G. R. (eds), Dynamics of Populations. Wageningen: Centre for Agricultural Publishing and Documentation, pp. 298–312.Google Scholar
Cope, D. R., Iason, G. R. & Gordon, I. J. (2004). Disease reservoirs in complex systems: a comment on recent work by Laurenson et al. Journal of Animal Ecology, 73, 807–810.CrossRefGoogle Scholar
Cory, J. S. & Hoover, K. (2006). Plant-mediated effects in insect–pathogen interactions. Trends in Ecology & Evolution, 21, 278–286.CrossRefGoogle ScholarPubMed
Cory, J. S. & Myers, J. H. (2000). Direct and indirect ecological effects of biological control. Trends in Ecology & Evolution, 15, 137–139.CrossRefGoogle Scholar
Cossentine, J. E. & Lewis, L. C. (1988). Impact of Nosema pyrausta, Nosema sp. and a nuclear polyhedrosis virus on Lydella thompsoni within infected Ostrinia nubilalis hosts. Journal of Invertebrate Pathology, 51, 126–132.CrossRefGoogle Scholar
Cottrell, T. E. & Yeargan, K. V. (1998). Intraguild predation between an introduced lady beetle, Harmonia axyridis (Coleoptera: Coccinellidae), and a native lady beetle, Coleomegilla maculata (Coleoptera: Coccinellidae). Journal of the Kansas Entomological Society, 71, 159–163.Google Scholar
Courchamp, F. & Sugihara, G. (1999). Modeling the biological control of an alien predator to protect island species from extinction. Ecological Applications, 9, 112–123.CrossRefGoogle Scholar
Craig, B. H., Pilkington, J. G., Kruuk, L. E. B. & Pemberton, J. M. (2007). Epidemiology of parasitic protozoan infections in Soay sheep (Ovis aries L.) on St Kilda. Parasitology, 134, 9–21.CrossRefGoogle ScholarPubMed
Cremer, S., Ugelvig, L. V., Drijfhout, F. P., Schlick-Steiner, B. C., Steiner, F. M., Seifert, B., Hughes, D. P., Schulz, A., Petersen, K. S., Konrad, H., Stauffer, C., Kiran, K., Espadaler, X., d'Ettorre, P., Aktac, N., Eilenberg, J., Jones, G. R., Nash, D. R., Pedersen, J. S. & Boomsma, J. J. (2008). The evolution of invasiveness in garden ants. PLoS One, 3, e3838.CrossRefGoogle ScholarPubMed
Cronin, J. T. (2007). Shared parasitoids in a metacommunity: indirect interactions inhibit herbivore membership in local communities. Ecology, 88, 2977–2990.CrossRefGoogle Scholar
Daszak, P., Cunningham, A. A. & Hyatt, A. D. (2000). Wildlife ecology: emerging infectious diseases of wildlife – threats to biodiversity and human health. Science, 287, 443–449.CrossRefGoogle ScholarPubMed
Daszak, P., Epstein, J. H., Kilpatrick, A. M., Aguirre, A. A., Karesh, W. B. & Cunningham, A. A. (2007). Collaborative research approaches to the role of wildlife in zoonotic disease emergence. In Childs, J. E., Mackenzie, J. S. & Richt, J. A. (eds) Wildlife and Emerging Zoonotic Diseases: The Biology, Circumstances and Consequences of Cross-species Transmission. Berlin: Springer, pp. 463–475.Google ScholarPubMed
Davies, C. M., Fairbrother, E. & Webster, J. P. (2002). Mixed strain schistosome infections of snails and the evolution of parasite virulence. Parasitology, 124, 31–38.CrossRefGoogle ScholarPubMed
Davies, D. M., Graves, J. D., Elias, C. O. & Williams, P. J. (1997). The impact of Rhinanthus spp. on sward productivity and composition: implications for the restoration of species-rich grasslands. Biological Conservation, 82, 87–93.CrossRefGoogle Scholar
Dawkins, R. (1982). The Extended Phenotype. Oxford: Oxford University Press.Google Scholar
Day, J. F. & Edman, J. D. (1983). Malaria renders mice susceptible to mosquito feeding when gametocytes are most infective. Journal of Parasitology, 69, 163–170.CrossRefGoogle ScholarPubMed
Day, T. & Gandon, S. (2007). Applying population–genetic models in theoretical evolutionary epidemiology. Ecology Letters, 10, 876–888.CrossRefGoogle ScholarPubMed
Castro, F. & Bolker, B. (2005). Mechanisms of disease-induced extinction. Ecology Letters, 8, 117–126.CrossRefGoogle Scholar
Moraes, C. M., Mescher, M. C. & Tumlinson, J. H. (2001). Caterpillar-induced nocturnal plant volatiles repel conspecific females. Nature, 410, 577–580.CrossRefGoogle ScholarPubMed
Roode, J. C., Helinski, M. E. H., Anwar, M. A. & Read, A. F. (2005a). Dynamics of multiple infection and within-host competition in genetically diverse malaria infections. American Naturalist, 166, 531–542.Google ScholarPubMed
Roode, J. C., Pansini, R., Cheesman, S. J., Helinski, M. E. H., Huijben, S., Wargo, A. R., Bell, A. S., Chan, B. H. K., Walliker, D. & Read, A. F. (2005b). Virulence and competitive ability in genetically diverse malaria infections. Proceedings of the National Academy of Sciences of the United States of America, 102, 7624–7628.CrossRefGoogle ScholarPubMed
Vos, M., Oosten, V. R., Poecke, R. M. P., Pelt, J. A., Pozo, M. J., Mueller, M. J., Buchala, A. J., Metraux, J. P., Loon, L. C., Dicke, M. & Pieterse, C. M. J. (2005). Signal signature and transcriptome changes of Arabidopsis during pathogen and insect attack. Molecular Plant–Microbe Interactions, 18, 923–937.CrossRefGoogle ScholarPubMed
Dennehy, J. J., Friedenberg, N. A., Holt, R. D. & Turner, P. E. (2006). Viral ecology and the maintenance of novel host use. American Naturalist, 167, 429–439.CrossRefGoogle ScholarPubMed
Diamond, J. M. (1997). Guns, Germs and Steel: The Fates of Human Societies. New York: W.W. Norton.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, 756–767.CrossRefGoogle Scholar
Dick, J. T. A. (2008). Role of behaviour in biological invasions and species distributions: lessons from interactions between the invasive Gammarus pulex and the native G. duebeni (Crustacea: Amphipoda). Contributions to Zoology, 77, 91–98.Google Scholar
Dick, J. T. A., Armstrong, M., Clarke, H. C., Farnesworth, K. D., Hatcher, M. J., Ennis, , , M., Kelly, A. & Dunn, A. M. (2010). Parasitism may enhance rather than reduce the predatory impact of an invader. Biology Letters, 6, 636–638.CrossRefGoogle ScholarPubMed
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, 79–88.CrossRefGoogle 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, 375–383.CrossRefGoogle Scholar
Dicke, M. & Baldwin, I. T. (2010). The evolutionary context for herbivore-induced plant volatiles: beyond the ‘cry for help’. Trends in Plant Science, 15, 167–175.CrossRefGoogle ScholarPubMed
Dicke, M. & Dijkman, H. (1992). Induced defense in detached uninfested plant leaves: effects on behavior of herbivores and their predators. Oecologia, 91, 554–560.CrossRefGoogle Scholar
Diehl, S. & Feissel, M. (2000). Effects of enrichment on three-level food chains with omnivory. American Naturalist, 155, 200–218.Google ScholarPubMed
Diekmann, O., Heesterbeek, J. A. P. & Metz, J. A. J. (1990). On the definition and the computation of the basic reproduction ratio R0 in models for infectious diseases in heterogeneous populations. Journal of Mathematical Biology, 28, 365–382.CrossRefGoogle ScholarPubMed
Dillon, R. J., Vennard, C. T., Buckling, A. & Charnley, A. K. (2005). Diversity of locust gut bacteria protects against pathogen invasion. Ecology Letters, 8, 1291–1298.CrossRefGoogle Scholar
Dobson, A. (2004). Population dynamics of pathogens with multiple host species. American Naturalist, 164, S64–S78.CrossRefGoogle ScholarPubMed
Dobson, A. (2009). Climate variability, global change, immunity, and the dynamics of infectious diseases. Ecology, 90, 920–927.CrossRefGoogle ScholarPubMed
Dobson, A., Cattadori, I., Holt, R. D., Ostfeld, R. S., Keesing, F., Krichbaum, K., Rohr, J. R., Perkins, S. E. & Hudson, P. J. (2006). Sacred cows and sympathetic squirrels: the importance of biological diversity to human health. PLoS Med, 3, e231.CrossRefGoogle ScholarPubMed
Dobson, A. & Crawley, W. (1994). Pathogens and the structure of plant communities. Trends in Ecology & Evolution, 9, 393–398.CrossRefGoogle ScholarPubMed
Dobson, A. & Foufopoulos, J. (2001). Emerging infectious pathogens of wildlife. Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 356, 1001–1012.CrossRefGoogle ScholarPubMed
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, 11482–11489.CrossRefGoogle ScholarPubMed
Dobson, A. P. (1985). The population dynamics of competition between parasites. Parasitology, 91, 317–347.CrossRefGoogle ScholarPubMed
Dobson, A. P. (1988). The population biology of parasite-induced changes in host behavior. Quarterly Review of Biology, 63, 139–165.CrossRefGoogle ScholarPubMed
Dobson, A. P. (1995). The ecology and epidemiology of rinderpest virus in Serengeti and Ngorongoro crater conservation area. In Sinclair, A. R. E. & Arcese, P. (eds), Serengeti II: Research, Management and Conservation of an Ecosystem. Chicago: University of Chicago Press, pp. 485–505.Google Scholar
Dobson, A. P. (2005). What links bats to emerging infectious diseases?Science, 310, 628–629.CrossRefGoogle ScholarPubMed
Dobson, A. P. & Carper, E. R. (1996). Infectious diseases and human population history: throughout history the establishment of disease has been a side effect of the growth of civilization. Bioscience, 46, 115–126.CrossRefGoogle Scholar
Dobson, A. P. & Hudson, P. J. (1986). Parasites, disease and the structure of ecological communities. Trends in Ecology & Evolution, 1, 11–15.CrossRefGoogle ScholarPubMed
Dobson, A. P. & Hudson, P. J. (1992). Regulation and stability of a free-living host–parasite system: Trichostrongylus tenuis in red grouse – 2. Population models. Journal of Animal Ecology, 61, 487–498.CrossRefGoogle Scholar
Dobson, A. P. & Keymer, A. E. (1985). Life history models. In Crompton, D. W. T. & Nickol, B. B. (eds), Biology of the Acanthocephala. Cambridge: Cambridge University Press, pp. 347–384.Google Scholar
Dobson, A. P. & May, R. M. (1987). The effects of parasites on fish populations: theoretical aspects. International Journal for Parasitology, 17, 363–370.CrossRefGoogle ScholarPubMed
Drake, J. M. (2003). The paradox of the parasites: implications for biological invasion. Proceedings of the Royal Society of London Series B – Biological Sciences, 270, S133–S135.CrossRefGoogle ScholarPubMed
Duffy, M. A. & Hall, S. R. (2008). Selective predation and rapid evolution can jointly dampen effects of virulent parasites on Daphnia populations. American Naturalist, 171, 499–510.CrossRefGoogle ScholarPubMed
Duffy, M. A., Hall, S. R., Tessier, A. J. & Huebner, M. (2005). Selective predators and their parasitized prey: are epidemics in zooplankton under top-down control?Limnology and Oceanography, 50, 412–420.CrossRefGoogle Scholar
Duffy, M. A. & Sivars-Becker, L. (2007). Rapid evolution and ecological host–parasite dynamics. Ecology Letters, 10, 44–53.CrossRefGoogle ScholarPubMed
Dumont, A. & Crete, M. (1996). The meningeal worm, Parelaphostrongylus tenuis, a marginal limiting factor for moose, Alces alces, in Southern Quebec. Canadian Field-Naturalist, 110, 413–418.Google Scholar
Dunn, A. M. (2009). Parasites and biological invasions. Advances in Parasitology, 68, 161–184.CrossRefGoogle ScholarPubMed
Dunn, A. M. & Dick, J. T. A. (1998). Parasitism and epibiosis in native and non-native gammarids in freshwater in Ireland. Ecography, 21, 593–598.CrossRefGoogle Scholar
Dunn, A. M., Terry, R. S. & Smith, J. E. (2001). Transovarial transmission in the microsporidia. Advances in Parasitology, 48, 57–100.CrossRefGoogle ScholarPubMed
Dunn, J. C., McClymont, H. E., Christmas, M. & Dunn, A. M. (2009). Competition and parasitism in the native white clawed crayfish Austropotamobius pallipes and the invasive signal crayfish Pacifastacus leniusculus in the UK. Biological Invasions, 11, 315–324.CrossRefGoogle Scholar
Dunne, J. A., Williams, R. J. & Martinez, N. D. (2002a). Food-web structure and network theory: the role of connectance and size. Proceedings of the National Academy of Sciences of the United States of America, 99, 12917–12922.CrossRefGoogle ScholarPubMed
Dunne, J. A., Williams, R. J. & Martinez, N. D. (2002b). Network structure and biodiversity loss in food webs: robustness increases with connectance. Ecology Letters, 5, 558–567.CrossRefGoogle Scholar
Duron, O., Bouchon, D., Boutin, S., Bellamy, L., Zhou, L. Q., Engelstadter, J. & Hurst, G. D. (2008). The diversity of reproductive parasites among arthropods: Wolbachia do not walk alone. BMC Biology, 6, e27.CrossRefGoogle Scholar
Dwosh, H. A., Hong, H. H. L., Austgarden, D., Herman, S. & Schabas, R. (2003). Identification and containment of an outbreak of SARS in a community hospital. Canadian Medical Association Journal, 168, 1415–1420.Google Scholar
Dwyer, G., Dushoff, J. & Yee, S. H. (2004). The combined effects of pathogens and predators on insect outbreaks. Nature, 430, 341–345.CrossRefGoogle ScholarPubMed
Dybdahl, M. F. & Lively, C. M. (1995). Host–parasite interactions: infection of common clones in natural populations of a fresh-water snail (Potamopyrgus antipodarum). Proceedings of the Royal Society of London Series B – Biological Sciences, 260, 99–103.CrossRefGoogle Scholar
Dye, C. & Gay, N. (2003). Modeling the SARS epidemic. Science, 300, 1884–1885.CrossRefGoogle ScholarPubMed
Edeline, E., Ben Ari, T., Vollestad, L. A., Winfield, I. J., Fletcher, J. M., Ben James, J. & Stenseth, N. C. (2008). Antagonistic selection from predators and pathogens alters food-web structure. Proceedings of the National Academy of Sciences of the United States of America, 105, 19792–19796.CrossRefGoogle ScholarPubMed
Elton, C. S. (1958). The Ecology of Invasions by Animals and Plants. London: Methuen.CrossRefGoogle Scholar
Engelberth, J., Alborn, H. T., Schmelz, E. A. & Tumlinson, J. H. (2004). Airborne signals prime plants against insect herbivore attack. Proceedings of the National Academy of Sciences of the United States of America, 101, 1781–1785.CrossRefGoogle ScholarPubMed
Enscore, R. E., Biggerstaff, B. J., Brown, T. L., Fulgham, R. F., Reynolds, P. J., Engelthaler, D. M., Levy, C. E., Parmenter, R. R., Montenieri, J. A., Cheek, J. E., Grinnell, R. K., Ettestad, P. J. & Gage, K. L. (2002). Modeling relationships between climate and the frequency of human plague cases in the southwestern United States, 1960–1997. American Journal of Tropical Medicine and Hygiene, 66, 186–196.CrossRefGoogle ScholarPubMed
Epstein, P. R. (2000). Is global warming harmful to health?Scientific American, 283 (2), 50–57.CrossRefGoogle Scholar
Epstein, P. R., Diaz, H. F., Elias, S., Grabherr, G., Graham, N. E., Martens, W. J. M., Mosley-Thompson, E. & Susskind, J. (1998). Biological and physical signs of climate change: focus on mosquito-borne diseases. Bulletin of the American Meteorological Society, 79, 409–417.2.0.CO;2>CrossRefGoogle Scholar
Evans, H. C. (2008). The endophyte–enemy release hypothesis: implications for classical biological control and plant invasions. In Julien, M. H., Sforza, R., Bon, M. C., Evans, H. C., Hatcher, P. E., Hinz, H. L. & Rector, B. G. (eds), Proceedings of the XII Symposium on Biological Control of Weeds. Wallingford: CAB International, pp. 20–25.Google Scholar
Ezenwa, V. O., Godsey, M. S., King, R. J. & Guptill, S. C. (2006). Avian diversity and West Nile virus: testing associations between biodiversity and infectious disease risk. Proceedings of the Royal Society of London Series B – Biological Sciences, 273, 109–117.CrossRefGoogle ScholarPubMed
Ezenwa, V. O., Milheim, L. E., Coffey, M. F., Godsey, M. S., King, R. J. & Guptill, S. C. (2007). Land cover variation and West Nile virus prevalence: patterns, processes, and implications for disease control. Vector-Borne and Zoonotic Diseases, 7, 173–180.CrossRefGoogle ScholarPubMed
Fa, J. E., Peres, C. A. & Meeuwig, J. (2002). Bushmeat exploitation in tropical forests: an intercontinental comparison. Conservation Biology, 16, 232–237.CrossRefGoogle Scholar
