Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-25T11:52:31.631Z Has data issue: false hasContentIssue false

Tracking year-to-year changes in intestinal nematode communities of rufous mouse lemurs (Microcebus rufus)

Published online by Cambridge University Press:  20 April 2015

TUOMAS AIVELO*
Affiliation:
Institute of Biotechnology, University of Helsinki, P.O. Box 56, FI-00014 Helsinki, Finland Department of Biosciences, University of Helsinki, Helsinki, Finland
ALAN MEDLAR
Affiliation:
Institute of Biotechnology, University of Helsinki, P.O. Box 56, FI-00014 Helsinki, Finland
ARI LÖYTYNOJA
Affiliation:
Institute of Biotechnology, University of Helsinki, P.O. Box 56, FI-00014 Helsinki, Finland
JUHA LAAKKONEN
Affiliation:
Department of Veterinary Biosciences, University of Helsinki, P.O. Box 66, FI-00014 Helsinki, Finland
JUKKA JERNVALL
Affiliation:
Institute of Biotechnology, University of Helsinki, P.O. Box 56, FI-00014 Helsinki, Finland
*
*Corresponding author. Institute of Biotechnology, University of Helsinki, P.O. Box 56, FI-00014 Helsinki, Finland. E-mail: tuomas.aivelo@helsinki.fi

Summary

While it is known that intestinal parasite communities vary in their composition over time, there is a lack of studies addressing how variation in component communities (between-hosts) manifests in infracommunities (within-host) during the host lifespan. In this study, we investigate the changes in the intestinal parasite infracommunities in wild-living rufous mouse lemurs (Microcebus rufus) from Ranomafana National Park in southeastern Madagascar from 2010 to 2012. We used high-throughput barcoding of the 18S rRNA gene to interrogate parasite community structure. Our results show that in these nematode communities, there were two frequently occurring putative species and four rarer putative species. All putative species were randomly distributed over host individuals and they did not occur in clear temporal patterns. For the individuals caught in at least two different years, there was high turnover of putative species and high variation in fecal egg counts. Our study shows that while there was remarkable variation in infracommunities over time, the component community was relatively stable. Nevertheless, the patterns of prevalence varied substantially between years in each component community.

Type
Research Article
Copyright
Crown Copyright © Cambridge University Press 2016 

