Hostname: page-component-7c8c6479df-7qhmt Total loading time: 0 Render date: 2024-03-27T09:40:17.429Z Has data issue: false hasContentIssue false

The relative importance of host characteristics and co-infection in generating variation in Heligmosomoides polygyrus fecundity

Published online by Cambridge University Press:  29 January 2010

L. T. LUONG*
Affiliation:
Center for Infectious Disease Dynamics, Penn State University, University Park, PA16802USA
S. E. PERKINS
Affiliation:
Center for Infectious Disease Dynamics, Penn State University, University Park, PA16802USA
D. A. GREAR
Affiliation:
Center for Infectious Disease Dynamics, Penn State University, University Park, PA16802USA
A. RIZZOLI
Affiliation:
Edmund Mach Foundation, IASMA Research and Innovation Centre, S. Michele all'Adige, Trento, Trentino, Italy
P. J. HUDSON
Affiliation:
Center for Infectious Disease Dynamics, Penn State University, University Park, PA16802USA
*
*Corresponding author: Center for Infectious Disease Dynamics, Penn State University, University Park, PA16802USA. Tel: +814 865 0522. Fax: +814 865 9131. E-mail: ltl1@psu.edu

Summary

We examined the relative importance of intrinsic host factors and microparasite co-infection in generating variation in Heligmosomoides polygyrus fecundity, a parameter that serves as a proxy for infectiousness. We undertook extensive trapping of Apodemus flavicollis, the yellow-necked mouse in the woodlands of the Italian Alps and recorded eggs in utero from the dominant nematode species H. polygyrus, and tested for the presence of five microparasite infections. The results showed that sex and breeding status interact, such that males in breeding condition harboured more fecund nematodes than other hosts; in particular, worms from breeding males had, on average, 52% more eggs in utero than worms from non-breeding males. In contrast, we found a weak relationship between intensity and body mass, and no relationship between intensity and sex or intensity and breeding condition. We did not find any evidence to support the hypothesis that co-infection with microparasites contributed to variation in worm fecundity in this system. The age-intensity profiles for mice singly-infected with H. polygyrus and those co-infected with the nematode and at least one microparasite were both convex and not statistically different from each other. We concluded that intrinsic differences between hosts, specifically with regard to sex and breeding condition, contribute relatively more to the variation in worm fecundity than parasite co-infection status.