Fauchier, J. & Thomas, F. (2001). Interaction between Gammarinema gammari (Nematoda), Microphallus papillorobustus (Trematoda) and their common host Gammarus insensibilis (Amphipoda). Journal of Parasitology, 87, 1479–1481.CrossRefGoogle Scholar
Feener, D. H. (1981). Competition between ant species: outcome controlled by parasitic flies. Science, 214, 815–817.CrossRefGoogle ScholarPubMed
Feener, D. H. (2000). Is the assembly of ant communities mediated by parasitoids?Oikos, 90, 79–88.CrossRefGoogle Scholar
Fenton, A. & Pedersen, A. B. (2005). Community epidemiology framework for classifying disease threats. Emerging Infectious Diseases, 11, 1815–1821.CrossRefGoogle ScholarPubMed
Fenton, A. & Rands, S. A. (2006). The impact of parasite manipulation and predator foraging behavior on predator–prey communities. Ecology, 87, 2832–2841.CrossRefGoogle ScholarPubMed
Ferguson, K. I. & Stiling, P. (1996). Non-additive effects of multiple natural enemies on aphid populations. Oecologia, 108, 375–379.CrossRefGoogle ScholarPubMed
Ferrari, N., Cattadori, I. M., Rizzoli, A. & Hudson, P. J. (2009). Heligmosomoides polygyrus reduces infestation of Ixodes ricinus in free-living yellow-necked mice, Apodemus flavicollis. Parasitology, 136, 305–316.CrossRefGoogle ScholarPubMed
Ferrer, M. & Negro, J. J. (2004). The near extinction of two large European predators: super specialists pay a price. Conservation Biology, 18, 344–349.CrossRefGoogle Scholar
Field, H. E., Mackenzie, J. S. & Daszak, P. (2007). Henipaviruses: emerging paramyxoviruses associated with fruit bats. In Childs, J. E., Mackenzie, J. S. & Richt, J. A. (eds) Wildlife and Emerging Zoonotic Diseases: The Biology, Circumstances and Consequences of Cross-species Transmission. Berlin: Springer, pp. 133–159.Google ScholarPubMed
Fielding, N. J., MacNeil, C., Dick, J. T. A., Elwood, R. W., Riddell, G. E. & Dunn, A. M. (2003). Effects of the acanthocephalan parasite Echinorhynchus truttae on the feeding ecology of Gammarus pulex (Crustacea: Amphipoda). Journal of Zoology, 261, 321–325.CrossRefGoogle Scholar
Fielding, N. J., MacNeil, C., Robinson, N., Dick, J. T. A., Elwood, R. W., Terry, R. S., Ruiz, Z. & Dunn, A. M. (2005). Ecological impacts of the microsporidian parasite Pleistophora mulleri on its freshwater amphipod host Gammarus duebeni celticus. Parasitology, 131, 331–336.CrossRefGoogle ScholarPubMed
Finkes, L. K., Cady, A. B., Mulroy, J. C., Clay, K. & Rudgers, J. A. (2006). Plant–fungus mutualism affects spider composition in successional fields. Ecology Letters, 9, 344–353.CrossRefGoogle ScholarPubMed
Firlej, A., Boivin, G., Lucas, E. & Coderre, D. (2005). First report of Harmonia axyridis Pallas being attacked by Dinocampus coccinellae Schrank in Canada. Biological Invasions, 7, 553–556.CrossRefGoogle Scholar
Frank, S. A. (1996). Models of parasite virulence. Quarterly Review of Biology, 71, 37–78.CrossRefGoogle ScholarPubMed
Fransen, J. J. & Vanlenteren, J. C. (1993). Host selection and survival of the parasitoid Encarsia formosa on greenhouse-whitefly, Trialeurodes vaporariorum, in the presence of hosts infected with the fungus Aschersonia aleyrodis. Entomologia experimentalis et Applicata, 69, 239–249.CrossRefGoogle Scholar
Fraser, C., Riley, S., Anderson, R. M. & Ferguson, N. M. (2004). Factors that make an infectious disease outbreak controllable. Proceedings of the National Academy of Sciences of the United States of America, 101, 6146–6151.CrossRefGoogle ScholarPubMed
Freckleton, R. P. & Lewis, O. T. (2006). Pathogens, density dependence and the coexistence of tropical trees. Proceedings of the Royal Society of London Series B – Biological Sciences, 273, 2909–2916.CrossRefGoogle ScholarPubMed
Fredensborg, B. L. & Poulin, R. (2005). Larval helminths in intermediate hosts: does competition early in life determine the fitness of adult parasites?International Journal for Parasitology, 35, 1061–1070.CrossRefGoogle ScholarPubMed
Freedman, H. I. (1990). A model of predator–prey dynamics as modified by the action of a parasite. Mathematical Biosciences, 99, 143–155.CrossRefGoogle ScholarPubMed
Furlong, M. J. & Pell, J. K. (1996). Interactions between the fungal entomopathogen Zoophthora radicans Brefeld (Entomophthorales) and two hymenopteran parasitoids attacking the diamondback moth, Plutella xylostella L. Journal of Invertebrate Pathology, 68, 15–21.CrossRefGoogle ScholarPubMed
Futerman, P. H., Layen, S. J., Kotzen, M. L., Franzen, C., Kraaijeveld, A. R. & Godfray, H. C. J. (2006). Fitness effects and transmission routes of a microsporidian parasite infecting Drosophila and its parasitoids. Parasitology, 132, 479–492.CrossRefGoogle ScholarPubMed
Galbreath, J., 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, 235–244.Google Scholar
Gallup, J. L. & Sachs, J. D. (2001). The economic burden of malaria. American Journal of Tropical Medicine and Hygiene, 64, 85–96.CrossRefGoogle ScholarPubMed
Gandon, S. (2002). Local adaptation and the geometry of host–parasite coevolution. Ecology Letters, 5, 246–256.CrossRefGoogle Scholar
Gandon, S., Capowiez, Y., Dubois, Y., Michalakis, Y. & Olivieri, I. (1996). Local adaptation and gene-for-gene coevolution in a metapopulation model. Proceedings of the Royal Society of London Series B – Biological Sciences, 263, 1003–1009.CrossRefGoogle Scholar
Garner, T. W. J., Perkins, M. W., Govindarajulu, P., Seglie, D., Walker, S., Cunningham, A. A. & Fisher, M. C. (2006). The emerging amphibian pathogen Batrachochytrium dendrobatidis globally infects introduced populations of the North American bullfrog, Rana catesbeiana. Biology Letters, 2, 455–459.CrossRefGoogle ScholarPubMed
Gatherer, D. (2009). The 2009 H1N1 influenza outbreak in its historical context. Journal of Clinical Virology, 45, 174–178.CrossRefGoogle ScholarPubMed
Genner, M. J., Michel, E. & Todd, J. A. (2008). Resistance of an invasive gastropod to an indigenous trematode parasite in Lake Malawi. Biological Invasions, 10, 41–49.CrossRefGoogle Scholar
George-Nascimento, M. A. & Marin, S. L. (1992). Effects induced by two host species, the South American fur-seal Arctocephalus australis (Zimmerman), and the South American sea lion Otaria Byronia (Blainville) (Carnivora, Otariidae), on the morphology and fecundity of Corynosoma sp. (Acanthocephala, Polymorphidae) in Uruguay. Revista Chilena De Historia Natural, 65, 183–193.Google Scholar
Georgiev, B. B., Sanchez, M. I., Vasileva, G. P., Nikolov, P. N. & Green, A. J. (2007). Cestode parasitism in invasive and native brine shrimps (Artemia spp.) as a possible factor promoting the rapid invasion of A. franciscana in the Mediterranean region. Parasitology Research, 101, 1647–1655.CrossRefGoogle ScholarPubMed
Gibbs, J. N. (1978). Intercontinental epidemiology of Dutch elm disease. Annual Review of Phytopathology, 16, 287–307.CrossRefGoogle Scholar
Gibson, C. C. & Watkinson, A. R. (1992). The role of the hemiparasitic annual Rhinanthus minor in determining grassland community structure. Oecologia, 89, 62–68.CrossRefGoogle ScholarPubMed
Gilbert, L., Norman, R., Laurenson, K. M., Reid, H. W. & Hudson, P. J. (2001). Disease persistence and apparent competition in a three-host community: an empirical and analytical study of large-scale, wild populations. Journal of Animal Ecology, 70, 1053–1061.CrossRefGoogle Scholar
Gilbert, M. T. P., Rambaut, A., Wlasiuk, G., Spira, T. J., Pitchenik, A. E. & Worobey, M. (2007). The emergence of HIV/AIDS in the Americas and beyond. Proceedings of the National Academy of Sciences of the United States of America, 104, 18566–18570.CrossRefGoogle ScholarPubMed
Gilligan, C. A. & Kleczkowski, A. (1997). Population dynamics of botanical epidemics involving primary and secondary infection. Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 352, 591–608.CrossRefGoogle Scholar
Gilligan, C. A. & Bosch, F. (2008). Epidemiological models for invasion and persistence of pathogens. Annual Review of Phytopathology, 46, 385–418.CrossRefGoogle ScholarPubMed
Godfree, R. C., Tinnin, R. O. & Forbes, R. B. (2002). The effects of dwarf mistletoe, witches' brooms, stand structure, and site characteristics on the crown architecture of lodgepole pine in Oregon. Canadian Journal of Forest Research, 32, 1360–1371.CrossRefGoogle Scholar
Graham, A. L. (2008). Ecological rules governing helminth–microparasite coinfection. Proceedings of the National Academy of Sciences of the United States of America, 105, 566–570.CrossRefGoogle ScholarPubMed
Greenman, J. V. & Hudson, P. J. (1997). Infected coexistence instability with and without density-dependent regulation. Journal of Theoretical Biology, 185, 345–356.CrossRefGoogle Scholar
Greenman, J. V. & Hudson, P. J. (1999). Host exclusion and coexistence in apparent and direct competition: an application of bifurcation theory. Theoretical Population Biology, 56, 48–64.CrossRefGoogle ScholarPubMed
Greenman, J. V. & Hudson, P. J. (2000). Parasite-mediated and direct competition in a two-host shared macroparasite system. Theoretical Population Biology, 57, 13–34.CrossRefGoogle Scholar
Greger, M. (2007). The human/animal interface: emergence and resurgence of zoonotic infectious diseases. Critical Reviews in Microbiology, 33, 243–299.CrossRefGoogle ScholarPubMed
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 Series B – Biological Sciences, 321, 541–563.CrossRefGoogle ScholarPubMed
Grenfell, B. T. (1992). Parasitism and the dynamics of ungulate grazing systems. American Naturalist, 139, 907–929.CrossRefGoogle Scholar
Grewell, B. J. (2008). Parasite facilitates plant species coexistence in a coastal wetland. Ecology, 89, 1481–1488.CrossRefGoogle Scholar
Grosholz, E. D. (1992). Interactions of intraspecific, interspecific, and apparent competition with host–pathogen population dynamics. Ecology, 73, 507–514.CrossRefGoogle Scholar
Grosholz, E. D. & Ruiz, G. M. (2003). Biological invasions drive size increases in marine and estuarine invertebrates. Ecology Letters, 6, 700–705.CrossRefGoogle Scholar
Grossman, Z., Meier-Schellersheim, M., Sousa, A. E., Victorino, R. M. M. & Paul, W. E. (2002). CD4(+) T-cell depletion in HIV infection: are we closer to understanding the cause?Nature Medicine, 8, 319–323.CrossRefGoogle ScholarPubMed
Gubbins, S., Gilligan, C. A. & Kleczkowski, A. (2000). Population dynamics of plant–parasite interactions: thresholds for invasion. Theoretical Population Biology, 57, 219–233.CrossRefGoogle ScholarPubMed
Guernier, V., Hochberg, M. E. & Guegan, J. F. O. (2004). Ecology drives the worldwide distribution of human diseases. PLoS Biology, 2, 740–746.CrossRefGoogle ScholarPubMed
Gumel, A. B., Ruan, S. G., Day, T., Watmough, J., Brauer, F., Driessche, P., Gabrielson, D., Bowman, C., Alexander, M. E., Ardal, S., Wu, J. H. & Sahai, B. M. (2004). Modelling strategies for controlling SARS outbreaks. Proceedings of the Royal Society of London Series B – Biological Sciences, 271, 2223–2232.CrossRefGoogle ScholarPubMed
Gurnell, J., Rushton, S. P., Lurz, P. W. W., Sainsbury, A. W., Nettleton, P., Shirley, M. D. F., Bruemmer, C. & Geddes, N. (2006). Squirrel poxvirus: landscape scale strategies for managing disease threat. Biological Conservation, 131, 287–295.CrossRefGoogle Scholar
Gurnell, J., Wauters, L. A., Lurz, P. W. W. & Tosi, G. (2004). Alien species and interspecific competition: effects of introduced eastern grey squirrels on red squirrel population dynamics. Journal of Animal Ecology, 73, 26–35.CrossRefGoogle Scholar
Hadeler, K. P. & Freedman, H. I. (1989). Predator–prey populations with parasitic infection. Journal of Mathematical Biology, 27, 609–631.CrossRefGoogle ScholarPubMed
Hahn, B. H., Shaw, G. M., Cock, K. M. & Sharp, P. M. (2000). AIDS as a zoonosis – scientific and public health implications. Science, 287, 607–614.CrossRefGoogle ScholarPubMed
Haine, E. R., Boucansaud, K. & Rigaud, T. (2005). Conflict between parasites with different transmission strategies infecting an amphipod host. Proceedings of the Royal Society of London Series B – Biological Sciences, 272, 2505–2510.CrossRefGoogle ScholarPubMed
Hall, S. R., Becker, C. R., Simonis, J. L., Duffy, M. A., Tessier, A. L. & Cáceres, C. E. (2009). Friendly competition: evidence for a dilution effect among competitors in a planktonic host–parasite system. Ecology, 90, 791–801.CrossRefGoogle Scholar
Hall, S. R., Duffy, M. A. & Caceres, C. E. (2005). Selective predation and productivity jointly drive complex behavior in host–parasite systems. American Naturalist, 165, 70–81.Google ScholarPubMed
Hall, S. R., Tessier, A. J., Duffy, M. A., Huebner, M. & Cáceres, C. E. (2006). Warmer does not have to mean sicker: temperature and predators can jointly drive timing of epidemics. Ecology, 87, 1684–1695.CrossRefGoogle Scholar
Hamback, P. A., Stenberg, J. A. & Ericson, L. (2006). Asymmetric indirect interactions mediated by a shared parasitoid: connecting species traits and local distribution patterns for two chrysomelid beetles. Oecologia, 148, 475–481.CrossRefGoogle ScholarPubMed
Hanley, K. A., Petren, K. & Case, T. J. (1998). An experimental investigation of the competitive displacement of a native gecko by an invading gecko: no role for parasites. Oecologia, 115, 196–205.CrossRefGoogle ScholarPubMed
Hanley, K. A., Vollmer, D. M. & Case, T. J. (1995). The distribution and prevalence of helminths, coccidia and blood parasites in two competing species of gecko: implications for apparent competition. Oecologia, 102, 220–229.CrossRefGoogle ScholarPubMed
Hanselmann, R., Rodriguez, A., Lampo, M., Fajardo-Ramos, L., Aguirre, A. A., Kilpatrick, A. M., Rodriguez, J. P. & Daszak, P. (2004). Presence of an emerging pathogen of amphibians in introduced bullfrogs Rana catesbiana in Venezuela. Biological Conservation, 120, 115–119.CrossRefGoogle Scholar
Haque, M. & Venturino, E. (2006). The role of transmissible diseases in the Holling–Tanner predator–prey model. Theoretical Population Biology, 70, 273–288.CrossRefGoogle ScholarPubMed
Harkonen, T., Harding, K., Rasmussen, T. D., Teilmann, J. & Dietz, R. (2007). Age- and sex-specific mortality patterns in an emerging wildlife epidemic: the phocine distemper in European harbour seals. PLoS One, 2, e887.CrossRefGoogle Scholar
Harmon, J. P. & Andow, D. A. (2004). Indirect effects between shared prey: predictions for biological control. Biocontrol, 49, 605–626.CrossRefGoogle Scholar
Hartemink, N. A., Davis, S. A., Reiter, P., Hubalek, Z. & Heesterbeek, J. A. P. (2007). Importance of bird-to-bird transmission for the establishment of West Nile virus. Vector-Borne and Zoonotic Diseases, 7, 575–584.CrossRefGoogle ScholarPubMed
Harvell, C. D., Mitchell, C. E., Ward, J. R., Altizer, S., Dobson, A. P., Ostfeld, R. S. & Samuel, M. D. (2002). Ecology: climate warming and disease risks for terrestrial and marine biota. Science, 296, 2158–2162.CrossRefGoogle ScholarPubMed
Hatcher, M. J., Dick, J. T. A. & Dunn, A. M. (2006). How parasites affect interactions between competitors and predators. Ecology Letters, 9, 1253–1271.CrossRefGoogle ScholarPubMed
Hatcher, M. J., Dick, J. T. A. & Dunn, A. M. (2008). A keystone effect for parasites in intraguild predation?Biology Letters, 4, 534–537.CrossRefGoogle ScholarPubMed
Hatcher, M. J., Taneyhill, D. E., Dunn, A. M. & Tofts, C. (1999). Population dynamics under parasitic sex ratio distortion. Theoretical Population Biology, 56, 11–28.CrossRefGoogle ScholarPubMed
Hatcher, M. J. & Tofts, , C. (2004). Reductionism isn't functional. HP Laboratories Technical Report HPL-2004-222.
