Hostname: page-component-8448b6f56d-qsmjn Total loading time: 0 Render date: 2024-04-25T01:02:49.502Z Has data issue: false hasContentIssue false

Phylogeographic structure of Teretrius nigrescens (Coleoptera: Histeridae) predator of the invasive post harvest pest Prostephanus truncatus (Coleoptera: Bostrichidae)

Published online by Cambridge University Press:  15 April 2011

B.A. Omondi*
Affiliation:
Molecular Biologyand Biotechnology Unit, International Centre of Insect Physiology and Ecology, PO Box 30772–00100, Nairobi, Kenya School of Environmental Sciences and Development, North West University, Private Bag X6001, Potchefstroom 2520South Africa Biological Control Unit, International Centre of Insect Physiology and Ecology, PO Box 30772–00100, Nairobi, Kenya
J. van den Berg
Affiliation:
School of Environmental Sciences and Development, North West University, Private Bag X6001, Potchefstroom 2520South Africa
D. Masiga
Affiliation:
Molecular Biologyand Biotechnology Unit, International Centre of Insect Physiology and Ecology, PO Box 30772–00100, Nairobi, Kenya
F. Schulthess
Affiliation:
Biological Control Unit, International Centre of Insect Physiology and Ecology, PO Box 30772–00100, Nairobi, Kenya
*
*Author for correspondence Fax: +46-40-461991 E-mail: amanLGB@gmail.com

Abstract

The invasive larger grain borer Prostephanus truncatus (Horn) is the most important pest of farm-stored maize in Africa. It was introduced into the continent from Mesoamerica in the late 1970s and by 2008 had spread to at least 18 countries. Classical biological control using two populations of the predator Teretrius nigrescens Lewis achieved long-term and cost effective control in warm-humid areas, but not in cool and hot-dry zones. The present study investigated the phylogenetic relationships between geographical populations of the predator. Ten populations of T. nigrescens were studied using randomly amplified polymorphic DNA polymerase chain reaction (RAPD-PCR), sequence analysis of mitochondrial Cytochrme oxydase 1 (mtCOI) gene and ribosomal internally transcribed spacers (ITS) 1, 5.8S and ITS2. The mtCOI variation revealed two clades associated with geographical regions in Central America. It also reveals a significant isolation by distance between populations and considerable genetic shifts in laboratory rearing. RAPD-PCR did not reveal any potential SCAR diagnostic markers. The ITS variation mainly involved insertions and deletions of simple sequence repeats even within individuals. This study reveals the existence of two different mitochondrial lineages of the predator, associated with the geographical origin of populations distinguishable by fixed mutations on the mtCOI gene. The populations of T. nigrescens released in Africa belonged to two different clades from Meso America, namely south (released in West Africa) and north (released in eastern Africa). However, more polymorphic markers are required to clarify the observations in demographic time scales.

Type
Research Paper
Copyright
Copyright © Cambridge University Press 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

