Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-ndmmz Total loading time: 0 Render date: 2024-05-02T15:02:20.017Z Has data issue: false hasContentIssue false

Chapter Five - The causes and consequences of parasite interactions: African buffalo as a case study

from Part I - Understanding within-host processes

Published online by Cambridge University Press:  28 October 2019

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

Summary

Parasites live and interact in multi-species communities. As these interactions are often hidden, the extent to which they occur, their relative strength and consequences are poorly understood. We review work on parasite interactions occurring in free-living African buffalo, which are distributed across the African continent and host a diversity of parasites, from bacteria and viruses to helminths. Three case studies of pairwise interactions between some of the most common and economically important parasites of buffalo shed new light on the effects of parasite interactions for individual hosts and population-level disease dynamics. Work on interactions between macro- and microparasites (common gastrointestinal worm infections and bovine tuberculosis, TB) suggests that immune responses underlie complex interactions. At individual host level, worms enhance TB infection severity, but at population level they can limit TB spread. Analysis of interactions between TB and Rift Valley Fever virus (RVFV) shows that TB presence makes increases RVFV effects. Work into how two dominant members of the worm community living in the buffalo gastrointestinal tract reassemble after perturbation reveals that the processes driving interactions between parasites can be dynamic over time. We use combined approaches to bridge the gap between individual and population scales and show how studies of natural populations can advance understanding of parasite interactions.

Type
Chapter
Information
Wildlife Disease Ecology
Linking Theory to Data and Application
, pp. 129 - 160
Publisher: Cambridge University Press
Print publication year: 2019