Hatcher, P. E. (1995). Three-way interactions between plant pathogenic fungi, herbivorous insects and their host plants. Biological Reviews, 70, 639–694.CrossRefGoogle Scholar
Hatcher, P. E., Moore, J., Taylor, J. E., Tinney, G. W. & Paul, N. D. (2004). Phytohormones and plant–herbivore–pathogen interactions: integrating the molecular with the ecological. Ecology, 85, 59–69.CrossRefGoogle Scholar
Hatcher, P. E. & Paul, N. D. (2000). Beetle grazing reduces natural infection of Rumex obtusifolius by fungal pathogens. New Phytologist, 146, 325–333.CrossRefGoogle Scholar
Hatcher, P. E. & Paul, N. D. (2001). Plant pathogen–herbivore interactions and their effects on weeds. In Jeger, M. J. & Spence, N. J. (eds), Biotic Interactions in Plant–Pathogen Associations. Wallingford: CAB International, pp. 193–225.CrossRefGoogle Scholar
Hatcher, P. E., Paul, N. D., Ayres, P. G. & Whittaker, J. B. (1994a). The effect of a foliar disease (rust) on the development of Gastrophysa viridula (Coleoptera, Chrysomelidae). Ecological Entomology, 19, 349–360.CrossRefGoogle Scholar
Hatcher, P. E., Paul, N. D., Ayres, P. G. & Whittaker, J. B. (1994b). Interactions between Rumex spp., herbivores and a rust fungus: Gastrophysa viridula grazing reduces subsequent infection by Uromyces rumicis. Functional Ecology, 8, 265–272.CrossRefGoogle Scholar
Hawkins, B. A. & Cornell, H. V. (eds) (1999). Theoretical Approaches to Biological Control. Cambridge: Cambridge University Press.CrossRef
Hay, S. I., Guerra, C. A., Tatem, A. J., Noor, A. M. & Snow, R. W. (2004). The global distribution and population at risk of malaria: past, present, and future. Lancet Infectious Diseases, 4, 327–336.CrossRefGoogle ScholarPubMed
Haydon, D. T., Randall, D. A., Matthews, L., Knobel, D. L., Tallents, L. A., Gravenor, M. B., Williams, S. D., Pollinger, J. P., Cleaveland, S., Woolhouse, M. E. J., Sillero-Zubiri, C., Marino, J., Macdonald, D. W. & Laurenson, M. K. (2006). Low-coverage vaccination strategies for the conservation of endangered species. Nature, 443, 692–695.CrossRefGoogle ScholarPubMed
Hayes, E. B., Komar, N., Nasci, R. S., Montgomery, S. P., O'Leary, D. R. & Campbell, G. L. (2005). Epidemiology and transmission dynamics of West Nile virus disease. Emerging Infectious Diseases, 11, 1167–1173.CrossRefGoogle ScholarPubMed
Hearne, S. J. (2009). Control: the Striga conundrum. Pest Management Science, 65, 603–614.CrossRefGoogle ScholarPubMed
Hechinger, R. F. & Lafferty, K. D. (2005). Host diversity begets parasite diversity: bird final hosts and trematodes in snail intermediate hosts. Proceedings of the Royal Society of London Series B – Biological Sciences, 272, 1059–1066.CrossRefGoogle ScholarPubMed
Hechinger, R. F., Lafferty, K. D., Huspeni, T. C., Brooks, A. J. & Kuris, A. M. (2007). Can parasites be indicators of free-living diversity? Relationships between species richness and the abundance of larval trematodes and of local benthos and fishes. Oecologia, 151, 82–92.CrossRefGoogle ScholarPubMed
Heil, M. (2008). Indirect defence via tritrophic interactions. New Phytologist, 178, 41–61.CrossRefGoogle ScholarPubMed
Heinz, K. M. & Nelson, J. M. (1996). Interspecific interactions among natural enemies of Bemisia in an inundative biological control program. Biological Control, 6, 384–393.CrossRefGoogle Scholar
Henneman, M. L. & Memmott, J. (2001). Infiltration of a Hawaiian community by introduced biological control agents. Science, 293, 1314–1316.CrossRefGoogle ScholarPubMed
Henttonen, H., Fuglei, E., Gower, C. N., Haukisalmi, V., Ims, R. A., Niemimaa, J. & Yoccoz, N. G. (2001). Echinococcus multilocularis on Svalbard: introduction of an intermediate host has enabled the local life-cycle. Parasitology, 123, 547–552.CrossRefGoogle ScholarPubMed
Hernandez, A. D. & Sukhdeo, M. V. K. (2008a). Parasite effects on isopod feeding rates can alter the host's functional role in a natural stream ecosystem. International Journal for Parasitology, 38, 683–690.CrossRefGoogle Scholar
Hernandez, A. D. & Sukhdeo, M. V. K. (2008b). Parasites alter the topology of a stream food web across seasons. Oecologia, 156, 613–624.CrossRefGoogle ScholarPubMed
Herrick, N. J., Reitz, S. R., Carpenter, J. E. & O'Brien, C. W. (2008). Predation by Podisus maculiventris (Hemiptera: Pentatomidae) on Plutella xylostella (Lepidoptera: Plutellidae) larvae parasitized by Cotesia plutellae (Hymenoptera: Braconidae) and its impact on cabbage. Biological Control, 45, 386–395.CrossRefGoogle Scholar
Hethcote, H. W., Wang, W. D., Han, L. T. & Zhien, M. (2004). A predator–prey model with infected prey. Theoretical Population Biology, 66, 259–268.CrossRefGoogle ScholarPubMed
Hilker, F. M. & Schmitz, K. (2008). Disease-induced stabilization of predator–prey oscillations. Journal of Theoretical Biology, 255, 299–306.CrossRefGoogle ScholarPubMed
Hoch, G., Schopf, A. & Maddox, J. V. (2000). Interactions between an entomopathogenic microsporidium and the endoparasitoid Glyptapanteles liparidis within their host, the gypsy moth larva. Journal of Invertebrate Pathology, 75, 59–68.CrossRefGoogle ScholarPubMed
Hochberg, M. E., Hassell, M. P. & May, R. M. (1990). The dynamics of host–parasitoid–pathogen interactions. American Naturalist, 135, 74–94.CrossRefGoogle Scholar
Hochberg, M. E. & Holt, R. D. (1990). The coexistence of competing parasites:1 – the role of cross-species infection. American Naturalist, 136, 517–541.CrossRefGoogle Scholar
Hochberg, M. E. & Lawton, J. H. (1990). Competition between kingdoms. Trends in Ecology & Evolution, 5, 367–371.CrossRefGoogle ScholarPubMed
Holdich, D. M. (2002). Biology of Freshwater Crayfish. Oxford: Blackwell Science.Google Scholar
Holdich, D. M. & Poeckl, M. (2007). Invasive crustaceans in European inland waters. In Gherardi, F. (ed.), Biological Invaders in Inland Waters: Profiles, Distribution and Threats. Netherlands: Springer, pp. 29–75.CrossRefGoogle Scholar
Holdo, R. M., Sinclair, A. R. E., Dobson, A. P., Metzger, K. L., Bolker, B. M., Ritchie, M. E. & Holt, R. D. (2009). A disease-mediated trophic cascade in the Serengeti and its implications for ecosystem C. PLoS Biology, 7, e1000210.CrossRefGoogle ScholarPubMed
Holt, R. D. (1977). Predation, apparent competition, and structure of prey communities. Theoretical Population Biology, 12, 197–229.CrossRefGoogle ScholarPubMed
Holt, R. D. (1984). Spatial heterogeneity, indirect interactions, and the coexistence of prey species. American Naturalist, 124, 377–406.CrossRefGoogle ScholarPubMed
Holt, R. D. (1997). Community modules. In Gange, A. C. & Brown, V. K. (eds), Multitrophic Interactions in Terrestrial Systems. London: Blackwell Science, pp. 333–350.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. & Ray, C. (eds), Disease Ecology: Community Structure and Pathogen Dynamics. Oxford: Oxford University Press, pp. 6–27.CrossRefGoogle Scholar
Holt, R. D., Dobson, A. P., Begon, M., Bowers, R. G. & Schauber, E. M. (2003). Parasite establishment in host communities. Ecology Letters, 6, 837–842.CrossRefGoogle Scholar
Holt, R. D., Grover, J. & Tilman, D. (1994). Simple rules for interspecific dominance in systems with exploitative and apparent competition. American Naturalist, 144, 741–771.CrossRefGoogle Scholar
Holt, R. D. & Hochberg, M. E. (1998). The coexistence of competing parasites: part II – hyperparasitism and food chain dynamics. Journal of Theoretical Biology, 193, 485–495.CrossRefGoogle ScholarPubMed
Holt, R. D. & Hochberg, M. E. (2001). Indirect interactions, community modules and biological control: a theoretical perspective. In Wajnberg, E., Scott, J. K. & Quimby, P. C. (eds), Evaluating Indirect Ecological Effects of Biological Control. Wallingford: CAB International, pp. 13–37.Google Scholar
Holt, R. D. & Huxel, G. R. (2007). Alternative prey and the dynamics of intraguild predation: theoretical perspectives. Ecology, 88, 2706–2712.CrossRefGoogle ScholarPubMed
Holt, R. D. & Kotler, B. P. (1987). Short-term apparent competition. American Naturalist, 130, 412–430.CrossRefGoogle Scholar
Holt, R. D. & Lawton, J. H. (1993). Apparent competition and enemy-free space in insect host–parasitoid communities. American Naturalist, 142, 623–645.CrossRefGoogle ScholarPubMed
Holt, R. D. & Pickering, J. (1985). Infectious disease and species coexistence: a model of Lotka–Volterra form. American Naturalist, 126, 196–211.CrossRefGoogle Scholar
Holt, R. D. & Polis, G. A. (1997). A theoretical framework for intraguild predation. American Naturalist, 149, 745–764.CrossRefGoogle Scholar
Holt, R. D. & Roy, M. (2007). Predation can increase the prevalence of infectious disease. American Naturalist, 169, 690–699.CrossRefGoogle ScholarPubMed
Hoogendoorn, M. & Heimpel, G. E. (2002). Indirect interactions between an introduced and a native ladybird beetle species mediated by a shared parasitoid. Biological Control, 25, 224–230.CrossRefGoogle Scholar
Horimoto, T. & Kawaoka, Y. (2005). Influenza: lessons from past pandemics, warnings from current incidents. Nature Reviews Microbiology, 3, 591–600.CrossRefGoogle ScholarPubMed
Hudault, S., Guignot, J. & Servin, A. L. (2001). Escherichia coli strains colonising the gastrointestinal tract protect germfree mice against Salmonella typhimurium infection. GUT, 49, 47–55.CrossRefGoogle ScholarPubMed
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, 381–385.CrossRefGoogle Scholar
Hudson, P. J., Dobson, A. P. & Newborn, D. (1992a). Do parasites make prey vulnerable to predation: red grouse and parasites. Journal of Animal Ecology, 61, 681–692.CrossRefGoogle Scholar
Hudson, P. J., Dobson, A. P. & Newborn, D. (1998). Prevention of population cycles by parasite removal. Science, 282, 2256–2258.CrossRefGoogle ScholarPubMed
Hudson, P. J. & Greenman, J. (1998). Competition mediated by parasites: biological and theoretical progress. Trends in Ecology & Evolution, 13, 387–390.CrossRefGoogle ScholarPubMed
Hudson, P. J., Newborn, D. & Dobson, A. P. (1992b). Regulation and stability of a free-living host–parasite system: Trichostrongylus tenuis in red grouse – 1. Monitoring and parasite reduction experiments. Journal of Animal Ecology, 61, 477–486.CrossRefGoogle Scholar
Huspeni, T. C. & Lafferty, K. D. (2004). Using larval trematodes that parasitize snails to evaluate a saltmarsh restoration project. Ecological Applications, 14, 795–804.CrossRefGoogle Scholar
Huxham, M., Beaney, S. & Raffaelli, D. (1996). Do parasites reduce the chances of triangulation in a real food web?Oikos, 76, 284–300.CrossRefGoogle Scholar
Huxham, M., Raffaelli, D. & Pike, A. (1995). Parasites and food-web patterns. Journal of Animal Ecology, 64, 168–176.CrossRefGoogle Scholar
Inbar, M. & Gerling, D. (2008). Plant-mediated interactions between whiteflies, herbivores, and natural enemies. Annual Review of Entomology, 53, 431–448.CrossRefGoogle ScholarPubMed
Ishtiaq, F., Beadell, J. S., Baker, A. J., Rahmani, A. R., Jhala, Y. V. & Fleischer, R. C. (2006). Prevalence and evolutionary relationships of haematozoan parasites in native versus introduced populations of common myna Acridotheres tristis. Proceedings of the Royal Society of London Series B – Biological Sciences, 273, 587–594.CrossRefGoogle ScholarPubMed
Ives, A. R. & Murray, D. L. (1997). Can sublethal parasitism destabilize predator–prey population dynamics? A model of snowshoe hares, predators and parasites. Journal of Animal Ecology, 66, 265–278.CrossRefGoogle Scholar
Jaenike, J. (1995). Interactions between mycophagous Drosophila and their nematode parasites: from physiological to community ecology. Oikos, 72, 235–244.CrossRefGoogle Scholar
James, T. Y., Litvintseva, A. P., Vilgalys, R., Morgan, J. A. T., Taylor, J. W., Fisher, M. C., Berger, L., Weldon, C., du Preez, L. & Longcore, J. E. (2009). Rapid global expansion of the fungal disease chytridiomycosis into declining and healthy amphibian populations. PLoS Pathogens, 5, e1000458.CrossRefGoogle ScholarPubMed
Janssen, A., Sabelis, M. W., Magalhaes, S., Montserrat, M. & Hammen, T. (2007). Habitat structure affects intraguild predation. Ecology, 88, 2713–2719.CrossRefGoogle ScholarPubMed
Janzen, D. H. (1970). Herbivores and number of tree species in tropical forests. American Naturalist, 104, 501–528.CrossRefGoogle Scholar
Jensen, T., Jensen, K. T. & Mouritsen, K. N. (1998). The influence of the trematode Microphallus claviformis on two congeneric intermediate host species (Corophium): infection characteristics and host survival. Journal of Experimental Marine Biology and Ecology, 227, 35–48.CrossRefGoogle Scholar
Johnson, N. C., Graham, J. H. & Smith, F. A. (1997). Functioning of mycorrhizal associations along the mutualism–parasitism continuum. New Phytologist, 135, 575–586.CrossRefGoogle Scholar
Johnson, P. T. J., Dobson, A., Lafferty, K. D., Marcogliese, D. J., Memmott, J., Orlofske, S. A., Poulin, R. & Thieltges, D. W. (2010). When parasites become prey: ecological and epidemiological significance of eating parasites. Trends in Ecology & Evolution, 25, 362–371.CrossRefGoogle ScholarPubMed
Joly, D. O. & Messier, F. (2004). The distribution of Echinococcus granulosus in moose: evidence for parasite-induced vulnerability to predation by wolves?Oecologia, 140, 586–590.CrossRefGoogle ScholarPubMed
Jones, C. G., Lawton, J. H. & Shachak, M. (1994). Organisms as ecosystem engineers. Oikos, 69, 373–386.CrossRefGoogle Scholar
Jones, K. E., Patel, N. G., Levy, M. A., Storeygard, A., Balk, D., Gittleman, J. L. & Daszak, P. (2008). Global trends in emerging infectious diseases. Nature, 451, 990–993.CrossRefGoogle ScholarPubMed
Joo, J., Gunny, M., Cases, M., Hudson, P., Albert, R. & Harvill, E. (2006). Bacteriophage-mediated competition in Bordetella bacteria. Proceedings of the Royal Society of London Series B – Biological Sciences, 273, 1843–1848.CrossRefGoogle ScholarPubMed
Joshi, J., Matthies, D. & Schmid, B. (2000). Root hemiparasites and plant diversity in experimental grassland communities. Journal of Ecology, 88, 634–644.CrossRefGoogle Scholar
Juliano, S. A. & Lounibos, L. P. (2005). Ecology of invasive mosquitoes: effects on resident species and on human health. Ecology Letters, 8, 558–574.CrossRefGoogle ScholarPubMed
Kaltz, O. & Shykoff, J. A. (1998). Local adaptation in host–parasite systems. Heredity, 81, 361–370.CrossRefGoogle Scholar
Kanno, H. & Fujita, Y. (2003). Induced systemic resistance to rice blast fungus in rice plants infested by white-backed planthopper. Entomologia experimentalis et Applicata, 107, 155–158.CrossRefGoogle Scholar
Kanno, H., Satoh, M., Kimura, T. & Fujita, Y. (2005). Some aspects of induced resistance to rice blast fungus, Magnaporthe grisea, in rice plant infested by white-backed planthopper, Sogatella furcifera. Applied Entomology and Zoology, 40, 91–97.CrossRefGoogle Scholar
Karban, R., Baldwin, I. T. (1997). Induced Responses to Herbivory. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Karban, R., Baldwin, I. T., Baxter, K. J., Laue, G. & Felton, G. W. (2000). Communication between plants: induced resistance in wild tobacco plants following clipping of neighboring sagebrush. Oecologia, 125, 66–71.CrossRefGoogle ScholarPubMed
Keane, R. M. & Crawley, M. J. (2002). Exotic plant invasions and the enemy release hypothesis. Trends in Ecology & Evolution, 17, 164–170.CrossRefGoogle Scholar
Keele, B. F., Jones, J. H., Terio, K. A., Estes, J. D., Rudicell, R. S., Wilson, M. L., Li, Y. Y., Learn, G. H., Beasley, T. M., Schumacher-Stankey, J., Wroblewski, E., Mosser, A., Raphael, J., Kamenya, S., Lonsdorf, E. V., Travis, D. A., Mlengeya, T., Kinsel, M. J., Else, J. G., Silvestri, G., Goodall, J., Sharp, P. M., Shaw, G. M., Pusey, A. E. & Hahn, B. H. (2009). Increased mortality and AIDS-like immunopathology in wild chimpanzees infected with SIVcpz. Nature, 460, 515–519.CrossRefGoogle ScholarPubMed
Keele, B. F., Heuverswyn, F., Li, Y. Y., Bailes, E., Takehisa, J., Santiago, M. L., Bibollet-Ruche, F., Chen, Y. L., Wain, L. V., Liegeois, F., Loul, S., Ngole, E. M., Bienvenue, Y., Delaporte, E., Brookfield, J. F. Y., Sharp, P. M., Shaw, G. M., Peeters, M. & Hahn, B. H. (2006). Chimpanzee reservoirs of pandemic and nonpandemic HIV-1. Science, 313, 523–526.CrossRefGoogle ScholarPubMed
Keesing, F., Holt, R. D. & Ostfeld, R. S. (2006). Effects of species diversity on disease risk. Ecology Letters, 9, 485–498.CrossRefGoogle ScholarPubMed
Kelly, D. W., Bailey, R. J., MacNeil, C., Dick, J. T. A. & McDonald, R. A. (2006). Invasion by the amphipod Gammarus pulex alters community composition of native freshwater macroinvertebrates. Diversity and Distributions, 12, 525–534.CrossRefGoogle Scholar