Allendorf, F.W. & Lundquist, L.L. (2003) Introduction: Population biology, evolution, and control of invasive species. Conservation Biology 17, 2430.CrossRefGoogle Scholar
Bacigalupe, D.L. (2009) Biological invasions and phenotypic evolution: a quantitative genetic perspective. Biological Invasions 11, 22432250.CrossRefGoogle Scholar
Bellows, T.S., Fisher, T.W. & Cattairone, L.E. (1999) Handbook of Biological Control: Principles and Application of Biological Control. San Diego, CA, USA, Academic Press.Google Scholar
Bensasson, D., Zhang, D.-X. & Hewitt, G.M. (2000) Frequent Assimilation of mitochondrial DNA by grasshopper nuclear genomes. Molecular Biology and Evolution 17, 406415.CrossRefGoogle ScholarPubMed
Benson, G. (1999) Tandem Repeats Finder: a program to analyze DNA sequences. Nucleic Acids Research 27, 573580.Google Scholar
Bigler, F. (1992) Quality Control in insect rearing systems. pp. 189210in Ochieng-Odero, J.P.R. (Ed.) Techniques of Insect Rearing for the Development of Integrated Pest and Vector Management Strategies. Proceedings of the International Group Training Course on Techniques of Insect Rearing for the Development of Integrated Pest and Vector Management, ICIPE, 16 March–3 April 1992, Nairobi, Kenya.Google Scholar
Böye, J.A. (1988) Autoökologische Untersuchungen zum Verhalten des Grossen Kornbohrers Prostephanus truncatus Horn (Coleoptera: Bostrichidae) in Costa Rica. PhD thesis, Christian-Albrechts University Kiel, Germany.Google Scholar
Böye, J. (1990) Ecological aspects of Prostephanus truncatus (Horn) (Coleoptera: Histeridae) in Costa Rica. pp. 7386in Markham, R.H. & Herren, H.R. (Eds) Proceedings of the IITA/FAO Coordination Meeting, 2–3 June 1989, Cotonou, Benin.Google Scholar
Clement, M., Posada, D. & Crandall, K. (2000) TCS: a computer program to estimate gene genealogies. Molecular Ecology 9, 16571660.CrossRefGoogle ScholarPubMed
Cognato, A.I., Harlin, A.D. & Fisher, M.L. (2003) Genetic structure among pinyon pine beetle populations (Scolytinae: Ips confusus). Environmental Entomology 32, 12621270.CrossRefGoogle Scholar
Giles, P.H., Hill, M.G., Nang'ayo, F.L.O., Farrell, G. & Kibata, G.N. (1996) Release and establishment of the predator Teretriosoma nigrescens Lewis for the biological control of Prostephanus truncatus (Horn) in Kenya. African Crop Science Journal 4, 325337.Google Scholar
Gomez-Zurita, J., Juan, C. & Petitpierre, E. (2000) Sequence, secondary structure and phylogenetic analyses of the ribosomal internal transcribed spacer 2 (ITS2) in the Timarcha leaf beetles (Coleoptera; Chrysomelidae). Insect Molecular Biology 9, 591604.CrossRefGoogle ScholarPubMed
Grevstad, F.S. (1999) Experimental invasions using biological control introductions: The influence of release size on the chance of population establishment. Biological Invasions 1, 313323.Google Scholar
Gueye, M.T., Goergen, G., Badiane, D., Hell, K. & Lamboni, L. (2008) First report on occurrence of the larger grain borer Prostephanus truncatus (Horn) (Coleoptera: Bostrychidae) in Senegal. African Entomology 16, 309311.CrossRefGoogle Scholar
Hall, T.A. (1999) BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symposium Series 4, 9598.Google Scholar
Hill, M.G., Nang'ayo, F.L.O. & Wright, D.J. (2003) Biological control of the larger grain borer Prostephanus truncatus (Coleoptera: Bostrichidae) in Kenya using a predatory beetle Teretrius nigrescens (Coleoptera: Histeridae). Bulletin of Entomological Research 93, 299306.CrossRefGoogle ScholarPubMed