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

Alizon, S. (2013). Parasite co-transmission and the evolutionary epidemiology of virulence. Evolution 67, 921933.Google Scholar
Anderson, R. C., Chabaud, A. G. and Willmott, S. (2009). Keys to the Nematode Parasites of Vertebrates, 1st Edn. CAB International, Wallingford, UK.CrossRefGoogle Scholar
Atsalis, S. (2008). A Natural History of the Brown Mouse Lemur. Pearson Education, Upper Saddle River, NJ, USA.Google Scholar
Baermann, G. (1917). Eine einfache methode zur auffindung von Ankylostomum (Nematoden) larven in erdproben. Tijdschrid Ergeneeskd 57, 131137.Google Scholar
Balmer, O. and Caccone, A. (2008). Multiple-strain infections of Trypanosoma brucei across Africa. Acta Tropica 107, 275279.CrossRefGoogle ScholarPubMed
Balmer, O. and Tanner, M. (2011). Prevalence and implications of multiple-strain infections. Lancet Infectious Diseases 11, 868878.Google Scholar
Barnes, E. M. (1970). Effect of hibernation intestinal flora. The American Journal of Clinical Nutrition 23, 15191524.Google Scholar
Bates, D., Maechler, M., Bolker, B. and Walker, S. (2014). lme4: linear mixed-effects models using Eigen and S4. R package version 1.1–7. http://CRAN.R-project.org/package=lme4 Google Scholar
Behnke, J. M., Bajer, A., Harris, P. D., Newington, L., Pidgeon, E., Rowlands, G., Sheriff, C., Kuliś-Malkowska, K., Siński, E., Gilbert, F. S. and Barnard, C. J. (2008 a). Temporal and between-site variation in helminth communities of bank voles (Myodes glareolus) from N.E. Poland. 2. The infracommunity level. Parasitology 135, 9991018.CrossRefGoogle ScholarPubMed
Behnke, J. M., Bajer, A., Harris, P. D., Newington, L., Pidgeon, E., Rowlands, G., Sheriff, C., Kuliś-Malkowska, K., Siński, E., Gilbert, F. S. and Barnard, C. J. (2008 b). Temporal and between-site variation in helminth communities of bank voles (Myodes glareolus) from N.E. Poland. 1. Regional fauna and component community levels. Parasitology 135, 985997.Google Scholar
Bhadury, P. and Austen, M. C. (2010). Barcoding marine nematodes: an improved set of nematode 18S rRNA primers to overcome eukaryotic co-interference. Hydrobiologia 641, 245251.Google Scholar
Blaxter, M., Mann, J., Chapman, T., Thomas, F., Whitton, C., Floyd, R. and Abebe, E. (2005). Defining operational taxonomic units using DNA barcode data. Philosophical Transactions of the Royal Society B: Biological Sciences 360, 19351943.Google Scholar
Bouma, H. R., Carey, H. V. and Kroese, F. G. M. (2010). Hibernation: the immune system at rest? Journal of Leukocyte Biology 88, 619624.Google Scholar
Bush, A. O., Lafferty, K. D., Lotz, J. M. and Shostak, A. W. (1997). Parasitology meets ecology on its own terms: Margolis et al. revisited. The Journal of Parasitology 83, 575583.Google Scholar
Callait, M. and Gauthier, D. (2000). Parasite adaptations to hibernation in Alpine marmots (Marmota marmota). In Life in the Cold (ed. Heldmaier, G. and Klingenspor, M.), pp. 139146. Springer-Verlag, Berlin.Google Scholar
Carey, H. V. (1990). Seasonal changes in mucosal structure and function in ground squirrel intestine. American Journal of Physiology – Regulatory, Integrative and Comparative Physiology 259, R385R392.CrossRefGoogle ScholarPubMed
Carey, H. V., Pike, A. C., Weber, J. L., Turner, A. K., Visser, S. C., Beijer-Liefers, H. R., Bouma, H. R. and Kroese, F. G. M. (2012). Impact of hibernation on gut microbiota and intestinal barrier function in ground squirrels. In Living in a Seasonal World: Thermoregulatory and Metabolic Adaptations (ed. Ruf, T., Bieber, C., Arnold, W. and Millesi, E.), pp. 281291. Springer, Heidelberg, Germany.Google Scholar
Chute, R. M. (1960). Overwintering of helminths in hibernating animals. Journal of Parasitology, 46, 539.CrossRefGoogle Scholar
Clutton-Brock, T. H. and Pemberton, J. M. (2004). Individuals and populations. In Soay Sheep: Dynamics and Selection in an Island Population (ed. Clutton-Brock, T. H. and Pemberton, J. M.), pp. 116. CUP, Cambridge, UK.Google Scholar
Coggings, J. R., Tedesco, J. L. and Ruppercht, C. E. (1982). Seasonal changes and overwintering of parasites in the bat, Myotis lucifugus (Le Conte), in a Wisconsin hibernaculum. American Midland Naturalist 107, 305315.Google Scholar