Kelly, D. W., Paterson, R. A., Townsend, C. R., Poulin, R. & Tompkins, D. M. (2009). Parasite spillback: a neglected concept in invasion ecology?Ecology, 90, 2047–2056.CrossRefGoogle ScholarPubMed
Kennedy, C. R. & Guegan, J. F. (1994). Regional versus local helminth parasite richness in British fresh-water fish: saturated or unsaturated parasite communities. Parasitology, 109, 175–185.CrossRefGoogle ScholarPubMed
Kessler, A. & Baldwin, I. T. (2001). Defensive function of herbivore-induced plant volatile emissions in nature. Science, 291, 2141–2144.CrossRefGoogle ScholarPubMed
Kfir, R., Gouws, J. & Moore, S. D. (1993). Biology of Tetrastichus howardi (Olliff) (Hymenoptera, Eulophidae): a facultative hyperparasitoid of stem borers. Biocontrol Science and Technology, 3, 149–159.CrossRefGoogle Scholar
Kiesecker, J. M. & Blaustein, A. R. (1999). Pathogen reverses competition between larval amphibians. Ecology, 80, 2442–2448.CrossRefGoogle Scholar
Kilpatrick, A. M., Daszak, P., Goodman, S. J., Rogg, H., Kramer, L. D., Cedeno, V. & Cunningham, A. A. (2006). Predicting pathogen introduction: West Nile virus spread to Galapagos. Conservation Biology, 20, 1224–1231.CrossRefGoogle Scholar
Kim, K. & Harvell, C. D. (2004). The rise and fall of a six-year coral–fungal epizootic. American Naturalist, 164, S52–S63.CrossRefGoogle ScholarPubMed
King, A. A. & Hastings, A. (2003). Spatial mechanisms for coexistence of species sharing a common natural enemy. Theoretical Population Biology, 64, 431–438.CrossRefGoogle ScholarPubMed
Kleczkowski, A., Gilligan, C. A. & Bailey, D. J. (1997). Scaling and spatial dynamics in plant–pathogen systems: from individuals to populations. Proceedings of the Royal Society of London Series B – Biological Sciences, 264, 979–984.CrossRefGoogle Scholar
Kloepper, J. W., Ryu, C. M. & Zhang, S. A. (2004). Induced systemic resistance and promotion of plant growth by Bacillus spp. Phytopathology, 94, 1259–1266.CrossRefGoogle ScholarPubMed
Knevel, I. C., Lans, T., Menting, F. B. J., Hertling, U. M. & Putten, W. H. (2004). Release from native root herbivores and biotic resistance by soil pathogens in a new habitat both affect the alien Ammophila arenaria in South Africa. Oecologia, 141, 502–510.CrossRefGoogle Scholar
Koella, J. C., Lynch, P. A., Thomas, M. B. & Read, A. F. (2009). Towards evolution-proof malaria control with insecticides. Evolutionary Applications, 2, 469–480.CrossRefGoogle ScholarPubMed
Kohler, S. L. (2008). The ecology of host–parasite interactions in aquatic insects. In Lancaster, J., Briers, R. & Macadam, C. (eds), Aquatic Insects: Challenges to Populations. Wallingford: CAB International, pp. 55–80.CrossRefGoogle Scholar
Kohler, S. L. & Wiley, M. J. (1997). Pathogen outbreaks reveal large-scale effects of competition in stream communities. Ecology, 78, 2164–2176.CrossRefGoogle Scholar
Kokko, H. & Lindstrom, J. (1998). Seasonal density dependence, timing of mortality, and sustainable harvesting. Ecological Modelling, 110, 293–304.CrossRefGoogle Scholar
Kolar, C. S. & Lodge, D. M. (2001). Progress in invasion biology: predicting invaders. Trends in Ecology & Evolution, 16, 199–204.CrossRefGoogle ScholarPubMed
Komar, N., Langevin, S., Hinten, S., Nemeth, N., Edwards, E., Hettler, D., Davis, B., Bowen, R. & Bunning, M. (2003). Experimental infection of North American birds with the New York 1999 strain of West Nile virus. Emerging Infectious Diseases, 9, 311–322.CrossRefGoogle ScholarPubMed
Kopp, K. & Jokela, J. (2007). Resistant invaders can convey benefits to native species. Oikos, 116, 295–301.CrossRefGoogle Scholar
Koprivnikar, J., Forbes, M. R. & Baker, R. L. (2008). Larval amphibian growth and development under varying density: are parasitized individuals poor competitors?Oecologia, 155, 641–649.CrossRefGoogle ScholarPubMed
Krkosek, M., Lewis, M. A., Morton, A., Frazer, L. N. & Volpe, J. P. (2006). Epizootics of wild fish induced by farm fish. Proceedings of the National Academy of Sciences of the United States of America, 103, 15506–15510.CrossRefGoogle ScholarPubMed
Kuiken, T., Kennedy, S., Barrett, T., Bildt, M. W. G., Borgsteede, F. H., Brew, S. D., Codd, G. A., Duck, C., Deaville, R., Eybatov, T., Forsyth, M. A., Foster, G., Jepson, P. D., Kydyrmanov, A., Mitrofanov, I., Ward, C. J., Wilson, S. & Osterhaus, A. (2006). The 2000 canine distemper epidemic in Caspian seals (Phoca caspica): pathology and analysis of contributory factors. Veterinary Pathology, 43, 321–338.CrossRefGoogle ScholarPubMed
Kulmatiski, A., Beard, K. H., Stevens, J. R. & Cobbold, S. M. (2008). Plant–soil feedbacks: a meta-analytical review. Ecology Letters, 11, 980–992.CrossRefGoogle ScholarPubMed
Kuo, C. H., Corby-Harris, V. & Promislow, D. E. L. (2008). The unavoidable costs and unexpected benefits of parasitism: population and metapopulation models of parasite-mediated competition. Journal of Theoretical Biology, 250, 244–256.CrossRefGoogle ScholarPubMed
Kuris, A. M., Hechinger, R. F., Shaw, J. C., Whitney, K. L., Aguirre-Macedo, L., Boch, C. A., Dobson, A. P., Dunham, E. J., Fredensborg, B. L., Huspeni, T. C., Lorda, J., Mababa, L., Mancini, F. T., Mora, A. B., Pickering, M., Talhouk, N. L., Torchin, M. E. & Lafferty, K. D. (2008). Ecosystem energetic implications of parasite and free-living biomass in three estuaries. Nature, 454, 515–518.CrossRefGoogle ScholarPubMed
Kuris, A. M. & Lafferty, K. D. (1992). Modeling crustacean fisheries: effects of parasites on management strategies. Canadian Journal of Fisheries and Aquatic Sciences, 49, 327–336.CrossRefGoogle Scholar
Kuris, A. M. & Lafferty, K. D. (1994). Community structure: larval trematodes in snail hosts. Annual Review of Ecology and Systematics, 25, 189–217.CrossRefGoogle Scholar
LaDeau, S. L., Kilpatrick, A. M. & Marra, P. P. (2007). West Nile virus emergence and large-scale declines of North American bird populations. Nature, 447, 710–713.CrossRefGoogle ScholarPubMed
Lafferty, K. D. (1992). Foraging on prey that are modified by parasites. American Naturalist, 140, 854–867.CrossRefGoogle Scholar
Lafferty, K. D. (2009). The ecology of climate change and infectious diseases. Ecology, 90, 888–900.CrossRefGoogle ScholarPubMed
Lafferty, K. D., Allesina, S., Arim, M., Briggs, C. J., Leo, G., Dobson, A. P., Dunne, J. A., Johnson, P. T. J., Kuris, A. M., Marcogliese, D. J., Martinez, N. D., Memmott, J., Marquet, P. A., McLaughlin, J. P., Mordecai, E. A., Pascual, M., Poulin, R. & Thieltges, D. W. (2008a). Parasites in food webs: the ultimate missing links. Ecology Letters, 11, 533–546.CrossRefGoogle ScholarPubMed
Lafferty, K. D., Dobson, A. P. & Kuris, A. M. (2006a). Parasites dominate food web links. Proceedings of the National Academy of Sciences of the United States of America, 103, 11211–11216.CrossRefGoogle ScholarPubMed
Lafferty, K. D., Hechinger, R. F., Shaw, J. C., Whitney, K. & Kuris, A. M. (2006b). Food webs and parasites in a salt marsh ecosystem. In Collinge, S. K. & Ray, C. (eds), Disease Ecology: Community Structure and Pathogen Dynamics. Oxford: Oxford University Press, pp. 119–134.CrossRefGoogle Scholar
Lafferty, K. D. & Kuris, A. M. (2009). Parasites reduce food web robustness because they are sensitive to secondary extinction as illustrated by an invasive estuarine snail. Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 364, 1659–1663.CrossRefGoogle ScholarPubMed
Lafferty, K. D. & Morris, A. K. (1996). Altered behavior of parasitized killifish increases susceptibility to predation by bird final hosts. Ecology, 77, 1390–1397.CrossRefGoogle Scholar
Lafferty, K. D., Porter, J. W. & Ford, S. E. (2004). Are diseases increasing in the ocean?Annual Review of Ecology Evolution and Systematics, 35, 31–54.CrossRefGoogle Scholar
Lafferty, K. D., Shaw, J. C. & Kuris, A. M. (2008b). Reef fishes have higher parasite richness at unfished Palmyra Atoll compared to fished Kiritimati Island. Ecohealth, 5, 338–345.CrossRefGoogle ScholarPubMed
Lagrue, C. & Poulin, R. (2008). Intra- and interspecific competition among helminth parasites: effects on Coitocaecum parvum life history, strategy, size and fecundity. International Journal for Parasitology, 38, 1435–1444.CrossRefGoogle ScholarPubMed
Lau, S. K. P., Woo, P. C. Y., Li, K. S. M., Huang, Y., Tsoi, H. W., Wong, B. H. L., Wong, S. S. Y., Leung, S. Y., Chan, K. H. & Yuen, K. Y. (2005). Severe acute respiratory syndrome coronavirus-like virus in Chinese horseshoe bats. Proceedings of the National Academy of Sciences of the United States of America, 102, 14040–14045.CrossRefGoogle ScholarPubMed
Laurenson, M. K., McKendrick, I. J., Reid, H. W., Challenor, R. & Mathewson, G. K. (2007). Prevalence, spatial distribution and the effect of control measures on louping-ill virus in the Forest of Bowland, Lancashire. Epidemiology and Infection, 135, 963–973.CrossRefGoogle ScholarPubMed
Laurenson, M. K., Norman, R. A., Gilbert, L., Reid, H. W. & Hudson, P. J. (2003). Identifying disease reservoirs in complex systems: mountain hares as reservoirs of ticks and louping-ill virus, pathogens of red grouse. Journal of Animal Ecology, 72, 177–185.CrossRefGoogle Scholar
Laurenson, M. K., Norman, R. A., Gilbert, L., Reid, H. W. & Hudson, P. J. (2004). Mountain hares, louping-ill, red grouse and harvesting: complex interactions but few data. Journal of Animal Ecology, 73, 811–813.CrossRefGoogle Scholar
Lavigne, D. M. & Schmitz, O. J. (1990). Global warming and increasing population densities: a prescription for seal plagues. Marine Pollution Bulletin, 21, 280–284.CrossRefGoogle Scholar
Leath, K. T. & Byers, R. A. (1977). Interaction of fusarium root-rot with pea aphid and potato leafhopper feeding on forage legumes. Phytopathology, 67, 226–229.CrossRefGoogle Scholar
Lee, K. A. & Klasing, K. C. (2004). A role for immunology in invasion biology. Trends in Ecology & Evolution, 19, 523–529.CrossRefGoogle ScholarPubMed
Lee, K. A., Martin, L. B. & Wikelski, M. C. (2005). Responding to inflammatory challenges is less costly for a successful avian invader, the house sparrow (Passer domesticus), than its less-invasive congener. Oecologia, 145, 244–251.CrossRefGoogle Scholar
Lefevre, T., Roche, B., Poulin, R., Hurd, H., Renaud, F. & Thomas, F. (2008). Exploiting host compensatory responses: the ‘must’ of manipulation?Trends in Parasitology, 24, 435–439.CrossRefGoogle Scholar
Lehtonen, P., Helander, M., Wink, M., Sporer, F. & Saikkonen, K. (2005). Transfer of endophyte-origin defensive alkaloids from a grass to a hemiparasitic plant. Ecology Letters, 8, 1256–1263.CrossRefGoogle Scholar
Lello, J., Boag, B., Fenton, A., Stevenson, I. R. & Hudson, P. J. (2004). Competition and mutualism among the gut helminths of a mammalian host. Nature, 428, 840–844.CrossRefGoogle ScholarPubMed
Lemons, A., Clay, K. & Rudgers, J. A. (2005). Connecting plant–microbial interactions above and belowground: a fungal endophyte affects decomposition. Oecologia, 145, 595–604.CrossRefGoogle ScholarPubMed
Lenski, R. E. & May, R. M. (1994). The evolution of virulence in parasites and pathogens: reconciliation between two competing hypotheses. Journal of Theoretical Biology, 169, 253–265.CrossRefGoogle ScholarPubMed
Leroy, E. M., Kumulungui, B., Pourrut, X., Rouquet, P., Hassanin, A., Yaba, P., Delicat, A., Paweska, J. T., Gonzalez, J. P. & Swanepoel, R. (2005). Fruit bats as reservoirs of Ebola virus. Nature, 438, 575–576.CrossRefGoogle Scholar
Leroy, E. M., Rouquet, P., Formenty, P., Souquiere, S., Kilbourne, A., Froment, J. M., Bermejo, M., Smit, S., Karesh, W., Swanepoel, R., Zaki, S. R. & Rollin, P. E. (2004). Multiple Ebola virus transmission events and rapid decline of central African wildlife. Science, 303, 387–390.CrossRefGoogle ScholarPubMed
Leung, T. L. F., Keeney, D. B. & Poulin, R. (2009). Cryptic species complexes in manipulative echinostomatid trematodes: when two become six. Parasitology, 136, 241–252.CrossRefGoogle ScholarPubMed
Li, W. D., Shi, Z. L., Yu, M., Ren, W. Z., Smith, C., Epstein, J. H., Wang, H. Z., Crameri, G., Hu, Z. H., Zhang, H. J., Zhang, J. H., McEachern, J., Field, H., Daszak, P., Eaton, B. T., Zhang, S. Y. & Wang, L. F. (2005). Bats are natural reservoirs of SARS-like coronaviruses. Science, 310, 676–679.CrossRefGoogle ScholarPubMed
Lips, K. R., Diffendorfer, J., Mendelson, J. R. & Sears, M. W. (2008). Riding the wave: reconciling the roles of disease and climate change in amphibian declines. PLoS Biology, 6, 441–454.CrossRefGoogle ScholarPubMed
Lipsitch, M., Cohen, T., Cooper, B., Robins, J. M., Ma, S., James, L., Gopalakrishna, G., Chew, S. K., Tan, C. C., Samore, M. H., Fisman, D. & Murray, M. (2003). Transmission dynamics and control of severe acute respiratory syndrome. Science, 300, 1966–1970.CrossRefGoogle ScholarPubMed
Lloyd-Smith, J. O., Galvani, A. P. & Getz, W. M. (2003). Curtailing transmission of severe acute respiratory syndrome within a community and its hospital. Proceedings of the Royal Society of London Series B – Biological Sciences, 270, 1979–1989.CrossRefGoogle ScholarPubMed
Lloyd-Smith, J. O., George, D., Pepin, K. M., Pitzer, V. E., Pulliam, J. R. C., Dobson, A. P., Hudson, P. J. & Grenfell, B. T. (2009). Epidemic dynamics at the human–animal interface. Science, 326, 1362–1367.CrossRefGoogle ScholarPubMed
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, 355–359.CrossRefGoogle ScholarPubMed
Lockwood, J. L., Hoopes, M. F. & Marchetti, M. P. (2007). Invasion Ecology. Oxford: Blackwell.Google Scholar
LoGiudice, K., Duerr, S. T. K., Newhouse, M. J., Schmidt, K. A., Killilea, M. E. & Ostfeld, R. S. (2008). Impact of host community composition on Lyme disease risk. Ecology, 89, 2841–2849.CrossRefGoogle ScholarPubMed
LoGiudice, K., Ostfeld, R. S., Schmidt, K. A. & Keesing, F. (2003). The ecology of infectious disease: effects of host diversity and community composition on Lyme disease risk. Proceedings of the National Academy of Sciences of the United States of America, 100, 567–571.CrossRefGoogle ScholarPubMed
Loo, J. (2009). Ecological impacts of non-indigenous invasive fungi as forest pathogens. Biological Invasions, 11, 81–96.CrossRefGoogle Scholar
Lopez, G., Lopez-Parra, M., Fernandez, L., Martinez-Granados, C., Martinez, F., Meli, M. L., Gil-Sanchez, J. M., Viqueira, N., Diaz-Portero, M. A., Cadenas, R., Lutz, H., Vargas, A. & Simon, M. A. (2009). Management measures to control a feline leukemia virus outbreak in the endangered Iberian lynx. Animal Conservation, 12, 173–182.CrossRefGoogle Scholar
Losey, J. E., Ives, A. R., Harmon, J., Ballantyne, F. & Brown, C. (1997). A polymorphism maintained by opposite patterns of parasitism and predation. Nature, 388, 269–272.CrossRefGoogle Scholar
Lubick, N. (2010). Emergency medicine for frogs. Nature, 465, 680–681.CrossRefGoogle ScholarPubMed
MacArthur, R. (1955). Fluctuations of animal populations, and a measure of community stability. Ecology, 36, 533–536.CrossRefGoogle Scholar
MacDonald, D. W. & Laurenson, M. K. (2006). Infectious disease: inextricable linkages between human and ecosystem health. Biological Conservation, 131, 143–150.CrossRefGoogle Scholar
MacDonald, D. W., Riordan, P. & Mathews, F. (2006). Biological hurdles to the control of TB in cattle: a test of two hypotheses concerning wildlife to explain the failure of control. Biological Conservation, 131, 268–286.CrossRefGoogle Scholar
Mack, K. M. L. & Rudgers, J. A. (2008). Balancing multiple mutualists: asymmetric interactions among plants, arbuscular mycorrhizal fungi, and fungal endophytes. Oikos, 117, 310–320.CrossRefGoogle Scholar
Mack, R. N., Simberloff, D., Lonsdale, W. M., Evans, H., Clout, M. & Bazzaz, F. A. (2000). Biotic invasions: causes, epidemiology, global consequences, and control. Ecological Applications, 10, 689–710.CrossRefGoogle 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, 516–527.CrossRefGoogle ScholarPubMed
MacNeil, C., Dick, J. T. A., Hatcher, M. J. & Dunn, A. M. (2003a). Differential drift and parasitism in invading and native Gammarus spp. (Crustacea: Amphipoda). Ecography, 26, 467–473.CrossRefGoogle Scholar
MacNeil, C., Dick, J. T. A., Hatcher, M. J., Fielding, N. J., Hume, K. D. & Dunn, A. M. (2003b). Parasite transmission and cannibalism in an amphipod (Crustacea). International Journal for Parasitology, 33, 795–798.CrossRefGoogle Scholar