Access options

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

References

Abu Samra, N., Jori, F., Xiao, L.H., Rikhotso, O. & Thompson, P.N. (2013) Molecular characterization of Cryptosporidium species at the wildlife/livestock interface of the Kruger National Park, South Africa. Comparative Immunology, Microbiology and Infectious Diseases, 36, 295302.Google Scholar
Allen, J.E. & Maizels, R.M. (2011) Diversity and dialogue in immunity to helminths. Nature Reviews Immunology, 11, 375388.Google Scholar
Anderson, E.C. & Rowe, L.W. (1998) The prevalence of antibody to the viruses of bovine virus diarrhoea, bovine herpes virus 1, rift valley fever, ephemeral fever and bluetongue and to Leptospira sp. in free-ranging wildlife in Zimbabwe. Epidemiology and Infection, 121, 441449.Google Scholar
Anderson, R.M. & May, R.M. (1991) Infectious Diseases of Humans:Dynamics and Control. Oxford: Oxford University Press.Google Scholar
Atherstone, C., Picozzi, K. & Kalema-Zikusoka, G. (2014) Short Report: Seroprevalence of Leptospira Hardjo in cattle and African buffalos in Southwestern Uganda. American Journal of Tropical Medicine and Hygiene, 90, 288290.Google Scholar
Ayebazibwe, C., Mwiine, F.N., Balinda, S.N., Tjornehoj, K. & Alexandersen, S. (2012) Application of the Ceditest (R) FMDV type O and FMDV-NS enzyme-linked immunosorbent assays for detection of antibodies against foot-and-mouth disease virus in selected livestock and wildlife species in Uganda. Journal of Veterinary Diagnostic Investigation, 24, 270276.CrossRefGoogle Scholar
Beechler, B.R., Bengis, R., Swanepoel, R., et al. (2015a) Rift Valley Fever in Kruger National Park: do buffalo play a role in the inter-epidemic circulation of virus? Transboundary and Emerging Diseases, 62, 2432.Google Scholar
Beechler, B.R., Manore, C.A., Reininghaus, B., et al. (2015b) Enemies and turncoats: bovine tuberculosis exposes pathogenic potential of Rift Valley fever virus in a common host, African buffalo (Syncerus caffer). Proceedings of the Royal Society of London B, 282, 20142942.Google Scholar
Bender, E.A., Case, T.J. & Gilpin, M.E. (1984) Perturbation experiments in community ecology: theory and practice. Ecology, 65, 113.Google Scholar
Beura, L.K., Hamilton, S.E., Bi, K., et al. (2016) Normalizing the environment recapitulates adult human immune traits in laboratory mice. Nature, 532, 512516.Google Scholar
Biron, C.A., Nguyen, K.B., Pien, G.C., Cousens, L.P. & Salazar-Mather, T.P. (1999) Natural killer cells in antiviral defense: function and regulation by innate cytokines. Annual Review of Immunology, 17, 189220.Google Scholar
Brown, J.H., Whitham, T.G., Morgan Ernest, S.K. & Gehring, C.A. (2001) Complex species interactions and the dynamics of ecological systems: long-term experiments. Science, 293, 643.Google Scholar
Budischak, S.A., Hoberg, E.P., Abrams, A., Jolles, A.E. & Ezenwa, V.O. (2015) A combined parasitological molecular approach for noninvasive characterization of parasitic nematode communities in wild hosts. Molecular Ecology Resources, 15, 11121119.Google Scholar
Budischak, S.A., Hoberg, E.P., Abrams, A., Jolles, A.E. & Ezenwa, V.O. (2016) Experimental insight into the process of parasite community assembly. Journal of Animal Ecology, 85, 12221233.Google Scholar
Budischak, S.A., Jolles, A.E. & Ezenwa, V.O. (2012) Direct and indirect costs of co-infection in the wild: linking gastrointestinal parasite communities, host hematology, and immune function. International Journal for Parasitology: Parasites and Wildlife, 1, 212.Google Scholar
Budischak, S.A., O’Neal, D., Jolles, A.E. & Ezenwa, V.O. (2018) Differential host responses to parasitism shape divergent fitness costs of infection. Functional Ecology, 32, 324333.Google Scholar
Callaway, R.M. & Walker, L.R. (1997) Competition and facilitation: a synthetic approach to interactions in plant communities. Ecology, 78, 19581965.Google Scholar
Chaisi, M.E., Sibeko, K.P., Collins, N.E., Potgieter, F.T. & Oosthuizen, M.C. (2011) Identification of Theileria parva and Theileria sp. (buffalo) 18S rRNA gene sequence variants in the African Buffalo (Syncerus caffer) in southern Africa. Veterinary Parasitology, 182, 150162.Google Scholar
Chenine, A.L., Shai-Kobiler, E., Steele, L.N., et al. (2008) Acute Schistosoma mansoni infection increases susceptibility to systemic SHIV clade C infection in rhesus macaques after mucosal virus exposure. PLoS Neglected Tropical Diseases, 2, e265.Google Scholar