Conover, W. J. (1999). Practical Nonparametric Statistics, 3rd Edn. John Wiley & Sons, Hoboken, NJ, USA.Google Scholar
Craig, B. H., Pilkington, J. G. and Pemberton, J. M. (2006). Gastrointestinal nematode species burdens and host mortality in a feral sheep population. Parasitology 133, 485496.Google Scholar
Craig, B. H., Jones, O. R., Pilkington, J. G. and Pemberton, J. M. (2009). Re-establishment of nematode infra-community and host survivorship in wild Soay sheep following anthelmintic treatment. Veterinary Parasitology 161, 4752.Google Scholar
Dausmann, K. H., Glos, J., Ganzhorn, J. U. and Heldmaier, G. (2005). Hibernation in the tropics: lessons from a primate. Journal of Comparative Physiology B: Biochemical, Systems, and Environmental Physiology 175, 147155.CrossRefGoogle ScholarPubMed
De Roij, J. and MacColl, A. D. C. (2012). Consistent differences in macroparasite community composition among populations of three-spined sticklebacks, Gasterosteus aculeatus L. Parasitology 139, 14781491.Google Scholar
Dobson, A. (1986). Inequalities in the individual reproductive success of parasites. Parasitology 92, 675682.Google Scholar
Dove, A. D. M. and Cribb, T. H. (2006). Species accumulation curves and their applications in parasite ecology. Trends in Parasitology 22, 568574.Google Scholar
Dreyer, G., Fernandes-Silva, E., Alves, S., Rocha, A., Albuquerque, R. and Addiss, D. (1996). Patterns of detection of Strongyloides stercoralis in stool specimens: implications for diagnosis and clinical trials. Journal of Clinical Microbiology 34, 25692571.Google Scholar
Dubey, J. P., Moura, L., Majumdar, D., Sundar, N., Velmurugan, G. V., Kwok, O. C. H., Kelly, P., Krecek, R. C. and Su, C. (2009). Isolation and characterization of viable Toxoplasma gondii isolates revealed possible high frequency of mixed infection in feral cats (Felis domesticus) from St Kitts, West Indies. Parasitology 136, 589594.Google Scholar
Ebbert, M. A., McGrew, W. C. and Marchant, L. F. (2013). Community composition, correlations among taxa, prevalence, and richness in gastrointestinal parasites of baboons in Senegal, West Africa. Primates 54, 183189.Google Scholar
Edgar, R. C., Haas, B. J., Clemente, J. C., Quince, C. and Knight, R. (2011). UCHIME improves sensitivity and speed of chimera detection. Bioinformatics 27, 21942200.Google Scholar
Fenton, A., Viney, M. E. and Lello, J. (2010). Detecting interspecific macroparasite interactions from ecological data: patterns and process. Ecology Letters 13, 606615.Google Scholar
Fenton, A., Knowles, S. C. L., Petchey, O. L. and Pedersen, A. (2014). The reliability of observational approaches for detecting interspecific parasite interactions: comparison with experimental results. International Journal of Parasitology 44, 437445.Google Scholar
Gau, R. J., Kutz, S. and Elkin, B. T. (1999). Parasites in grizzly bears from the central Canadian Arctic. Journal of Wildlife Diseases 35, 618621.Google Scholar
Gibbons, L. M., Jacobs, D. E., Fox, M. T. and Hansen, J. (2014). The RVC/FAO Guide to Veterinary Diagnostic Parasitology. http://www.rvc.ac.uk/review/Parasitology/Index/Index.htm.Google Scholar
Gillespie, T. R. (2006). Noninvasive assessment of gastrointestinal parasite infections in free-ranging primates. International Journal of Primatology 27, 11291143.Google Scholar
Gotelli, N. J. and Colwell, R. K. (2001). Quantifying biodiversity: procedures and pitfalls in the measurement and comparison of species richness. Ecology Letters 4, 379391.Google Scholar
Grenfell, B. T., Wilson, K., Isham, V. S., Boyd, H. E. and Dietz, K. (1995). Modelling patterns of parasite aggregation in natural populations: trichostrongylid nematode-ruminant interactions as a case study. Parasitology 111 (Suppl.), S135S151.Google Scholar
Griffiths, E. C., Pedersen, A. B., Fenton, A. and Petchey, O. L. (2011). The nature and consequences of coinfection in humans. The Journal of Infection 63, 200206.Google Scholar
Haukisalmi, V., Henttonen, H. and Tenora, F. (1988). Population dynamics of common and rare helminths in cyclic vole populations. Journal of Animal Ecology 57, 807825.Google Scholar