MacNeil, C., Dick, J. T. A., Hatcher, M. J., Terry, R. S., Smith, J. E. & Dunn, A. M. (2003c). Parasite-mediated predation between native and invasive amphipods. Proceedings of the Royal Society of London Series B – Biological Sciences, 270, 1309–1314.CrossRefGoogle ScholarPubMed
MacNeil, C., Fielding, N. J., Dick, J. T. A., Briffa, M., Prenter, J., Hatcher, M. J. & Dunn, A. M. (2003d). An acanthocephalan parasite mediates intraguild predation between invasive and native freshwater amphipods (Crustacea). Freshwater Biology, 48, 2085–2093.CrossRefGoogle Scholar
MacNeil, C., Fielding, N. J., Hume, K. D., Dick, J. T. A., Elwood, R. W., Hatcher, M. J. & Dunn, A. M. (2003e). Parasite altered micro-distribution of Gammarus pulex (Crustacea: Amphipoda). International Journal for Parasitology, 33, 57–64.CrossRefGoogle Scholar
Malmstrom, C. M., Hughes, C. C., Newton, L. A. & Stoner, C. J. (2005a). Virus infection in remnant native bunchgrasses from invaded California grasslands. New Phytologist, 168, 217–230.CrossRefGoogle ScholarPubMed
Malmstrom, C. M., McCullough, A. J., Johnson, H. A., Newton, L. A. & Borer, E. T. (2005b). Invasive annual grasses indirectly increase virus incidence in California native perennial bunchgrasses. Oecologia, 145, 153–164.CrossRefGoogle ScholarPubMed
Malmstrom, C. M., Stoner, C. J., Brandenburg, S. & Newton, L. A. (2006). Virus infection and grazing exert counteracting influences on survivorship of native bunchgrass seedlings competing with invasive exotics. Journal of Ecology, 94, 264–275.CrossRefGoogle ScholarPubMed
March, W. A. & Watson, D. M. (2007). Parasites boost productivity: effects of mistletoe on litterfall dynamics in a temperate Australian forest. Oecologia, 154, 339–347.CrossRefGoogle Scholar
Marcogliese, D. J. & Cone, D. K. (1997). Food webs: a plea for parasites. Trends in Ecology & Evolution, 12, 394–394.CrossRefGoogle ScholarPubMed
Marr, S. R., Mautz, W. J. & Hara, A. H. (2008). Parasite loss and introduced species: a comparison of the parasites of the Puerto Rican tree frog (Eleutherodactylus coqui) in its native and introduced ranges. Biological Invasions, 10, 1289–1298.CrossRefGoogle Scholar
Marra, P. P., Griffing, S., Caffrey, C., Kilpatrick, A. M., McLean, R., Brand, C., Saito, E., Dupuis, A. P., Kramer, L. & Novak, R. (2004). West Nile virus and wildlife. Bioscience, 54, 393–402.CrossRefGoogle Scholar
Marshall, K., Mamone, M. & Barclay, R. (2003). A survey of Douglas-fir dwarf mistletoe brooms used for nests by northern spotted owls on the Applegate Ranger District and Ashland Resource area in southwest Oregon. Western Journal of Applied Forestry, 18, 115–117.Google Scholar
Martens, P., Kovats, R. S., Nijhof, S., Vries, P., Livermore, M. T. J., Bradley, D. J., Cox, J. & McMichael, A. J. (1999). Climate change and future populations at risk of malaria. Global Environmental Change: Human and Policy Dimensions, 9, S89–S107.CrossRefGoogle Scholar
Martens, W. J. M., Jetten, T. H. & Focks, D. A. (1997). Sensitivity of malaria, schistosomiasis and dengue to global warming. Climatic Change, 35, 145–156.CrossRefGoogle Scholar
Massey, R. C., Buckling, A. & French-Constant, R. (2004). Interference competition and parasite virulence. Proceedings of the Royal Society of London Series B – Biological Sciences, 271, 785–788.CrossRefGoogle ScholarPubMed
Mathews, F. (2009). Zoonoses in wildlife: integrating ecology into management. Advances in Parasitology, 68, 185–209.CrossRefGoogle ScholarPubMed
Mathiasen, R. L., Nickrent, D. L., Shaw, D. C. & Watson, D. M. (2008). Mistletoes: pathology, systematics, ecology, and management. Plant Disease, 92, 988–1006.CrossRefGoogle Scholar
Matson, K. D. (2006). Are there differences in immune function between continental and insular birds?Proceedings of the Royal Society of London Series B – Biological Sciences, 273, 2267–2274.CrossRefGoogle ScholarPubMed
Matthies, D. (1995). Parasitic and competitive interactions between the hemiparasites Rhinanthus serotinus and Odontites rubra and their host Medicago sativa. Journal of Ecology, 83, 245–251.CrossRefGoogle Scholar
May, R. M. (1973). Stability and complexity in model ecosystems. Monographs in Population Biology, 6, 1–235.Google ScholarPubMed
May, R. M. & Anderson, R. M. (1987). Transmission dynamics of HIV infection. Nature, 326, 137–142.CrossRefGoogle ScholarPubMed
May, R. M. & Anderson, R. M. (1990). Parasite–host coevolution. Parasitology, 100, S89–S101.CrossRefGoogle ScholarPubMed
May, R. M., Gupta, S. & McLean, A. R. (2001). Infectious disease dynamics: what characterizes a successful invader?Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 356, 901–910.CrossRefGoogle ScholarPubMed
McCallum, H., Barlow, N. & Hone, J. (2001). How should pathogen transmission be modelled?Trends in Ecology & Evolution, 16, 295–300.CrossRefGoogle ScholarPubMed
McCallum, H., Gerber, L. & Jani, A. (2005). Does infectious disease influence the efficacy of marine protected areas? A theoretical framework. Journal of Applied Ecology, 42, 688–698.CrossRefGoogle Scholar
McCallum, H., Tompkins, D. M., Jones, M., Lachish, S., Marvanek, S., Lazenby, B., Hocking, G., Wiersma, J. & Hawkins, C. E. (2007). Distribution and impacts of Tasmanian devil facial tumor disease. Ecohealth, 4, 318–325.CrossRefGoogle Scholar
McCann, K., Hastings, A. & Huxel, G. R. (1998). Weak trophic interactions and the balance of nature. Nature, 395, 794–798.CrossRefGoogle Scholar
McCann, K. S. (2000). The diversity–stability debate. Nature, 405, 228–233.CrossRefGoogle ScholarPubMed
McGeoch, M. A., Butchart, S. H. M., Spear, D., Marais, E., Kleynhans, E. J., Symes, A., Chanson, J. & Hoffmann, M. (2010). Global indicators of biological invasion: species numbers, biodiversity impact and policy responses. Diversity and Distributions, 16, 95–108.CrossRefGoogle Scholar
McMichael, A. J., Bouma, M. J., Mooney, H. A. & Hobbs, R. J. (2000). Global changes, invasive species, and human health. In Mooney, H. A. & Hobbs, R. J. (eds), Invasive Species in a Changing World. Washington, DC: Island Press, pp. 191–240.Google Scholar
Medoc, V. & Beisel, J. N. (2008). An acanthocephalan parasite boosts the escape performance of its intermediate host facing non-host predators. Parasitology, 135, 977–984.CrossRefGoogle ScholarPubMed
Medoc, V., Bollache, L. & Beisel, J. N. (2006). Host manipulation of a freshwater crustacean (Gammarus roeseli) by an acanthocephalan parasite (Polymorphus minutus) in a biological invasion context. International Journal for Parasitology, 36, 1351–1358.CrossRefGoogle Scholar
Memmott, J., Fowler, S. V., Paynter, Q., Sheppard, A. W. & Syrett, P. (2000). The invertebrate fauna on broom, Cytisus scoparius, in two native and two exotic habitats. Acta Oecologica: International Journal of Ecology, 21, 213–222.CrossRefGoogle Scholar
Meyling, N. V. & Hajek, A. E. (2010). Principles from community and metapopulation ecology: application to fungal entomopathogens. Biocontrol, 55, 39–54.CrossRefGoogle Scholar
Michaelis, M., Doerr, H. W. & Cinatl, J. (2009). Novel swine-origin influenza A virus in humans: another pandemic knocking at the door. Medical Microbiology and Immunology, 198, 175–183.CrossRefGoogle Scholar
Milinski, M. (1985). Risk of predation of parasitized sticklebacks (Gasterosteus aculeatus L.) under competition for food. Behaviour, 93, 203–215.CrossRefGoogle Scholar
Mills, N. J. & Gutierrez, A. P. (1996). Prospective modelling in biological control: an analysis of the dynamics of heteronomous hyperparasitism in a cotton–whitefly–parasitoid system. Journal of Applied Ecology, 33, 1379–1394.CrossRefGoogle Scholar
Minchella, D. J. & Scott, M. E. (1991). Parasitism: a cryptic determinant of animal community structure. Trends in Ecology & Evolution, 6, 250–254.CrossRefGoogle ScholarPubMed
Mitchell, C. E. & Power, A. G. (2003). Release of invasive plants from fungal and viral pathogens. Nature, 421, 625–627.CrossRefGoogle ScholarPubMed
Mitchell, C. E., Tilman, D. & Groth, J. V. (2002). Effects of grassland plant species diversity, abundance, and composition on foliar fungal disease. Ecology, 83, 1713–1726.CrossRefGoogle Scholar
Mittelbach, G. G. (2005). Parasites, communities, and ecosystems: conclusions and perspectives. In Thomas, F., Renaud, F. & Geugan, J. F. (eds) Parasitism and Ecosystems, Oxford: Oxford University Press, pp. 171–176.CrossRefGoogle Scholar
Moleon, M., Almaraz, P. & Sanchez-Zapata, J. A. (2008). An emerging infectious disease triggering large-scale hyperpredation. PLoS One, 3, e2307.CrossRefGoogle ScholarPubMed
Moller, A. P. (2008). Interactions between interactions: predator–prey, parasite–host, and mutualistic interactions. Annals of the New York Academy of Science, 1133, 180–186.CrossRefGoogle ScholarPubMed
Molofsky, J., Bever, J. D. & Antonovics, J. (2001). Coexistence under positive frequency dependence. Proceedings of the Royal Society of London Series B – Biological Sciences, 268, 273–277.CrossRefGoogle ScholarPubMed
Molofsky, J., Bever, J. D., Antonovics, J. & Newman, T. J. (2002). Negative frequency dependence and the importance of spatial scale. Ecology, 83, 21–27.CrossRefGoogle Scholar
Money, N. P. (2007). Triumph of the Fungi: A Rotten History. Oxford: Oxford University Press.Google Scholar
Moore, J. (2002). Parasites and the Behavior of Animals. Oxford: Oxford University Press.Google Scholar
Moore, S. M., Borer, E. T. & Hosseini, P. R. (2010). Predators indirectly control vector-borne disease: linking predator–prey and host–pathogen models. Journal of the Royal Society Interface, 7, 161–176.CrossRefGoogle ScholarPubMed
Morens, D. M., Folkers, G. K. & Fauci, A. S. (2008). Emerging infections: a perpetual challenge. Lancet Infectious Diseases, 8, 710–719.CrossRefGoogle ScholarPubMed
Moret, Y., Bollache, L., Wattier, R. & Rigaud, T. (2007). Is the host or the parasite the most locally adapted in an amphipod–acanthocephalan relationship? A case study in a biological invasion context. International Journal for Parasitology, 37, 637–644.CrossRefGoogle ScholarPubMed
Morris, R. J., Lewis, O. T. & Godfray, H. C. J. (2004). Experimental evidence for apparent competition in a tropical forest food web. Nature, 428, 310–313.CrossRefGoogle Scholar
Morris, R. J., Muller, C. B. & Godfray, H. C. J. (2001). Field experiments testing for apparent competition between primary parasitoids mediated by secondary parasitoids. Journal of Animal Ecology, 70, 301–309.CrossRefGoogle Scholar
Morris, W. F., Hufbauer, R. A., Agrawal, A. A., Bever, J. D., Borowicz, V. A., Gilbert, G. S., Maron, J. L., Mitchell, C. E., Parker, I. M., Power, A. G., Torchin, M. E. & Vazquez, D. P. (2007). Direct and interactive effects of enemies and mutualists on plant performance: a meta-analysis. Ecology, 88, 1021–1029.CrossRefGoogle ScholarPubMed
Morrison, L. W. (1999). Indirect effects of phorid fly parasitoids on the mechanisms of interspecific competition among ants. Oecologia, 121, 113–122.CrossRefGoogle ScholarPubMed
Mouillot, D., Krasnov, B. R. & Poulin, R. (2008a). High intervality explained by phylogenetic constraints in host–parasite webs. Ecology, 89, 2043–2051.CrossRefGoogle ScholarPubMed
Mouillot, D., Krasnov, B. R., Shenbrot, G. I. & Poulin, R. (2008b). Connectance and parasite diet breadth in flea–mammal webs. Ecography, 31, 16–20.CrossRefGoogle Scholar
Mouritsen, K. N. (2001). Hitch-hiking parasite: a dark horse may be the real rider. International Journal for Parasitology, 31, 1417–1420.CrossRefGoogle ScholarPubMed
Mouritsen, K. N. & Haun, S. C. B. (2008). Community regulation by herbivore parasitism and density: trait-mediated indirect interactions in the intertidal. Journal of Experimental Marine Biology and Ecology, 367, 236–246.CrossRefGoogle Scholar
Mouritsen, K. N. & Jensen, T. (2006). The effect of Sacculina carcini infections on the fouling, burying behaviour and condition of the shore crab, Carcinus maenas. Marine Biology Research, 2, 270–275.CrossRefGoogle Scholar
Mouritsen, K. N., Mouritsen, L. T. & Jensen, K. T. (1998). Change of topography and sediment characteristics on an intertidal mud-flat following mass-mortality of the amphipod Corophium volutator. Journal of the Marine Biological Association of the United Kingdom, 78, 1167–1180.CrossRefGoogle Scholar
Mouritsen, K. N. & Polml, R. (2006). A parasite indirectly impacts both abundance of primary producers and biomass of secondary producers in an intertidal benthic community. Journal of the Marine Biological Association of the United Kingdom, 86, 221–226.CrossRefGoogle Scholar
Mouritsen, K. N. & Poulin, R. (2003). Parasite-induced trophic facilitation exploited by a non-host predator: a manipulator's nightmare. International Journal for Parasitology, 33, 1043–1050.CrossRefGoogle ScholarPubMed
Mouritsen, K. N. & Poulin, R. (2005). Parasitism can influence the intertidal zonation of non-host organisms. Marine Biology, 148, 1–11.CrossRefGoogle Scholar
Mueller, R. C. & Gehring, C. A. (2006). Interactions between an above-ground plant parasite and below-ground ectomycorrhizal fungal communities on pinyon pine. Journal of Ecology, 94, 276–284.CrossRefGoogle Scholar
Muller, C. B. & Brodeur, J. (2002). Intraguild predation in biological control and conservation biology. Biological Control, 25, 216–223.CrossRefGoogle 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, 250–264.CrossRefGoogle Scholar
Musselman, L. J. & Press, M. C. (1995). Introduction to parasitic plants. In Press, M. C. & Graves, J. D. (eds), Parasitic Plants. London: Chapman & Hall, pp. 1–13.Google Scholar
Myers, J. H. & Bazely, D. R. (2003). Ecology and Control of Introduced Plants. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Mylius, S. D., Klumpers, K., Roos, A. M. & Persson, L. (2001). Impact of intraguild predation and stage structure on simple communities along a productivity gradient. American Naturalist, 158, 259–276.CrossRefGoogle ScholarPubMed
Naugle, D. E., Aldridge, C. L., Walker, B. L., Cornish, T. E., Moynahan, B. J., Holloran, M. J., Brown, K., Johnson, G. D., Schmidtmann, E. T., Mayer, R. T., Kato, C. Y., Matchett, M. R., Christiansen, T. J., Cook, W. E., Creekmore, T., Falise, R. D., Rinkes, E. T. & Boyce, M. S. (2004). West Nile virus: pending crisis for greater sage-grouse. Ecology Letters, 7, 704–713.CrossRefGoogle Scholar
Nel, L. H. & Rupprecht, C. E. (2007). Emergence of Lyssaviruses in the old world: the case of Africa. In Childs, J. E., Mackenzie, J. S. & Richt, J. A. (eds), Wildlife and Emerging Zoonotic Diseases: The Biology, Circumstances and Consequences of Cross-species Transmission. Berlin: Springer, pp. 161–193.CrossRefGoogle Scholar
Neutel, A. M., Heesterbeek, J. A. P. & Ruiter, P. C. (2002). Stability in real food webs: weak links in long loops. Science, 296, 1120–1123.CrossRefGoogle ScholarPubMed
New, T. R. (2005). Invertebrate Conservation and Agricultural Ecosystems. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Nickrent, D. L., Malecot, V., Vidal-Russell, R. & Der, J. P. (2010). A revised classification of Santalales. Taxon, 59, 538–558.Google Scholar
Nogales, M., Martin, A., Tershy, B. R., Donlan, C. J., Witch, D., Puerta, N., Wood, B. & Alonso, J. (2004). A review of feral cat eradication on islands. Conservation Biology, 18, 310–319.CrossRefGoogle Scholar
Norman, R., Bowers, R. G., Begon, M. & Hudson, P. J. (1999). Persistence of tick-borne virus in the presence of multiple host species: tick reservoirs and parasite mediated competition. Journal of Theoretical Biology, 200, 111–118.CrossRefGoogle ScholarPubMed
Norton, D. A. & Reid, N. (1997). Lessons in ecosystem management from management of threatened and pest loranthaceous mistletoes in New Zealand and Australia. Conservation Biology, 11, 759–769.CrossRefGoogle Scholar
Novembre, J., Galvani, A. P. & Slatkin, M. (2005). The geographic spread of the CCR5 Delta32 HIV-resistance allele. PLoS Biology, 3, e339.CrossRefGoogle ScholarPubMed
Odum, E. P. (1953). Fundamentals of Ecology. Philadelphia: Saunders.Google Scholar
Oidtmann, B., El-Matbouli, M., Fischer, H., Hoffmann, R., Klarding, K., Schmid, I. & Schmidt, R. (1997). Light Microscopy of Astacus astacus L. Under Normal and Selected Pathological Conditions, with Special Emphasis to Porcelain Disease and Crayfish Plague. Lafayette: International Association of Astacology.Google Scholar
Oksanen, L., Fretwell, S. D., Arruda, J. & Niemela, P. (1981). Exploitation ecosystems in gradients of primary productivity. American Naturalist, 118, 240–261.CrossRefGoogle Scholar