Cousins, D.V., Huchzermeyer, H., Griffin, J., Van Rensburg, I.B.J.B.G. & Kriek, N. (2004) Tuberculosis. In: Coetzer, J. & Tustin, R.C. (eds.), Infectious Diseases of Livestock. (pp. 19731991). Oxford: Oxford University Press.Google Scholar
Cox, F.E.G. (2001) Concomitant infections, parasites and immune responses. Parasitology, 122, S23S38.Google Scholar
Dini-Andreote, F., Stegen, J.C., van Elsas, J.D. & Salles, J.F. (2015) Disentangling mechanisms that mediate the balance between stochastic and deterministic processes in microbial succession. Proceedings of the National Academy of Sciences of the United States of America, 112, E1326E1332.Google ScholarPubMed
Else, K.J. & Finkelman, F.D. (1998) Intestinal nematode parasites, cytokines and effector mechanisms. International Journal for Parasitology, 28, 11451158.CrossRefGoogle ScholarPubMed
Entrican, G. (2002) Immune regulation during pregnancy and host–pathogen interactions in infectious abortion. Journal of Comparative Pathology, 126, 7994.Google Scholar
Eygelaar, D., Jori, F., Mokopasetso, M., et al. (2015) Tick-borne haemoparasites in African buffalo (Syncerus caffer) from two wildlife areas in Northern Botswana. Parasites & Vectors, 8, 26.Google Scholar
Ezenwa, V.O. (2004) Host social behavior and parasitic infection: a multifactorial approach. Behavioral Ecology, 15, 446454.Google Scholar
Ezenwa, V.O., Etienne, R.S., Luikart, G., Beja-Pereira, A. & Jolles, A.E. (2010) Hidden consequences of living in a wormy world: nematode-induced immune suppression facilitates tuberculosis invasion. American Naturalist, 176, 613624.Google Scholar
Ezenwa, V.O. & Jolles, A.E. (2015) Opposite effects of anthelmintic treatment on microbial infection at individual versus population scales. Science, 347, 175177.Google Scholar
Ezenwa, V.O., Price, S.A., Altizer, S., Vitone, N.D. & Cook, K.C. (2006) Host traits and parasite species richness in even and odd‐toed hoofed mammals, Artiodactyla and Perissodactyla. Oikos, 115, 526536.Google Scholar
Fagbo, S., Coetzer, J.A.W. & Venter, E.H. (2014) Seroprevalence of Rift Valley fever and lumpy skin disease in African buffalo (Syncerus caffer) in the Kruger National Park and Hluhluwe-iMfolozi Park, South Africa. Journal of the South African Veterinary Association, 85, e1e7.Google Scholar
Fayle, T.M., Eggleton, P., Manica, A., Yusah, K.M. & Foster, W.A. (2015) Experimentally testing and assessing the predictive power of species assembly rules for tropical canopy ants. Ecology Letters, 18, 254262.Google Scholar
Fenton, A., Knowles, S.C.L., Petchey, O.L. & Pedersen, A.B. (2014) The reliability of observational approaches for detecting interspecific parasite interactions: comparison with experimental results. International Journal for Parasitology, 44, 437445.Google Scholar
Fischer, N., Indenbirken, D., Meyer, T., et al. (2015) Evaluation of unbiased next-generation sequencing of RNA (RNA-seq) as a diagnostic method in influenza virus-positive respiratory samples. Journal of Clinical Microbiology, 53, 22382250.Google Scholar
Fitzgerald, S.D. & Kaneene, J.B. (2013) Wildlife reservoirs of bovine tuberculosis worldwide: hosts, pathology, surveillance, and control. Veterinary Pathology, 50, 488499.Google Scholar
Flick, R. & Bouloy, M. (2005) Rift Valley fever virus. Current Molecular Medicine, 5, 827834.CrossRefGoogle ScholarPubMed
Flynn, J.L. & Chan, J. (2001) Immunology of tuberculosis. Annual Review of Immunology, 19, 93129.CrossRefGoogle ScholarPubMed
Gomo, C., de Garine-Wichatitsky, M., Caron, A. & Pfukenyi, D.M. (2012) Survey of brucellosis at the wildlife–livestock interface on the Zimbabwean side of the Great Limpopo Transfrontier Conservation Area. Tropical Animal Health and Production, 44, 7785.CrossRefGoogle ScholarPubMed
Gradwell, D.V., Schutte, A.P., Vanniekerk, C.A. & Roux, D.J. (1977) Isolation of Brucella abortus biotype 1 from African buffalo in Kruger National Park. Journal of the South African Veterinary Association, 48, 4143.Google ScholarPubMed
Graham, A.L. (2008) Ecological rules governing helminth–microparasite coinfection. Proceedings of the National Academy of Sciences of the United States of America, 105, 566570.Google Scholar
Griffiths, E.C., Pedersen, A.B., Fenton, A. & Petchey, O.L. (2014) Analysis of a summary network of co-infection in humans reveals that parasites interact most via shared resources. Proceedings of the Royal Society of London B, 281, 20132286.Google ScholarPubMed