Type
Research Article
Copyright
Copyright © Cambridge University Press 2010

Access options

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

References

REFERENCES

Anderson, M. (1994). Sexual Selection. Princeton University Press. Princeton, NJ, USA.Google Scholar
Anderson, R. M. and Gordon, D. M. (1982). Processes influencing the distribution of parasite numbers within host populations with special emphasis on parasite-induced host mortalities. Parasitology 85, 373398.Google Scholar
Anderson, R. M. and May, R. M. (1991). Infectious Diseases of Humans: Dynamics and Control. Oxford University Press, Oxford, UK.Google Scholar
Anderson, R. M. and Schad, G. A. (1985). Hookworm burdens and fecal egg counts: an analysis of the biological basis of variation. Transactions of the Royal Society of Tropical Medicine and Hygiene 79, 812825.Google Scholar
Bajer, A., Behnke, J., Pawelczyk, A., Kulis, K., Sereda, M. J. and Sinski, E. (2005). Medium-term temporal stability of the helminth component community structure in bank voles (Clethrionomys glareolus) from the Mazury Lake District region of Poland. Parasitology 130, 213228.Google Scholar
Bassetti, S., Bischoff, W. E. and Sherertz, R. J. (2005). Are SARS superspreaders cloud adults? Emerging Infectious Disease: http://www.cdc.gov/ncidod/EID/vol11no04/04-0639.htm.Google Scholar
Burnham, K. P. and Anderson, D. R. (2002). Model Selection and Inference: a Practical Information-Theorectical Approach, 2nd Edn. Springer-Verlag, New York, USA.Google Scholar
Cattadori, I. M., Albert, R. and Boag, B. (2007). Variation in host susceptibility and infectiousness generated by co-infection: the myxoma-Trichostrongylus retortaeformis case in wild rabbits. Journal of the Royal Society Interface 4, 831840.Google Scholar
Chase-Topping, M., Gally, D., Low, C., Matthews, L. and Woolhouse, M. (2008). Super-shedding and the link between human infection and livestock carriage of Escherichia coli O157. Nature Reviews Microbiology 6, 904912.Google Scholar
Chylinski, C., Boag, B., Stear, M. J. and Cattadori, I. M. (2009). Effects of host characteristics and parasite intensity on growth and fecundity of Trichostrongylus retortaeformis infections in rabbits. Parasitology 136, 117123.Google Scholar
Cox, F. E. G. (2001). Concomitant infections, parasites and immune responses. Parasitology 122, S23S38.Google Scholar
Crouch, A. C., Baxby, D., McCracken, C. M., Gaskell, R. M. and Bennett, M. (1995). Serological evidence for the reservoir hosts of cowpox virus in British wildlife. Epidemiology and Infection 115, 185191.Google Scholar
Dare, O. K. and Forbes, M. R. (2008). Rates of development in male and female Wood Frogs and patterns of parasitism by lung nematodes. Parasitology 135, 385393.Google Scholar
Dezfuli, B. S., Giari, L. and Poulin, R. (2001). Costs of intraspecific and interspecific host sharing in acanthocephalan cystacanths. Parasitology 122, 483489.Google Scholar
Elftman, M. D., Norbury, C. C., Bonneau, R. H. and Truckenmillar, M. E. (2007). Corticosterone impairs dendritic cell maturation and function. Immunology 122, 279290.Google Scholar
Elton, C. (1931). The study of epidemic diseases among wild animals. Journal of Hygiene 31, 435456.Google Scholar
Euzeby, J. (1982). From Echinococcus-Taenia biology in carnivore to man hydatidosis etiology and epidemiology. Revue De Medecine Veterinaire 133, 8394.Google Scholar
Ferrari, N., Cattadori, I. M., Nespereira, J., Rizzoli, A. and Hudson, P. J. (2004). The role of host sex in parasite dynamics: field experiments on the yellow-necked mouse Apodemus flavicollis. Ecology Letters 7, 8894.Google Scholar
Ferrari, N., Rosa, R., Pugliesa, A. and Hudson, P. J. (2007). The role of sex in parasite dynamics: Model simulations on transmission of Heligmosomoides polygyrus in populations of yellow-necked mice, Apodemus flavicollis. International Journal for Parasitology 37, 341349.Google Scholar