Olff, H., Hoorens, B., Goede, R. G. M., Putten, W. H. & Gleichman, J. M. (2000). Small-scale shifting mosaics of two dominant grassland species: the possible role of soil-borne pathogens. Oecologia, 125, 45–54.CrossRefGoogle ScholarPubMed
Oliveira, N. M. & Hilker, F. M. (2010). Modelling disease introduction as biological control of invasive predators to preserve endangered prey. Bulletin of Mathematical Biology, 72, 444–468.CrossRefGoogle ScholarPubMed
Olsen, B., Munster, V. J., Wallensten, A., Waldenstrom, J., Osterhaus, A. & Fouchier, R. A. M. (2006). Global patterns of influenza A virus in wild birds. Science, 312, 384–388.CrossRefGoogle ScholarPubMed
Omacini, M., Chaneton, E. J., Ghersa, C. M. & Muller, C. B. (2001). Symbiotic fungal endophytes control insect host–parasite interaction webs. Nature, 409, 78–81.CrossRefGoogle ScholarPubMed
Omacini, M., Chaneton, E. J., Ghersa, C. M. & Otero, P. (2004). Do foliar endophytes affect grass litter decomposition? A microcosm approach using Lolium multiflorum. Oikos, 104, 581–590.CrossRefGoogle Scholar
Omacini, M., Eggers, T., Bonkowski, M., Gange, A. C. & Jones, T. H. (2006). Leaf endophytes affect mycorrhizal status and growth of co-infected and neighbouring plants. Functional Ecology, 20, 226–232.CrossRefGoogle Scholar
Onstad, D. W. (1992). Evaluation of epidemiological thresholds and asymptotes with variable plant densities. Phytopathology, 82, 1028–1032.CrossRefGoogle Scholar
Onstad, D. W. & Kornkven, E. A. (1992). Persistence and endemicity of pathogens in plant populations over time and space. Phytopathology, 82, 561–566.CrossRefGoogle Scholar
Orr, M. R., Camargo, R. X. & Benson, W. W. (2003). Interactions between ant species increase arrival rates of an ant parasitoid. Animal Behaviour, 65, 1187–1193.CrossRefGoogle Scholar
Ostfeld, R. S. (2009a). Biodiversity loss and the rise of zoonotic pathogens. Clinical Microbiology and Infection, 15, 40–43.CrossRefGoogle ScholarPubMed
Ostfeld, R. S. (2009b). Climate change and the distribution and intensity of infectious diseases. Ecology, 90, 903–905.CrossRefGoogle ScholarPubMed
Ostfeld, R. S., Canham, C. D., Oggenfuss, K., Winchcombe, R. J. & Keesing, F. (2006). Climate, deer, rodents, and acorns as determinants of variation in Lyme-disease risk. PloS Biology, 4, 1058–1068.CrossRefGoogle ScholarPubMed
Ostfeld, R. S., Glass, G. E. & Keesing, F. (2005). Spatial epidemiology: an emerging (or re-emerging) discipline. Trends in Ecology & Evolution, 20, 328–336.CrossRefGoogle ScholarPubMed
Ostfeld, R. S. & Holt, R. D. (2004). Are predators good for your health? Evaluating evidence for top-down regulation of zoonotic disease reservoirs. Frontiers in Ecology and the Environment, 2, 13–20.CrossRefGoogle Scholar
Ostfeld, R. S. & Keesing, F. (2000). Biodiversity and disease risk: the case of Lyme disease. Conservation Biology, 14, 722–728.CrossRefGoogle Scholar
Ostfeld, R. S. & LoGiudice, K. (2003). Community disassembly, biodiversity loss, and the erosion of an ecosystem service. Ecology, 84, 1421–1427.CrossRefGoogle Scholar
Otterstatter, M. C. & Thomson, J. D. (2008). Does pathogen spillover from commercially reared bumble bees threaten wild pollinators?PLoS One, 3, e2771.CrossRefGoogle ScholarPubMed
Ouellet, M., Mikaelian, I., Pauli, B. D., Rodrigue, J. & Green, D. M. (2005). Historical evidence of widespread chytrid infection in North American amphibian populations. Conservation Biology, 19, 1431–1440.CrossRefGoogle Scholar
Paaijmans, K. P., Read, A. F. & Thomas, M. B. (2009). Understanding the link between malaria risk and climate. Proceedings of the National Academy of Sciences of the United States of America, 106, 13844–13849.CrossRefGoogle ScholarPubMed
Packer, A. & Clay, K. (2000). Soil pathogens and spatial patterns of seedling mortality in a temperate tree. Nature, 404, 278–281.CrossRefGoogle Scholar
Packer, C., Altizer, S., Appel, M., Brown, E., Martenson, J., O'Brien, S. J., Roelke-Parker, M., Hofmann-Lehmann, R. & Lutz, H. (1999). Viruses of the Serengeti: patterns of infection and mortality in African lions. Journal of Animal Ecology, 68, 1161–1178.CrossRefGoogle Scholar
Packer, C., Holt, R. D., Hudson, P. J., Lafferty, K. D. & Dobson, A. P. (2003). Keeping the herds healthy and alert: implications of predator control for infectious disease. Ecology Letters, 6, 797–802.CrossRefGoogle Scholar
Paddock, C. D. & Yabsley, M. J. (2007). Ecological havoc, the rise of white-tailed deer, and the emergence of Amblyomma americanum – associated zoonoses in the United States. Current Topics in Microbiology and Immunology, 315, 289–324.Google ScholarPubMed
Park, T. (1948). Experimental studies of interspecies competition: 1 – competition between populations of the flour beetles, Tribolium confusum Duval and Tribolium castaneum Herbst. Ecological Monographs, 18, 265–307.CrossRefGoogle Scholar
Parker, C. (2009). Observations on the current status of Orobanche and Striga problems worldwide. Pest Management Science, 65, 453–459.CrossRefGoogle ScholarPubMed
Parker, G. A., Chubb, J. C., Ball, M. A. & Roberts, G. N. (2003). Evolution of complex life cycles in helminth parasites. Nature, 425, 480–484.CrossRefGoogle ScholarPubMed
Parrish, C. R., Holmes, E. C., Morens, D. M., Park, E. C., Burke, D. S., Calisher, C. H., Laughlin, C. A., Saif, L. J. & Daszak, P. (2008). Cross-species virus transmission and the emergence of new epidemic diseases. Microbiology and Molecular Biology Reviews, 72, 457–470.CrossRefGoogle ScholarPubMed
Pascual, M., Ahumada, J. A., Chaves, L. F., Rodo, X. & Bouma, M. (2006). Malaria resurgence in the East African highlands: temperature trends revisited. Proceedings of the National Academy of Sciences of the United States of America, 103, 5829–5834.CrossRefGoogle ScholarPubMed
Pascual, M. & Bouma, M. J. (2009). Do rising temperatures matter?Ecology, 90, 906–912.CrossRefGoogle ScholarPubMed
Pasternak, Z., Diamant, A. & Abelson, A. (2007). Co-invasion of a Red Sea fish and its ectoparasitic monogenean, Polylabris cf. mamaevi into the Mediterranean: observations on oncomiracidium behavior and infection levels in both seas. Parasitology Research, 100, 721–727.CrossRefGoogle ScholarPubMed
Pattison, J. (1998). The emergence of bovine spongiform encephalopathy and related diseases. Emerging Infectious Diseases, 4, 390–394.CrossRefGoogle ScholarPubMed
Paul, N. D., Hatcher, P. E. & Taylor, J. E. (2000). Coping with multiple enemies: an integration of molecular and ecological perspectives. Trends in Plant Science, 5, 220–225.CrossRefGoogle ScholarPubMed
Pedersen, A. B. & Fenton, A. (2007). Emphasizing the ecology in parasite community ecology. Trends in Ecology & Evolution, 22, 133–139.CrossRefGoogle ScholarPubMed
Pedersen, A. B., Jones, K. E., Nunn, C. L. & Altizer, S. (2007). Infectious diseases and extinction risk in wild mammals. Conservation Biology, 21, 1269–1279.CrossRefGoogle ScholarPubMed
Pell, J. K., Hannam, J. J. & Steinkraus, D. C. (2010). Conservation biological control using fungal entomopathogens. Biocontrol, 55, 187–198.CrossRefGoogle Scholar
Pennings, S. C. & Callaway, R. M. (1996). Impact of a parasitic plant on the structure and dynamics of salt marsh vegetation. Ecology, 77, 1410–1419.CrossRefGoogle Scholar
Pennings, S. C. & Callaway, R. M. (2002). Parasitic plants: parallels and contrasts with herbivores. Oecologia, 131, 479–489.CrossRefGoogle ScholarPubMed
Petermann, J. S., Fergus, A. J. F., Turnbull, L. A. & Schmid, B. (2008). Janzen–Connell effects are widespread and strong enough to maintain diversity in grasslands. Ecology, 89, 2399–2406.CrossRefGoogle ScholarPubMed
Peterson, R. O., Thomas, N. J., Thurber, J. M., Vucetich, J. A. & Waite, T. A. (1998). Population limitation and the wolves of Isle Royale. Journal of Mammalogy, 79, 828–841.CrossRefGoogle Scholar
Philpott, S. M. (2005). Trait-mediated effects of parasitic phorid flies (Diptera: Phoridae) on ant (Hymenoptera: Formicidae) competition and resource access in coffee agro-ecosystems. Environmental Entomology, 34, 1089–1094.CrossRefGoogle Scholar
Phoenix, G. K. & Press, M. C. (2005). Linking physiological traits to impacts on community structure and function: the role of root hemiparasitic Orobanchaceae (ex-Scrophulariaceae). Journal of Ecology, 93, 67–78.CrossRefGoogle Scholar
Pieterse, C. M. J., Wees, S. C. M., Ton, J., Pelt, J. A. & Loon, L. C. (2002). Signalling in rhizobacteria-induced systemic resistance in Arabidopsis thaliana. Plant Biology, 4, 535–544.CrossRefGoogle Scholar
Pimentel, D., Lach, L., Zuniga, R. & Morrison, D. (2000). Environmental and economic costs of nonindigenous species in the United States. Bioscience, 50, 53–65.CrossRefGoogle Scholar
Pimm, S. L. (1982). Food Webs. London: Chapman & Hall.CrossRefGoogle Scholar
Pimm, S. L. & Lawton, J. H. (1978). Feeding on more than one trophic level. Nature, 275, 542–544.CrossRefGoogle Scholar
Pinzon, J. E., Wilson, J. M., Tucker, C. J., Arthur, R., Jahrling, P. B. & Formenty, P. (2004). Trigger events: enviroclimatic coupling of Ebola hemorrhagic fever outbreaks. American Journal of Tropical Medicine and Hygiene, 71, 664–674.Google ScholarPubMed
Plantier, J. C., Leoz, M., Dickerson, J. E., Oliveira, F., Cordonnier, F., Lemee, V., Damond, F., Robertson, D. L. & Simon, F. (2009). A new human immunodeficiency virus derived from gorillas. Nature Medicine, 15, 871–872.CrossRefGoogle ScholarPubMed
Ploetz, R. C. (1994). Panama disease: return of the first banana menace. International Journal of Pest Management, 40, 326–336.CrossRefGoogle Scholar
Polis, G. A. & Holt, R. D. (1992). Intraguild predation: the dynamics of complex trophic interactions. Trends in Ecology & Evolution, 7, 151–154.CrossRefGoogle ScholarPubMed
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, 297–330.CrossRefGoogle Scholar
Poulin, R. (1995). ‘Adaptive’ changes in the behaviour of parasitized animals: a critical review. International Journal for Parasitology, 25, 1371–1383.CrossRefGoogle ScholarPubMed
Poulin, R. (1997). Species richness of parasite assemblages: evolution and patterns. Annual Review of Ecology and Systematics, 28, 341–358.CrossRefGoogle Scholar
Poulin, R. (1999). The functional importance of parasites in animal communities: many roles at many levels?International Journal for Parasitology, 29, 903–914.CrossRefGoogle ScholarPubMed
Poulin, R. & Mouritsen, K. N. (2006). Climate change, parasitism and the structure of intertidal ecosystems. Journal of Helminthology, 80, 183–191.CrossRefGoogle ScholarPubMed
Poulin, R., Nichol, K. & Latham, A. D. A. (2003). Host sharing and host manipulation by larval helminths in shore crabs: cooperation or conflict?International Journal for Parasitology, 33, 425–433.CrossRefGoogle ScholarPubMed
Pounds, J. A., Bustamante, M. R., Coloma, L. A., Consuegra, J. A., Fogden, M. P. L., Foster, P. N., Marca, E., Masters, K. L., Merino-Viteri, A., Puschendorf, R., Ron, S. R., Sanchez-Azofeifa, G. A., Still, C. J. & Young, B. E. (2006). Widespread amphibian extinctions from epidemic disease driven by global warming. Nature, 439, 161–167.CrossRefGoogle ScholarPubMed
Power, A. G. & Mitchell, C. E. (2004). Pathogen spillover in disease epidemics. American Naturalist, 164, S79–S89.CrossRefGoogle ScholarPubMed
Prenter, J., MacNeil, C., Dick, J. T. A. & Dunn, A. M. (2004). Roles of parasites in animal invasions. Trends in Ecology & Evolution, 19, 385–390.CrossRefGoogle ScholarPubMed
Press, M. C., Nour, J. J., Bebawi, F. F. & Stewart, G. R. (1989). Antitranspirant-induced heat-stress in the parasitic plant Striga hermonthica – a novel method of control. Journal of Experimental Botany, 40, 585–591.CrossRefGoogle Scholar
Press, M. C. & Phoenix, G. K. (2005). Impacts of parasitic plants on natural communities. New Phytologist, 166, 737–751.CrossRefGoogle ScholarPubMed
Price, P. W., Bouton, C. E., Gross, P., McPheron, B. A., Thompson, J. N. & Weis, A. E. (1980). Interactions among three trophic levels – influence of plants on interactions between insect herbivores and natural enemies. Annual Review of Ecology and Systematics, 11, 41–65.CrossRefGoogle Scholar
Price, P. W., Westoby, M., Rice, B., Atsatt, P. R., Fritz, R. S., Thompson, J. N. & Mobley, K. (1986). Parasite mediation in ecological interactions. Annual Review of Ecology and Systematics, 17, 487–505.CrossRefGoogle Scholar
Prins, H. H. T. & Vanderjeugd, H. P. (1993). Herbivore population crashes and woodland structure in East Africa. Journal of Ecology, 81, 305–314.CrossRefGoogle Scholar
Prins, H. H. T. & Weyerhaeuser, F. J. (1987). Epidemics in populations of wild ruminants: anthrax and impala, rinderpest and buffalo in Lake Manyara national park, Tanzania. Oikos, 49, 28–38.CrossRefGoogle Scholar
Pywell, R. F., Bullock, J. M., Walker, K. J., Coulson, S. J., Gregory, S. J. & Stevenson, M. J. (2004). Facilitating grassland diversification using the hemiparasitic plant Rhinanthus minor. Journal of Applied Ecology, 41, 880–887.CrossRefGoogle Scholar
Quested, H. M. (2008). Parasitic plants: impacts on nutrient cycling. Plant and Soil, 311, 269–272.CrossRefGoogle Scholar
Rand, T. A. & Louda, S. A. (2006). Spillover of agriculturally subsidized predators as a potential threat to native insect herbivores in fragmented landscapes. Conservation Biology, 20, 1720–1729.CrossRefGoogle ScholarPubMed
Randolph, S. E. (2009). Perspectives of climate change impacts on infectious diseases. Ecology, 90, 927–931.CrossRefGoogle ScholarPubMed
Rasmann, S., Kollner, T. G., Degenhardt, J., Hiltpold, I., Toepfer, S., Kuhlmann, U., Gershenzon, J. & Turlings, T. C. J. (2005). Recruitment of entomopathogenic nematodes by insect-damaged maize roots. Nature, 434, 732–737.CrossRefGoogle ScholarPubMed
Rasmann, S. & Turlings, T. C. J. (2007). Simultaneous feeding by aboveground and belowground herbivores attenuates plant-mediated attraction of their respective natural enemies. Ecology Letters, 10, 926–936.CrossRefGoogle ScholarPubMed
Ray, C. & Collinge, S. K. (2006). Potential effects of a keystone species on the dynamics of sylvatic plague. In Collinge, S. K. & Ray, C. (eds), Disease Ecology: Community Structure and Pathogen Dynamics. Oxford:Oxford University Press, pp. 202–216.CrossRefGoogle Scholar
Read, A. F. & Taylor, L. H. (2001). The ecology of genetically diverse infections. Science, 292, 1099–1102.CrossRefGoogle ScholarPubMed
Reinhart, K. O. & Callaway, R. M. (2006). Soil biota and invasive plants. New Phytologist, 170, 445–457.CrossRefGoogle ScholarPubMed
Reinhart, K. O., Packer, A., Putten, W. H. & Clay, K. (2003). Plant–soil biota interactions and spatial distribution of black cherry in its native and invasive ranges. Ecology Letters, 6, 1046–1050.CrossRefGoogle Scholar
Reiter, P. (2010). West Nile virus in Europe: understanding the present to gauge the future. Eurosurveillance, 15, 19508.Google ScholarPubMed
Rejmanek, M. & Richardson, D. M. (1996). What attributes make some plant species more invasive?Ecology, 77, 1655–1661.CrossRefGoogle Scholar
Reynolds, H. L., Packer, A., Bever, J. D. & Clay, K. (2003). Grassroots ecology: plant–microbe–soil interactions as drivers of plant community structure and dynamics. Ecology, 84, 2281–2291.CrossRefGoogle Scholar
Reynolds, J. C. (1985). Details of the geographic replacement of the red squirrel (Sciurus vulgaris) by the grey squirrel (Sciurus carolinensis) in eastern England. Journal of Animal Ecology, 54, 149–162.CrossRefGoogle Scholar
Rigaud, T. & Moret, Y. (2003). Differential phenoloxidase activity between native and invasive gammarids infected by local acanthocephalans: differential immunosuppression?Parasitology, 127, 571–577.CrossRefGoogle ScholarPubMed
Riley, S., Fraser, C., Donnelly, C. A., Ghani, A. C., Abu-Raddad, L. J., Hedley, A. J., Leung, G. M., Ho, L. M., Lam, T. H., Thach, T. Q., Chau, P., Chan, K. P., Leung, P. Y., Tsang, T., Ho, W., Lee, K. H., Lau, E. M. C., Ferguson, N. M. & Anderson, R. M. (2003). Transmission dynamics of the etiological agent of SARS in Hong Kong: impact of public health interventions. Science, 300, 1961–1966.CrossRefGoogle Scholar
Roberts, M. G. (1995). A pocket guide to host–parasite models. Parasitology Today, 11, 172–177.CrossRefGoogle ScholarPubMed
Roelke-Parker, M. E., Munson, L., Packer, C., Kock, R., Cleaveland, S., Carpenter, M., Obrien, S. J., Pospischil, A., Hofmann-Lehmann, R., Lutz, H., Mwamengele, G. L. M., Mgasa, M. N., Machange, G. A., Summers, B. A. & Appel, M. J. G. (1996). A canine distemper virus epidemic in Serengeti lions (Panthera leo). Nature, 379, 441–445.CrossRefGoogle Scholar