Hamblin, C., Anderson, E.C., Jago, M., Mlengeya, T. & Hipji, K. (1990) Antibodies to some pathogenic agents in free-living wild species in Tanzania. Epidemiology and Infection, 105, 585594.Google Scholar
Hamblin, C., Hedger, R.S. & Condy, J.B. (1980) The isolation of parainfluenza 3 virus from free-living African buffalo (Syncerus caffer). Veterinary Record, 107, 18.Google Scholar
Henrichs, B., Oosthuizen, M.C., Troskie, M., et al. (2016) Within guild co-infections influence parasite community membership: a longitudinal study in African Buffalo. Journal of Animal Ecology, 85, 10251034.Google Scholar
Hoberg, E.P., Abrams, A. & Ezenwa, V.O. (2008) An exploration of diversity among the Ostertagiinae (Nematoda: Trichostrongyloidea) in ungulates from sub-Saharan Africa with a proposal for a new genus. Journal of Parasitology, 94, 230251.Google Scholar
Hoberg, E.P., Abrams, A. & Pilitt, P.A. (2010) A new species of trichostrongyloid in African buffalo (Syncerus caffer). The Journal of Parasitology, 96, 129136.Google Scholar
Hoverman, J.T., Hoye, B.J. & Johnson, P.T.J. (2013) Does timing matter? How priority effects influence the outcome of parasite interactions within hosts. Oecologia, 173, 14711480.Google Scholar
Jolles, A.E., Etienne, R.S. & Olff, H. (2006) Independent and competing disease risks: implications for host populations in variable environments. American Naturalist, 167, 745757.Google Scholar
Jolles, A.E., Ezenwa, V.O., Etienne, R.S., Turner, W.C. & Olff, H. (2008) Interactions between macroparasites and microparasites drive infection patterns in free-ranging African buffalo. Ecology, 89, 22392250.Google Scholar
Jolles, A.E., Le Roex, N.I.C.K.I., Flacke, G., et al. (2017) Wildlife disease dynamics in carnivore and herbivore hosts in the Hluhluwe-iMfolozi Park. In: Cromsigt, J.P., Archibald, S. & Owen-Smith, N. (eds.), Conserving Africa’s Mega-Diversity in the Anthropocene: The Hluhluwe-iMfolozi Park Story. Cambridge: Cambridge University Press.Google Scholar
Knowles, S.C., Fenton, A., Petchey, O.L., et al. (2013) Stability of within-host–parasite communities in a wild mammal system. Proceedings of the Royal Society of London B, 280, 20130598.Google Scholar
Krasnov, B.R., Pilosof, S., Stanko, M., et al. (2014) Co-occurrence and phylogenetic distance in communities of mammalian ectoparasites: limiting similarity versus environmental filtering. Oikos, 123, 6370.Google Scholar
Kreisinger, J., Bastien, G., Hauffe, H.C., Marchesi, J. & Perkins, S.E. (2015) Interactions between multiple helminths and the gut microbiota in wild rodents. Philosophical Transactions of the Royal Society of London B:Biological Sciences, 370, 20140295.Google Scholar
Kuris, A.M. (1990) Guild structure of larval trematodes in molluscan hosts: Prevalence, dominance, and significance of competition. In: Esch, G.W., Bush, A.O. & Aho, J.M. (eds.), Parasite Communities: Patterns and Processes (pp. 69100). London: Chapman and Hall.Google 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, 840844.Google Scholar
Levitt, B., Obala, A., Langdon, S., et al. (2017) Overlap extension barcoding for the next generation sequencing and genotyping of Plasmodium falciparum in individual patients in Western Kenya. Scientific Reports, 7, 41108.Google Scholar
Lowry, E., Rollinson, E.J., Laybourn, A.J., et al. (2013) Biological invasions: a field synopsis, systematic review, and database of the literature. Ecology & Evolution, 3, 182196.Google Scholar
Maggioli, M.F., Palmer, M.V., Thacker, T.C., Vordermeier, H.M. & Waters, W.R. (2015) Characterization of effector and memory T cell subsets in the immune response to bovine tuberculosis in cattle. PLoS ONE, 10, e0122571.Google Scholar
Maizels, R.M., Hewitson, J.P., Murray, J., et al. (2012) Immune modulation and modulators in Heligmosomoides polygyrus infection. Experimental Parasitology, 132, 7689.Google Scholar
Manore, C.A. & Beechler, B.R. (2015) Inter-epidemic and between-season persistence of rift valley fever: vertical transmission or cryptic cycling? Transboundary Emerging Diseases, 62, 1323.Google Scholar
McSorley, H.J. & Maizels, R.M. (2012) Helminth infections and host immune regulation. Clinical Microbiology Reviews, 25, 585608.Google Scholar
Michel, A.L. & Bengis, R.G. (2012) The African buffalo: a villain for inter-species spread of infectious diseases in southern Africa. Onderstepoort Journal of Veterinary Research, 79(2).Google Scholar