Hodges, R.J., Dunstan, W.R., Magazini, I. & Golob, P. (1983) An outbreak of Prostephanus truncatus (Horn) (Coleoptera: Bostrichidae) in East Africa. Protection Ecology 5, 183194.Google Scholar
Hodges, R.J., Addo, S. & Birkinshaw, L. (2003) Can observation of climatic variables be used to predict the flight dispersal rates of Prostephanus truncatus? Agricultural and Forest Entomology 5, 123135.CrossRefGoogle Scholar
IITA (1999) Integrated Management of Maize Pests and Diseases. Annual Report, Plant Health Management Division, 1999. International Institute of Tropical Agriculture, Cotonou, Benin.Google Scholar
Jensen, J.L., Bohonak, A.J. & Kelley, S.T. (2005). Isolation by distance, web service. BMC Genetics 6, 13. v.3.16. Available online at http://ibdws.sdsu.edu/ (accessed 28 February 2011).Google Scholar
Kimura, M. (1980) A simple method for estimating evolutionary rate of base substitutions through comparative studies of nucleotide sequences. Journal of Molecular Evolution 16, 111120.CrossRefGoogle ScholarPubMed
Kolar, C.S. & Lodge, D.M. (2001) Progress in invasion biology: predicting invaders. Trends in Ecology and Evolution 16, 199204.Google Scholar
Lindell, J., Ngo, A. & Murphy, R.T.W. (2006) Deep genealogies and the mid-peninsular seaway of Baja California. Journal of Biogeography 33, 13271331.Google Scholar
Lindell, J. & Murphy, R.T.W. (2008) Simple identification of mitochondrial lineages in contact zones based on lineage-selective primers. Molecular Ecology Resources 8, 6673.CrossRefGoogle ScholarPubMed
Lloyd, C.J., Hufbauer, R.A., Jackson, A., Nissen, S.J. & Norton, P.N. (2005) Pre- and post-introduction patterns in neutral genetic diversity in the leafy spurge gall midge, Spurgia capitigena (Bremi) (Diptera: Cecidomyiidae). Biological Control 33, 153164.CrossRefGoogle Scholar
Lockwood, J.L., Cassey, P. & Blackburn, T. (2005) The role of propagule pressure in explaining species invasions. Trends in Ecology and Evolution 20, 223228.Google Scholar
Lunt, D.H., Zhang, D.-X., Szymura, J.M. & Hewitt, G.M. (1996) The insect cytochrome oxidase 1 gene: evolutionary patterns and conserved primers for phylogenetic studies. Insect Molecular Biology 5, 153165.CrossRefGoogle Scholar
Markham, R.H., Wright, V.F. & Ríos-Ibarra, R.M. (1991) A selective review of research on Prostephanus truncatus (Coleoptera: Bostrichidae) with an annotated and updated bibliography. Ceiba, 321:v + 90 pp.Google Scholar
Meikle, W.G., Rees, D. & Markham, R.H. (2002) Biological control of the larger grain borer, Prostephanus truncatus (Horn) (Coleoptera: Bostrichidae). Integrated Pest Management Reviews 7, 123138.Google Scholar
Memmott, J., Craze, P.G., Harman, H.M., Syrett, P. & Fowler, S.V. (2005) The effect of propagule size on the invasion of an alien insect. Journal of Animal Ecology 74, 5062.Google Scholar
Mynhardt, G., Harris, M.K. & Cognato, A.I. (2007) Population genetics of the pecan weevil (Coleoptera: Curculionidae) Inferred from mitochondrial nucleotide data. Annals of the Entomological Society of America 100, 582590.Google Scholar
Nang'ayo, F.L.O. (1996) Ecological studies on Larger Grain Borer in savanna woodlands of Kenya. PhD thesis, Imperial College, London, UK.Google Scholar
Navajas, M., Lagnel, J., Gutiérrez, J. & Boursot, P. (1998) Species-wide homogeneity of nuclear ribosomal ITS2 sequences in the spider mite Tetranychus urticae contrasts with extensive mitochondrial COI polymorphism. Heredity 80, 742752.Google Scholar
Omwega, C.O. & Overholt, W.A. (1996) Genetic changes occurring during laboratory rearing of Cotesia flavipes Cameron (Hymenoptera: Braconidae) an imported parasitoid for the control of graminaceous stem borers in Africa. African Entomology 4, 231237.Google Scholar