Galvani, A. P. and May, R. M. (2005). Epidemiology – dimensions of superspreading. Nature, London 438, 293295.Google Scholar
Goater, C. P. (1992). Experimental population dynamics of Rhabdias bufonis (Nematoda) in toads (Bufo bufo) – density-dependence in the primary infection. Parasitology 104, 179187.Google Scholar
Graham, A. L. (2008). Ecological rules governing helminth-microparasite co-infection. Proceedings of the National Academy of Sciences, USA 105, 566570.Google Scholar
Gregory, R. D. (1991). Parasite epidemiology and host population growth – Heligmosomoides polygyrus (Nematoda) in enclosed wood mouse populations. Journal of Animal Ecology 60, 805821.Google Scholar
Gregory, R. D., Keymer, A. E. and Clarke, J. R. (1990). Genetics, sex and exposure – the ecology of Heligmosomoides poylgyrus (Nematoda) in the wood mouse. Journal of Animal Ecology 59, 363378.Google Scholar
Gregory, R. D., Montgomery, S. S. J. and Montgomery, W. I. (1992). Population biology of Heligmosomoides polygyrus (Nematoda) in the wood mouse. Journal of Animal Ecology 61, 749757.Google Scholar
Harder, A., Wunderlich, F. and Marinovski, P. (1992). Effects of testosterone on Heterakis spumosa infections in mice. Parasitology 105, 335342.Google Scholar
Harvey, S. C., Paterson, S. and Viney, M. E. (1999). Heterogeneity in the distribution of Strongyloides ratti infective stages among the faecal pellets of rats. Parasitology 119, 227235.Google Scholar
Hudson, P. J. and Dobson, A. P. (1995). Macroparasites: observed patterns. In Ecology of Infectious Diseases in Natural Population (ed. Grenfell, B. T. and Dobson, A. P.), pp. 144176. Cambridge University Press, Cambridge, UK.Google Scholar
Janeway, C. A. and Travers, P. (1996). Immunobiology: The Immune System in Health and Disease. 2nd Edn. Garland Publishing, New York, NY, USA.Google Scholar
Keeling, M. J., Woolhouse, M. E. J., Shaw, D. J., Matthews, L., Chase-Topping, M., Haydon, D. T., Cornell, S. J., Kappey, J., Wilesmith, J. and Grenfell, B. T. (2001). Dynamics of the 2001 UK foot and mouth epidemic: Stochastic dispersal in a heterogeneous landscape. Science 294, 813817.Google Scholar
Keymer, A. E. and Hiorns, R. W. (1986). Heligmosomoides polygyrus (Nematoda) – the dynamics of primary and repeated infection in outbred mice. Proceedings of the Royal Society of London, B 229, 4767.Google Scholar
Keymer, A. E. and Slater, A. F. G. (1987). Helminth fecundity – density dependence or statistical illusion. Parasitology Today 3, 5658.Google Scholar
Keymer, A. E. and Tarlton, A. B. (1991). The population dynamics of acquired immunity to Heligmosomoides polygyrus in the laboratory mouse – strain, diet and exposure. Parasitology 103, 121126.Google Scholar
Lello, J., Boag, B., Fenton, A., Stevenson, I. R. and Hudson, P. J. (2004). Competition and mutualism among the gut helminths of a mammalian host. Nature, London 428, 840844.Google Scholar
Liz, J. S., Anderes, L., Sumner, J. W., Massung, R. F., Gern, L., Rutti, B. and Brossard, M. (2000). PCR detection of granulocytic ehrlichiae in Ixodes ricinus ticks and wild small mammals in western Switzerland. Journal of Clinical Microbiology 38, 10021007.Google Scholar
Luong, L. T., Grear, D. A. and Hudson, P. J. (2009). Male hosts are responsible for the transmission of a trophically transmitted parasite, Pterygodermatites peromysci, to the intermediate host in the absence of sex-biased infection. International Journal for Parasitology 39, 12631268.Google Scholar
Marcogliese, D. J. (1997). Fecundity of sealworm (Pseudoterranova decipiens) infecting grey seals (Halichoerus grypus) in the Gulf of St. Lawrence, Canada: Lack of density-dependent effects. International Journal for Parasitology 27, 14011409.Google Scholar
Matthews, L., McKendrick, I. J., Ternent, H., Gunn, G. J., Synge, B. and Woolhouse, M. E. J. (2006). Super-shedding cattle and the transmission dynamics of Escherichia coli O157. Epidemiology and Infection 134, 131142.Google Scholar