Miguel, E., Grosbois, V., Caron, A., et al. (2013) Contacts and foot and mouth disease transmission from wild to domestic bovines in Africa. Ecosphere, 4, art51.Google Scholar
Mollot, G., Pantel, J.H. & Romanuk, T.N. (2017) The effects of invasive species on the decline in species richness: a global meta-analysis. In: Bohan, D.A., Dumbrell, A.J. & Massol, F. (eds.), Advances in Ecological Research (pp. 6183). New York, NY: Academic Press.Google Scholar
Morton, E.R., Lynch, J., Froment, A., et al. (2015) Variation in rural African gut microbiota is strongly correlated with colonization by Entamoeba and subsistence. PLOS Genetics, 11, e1005658.Google Scholar
Mosmann, T.R. & Sad, S. (1996) The expanding universe of T-cell subsets: Th1, Th2 and more. Immunology Today, 17, 138146.Google Scholar
Nanyingi, M.O., Munyua, P., Kiama, S.G., et al. (2015) A systematic review of Rift Valley Fever epidemiology 1931–2014. Infection Ecology & Epidemiology, 5, 28024.Google Scholar
Newbold, L.K., Burthe, S.J., Oliver, A.E., et al. (2017) Helminth burden and ecological factors associated with alterations in wild host gastrointestinal microbiota. ISME Journal, 11, 663675.Google Scholar
Osborne, L.C., Monticelli, L.A., Nice, T.J., et al. (2014) Virus–helminth coinfection reveals a microbiota-independent mechanism of immunomodulation. Science, 345, 578.Google Scholar
Pawlowski, A., Jansson, M., Sköld, M., Rottenberg, M.E. & Källenius, G. (2012) Tuberculosis and HIV co-infection. PLOS Pathogens, 8, e1002464.Google Scholar
Pedersen, A.B. & Fenton, A. (2007) Emphasizing the ecology in parasite community ecology. Trends in Ecology & Evolution, 22, 133139.Google Scholar
Pedersen, A.B. & Fenton, A. (2015) The role of antiparasite treatment experiments in assessing the impact of parasites on wildlife. Trends in Parasitology, 31, 200211.CrossRefGoogle ScholarPubMed
Pepin, M., Bouloy, M., Bird, B.H., Kemp, A. & Paweska, J. (2010) Rift Valley fever virus (Bunyaviridae: Phlebovirus): an update on pathogenesis, molecular epidemiology, vectors, diagnostics and prevention. Veterinary Research, 41, 61.Google Scholar
Petney, T.N. & Andrews, R.H. (1998) Multiparasite communities in animals and humans: frequency, structure and pathogenic significance. International Journal for Parasitology, 28, 377393.Google Scholar
Pienaar, N.J. & Thompson, P.N. (2013) Temporal and spatial history of Rift Valley fever in South Africa: 1950 to 2011. Onderstepoort Journal of Veterinary Research, 80(1).Google Scholar
Pilosof, S., Morand, S., Krasnov, B.R. & Nunn, C.L. (2015) Potential parasite transmission in multi-host networks based on parasite sharing. PLoS ONE, 10, e0117909.Google Scholar
Pirson, C., Jones, G.J., Steinbach, S., Besra, G.S. & Vordermeier, H.M. (2012) Differential effects of Mycobacterium bovis-derived polar and apolar lipid fractions on bovine innate immune cells. Veterinary Research, 43, 54.Google Scholar
Pitchford, R.J. (1976) Preliminary observations on the distribution, definitive hosts and possible relation with other schistosomes, of Schistosoma margrebowiei, Le Roux, 1933 and Schistosoma leiperi, Le Roux, 1955. Journal of Helminthology, 50, 111123.Google Scholar
Pollock, J.M., Rodgers, J.D., Welsh, M.D. & McNair, J. (2006) Pathogenesis of bovine tuberculosis: experimental models of infection. Veterinary Microbiology, 112, 141150.Google Scholar
Pospischil, A., Kaiser, C., Hofmann-Lehmann, R., et al. (2012) Evidence for Chlamydia in wild mammals of the Serengeti. Journal of Wildlife Diseases, 48, 10741078.Google Scholar
Poulin, R. (1996) Richness, nestedness, and randomness in parasite infracommunity structure. Oecologia, 105, 545551.Google Scholar
Prins, H.H.T. (1996) Ecology and Behaviour of the African Buffalo: Social Inequality and Decision Making. London: Chapman and Hall.Google Scholar
Reese, T.A., Bi, K., Kambal, A., et al. (2016) Sequential infection with common pathogens promotes human-like immune gene expression and altered vaccine response. Cell Host & Microbe, 19, 713719.CrossRefGoogle ScholarPubMed
Regidor-Cerrillo, J., Arranz-Solís, D., Benavides, J., et al. (2014) Neospora caninum infection during early pregnancy in cattle: how the isolate influences infection dynamics, clinical outcome and peripheral and local immune responses. Veterinary Research, 45, 10.Google Scholar