Hayward, A. D. (2013). Causes and consequences of intra- and inter-host heterogeneity in defense against nematodes. Parasite Immunology 35, 362373.CrossRefGoogle ScholarPubMed
Hayward, A. D., Wilson, A. J., Pilkington, J. G., Pemberton, J. M. and Kruuk, L. E. B. (2009). Ageing in a variable habitat: environmental stress affects senescence in parasite resistance in St Kilda Soay sheep. Proceedings of the Royal Society B: Biological Sciences 276, 34773485.Google Scholar
Højsgaard, S., Halekoh, U. and Yan, J. (2006). The R package geepack for generalized estimating equations. Journal of Statistical Software 15, 111.Google Scholar
Jones, O. R., Crawley, M. J., Pilkington, J. G. and Pemberton, J. M. (2005). Predictors of early survival in Soay sheep: cohort-, maternal- and individual-level variation. Proceedings of the Royal Society B: Biological sciences 272, 26192625.Google Scholar
Klein, S. (2004). Hormonal and immunological mechanisms mediating sex differences in parasite infection. Parasite Immunology 26, 247264.Google Scholar
Kurtz, C. C. and Carey, H. V. (2007). Seasonal changes in the intestinal immune system of hibernating ground squirrels. Developmental and Comparative Immunology 31, 415428.CrossRefGoogle ScholarPubMed
Liang, K.-Y. and Zeger, S. L. (1986). Longitudinal data analysis using generalized linear models. Biometrika 73, 1322.Google Scholar
Löytynoja, A., Vilella, A. J. and Goldman, N. (2012). Accurate extension of multiple sequence alignments using a phylogeny-aware graph algorithm. Bioinformatics 28, 16841691.Google Scholar
Medlar, A., Aivelo, T. and Löytynoja, A. (2014). Séance: reference-based phylogenetic analysis for 18S rRNA studies. BMC Evolutionary Biology 14, 235.Google Scholar
Nakagawa, S. and Schielzeth, H. (2010). Repeatability for Gaussian and non-Gaussian data: a practical guide for biologists. Biological Reviews of the Cambridge Philosophical Society 85, 935956.Google Scholar
Nishigori, M. (1928). The factors which influence the external development of Strongyloides stercoralis and on auto-infection with this parasite. Journal of Formosan Medical Association 247, 156.Google Scholar
Oksanen, J., Blanchet, F. G., Kindt, R., Legendre, P., Minchin, P. R., O'Hara, R. B., Simpson, G. L., Solymos, P., Stevens, M. H. H. and Wagner, H. (2013). Vegan: Community Ecology Package. R package version 2.0–1o0. http://CRAN.R-project.org/package=vegan Google Scholar
Pilosof, S., Fortuna, M. A., Vinarski, M. V., Korallo-Vinarskaya, N. P. and Krasnov, B. R. (2013). Temporal dynamics of direct reciprocal and indirect effects in a host-parasite network. The Journal of Animal Ecology 82, 987996.Google Scholar
Poulin, R. (1997). Population abundance and sex ratio in dioecious helminth parasites. Oecologia 111, 375380.Google Scholar
Quast, C., Pruesse, E., Yilmaz, P., Gerken, J., Schweer, T., Yarza, P., Peplies, J. and Glöckner, F. O. (2013). The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. Nucleic Acids Research 41, D590D596.CrossRefGoogle ScholarPubMed
Quince, C., Lanzen, A., Davenport, R. J. and Turnbaugh, P. J. (2011). Removing noise from pyrosequenced amplicons. BMC Bioinformatics 12, 38.Google Scholar
R Core Team (2013). R: a Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna, Austria. http://www.R-project.org/ Google Scholar
Raharivololona, B. M. and Ganzhorn, J. U. (2009). Gastrointestinal parasite infection of the gray mouse lemur (Microcebus murinus) in the littoral forest of Mandena, Madagascar: effects of forest fragmentation and degradation. Madagascar Conservation and Development 4, 103112.Google Scholar
Raharivololona, B. and Ganzhorn, J. (2010). Seasonal variations in gastrointestinal parasites excreted by the gray mouse lemur Microcebus murinus in Madagascar. Endangered Species Research 11, 113122.Google Scholar
Revelle, W. (2015). Psych: Procedures for Psychological, Psychometric and Personality Research. R package version 2.0–10. http://personality-project.org/r/psych/ Google Scholar