McClelland, G. (1980). Phocanema decipiens – growth, reproduction, and survival in seals. Experimental Parasitology 49, 175187.Google Scholar
Michel, J. F. (1963). Phenomena of host resistance and course of infection of Ostertagia osteragi in calves. Parasitology 53, 6384.Google Scholar
Montes, M., Sanchez, C., Verdonck, K., Lake, J. E., Gonzalez, E., et al. (2009). Regulatory T cell expansion in HTLV-1 and strongyloidiasis co-infection is associated with reduced IL-5 responses to Strongyloides stercoralis antigen. PLoS Neglected Tropical Diseases 3, e456. doi:10.1371/journal.pntd.0000456.Google Scholar
Noble, G. A. (1961). Stress and parasitism. I. A preliminary investigation of effects of stress on ground squirrels and their parasites. Experimental Parasitology 11, 6367.Google Scholar
Noble, G. A. (1962). Stress and parasitism. II. Effect of crowding and fighting among ground squirrels on their coccidia and trichomonads. Experimental Parasitology 12, 368371.Google Scholar
Noble, G. A. (1966). Stress and parasitism. III. Reduced night temperature and effect on pinworms of ground squirrels. Experimental Parasitology 18, 6162.Google Scholar
O'sullivan, H. M., Smal, C. M. and Fairley, J. S. (1984). A study of parasite infestations in populations of small rodents (Apodemus sylvaticus and Clethrionomys glareolus) on Ross Island, Killarney. Journal of Life Sciences Royal Dublin Society 5, 2942.Google Scholar
Paterson, S. and Barber, R. (2007). Experimental evolution of parasite life-history traits in Strongyloides ratti (Nematoda). Proceedings of the Royal Society London, B 274, 14671474.Google Scholar
Paterson, S. and Viney, M. E. (2002). Host immune responses are necessary for density dependence in nematode infections. Parasitology 125, 283292.Google Scholar
Perkins, S. E., Cagnacci, F., Stradiotto, A., Arnoldi, D. and Hudons, P. J. (2009). Comparison of social networks derived from ecological data: implications for inferring infectious disease dynamics. Journal of Animal Ecology 78, 10151022.Google Scholar
Perkins, S. E., Cattadori, I. M., Tagliapietra, V., Rizzoli, A. P. and Hudson, P. J. (2003). Empirical evidence for key hosts in persistence of a tick-borne disease. International Journal for Parasitology 33, 909917.Google Scholar
Perkins, S. E., Cattadori, I. M., Tagliapietra, V., Rizzoli, A. P. and Hudson, P. J. (2006). Localized deer absence leads to tick amplification. Ecology 87, 19811986.Google Scholar
Poulin, R. (1996). Helminth growth in vertebrate hosts: Does host sex matter? International Journal for Parasitology 26, 13111315.Google Scholar
Quinnell, R. J. (1992). The population dynamics of Heligmosomoides polygyrus in an enclosure population of wood mice. Journal of Animal Ecology 61, 669679.Google Scholar
Ractliffe, L. H. and LeJambre, L. F. (1971). Increase of rate of egg production with growth in some intestinal nematodes of sheep and horses. International Journal for Parasitology 1, 153156.Google Scholar
Rasband, W. (2007). ImageJ. U.S. National Institutes of Health, Bethesda, MD, USA.Google Scholar
Rizzoli, A., Neteler, M., Rosa, R., Versini, W., Cristofolini, A., Bregoli, M., Buckley, A. and Gould, E. A. (2007). Early detection of tick-borne encephalitis virus spatial distribution and activity in the province of Trento, northern Italy. Geospatial Health 1, 169176.Google Scholar
Scott, M. E. (1987). Regulation of mouse colony abundance by Heligmosomoides polygyrus. Parasitology 95, 111124.Google Scholar
Scott, M. E. (1990). An experimental and theoretical study of the dynamics of mouse nematode (Heligmosomoides polygyrus) interaction. Parasitology 101, 7592.Google Scholar
Shaw, D. J. and Dobson, A. P. (1995). Patterns of macroparasite abundance and aggregation in wildlife populations: A quantitative review. Parasitology 111, S111S133.Google Scholar