Renwick, A.R., White, P.C.L. & Bengis, R.G. (2007) Bovine tuberculosis in southern African wildlife: a multi-species host–pathogen system. Epidemiology and Infection, 135, 529540.Google Scholar
Rodwell, T.C., Kriek, N.P., Bengis, R.G., et al. (2001) Prevalence of bovine tuberculosis in African buffalo at Kruger National Park. Journal of Wildlife Diseases, 37, 258264.Google Scholar
Romano, A., Doria, N.A., Mendez, J., Sacks, D.L. & Peters, N.C. (2015) Cutaneous infection with Leishmania major mediates heterologous protection against visceral infection with Leishmania infantum. The Journal of Immunology, 195, 3816.CrossRefGoogle ScholarPubMed
Roug, A., Muse, E.A., Smith, W.A., et al. (2016) Demographics and parasites of African buffalo (Syncerus caffer Sparrman, 1779) in Ruaha National Park, Tanzania. African Journal of Ecology, 54, 146153.Google Scholar
Schmid-Hempel, P. (2017) Parasites and their social hosts. Trends in Parasitology, 33, 453462.Google Scholar
Segre, H., Ron, R., De Malach, N., et al. (2014) Competitive exclusion, beta diversity, and deterministic vs. stochastic drivers of community assembly. Ecology Letters, 17, 14001408.Google Scholar
Sinclair, A.R.E. (1977) The African Buffalo, A Study of Resource Limitation of Populations. Chicago, IL: Chicago University Press.Google Scholar
Stock, T. & Holmes, J.C. (1987) Dioecocestus asper (Cestoda: Dioecocestidae): an interference competitor in an enteric helminth community. The Journal of Parasitology, 73, 11161123.Google Scholar
Stock, T. & Holmes, J.C. (1988) Functional relationships and microhabitat distributions of enteric helminths of grebes (Podicipedidae): the evidence for interactive communities. The Journal of Parasitology, 74, 214227.Google Scholar
Su, Z., Segura, M., Morgan, K., Loredo-Osti, J.C. & Stevenson, M.M. (2005) Impairment of protective immunity to blood-stage malaria by concurrent nematode infection. Infection and Immunity, 73, 35313539.Google Scholar
Swanepoel, R. & Coetzer, J. (2004) Rift valley fever. Infectious Diseases of Livestock, 2, 10371070.Google Scholar
Taylor, M.D., van der Werf, N. & Maizels, R.M. (2012) T cells in helminth infection: the regulators and the regulated. Trends in Immunology, 33, 181189.Google Scholar
Taylor, W.A., Skinner, J.D. & Boomker, J. (2013) Nematodes of the small intestine of African buffaloes, Syncerus caffer, in the Kruger National Park, South Africa. Onderstepoort Journal of Veterinary Research, 80, 14.Google Scholar
Telfer, S. & Bown, K. (2012) The effects of invasion on parasite dynamics and communities. Functional Ecology, 26, 12881299.Google Scholar
Telfer, S., Lambin, X., Birtles, R., et al. (2010) Species interactions in a parasite community drive infection risk in a wildlife population. Science (New York, N.Y.), 330, 243246.Google Scholar
Thacker, T.C., Palmer, M.V. & Waters, W.R. (2007) Associations between cytokine gene expression and pathology in Mycobacterium bovis infected cattle. Veterinary Immunology and Immunopathology, 119, 204213.Google Scholar
Tilman, D. (1994) Competition and biodiversity in spatially structured habitats. Ecology, 75, 216.Google Scholar
Walsh, J.R., Carpenter, S.R. & Vander Zanden, M.J. (2016) Invasive species triggers a massive loss of ecosystem services through a trophic cascade. Proceedings of the National Academy of Sciences of the United States of America, 113, 40814085.Google Scholar
Walson, J.L., Otieno, P.A., Mbuchi, M., et al. (2008) Albendazole treatment of HIV-1 and helminth co-infection: a randomized, double blind, placebo-controlled trial. AIDS (London, England), 22, 16011609.Google Scholar
Waters, W.R., Palmer, M.V., Thacker, T.C., et al. (2011) Tuberculosis immunity: opportunities from studies with cattle. Clinical Developmental Immunology, 2011, 768542.Google Scholar
Welsh, M.D., Cunningham, R.T., Corbett, D.M., et al. (2005) Influence of pathological progression on the balance between cellular and humoral immune responses in bovine tuberculosis. Immunology, 114, 101111.Google Scholar
Xu, D.-H., Pridgeon, J.W., Klesius, P.H. & Shoemaker, C.A. (2012) Parasitism by protozoan Ichthyophthirius multifiliis enhanced invasion of Aeromonas hydrophila in tissues of channel catfish. Veterinary Parasitology, 184, 101107.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

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

Available formats
×

Save book to Google Drive

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

Available formats
×