Otranto, D., Traversa, D., Guida, B., Tarasitano, E., Fiorente, P. & Stevens, J.R. (2003) Molecular characterisation of the mitochondrial cytochrome oxidase 1 gene of Oestridae species causing myiasis. Medical and Veterinary Entomology 17, 307315.Google Scholar
Phillips, C.B., Baird, D.B., Iline, I.I., McNeill, M.R., Proffitt, J.R., Goldson, S.L.J.M. & Kean, J.M. (2008) East meets west: adaptive evolution of an insect introduced for biological control. Journal of Applied Ecology 45, 948956.CrossRefGoogle Scholar
Rees, D.P., Rivera, R.R. & Rodriguez, F.S.H. (1990) Observations on the ecology of Teretriosoma nigrescens (Lewis) (Col., Histeridae) and its prey Prostephanus truncatus (Horn) (Col., Bostrichidae) in the Yucatan peninsula, Mexico. Tropical Science 30, 153165.Google Scholar
Roderick, G.K. & Navajas, M. (2003) Genes in new environments: Genetics and evolution in biological control. Nature Reviews Genetics 4, 889899.CrossRefGoogle ScholarPubMed
Saitou, N. & Nei, M. (1987) The neighbor-joining method: A new method for reconstructing phylogenetic trees. Molecular Biology and Evolution 4, 406425.Google Scholar
Sakai, A.K., Allendorf, F.W., Holt, J.S.M., Lodge, D.M., Molofsky, J., With, K.A., Baughman, S., Cabin, R.J., Cohen, J.E., Ellstrand, N.C., McCauley, D.E., O'Neil, P., Parker, I.M., Thompson, J.N. & Weller, S.G. (2001) The population biology of invasive species. Annual Review of Ecology and Systematics 32, 305332.Google Scholar
Sambrook, J., Fritsch, E.F. & Maniatis, T. (1989) Molecular Cloning, A Laboratory Manual. 2nd edn.New York, NY, USA, Cold Spring Harbour Laboratory Press.Google Scholar
Schneider, H., Borgemeister, C., Setamou, M., Affognon, H., Bell, A., Zweigert, M.E., Poehling, H. & Schulthess, F. (2004) Biological control of the larger grain borer Prostephanus truncatus (Horn) (Coleoptera: Bostrichidae) by its predator Teretrius nigrescens (Lewis) (Coleoptera: Histeridae) in Togo and Benin. Biological Control 30, 241255.CrossRefGoogle Scholar
Tamura, K., Nei, M. & Kumar, S. (2004) Prospects for inferring very large phylogenies by using the neighbor-joining method. Proceedings of the National Academy of Sciences, USA 101, 1103011035.Google Scholar
Tamura, K., Dudley, J., Nei, M. & Kumar, S. (2007) MEGA4: Molecular Evolutionary Genetics Analysis (MEGA) software version 4.0. Molecular Biology and Evolution 24, 15961599.Google Scholar
Tigar, B.J., Osborne, P.E., Key, G.E., FloresS, M.E. & Vazquez-Arista, M. (1994) Distribution and abundance of Prostephanus truncatus (Coleoptera: Bostrichidae) by its predator Teretriosoma nigrescens (Coleoptera: Histeridae) in Mexico. Bulletin of Entomological Research 84, 555565.CrossRefGoogle Scholar
Vincze, T., Posfai, J. & Roberts, R.J. (2003) NEBCutter: a program to cleave DNA with restriction enzymes. Nucleic Acids Research 31, 36883691.CrossRefGoogle ScholarPubMed
White, T.J., Bruns, T., Lee, S. & Taylor, J. (1990) Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. pp. 315322in Innes, M., Gelfand, D., Sninsky, J. & White, T. (Eds) PCR Protocols: A Guide to Methods and Applications. San Diego, CA, USA, Academic Press.Google Scholar
Zayed, A., Constantin, S.A. & Packer, L. (2007) Successful biological invasion despite a severe genetic load. PLoS ONE 2, e868.Google Scholar
Zhang, D.-X. & Hewitt, G.M. (1997) Assessment of the universality of a set of conserved mitochondrial COI primers. Insect Molecular Biology 6, 143150.CrossRefGoogle ScholarPubMed