Rinaldi, L., Maurelli, M. P., Musella, V., Santaniello, A., Coles, G. C. and Cringoli, G. (2010). FLOTAC: an improved method for diagnosis of lungworm infection in sheep. Veterinary Parasitology 11, 395398.Google Scholar
Sambrook, J., Fritsch, E. F. and Maniatis, T. (1989). Molecular Cloning: a Laboratory Manual, 2nd Edn. Cold Spring Harbor Laboratory Press, Long Island, NY, USA.Google Scholar
Sandground, B. J. H. (1926). Biological studies on the life-cycle in the genus Strongyloides grassi, 1879. American Journal of Epidemiology 6, 337388.CrossRefGoogle 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
Simitch, T. and Petrovitch, Z. (1954). Ce qu'il advient avec les helminthes du Citellus citellus au cours du sommeil hibernal de ce rongeur. Rivista di Parassitologia 15, 655662.Google Scholar
Sokal, R. R. and Rohfl, F. J. (1995). The Principles and Practice of Statistics in Biological Research, 3rd Edn. W.H. Freeman, New York, USA.Google Scholar
Sonoyama, K., Fujiwara, R., Takemura, N., Ogasawara, T., Watanabe, J., Ito, H. and Morita, T. (2009). Response of gut microbiota to fasting and hibernation in Syrian hamsters. Applied and Environmental Microbiology 75, 64516456.Google Scholar
Stamatakis, A. (2006). RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22, 26882690.Google Scholar
Stear, M. J., Bishop, S. C., Duncan, J. L., Mckellar, Q. A. and Murray, M. (1995). The repeatability of faecal egg counts, peripheral eosinophil counts, and plasma pepsinogen concentrations during deliberate infection with Ostertagia circumcincta . International Journal for Parasitology 25, 375380.CrossRefGoogle ScholarPubMed
Stear, M. J., Bairden, K., Bishop, S. C., Gettinby, G., Mckellar, Q. A., Park, M., Strain, S. and Wallace, D. S. (1998). The processes influencing the distribution of parasitic nematodes among naturally infected lambs. Parasitology 117, 165171.Google Scholar
Stear, M. J., Singleton, D. and Matthews, L. (2011). An evolutionary perspective on gastrointestinal nematodes of sheep. Journal of Helminthology 85, 113120.Google Scholar
Stephenson, L. S., Latham, M. C. and Ottesen, E. A. (2000). Malnutrition and parasitic helminth infections. Parasitology 121, S23S38.Google Scholar
Telfer, S., Lambin, X., Birtles, R., Beldomenico, P., Burthe, S., Paterson, S. and Begon, M. (2010). Species interactions in a parasite community drive infection risk in a wildlife population. Science 330, 243246.Google Scholar
Uparanukraw, P., Phongsri, S. and Morakote, N. (1999). Fluctuations of larval excretion in Strongyloides stercoralis infection. The American Journal of Tropical Medicine and Hygiene 60, 967973.Google Scholar
Viney, M. E. and Graham, A. L. (2013). Patterns and processes in parasite co-infection. Advances in Parasitology 82, 321369.Google Scholar
Wickham, H. and Chang, W. (2014). ggplot2: an Implementation of the Grammar of Graphics. R package version 1.0.0. http://ggplot2.org/ Google Scholar
Wilson, K., Grenfell, B. T., Pilkington, J. G., Boyd, H. E. and Gulland, F. M. (2014). Parasites and their impact. In Soay Sheep: Dynamics and Selection in an Island Population (ed. Clutton-Brock, T. H. and Pemberton, J. M.), pp. 113165. Cambridge University Press, Cambridge, UK.Google Scholar
Wood, E. L. D., Matthews, J. B., Stephenson, S., Slote, M. and Nussey, D. H. (2013). Variation in fecal egg counts in horses managed for conservation purposes: individual egg shedding consistency, age effects and seasonal variation. Parasitology 140, 115128.Google Scholar
Wright, P. C. and Andriamihaja, B. (2002). Making a rain forest national park work in Madagascar: Ranomafana National Park and its long-term research commitment. In Making Parks Work: Strategies for Preserving Tropical Nature (ed. Terborgh, J., van Schaik, C., Davenport, L. and Rao, M.), pp. 112136. Island Press, Washington, DC, USA.Google Scholar
Zohdy, S., Gerber, B. D., Tecot, S., Blanco, M. B., Winchester, J. M., Wright, P. C. and Jernvall, J. (2014). Teeth, sex, and testosterone: aging in the world's smallest primate. PLoS ONE 9, e109528.CrossRefGoogle ScholarPubMed
Supplementary material: File

Aivelo supplementary material

Tables S1-S12

Download Aivelo supplementary material(File)
File 102.4 KB