Smith, J. W. (1988). An electronic method for estimating the vaginal and uterine egg content of nematodes, with special reference to ascaridoids. Canadian Journal of Zoology 66, 22532254.Google Scholar
Smith, D. L., McKenzie, F. E., Snow, R. W. and Hay, S. I. (2007). Revisiting the basic reproductive number for malaria and its implications for malaria control. PLoS Biology 5, e42. doi:10.1371/journal.pbio.0050042.Google Scholar
Stear, M. J., Bairden, K., Duncan, J. L., Holmes, P. H., McKellar, Q. A., Park, M., Strain, S., Murray, M., Bishop, S. C. and Gettinby, G. (1997). How hosts control worms. Nature 389, 27.Google Scholar
Stear, M. J. and Bishop, S. C. (1999). The curvilinear relationship between worm length and fecundity of Teladorsagia circumcincta. International Journal for Parasitology 29, 777780.Google Scholar
Stear, M. J., Bishop, S. C., Doligalska, M., Duncan, J. L., Holmes, P. H., Irvine, J., McCririe, L., McKellar, Q. A., Sinski, E. and Murray, M. (1995). Regulation of egg production, worm burden, worm length and worm fecundity by host responses in sheep infected with Ostertagia circumcincta. Parasite Immunology 17, 643652.Google Scholar
Stein, M. and Schleifer, S. J. (1985). Frontiers of stress research: stress and immunity. In Stress in Helath and Disease (ed. Zales, M.), pp. 97–114. Brunner/Mazel, New York, USA.Google Scholar
Swanson, J. A., Falvo, R. and Bone, L. W. (1984). Nippostrongylus brasiliensis – effects of testosterone on reproduction and establishment. International Journal for Parasitology 14, 241247.Google Scholar
Szalai, A. J. and Dick, T. A. (1989). Differences in numbers and inequalities in mass and fecundity during the egg producing period for raphid Ascaris acus (Nematoda, Anisakidae). Parasitology 98, 489495.Google Scholar
Tagliapietra, V., Rosa, R., Hauffe, H. C., Laakkonen, J., Voutilainen, L., Vapalahti, O., Vaheri, A., Henttonen, H. and Rizzoli, A. (2009). Spatial and temporal dynamics of lymphocytic choriomeningitis virus in wild rodents, Northern Italy. Emerging Infectious Diseases 15, 10191025.Google Scholar
Tanguay, G. V. and Scott, M. E. (1992). Factors generating aggregation of Heligmosomoides polygyrus (Nematoda) in laboratory mice. Parasitology 104, 519529.Google Scholar
Telfer, S., Bennett, M., Carslake, D., Helyar, S. and Begon, M. (2007). The dynamics of murid gammaherpesvirus 4 within wild, sympatric populations of bank voles and wood mice. Journal of Wildlife Diseases 43, 3239.Google Scholar
Tompkins, D. and Hudson, P. J. (1999). Regulation of nematode fecundity in the ring-necked pheasant (Phasianus colchicus): not just density dependence. Parasitology 118, 417423.Google Scholar
Viney, M. (2002). How do host immune responses affect nematode infections? Trends in Parasitology 18, 6366.Google Scholar
Wilson, K., Bjornstad, O. N., Dobson, A. P., Merler, S., Poglayen, G., Randolf, S. E., Read, A. F. and Skorping, A. (2001). Heterogeneities in macroparasite infections: patterns and processes. In The Ecology of Wildlife Disease (ed. Hudson, P. J., Rizzoli, A., Grenfell, B. T., Heesterbeek, H. and Dobson, A. P.), pp. 6–44. Oxford University Press, Oxford, UK.Google Scholar
Woolhouse, M. E. J., Dye, C., Etard, J. F., Smith, T., Charlwood, J. D., Garnett, G. P., Hagan, P., Hii, J. L. K., Ndhlovu, P. D., Quinnell, R. J., Watts, C. H., Chandiwana, S. K. and 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, USA 94, 338342.Google Scholar
Woolhouse, M. E. J., Watts, C. H. and Chandiwana, S. K. (1991). Heterogeneities in transmission rates and the epidemiology of Schistosome infection. Proceedings of the Royal Society of London. B 245, 109114.Google Scholar
Zuk, M. (1990). Reproductive strategies and disease susceptibility: an evolutionary viewpoint. Parasitology Today 6, 231233.Google Scholar
Zuk, M. and McKean, K. A. (1996). Sex differences in parasite infections: Patterns and processes. International Journal for Parasitology 26, 10091023.Google Scholar