Rogers, D. J. & Randolph, S. E. (2000). The global spread of malaria in a future, warmer world. Science, 289, 2283–2284.CrossRefGoogle Scholar
Rohde, K. (1998). Is there a fixed number of niches for endoparasites of fish?International Journal for Parasitology, 28, 1861–1865.CrossRefGoogle Scholar
Rohlfs, M. (2008). Host–parasitoid interaction as affected by interkingdom competition. Oecologia, 155, 161–168.CrossRefGoogle ScholarPubMed
Rohr, J. R., Raffel, T. R., Romansic, J. M., McCallum, H. & Hudson, P. J. (2008). Evaluating the links between climate, disease spread, and amphibian declines. Proceedings of the National Academy of Sciences of the United States of America, 105, 17436–17441.CrossRefGoogle ScholarPubMed
Rosenblum, E. B., Voyles, J., Poortne, T. J. & Stajich, J. E. (2010). The deadly chytrid fungus: a story of an emerging pathogen. PLoS Pathogens, 6, e1000550.CrossRefGoogle ScholarPubMed
Rosenheim, J. A. (1998). Higher-order predators and the regulation of insect herbivore populations. Annual Review of Entomology, 43, 421–447.CrossRefGoogle ScholarPubMed
Rosenheim, J. A. (2007). Intraguild predation: new theoretical and empirical perspectives. Ecology, 88, 2679–2680.CrossRefGoogle Scholar
Rosenheim, J. A., Kaya, H. K., Ehler, L. E., Marois, J. J. & Jaffee, B. A. (1995). Intraguild predation among biological-control agents: theory and evidence. Biological Control, 5, 303–335.CrossRefGoogle Scholar
Rosenzweig, M. L. (1973). Exploitation in three trophic levels. American Naturalist, 107, 275–294.CrossRefGoogle Scholar
Roy, H. E. & Cottrell, T. E. (2008). Forgotten natural enemies: interactions between coccinellids and insect-parasitic fungi. European Journal of Entomology, 105, 391–398.CrossRefGoogle Scholar
Roy, H. E. & Pell, J. K. (2000). Interactions between entomopathogenic fungi and other natural enemies: implications for biological control. Biocontrol Science and Technology, 10, 737–752.CrossRefGoogle Scholar
Roy, H. E., Pell, J. K. & Alderson, P. G. (1999). Effects of fungal infection on the alarm response of pea aphids. Journal of Invertebrate Pathology, 74, 69–75.CrossRefGoogle ScholarPubMed
Roy, H. E., Pell, J. K. & Alderson, P. G. (2001). Targeted dispersal of the aphid pathogenic fungus Erynia neoaphidis by the aphid predator Coccinella septempunctata. Biocontrol Science and Technology, 11, 99–110.CrossRefGoogle Scholar
Roy, H. E., Pell, J. K., Clark, S. J. & Alderson, P. G. (1998). Implications of predator foraging on aphid pathogen dynamics. Journal of Invertebrate Pathology, 71, 236–247.CrossRefGoogle ScholarPubMed
Roy, M. & Holt, R. D. (2008). Effects of predation on host–pathogen dynamics in SIR models. Theoretical Population Biology, 73, 319–331.CrossRefGoogle ScholarPubMed
Rudgers, J. A., Holah, J., Orr, S. P. & Clay, K. (2007). Forest succession suppressed by an introduced plant–fungal symbiosis. Ecology, 88, 18–25.CrossRefGoogle ScholarPubMed
Rudgers, J. A., Koslow, J. M. & Clay, K. (2004). Endophytic fungi alter relationships between diversity and ecosystem properties. Ecology Letters, 7, 42–51.CrossRefGoogle Scholar
Rudgers, J. A., Mattingly, W. B. & Koslow, J. M. (2005). Mutualistic fungus promotes plant invasion into diverse communities. Oecologia, 144, 463–471.CrossRefGoogle ScholarPubMed
Rudgers, J. A. & Orr, S. (2009). Non-native grass alters growth of native tree species via leaf and soil microbes. Journal of Ecology, 97, 247–255.CrossRefGoogle Scholar
Rudolf, V. H. W. (2007). The interaction of cannibalism and omnivory: consequences for community dynamics. Ecology, 88, 2697–2705.CrossRefGoogle ScholarPubMed
Rudolf, V. H. W. (2008). Consequences of size structure in the prey for predator–prey dynamics: the composite functional response. Journal of Animal Ecology, 77, 520–528.CrossRefGoogle ScholarPubMed
Rudolf, V. H. W. & Antonovics, J. (2005). Species coexistence and pathogens with frequency-dependent transmission. American Naturalist, 166, 112–118.CrossRefGoogle ScholarPubMed
Ruggieri, E. & Schreiber, S. J. (2005). The dynamics of the Schoener–Polis–Holt model of intra-guild predation. Mathematical Biosciences and Engineering, 2, 279–288.CrossRefGoogle ScholarPubMed
Runyon, J. B., Mescher, M. C. & Moraes, C. M. (2008). Parasitism by Cuscuta pentagona attenuates host plant defenses against insect herbivores. Plant Physiology, 146, 987–995.CrossRefGoogle ScholarPubMed
Rushton, S. P., Lurz, P. W. W., Gurnell, J. & Fuller, R. (2000). Modelling the spatial dynamics of parapoxvirus disease in red and grey squirrels: a possible cause of the decline in the red squirrel in the UK?Journal of Applied Ecology, 37, 997–1012.CrossRefGoogle Scholar
Rushton, S. P., Lurz, P. W. W., Gurnell, J., Nettleton, P., Bruemmer, C., Shirley, M. D. F. & Sainsbury, A. W. (2006). Disease threats posed by alien species: the role of a poxvirus in the decline of the native red squirrel in Britain. Epidemiology and Infection, 134, 521–533.CrossRefGoogle ScholarPubMed
Sachs, J. & Malaney, P. (2002). The economic and social burden of malaria. Nature, 415, 680–685.CrossRefGoogle ScholarPubMed
Sage, R. B., Woodburn, M. I. A., Davis, C. & Aebischer, N. J. (2002). The effect of an experimental infection of the nematode Heterakis gallinarum on hand-reared grey partridges Perdix perdix. Parasitology, 124, 529–535.CrossRefGoogle ScholarPubMed
Saikkonen, K., Faeth, S. H., Helander, M. & Sullivan, T. J. (1998). Fungal endophytes: a continuum of interactions with host plants. Annual Review of Ecology and Systematics, 29, 319–343.CrossRefGoogle Scholar
Sainsbury, A. W., Nettleton, P., Gilray, J. & Gurnell, J. (2000). Grey squirrels have high seroprevalence to a parapoxvirus associated with deaths in red squirrels. Animal Conservation, 3, 229–233.CrossRefGoogle Scholar
Sato, T., Arizono, M., Sone, R. & Harada, Y. (2008). Parasite-mediated allochthonous input: do hairworms enhance subsidized predation of stream salmonids on crickets?Canadian Journal of Zoology, 86, 231–235.CrossRefGoogle Scholar
Schall, J. J. (1992). Parasite-mediated competition in anolis lizards. Oecologia, 92, 58–64.CrossRefGoogle ScholarPubMed
Schloegel, L. M., Hero, J. M., Berger, L., Speare, R., McDonald, K. & Daszak, P. (2006). The decline of the sharp-snouted day frog (Taudactylus acutirostris): the first documented case of extinction by infection in a free-ranging wildlife species?Ecohealth, 3, 35–40.CrossRefGoogle Scholar
Schmitz, O. J. & Nudds, T. D. (1994). Parasite-mediated competition in deer and moose: how strong is the effect of meningeal worm on moose? Ecological Applications, 4, 91–103.CrossRefGoogle Scholar
Settle, W. H. & Wilson, L. T. (1990). Invasion by the variegated leafhopper and biotic interactions: parasitism, competition, and apparent competition. Ecology, 71, 1461–1470.CrossRefGoogle Scholar
Sharp, P. M. & Hahn, B. H. (2008). Aids: prehistory of HIV-1. Nature, 455, 605–606.CrossRefGoogle ScholarPubMed
Shoemaker, D. D., Ross, K. G., Keller, L., Vargo, E. L. & Werren, J. H. (2000). Wolbachia infections in native and introduced populations of fire ants (Solenopsis spp.). Insect Molecular Biology, 9, 661–673.CrossRefGoogle Scholar
Siegel, J. P., Maddox, J. V. & Ruesink, W. G. (1986). Impact of Nosema pyrausta on a braconid, Macrocentrus grandii, in central Illinois. Journal of Invertebrate Pathology, 47, 271–276.CrossRefGoogle Scholar
Sisterson, M. S. & Averill, A. L. (2003). Interactions between parasitized and unparasitized conspecifics: parasitoids modulate competitive dynamics. Oecologia, 135, 362–371.CrossRefGoogle ScholarPubMed
Slippers, B., Stenlid, J. & Wingfield, M. J. (2005). Emerging pathogens: fungal host jumps following anthropogenic introduction. Trends in Ecology & Evolution, 20, 420–421.CrossRefGoogle ScholarPubMed
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, 191–209.CrossRefGoogle Scholar
Smith, D. (2000). The population dynamics and community ecology of root hemiparasitic plants. American Naturalist, 155, 13–23.CrossRefGoogle ScholarPubMed
Smith, G. J. D., Vijaykrishna, D., Bahl, J., Lycett, S. J., Worobey, M., Pybus, O. G., Ma, S. K., Cheung, C. L., Raghwani, J., Bhatt, S., Peiris, J. S. M., Guan, Y. & Rambaut, A. (2009a). Origins and evolutionary genomics of the 2009 swine-origin H1N1 influenza A epidemic. Nature, 459, 1122–1125.CrossRefGoogle ScholarPubMed
Smith, K. F., Acevedo-Whitehouse, K. & Pedersen, A. B. (2009b). The role of infectious diseases in biological conservation. Animal Conservation, 12, 1–12.CrossRefGoogle Scholar
Smith, K. F., Behrens, M., Schloegel, L. M., Marano, N., Burgiel, S. & Daszak, P. (2009c). Reducing the risks of the wildlife trade. Science, 324, 594–595.CrossRefGoogle ScholarPubMed
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, 1349–1357.CrossRefGoogle ScholarPubMed
Smith, R. M. M., Drobniewski, F., Gibson, A., Montague, J. D. E., Logan, M. N., Hunt, D., Hewinson, G., Salmon, R. L. & O'Neill, B. (2004). Mycobacterium bovis infection, United Kingdom. Emerging Infectious Diseases, 10, 539–541.CrossRefGoogle Scholar
Snyder, W. E., Ballard, S. N., Yang, S., Clevenger, G. M., Miller, T. D., Ahn, J. J., Hatten, T. D. & Berryman, A. A. (2004). Complementary biocontrol of aphids by the ladybird beetle Harmonia axyridis and the parasitoid Aphelinus asychis on greenhouse roses. Biological Control, 30, 229–235.CrossRefGoogle Scholar
Sodora, D. L., Allan, J. S., Apetrei, C., Brenchley, J. M., Douek, D. C., Else, J. G., Estes, J. D., Hahn, B. H., Hirsch, V. M., Kaur, A., Kirchhoff, F., Muller-Trutwin, M., Pandrea, I., Schmitz, J. E. & Silvestri, G. (2009). Toward an AIDS vaccine: lessons from natural simian immunodeficiency virus infections of African nonhuman primate hosts. Nature Medicine, 15, 861–865.CrossRefGoogle ScholarPubMed
Sokolow, S. (2009). Effects of a changing climate on the dynamics of coral infectious disease: a review of the evidence. Diseases of Aquatic Organisms, 87, 5–18.CrossRefGoogle Scholar
Sorensen, R. E. & Minchella, D. J. (2001). Snail–trematode life history interactions: past trends and future directions. Parasitology, 123, S3–S18.CrossRefGoogle ScholarPubMed
Stahlhut, J. K., Liebert, A. E., Starks, P. T., Dapporto, L. & Jaenike, J. (2006). Wolbachia in the invasive European paper wasp Polistes dominulus. Insectes Sociaux, 53, 269–273.CrossRefGoogle Scholar
Steen, H., Taitt, M. & Krebs, C. J. (2002). Risk of parasite-induced predation: an experimental field study on Townsend's voles (Microtus townsendii). Canadian Journal of Zoology, 80, 1286–1292.CrossRefGoogle Scholar
Stout, M. J., Fidantsef, A. L., Duffey, S. S. & Bostock, R. M. (1999). Signal interactions in pathogen and insect attack: systemic plant-mediated interactions between pathogens and herbivores of the tomato, Lycopersicon esculentum. Physiological and Molecular Plant Pathology, 54, 115–130.CrossRefGoogle Scholar
Stout, M. J., Thaler, J. S. & Thomma, B. (2006). Plant-mediated interactions between pathogenic microorganisms and herbivorous arthropods. Annual Review of Entomology, 51, 663–689.CrossRefGoogle ScholarPubMed
Straub, C. S., Finke, D. L. & Snyder, W. E. (2008). Are the conservation of natural enemy biodiversity and biological control compatible goals?Biological Control, 45, 225–237.CrossRefGoogle Scholar
Stuart, S. N., Chanson, J. S., Cox, N. A., Young, B. E., Rodrigues, A. S. L., Fischman, D. L. & Waller, R. W. (2004). Status and trends of amphibian declines and extinctions worldwide. Science, 306, 1783–1786.CrossRefGoogle ScholarPubMed
Sumption, K. J. & Flowerdew, J. R. (1985). The ecological effects of the decline in rabbits (Oryctolagus cuniculus L.) due to myxomatosis. Mammal Review, 15, 151–186.CrossRefGoogle Scholar
Tain, L., Perrot-Minnot, M. J. & Cezilly, F. (2007). Differential influence of Pomphorhynchus laevis (Acanthocephala) on brain serotonergic activity in two congeneric host species. Biology Letters, 3, 68–71.CrossRefGoogle ScholarPubMed
Tanabe, K. & Namba, T. (2005). Omnivory creates chaos in simple food web models. Ecology, 86, 3411–3414.CrossRefGoogle Scholar
Tatem, A. J., Hay, S. I. & Rogers, D. J. (2006). Global traffic and disease vector dispersal. Proceedings of the National Academy of Sciences of the United States of America, 103, 6242–6247.CrossRefGoogle ScholarPubMed
Tattoni, C., Preatoni, D. G., Lurz, P. W. W., Rushton, S. P., Tosi, G., Bertolino, S., Martinoli, A. & Wauters, L. A. (2006). Modelling the expansion of a grey squirrel population: implications for squirrel control. Biological Invasions, 8, 1605–1619.CrossRefGoogle Scholar
Taylor, J. E., Hatcher, P. E. & Paul, N. D. (2004). Crosstalk between plant responses to pathogens and herbivores: a view from the outside in. Journal of Experimental Botany, 55, 159–168.CrossRefGoogle ScholarPubMed
Taylor, L. H., Latham, S. M. & Woolhouse, M. E. J. (2001). Risk factors for human disease emergence. Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 356, 983–989.CrossRefGoogle ScholarPubMed
Telfer, S., Birtles, R., Bennett, M., Lambin, X., Paterson, S. & Begon, M. (2008). Parasite interactions in natural populations: insights from longitudinal data. Parasitology, 135, 767–781.CrossRefGoogle ScholarPubMed
Telfer, S., Bown, K. J., Sekules, R., Begon, I., Hayden, T. & Birtles, R. (2005). Disruption of a host–parasite system following the introduction of an exotic host species. Parasitology, 130, 661–668.CrossRefGoogle ScholarPubMed
Terry, R. S., Smith, J. E., Sharpe, R. G., Rigaud, T., Littlewood, D. T. J., Ironside, J. E., Rollinson, D., Bouchon, D., MacNeil, C., Dick, J. T. A. & Dunn, A. M. (2004). Widespread vertical transmission and associated host sex-ratio distortion within the eukaryotic phylum Microspora. Proceedings of the Royal Society of London Series B – Biological Sciences, 271, 1783–1789.CrossRefGoogle ScholarPubMed
Thieltges, D. W., Montaudouin, X., Fredensborg, B., Jensen, K. T., Koprivnikar, J. & Poulin, R. (2008a). Production of marine trematode cercariae: a potentially overlooked path of energy flow in benthic systems. Marine Ecology Progress Series, 372, 147–155.CrossRefGoogle Scholar
Thieltges, D. W., Jensen, K. T. & Poulin, R. (2008b). The role of biotic factors in the transmission of free-living endohelminth stages. Parasitology, 135, 407–426.CrossRefGoogle ScholarPubMed
Thomas, F., Adamo, S. & Moore, J. (2005a). Parasitic manipulation: where are we and where should we go?Behavioural Processes, 68, 185–199.CrossRefGoogle Scholar
Thomas, F., Fauchier, J. & Lafferty, K. D. (2002a). Conflict of interest between a nematode and a trematode in an amphipod host: test of the ‘sabotage’ hypothesis. Behavioral Ecology and Sociobiology, 51, 296–301.CrossRefGoogle Scholar
Thomas, F., Mete, K., Helluy, S., Santalla, F., Verneau, O., DeMeeus, T., Cezilly, F. & Renaud, F. (1997). Hitch-hiker parasites or how to benefit from the strategy of another parasite. Evolution, 51, 1316–1318.CrossRefGoogle ScholarPubMed
Thomas, F., Poulin, R., Meeus, T., Guegan, J. F. & Renaud, F. (1999). Parasites and ecosystem engineering: what roles could they play?Oikos, 84, 167–171.CrossRefGoogle Scholar
Thomas, F., Poulin, R. & Renaud, F. (1998). Nonmanipulative parasites in manipulated hosts: ‘hitch-hikers’ or simply ‘lucky passengers’?Journal of Parasitology, 84, 1059–1061.CrossRefGoogle ScholarPubMed
Thomas, F., Renaud, F. & Guegan, J. F. (2005b). Parasitism and Ecosystems. Oxford: Oxford University Press.CrossRefGoogle Scholar
Thomas, F., Schmidt-Rhaesa, A., Martin, G., Manu, C., Durand, P. & Renaud, F. (2002b). Do hairworms (Nematomorpha) manipulate the water seeking behaviour of their terrestrial hosts?Journal of Evolutionary Biology, 15, 356–361.CrossRefGoogle Scholar
Thomas, K., Tompkins, D. M., Sainsbury, A. W., Wood, A. R., Dalziel, R., Nettleton, P. F. & McInnes, C. J. (2003). A novel poxvirus lethal to red squirrels (Sciurus vulgaris). Journal of General Virology, 84, 3337–3341.CrossRefGoogle Scholar
Thompson, R. M., Hemberg, M., Starzomski, B. M. & Shurin, J. B. (2007). Trophic levels and trophic tangles: the prevalence of omnivory in real food webs. Ecology, 88, 612–617.CrossRefGoogle ScholarPubMed
Thompson, R. M., Mouritsen, K. N. & Poulin, R. (2005). Importance of parasites and their life cycle characteristics in determining the structure of a large marine food web. Journal of Animal Ecology, 74, 77–85.CrossRefGoogle Scholar
Thrall, P. H. & Burdon, J. J. (2003). Evolution of virulence in a plant host–pathogen metapopulation. Science, 299, 1735–1737.CrossRefGoogle Scholar
Thrall, P. H. & Jarosz, A. M. (1994). Host–pathogen dynamics in experimental populations of Silene alba and Ustilago violacea: II – Experimental tests of theoretical models. Journal of Ecology, 82, 561–570.CrossRefGoogle Scholar
Tilman, D. (1982). Resource Competition and Community Structure. Princeton: Princeton University Press.Google ScholarPubMed
Tilman, D., May, R. M., Lehman, C. L. & Nowak, M. A. (1994). Habitat destruction and the extinction debt. Nature, 371, 65–66.CrossRefGoogle Scholar
Tompkins, D. M., Dickson, G. & Hudson, P. J. (1999). Parasite-mediated competition between pheasant and grey partridge: a preliminary investigation. Oecologia, 119, 378–382.CrossRefGoogle ScholarPubMed
Tompkins, D. M., Draycott, R. A. H. & Hudson, P. J. (2000a). Field evidence for apparent competition mediated via the shared parasites of two gamebird species. Ecology Letters, 3, 10–14.CrossRefGoogle Scholar
Tompkins, D. M., Dunn, A. M., Smith, M. J. & Telfer, S. (2010). Wildlife diseases: from individuals to ecosystems. Journal of Animal Ecology, 80, 19–38.CrossRefGoogle Scholar
Tompkins, D. M., Greenman, J. V. & Hudson, P. J. (2001). Differential impact of a shared nematode parasite on two gamebird hosts: implications for apparent competition. Parasitology, 122, 187–193.CrossRefGoogle ScholarPubMed
Tompkins, D. M., Greenman, J. V., Robertson, P. A. & Hudson, P. J. (2000b). The role of shared parasites in the exclusion of wildlife hosts: Heterakis gallinarum in the ring-necked pheasant and the grey partridge. Journal of Animal Ecology, 69, 829–840.CrossRefGoogle ScholarPubMed
Tompkins, D. M., Sainsbury, A. W., Nettleton, P., Buxton, D. & Gurnell, J. (2002). Parapoxvirus causes a deleterious disease in red squirrels associated with UK population declines. Proceedings of the Royal Society of London Series B – Biological Sciences, 269, 529–533.CrossRefGoogle ScholarPubMed
Tompkins, D. M., White, A. R. & Boots, M. (2003). Ecological replacement of native red squirrels by invasive greys driven by disease. Ecology Letters, 6, 189–196.CrossRefGoogle Scholar
Torchin, M. E., Byers, J. E. & Huspeni, T. C. (2005). Differential parasitism of native and introduced snails: replacement of a parasite fauna. Biological Invasions, 7, 885–894.CrossRefGoogle 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, 628–630.CrossRefGoogle ScholarPubMed
Torchin, M. E., Lafferty, K. D. & Kuris, A. M. (2002). Parasites and marine invasions. Parasitology, 124, S137–S151.CrossRefGoogle Scholar
Tounou, A. K., Kooyman, C., Douro-Kpindou, O. K. & Poehling, H. M. (2008a). Combined field efficacy of Paranosema locustae and Metarhizium anisopliae var. acridum for the control of Sahelian grasshoppers. Biocontrol, 53, 813–828.CrossRefGoogle Scholar
Tounou, A. K., Kooyman, C., Douro-Kplndou, O. K. & Poehling, H. M. (2008b). Interaction between Paranosema locustae and Metarhizium anisopliae var. acridum, two pathogens of the desert locust, Schistocerca gregaria under laboratory conditions. Journal of Invertebrate Pathology, 97, 203–210.CrossRefGoogle ScholarPubMed
Tsutsui, N. D., Kauppinen, S. N., Oyafuso, A. F. & Grosberg, R. K. (2003). The distribution and evolutionary history of Wolbachia infection in native and introduced populations of the invasive Argentine ant (Linepithema humile). Molecular Ecology, 12, 3057–3068.CrossRefGoogle Scholar
Tumpey, T. M., Basler, C. F., Aguilar, P. V., Zeng, H., Solorzano, A., Swayne, D. E., Cox, N. J., Katz, J. M., Taubenberger, J. K., Palese, P. & Garcia-Sastre, A. (2005). Characterization of the reconstructed 1918 Spanish influenza pandemic virus. Science, 310, 77–80.CrossRefGoogle ScholarPubMed
Turlings, T. C. J. & Tumlinson, J. H. (1992). Systemic release of chemical signals by herbivore-injured corn. Proceedings of the National Academy of Sciences of the United States of America, 89, 8399–8402.CrossRefGoogle ScholarPubMed
Twu, S. J., Chen, T. J., Chen, C. J., Olsen, S. J., Lee, L. T., Fisk, T., Hsu, K. H., Chang, S. C., Chen, K. T., Chiang, I. H., Wu, Y. C., Wu, J. S. & Dowell, S. F. (2003). Control measures for severe acute respiratory syndrome (SARS) in Taiwan. Emerging Infectious Diseases, 9, 718–720.CrossRefGoogle Scholar
Dam, N. M., Raaijmakers, C. E. & Putten, W. H. (2005). Root herbivory reduces growth and survival of the shoot feeding specialist Pieris rapae on Brassica nigra. Entomologia experimentalis et Applicata, 115, 161–170.Google Scholar
Putten, W. H., Van Dijk, C. & Peters, B. A. M. (1993). Plant-specific soil-borne diseases contribute to succession in foredune vegetation. Nature, 362, 53–56.CrossRefGoogle Scholar
Putten, W. H. & Peters, B. A. M. (1997). How soil-borne pathogens may affect plant competition. Ecology, 78, 1785–1795.CrossRefGoogle Scholar
Putten, W. H., Vet, L. E. M., Harvey, J. A. & Wackers, F. L. (2001). Linking above- and belowground multitrophic interactions of plants, herbivores, pathogens, and their antagonists. Trends in Ecology & Evolution, 16, 547–554.CrossRefGoogle Scholar
Loon, L. C., Bakker, P. & Pieterse, C. M. J. (1998). Systemic resistance induced by rhizosphere bacteria. Annual Review of Phytopathology, 36, 453–483.CrossRefGoogle ScholarPubMed
Nouhuys, S. & Hanski, I. (2000). Apparent competition between parasitoids mediated by a shared hyperparasitoid. Ecology Letters, 3, 82–84.CrossRefGoogle Scholar
Ommeren, R. J. & Whitham, T. G. (2002). Changes in interactions between juniper and mistletoe mediated by shared avian frugivores: parasitism to potential mutualism. Oecologia, 130, 281–288.CrossRefGoogle ScholarPubMed
Reeth, K. (2007). Avian and swine influenza viruses: our current understanding of the zoonotic risk. Veterinary Research, 38, 243–260.CrossRefGoogle ScholarPubMed
Veen, F. J. F., Morris, R. J. & Godfray, H. C. J. (2006). Apparent competition, quantitative food webs, and the structure of phytophagous insect communities. Annual Review of Entomology, 51, 187–208.CrossRefGoogle ScholarPubMed
Veen, F. J. F., Mueller, C. B., Pell, J. K. & Godfray, H. C. J. (2008). Food web structure of three guilds of natural enemies: predators, parasitoids and pathogens of aphids. Journal of Animal Ecology, 77, 191–200.CrossRefGoogle ScholarPubMed
Vance-Chalcraft, H. D., Rosenheim, J. A., Vonesh, J. R., Osenberg, , , C. W. & Sih, A. (2007). The influence of intraguild predation on prey suppression and prey release: a meta-analysis. Ecology, 88, 2689–2696.CrossRefGoogle ScholarPubMed
Vial, F., Cleaveland, S., Rasmussen, G. & Haydon, D. T. (2006). Development of vaccination strategies for the management of rabies in African wild dogs. Biological Conservation, 131, 180–192.CrossRefGoogle Scholar
Vicari, M., Hatcher, P. E. & Ayres, P. G. (2002). Combined effect of foliar and mycorrhizal endophytes on an insect herbivore. Ecology, 83, 2452–2464.CrossRefGoogle Scholar
Vidal, N., Peeters, M., Mulanga-Kabeya, C., Nzilambi, N., Robertson, D., Ilunga, W., Sema, H., Tshimanga, K., Bongo, B. & Delaporte, E. (2000). Unprecedented degree of human immunodeficiency virus type 1 (HIV-1) group M genetic diversity in the Democratic Republic of Congo suggests that the HIV-1 pandemic originated in Central Africa. Journal of Virology, 74, 10498–10507.CrossRefGoogle Scholar
Vinale, F., Sivasithamparam, K., Ghisalberti, E. L., Marra, R., Woo, S. L. & Lorito, M. (2008). Trichoderma–plant–pathogen interactions. Soil Biology & Biochemistry, 40, 1–10.CrossRefGoogle Scholar
Vredenburg, V. T., Knapp, R. A., Tunstall, T. S. & Briggs, C. J. (2010). Dynamics of an emerging disease drive large-scale amphibian population extinctions. Proceedings of the National Academy of Sciences of the United States of America, 107, 9689–9694.CrossRefGoogle ScholarPubMed
Vucetich, J. A. & Peterson, R. O. (2004). The influence of top-down, bottom-up and abiotic factors on the moose (Alces alces) population of Isle Royale. Proceedings of the Royal Society of London Series B – Biological Sciences, 271, 183–189.CrossRefGoogle ScholarPubMed
Wang, L. F. & Eaton, B. T. (2007). Bats, civets and the emergence of SARS. Current Topics in Microbiology and Immunology, 315, 325–344.Google ScholarPubMed
Ward, J. R., Kim, K. & Harvell, C. D. (2007). Temperature affects coral disease resistance and pathogen growth. Marine Ecology Progress Series, 329, 115–121.CrossRefGoogle Scholar
Ward, J. R. & Lafferty, K. D. (2004). The elusive baseline of marine disease: are diseases in ocean ecosystems increasing?PLoS Biology, 2, 542–547.CrossRefGoogle ScholarPubMed
Warkentin, I. G., Bickford, D., Sodhi, N. S. & Bradshaw, C. J. A. (2009). Eating frogs to extinction. Conservation Biology, 23, 1056–1059.CrossRefGoogle Scholar
Washburn, J. O., Mercer, D. R. & Anderson, J. R. (1991). Regulatory role of parasites: impact on host population shifts with resource availability. Science, 253, 185–188.CrossRefGoogle ScholarPubMed
Watson, D. M. (2001). Mistletoe: a keystone resource in forests and woodlands worldwide. Annual Review of Ecology and Systematics, 32, 219–249.CrossRefGoogle Scholar
Wattier, R. A., Haine, E. R., Beguet, J., Martin, G., Bollache, L., Musko, I. B., Platvoet, D. & Rigaud, T. (2007). No genetic bottleneck or associated microparasite loss in invasive populations of a freshwater amphipod. Oikos, 116, 1941–1953.CrossRefGoogle Scholar
Webster, R. G., Bean, W. J., Gorman, O. T., Chambers, T. M. & Kawaoka, Y. (1992). Evolution and ecology of influenza-A viruses. Microbiological Reviews, 56, 152–179.Google ScholarPubMed
Werner, E. E. & Peacor, S. D. (2003). A review of trait-mediated indirect interactions in ecological communities. Ecology, 84, 1083–1100.CrossRefGoogle Scholar
Werren, J. H., Baldo, L. & Clark, M. E. (2008). Wolbachia: master manipulators of invertebrate biology. Nature Reviews Microbiology, 6, 741–751.CrossRefGoogle ScholarPubMed
Werren, J. H. & Beukeboom, L. W. (1993). Population genetics of a parasitic chromosome: theoretical analysis of PSR in subdivided populations. American Naturalist, 142, 224–241.CrossRefGoogle ScholarPubMed
Wertheim, J. O. (2009). When pigs fly: the avian origin of a ‘swine flu’. Environmental Microbiology, 11, 2191–2192.CrossRefGoogle ScholarPubMed
Westbury, D. B. & Dunnett, N. P. (2008). The promotion of grassland forb abundance: a chemical or biological solution?Basic and Applied Ecology, 9, 653–662.CrossRefGoogle Scholar
White, S. M., Sait, S. M. & Rohani, P. (2007). Population dynamic consequences of parasitised-larval competition in stage-structured host–parasitoid systems. Oikos, 116, 1171–1185.Google Scholar
Whitehouse, A. T., Peay, S. & Kindemba, V. (2009). Ark Sites for White-clawed Crayfish: Guidance for the Aggregates Industry. Peterborough: Buglife – The Invertebrate Conservation Trust.Google Scholar
Whitlaw, H. A. & Lankester, M. W. (1994). The cooccurrence of moose, white-tailed deer, and Parelaphostrongylus tenuis in Ontario. Canadian Journal of Zoology, 72, 819–825.CrossRefGoogle Scholar
Wilcove, D. S., Rothstein, D., Dubow, J., Phillips, A. & Losos, E. (1998). Quantifying threats to imperiled species in the United States. Bioscience, 48, 607–615.CrossRefGoogle Scholar
Williams, P. D. & Day, T. (2001). Interactions between sources of mortality and the evolution of parasite virulence. Proceedings of the Royal Society of London Series B – Biological Sciences, 268, 2331–2337.CrossRefGoogle ScholarPubMed
Wilmers, C. C., Post, E., Peterson, R. O. & Vucetich, J. A. (2006). Predator disease out-break modulates top-down, bottom-up and climatic effects on herbivore population dynamics. Ecology Letters, 9, 383–389.CrossRefGoogle ScholarPubMed
Wodarz, D. & Sasaki, A. (2004). Apparent competition and recovery from infection. Journal of Theoretical Biology, 227, 403–412.CrossRefGoogle ScholarPubMed
Wolfe, L. M. (2002). Why alien invaders succeed: support for the escape-from-enemy hypothesis. American Naturalist, 160, 705–711.Google ScholarPubMed
Wolfe, N. D., Daszak, P., Kilpatrick, A. M. & Burke, D. S. (2005). Bushmeat hunting deforestation, and prediction of zoonoses emergence. Emerging Infectious Diseases, 11, 1822–1827.CrossRefGoogle ScholarPubMed
Wolfe, N. D., Dunavan, C. P. & Diamond, J. (2007). Origins of major human infectious diseases. Nature, 447, 279–283.CrossRefGoogle ScholarPubMed
Wood, C. L., Byers, J. E., Cottingham, K. L., Altman, I., Donahue, M. J. & Blakeslee, A. M. H. (2007). Parasites alter community structure. Proceedings of the National Academy of Sciences of the United States of America, 104, 9335–9339.CrossRefGoogle ScholarPubMed
Woolhouse, M. & Gaunt, E. (2007). Ecological origins of novel human pathogens. Critical Reviews in Microbiology, 33, 231–242.CrossRefGoogle ScholarPubMed
Woolhouse, M. E. J. (2002). Population biology of emerging and re-emerging pathogens. Trends in Microbiology, 10, S3–S7.CrossRefGoogle ScholarPubMed
Woolhouse, M. E. J., Dye, C., Etard, J. F., Smith, T., Charlesworth, J. D., Garnett, G. P., Hagan, P., Hii, J. L., Ndhlovu, P. D., Quinnell, R. J., Watts, C. H., Chandiwana, S. K. & Anderson, R. M. (1997). Heterogeneities in the transmission of infectious agents: implications for the design of control programs. Proceedings of the National Academy of Sciences of the United States of America, 94, 338–342.CrossRefGoogle ScholarPubMed
Woolhouse, M. E. J. & Gowtage-Sequeria, S. (2005). Host range and emerging and reemerging pathogens. Emerging Infectious Diseases, 11, 1842–1847.CrossRefGoogle ScholarPubMed
Woolhouse, M. E. J., Haydon, D. T. & Antia, R. (2005). Emerging pathogens: the epidemiology and evolution of species jumps. Trends in Ecology & Evolution, 20, 238–244.CrossRefGoogle ScholarPubMed
Woolhouse, M. E. J., Webster, J. P., Domingo, E., Charlesworth, B. & Levin, B. R. (2002). Biological and biomedical implications of the co-evolution of pathogens and their hosts. Nature Genetics, 32, 569–577.CrossRefGoogle ScholarPubMed
Worobey, M., Gemmel, M., Teuwen, D. E., Haselkorn, T., Kunstman, K., Bunce, M., Muyembe, J. J., Kabongo, J. M. M., Kalengayi, R. M., Marck, E., Gilbert, M. T. P. & Wolinsky, S. M. (2008). Direct evidence of extensive diversity of HIV-1 in Kinshasa by 1960. Nature, 455, 661–664.CrossRefGoogle ScholarPubMed
Wright, H. A., Wootton, R. J. & Barber, I. (2006). The effect of Schistocephalus solidus infection on meal size of three-spined stickleback. Journal of Fish Biology, 68, 801–809.CrossRefGoogle Scholar
Yan, G. Y. (1996). Parasite-mediated competition: a model of directly transmitted macroparasites. American Naturalist, 148, 1089–1112.CrossRefGoogle Scholar
Yan, G. Y., Stevens, L., Goodnight, C. J. & Schall, J. J. (1998). Effects of a tapeworm parasite on the competition of Tribolium beetles. Ecology, 79, 1093–1103.CrossRefGoogle Scholar
Yates, T. L., Mills, J. N., Parmenter, C. A., Ksiazek, T. G., Parmenter, R. R., Vande Castle, J. R., Calisher, C. H., Nichol, S. T., Abbott, K. D., Young, J. C., Morrison, M. L., Beaty, B. J., Dunnum, J. L., Baker, R. J., Salazar-Bravo, J. & Peters, C. J. (2002). The ecology and evolutionary history of an emergent disease: hantavirus pulmonary syndrome. Bioscience, 52, 989–998.CrossRefGoogle Scholar
Yoshida, T., Ellner, S. P., Jones, L. E., Bohannan, B. J. M., Lenski, R. E. & Hairston, N. G. (2007). Cryptic population dynamics: rapid evolution masks trophic interactions. PLoS Biology, 5, 1868–1879.CrossRefGoogle ScholarPubMed
Zhang, P. J., Zheng, S. J., Loon, J. J. A., Boland, W., David, A., Mumm, R. & Dicke, M. (2009). Whiteflies interfere with indirect plant defense against spider mites in Lima bean. Proceedings of the National Academy of Sciences of the United States of America, 106, 21202–21207.CrossRefGoogle ScholarPubMed

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.

  • References
  • Melanie J. Hatcher, University of Bristol, Alison M. Dunn, University of Leeds
  • Book: Parasites in Ecological Communities
  • Online publication: 05 August 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511987359.010
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.

  • References
  • Melanie J. Hatcher, University of Bristol, Alison M. Dunn, University of Leeds
  • Book: Parasites in Ecological Communities
  • Online publication: 05 August 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511987359.010
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.

  • References
  • Melanie J. Hatcher, University of Bristol, Alison M. Dunn, University of Leeds
  • Book: Parasites in Ecological Communities
  • Online publication: 05 August 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511987359.010
Available formats
×