Hostname: page-component-76fb5796d-skm99 Total loading time: 0 Render date: 2024-04-30T04:25:40.576Z Has data issue: false hasContentIssue false

Resolving multiple host use of an emergent pest of cotton with microsatellite data and chloroplast markers (Creontiades dilutus Stål; Hemiptera, Miridae)

Published online by Cambridge University Press:  23 May 2013

James P. Hereward*
Affiliation:
School of Biological Sciences, The University of Queensland, Brisbane, Queensland, Australia Cotton Catchment Communities Cooperative Research Centre, Australian Cotton Research Institute, Narrabri, New South Wales, Australia
Paul J. DeBarro
Affiliation:
CSIRO Ecosystem Sciences, GPO Box 2583, Brisbane, Queensland, Australia
Gimme H. Walter
Affiliation:
School of Biological Sciences, The University of Queensland, Brisbane, Queensland, Australia
*
*Author for correspondence Phone: +61 73365 2755 Fax: N/A E-mail: j.hereward@uq.edu.au

Abstract

Following the global uptake of transgenic cotton several Hemipteran pests have emerged as primary targets for pesticide control. Previous research on one such emergent pest: the green mirid, Creontiades dilutus, indicated differential use of two crop hosts, cotton (Gossypium hirsutum, Malvaceae) and lucerne (alfalfa) (Medicago sativa, Fabaceae). We tested the hypothesis that this apparent demographic independence of lucerne and cotton inhabiting mirids is the result of cryptic species being associated with these two crops. We assessed gene flow using microsatellite markers across adjacent cotton and lucerne crops at three geographically separated sites (up to 900 km apart). We also analysed the recent feeding behaviour of these insects by amplifying chloroplast markers from their gut contents. We find high gene flow between these two crops (mean pair wise FST between host plants=0.0141 within localities), and no evidence of cryptic species. Furthermore, the gut analyses revealed evidence of substantial recent movement between these two crops. We discuss the implications of these results for interpreting multiple host use in this species and setting future research priorities for this economically important pest.

Type
Research Paper
Copyright
Copyright © Cambridge University Press 2013 

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

Anderson, E.C. & Thompson, E.A. (2002) A model-based method for identifying species hybrids using multilocus genetic data. Genetics 160, 12171229.Google Scholar
Andris, M., Aradottir, G.I., Arnau, G., Audzijonyte, A., Bess, E.C., Bonadonna, F. et al. (2010) Permanent genetic resources added to molecular ecology resources database 1 June 2010–31 July 2010. Molecular Ecology Resources 10, 11061108.Google Scholar
Armstrong, J.S., Coleman, R.J. & Duggan, B.L. (2010) Actual and simulated injury of Creontiades signatus (Heteroptera: Miridae) feeding on cotton bolls. Journal of Entomological Science 45, 170177.Google Scholar
Armstrong, J.S., Coleman, R.J. & Adamczyk, J.J. (2011) Baseline susceptibility of Creontiades signatus to cotton insecticides, with emphasis on Malathion. Southwestern Entomologist 36, 145153.CrossRefGoogle Scholar
Bell, L., Ryan, M., Bennett, R., Collins, M. & Clarke, H. (2012) Growth, yield and seed composition of native Australian legumes with potential as grain crops. Journal of the Science of Food and Agriculture 92, 13541361.CrossRefGoogle ScholarPubMed
Bennett, R.G., Ryan, M.H., Colmer, T.D. & Real, D. (2010) Prioritisation of novel pasture species for use in water-limited agriculture: a case study of Cullen in the Western Australian wheatbelt. Genetic Resources and Crop Evolution 58, 83100.Google Scholar
Berthier, P., Excoffier, L. & Ruedi, M. (2006) Recurrent replacement of mtDNA and cryptic hybridization between two sibling bat species Myotis myotis and Myotis blythii. Proceedings of the Royal Society B—Biological Sciences 273, 31013109.Google Scholar
Bickford, D., Lohman, D., Sodhi, N., Ng, P., Meier, R., Winker, K., Ingram, K. & Das, I. (2007) Cryptic species as a window on diversity and conservation. Trends in Ecology and Evolution 22, 148203.CrossRefGoogle ScholarPubMed
Bonebrake, T.C., Watt, W.B., Perez, A. & Boggs, C.L. (2011) One variable species or multiple cryptic? Mitochondrial phylogeny of Central and North American Chlosyne lacinia (Lepidoptera: Nymphalidae). European Journal of Entomology 108, 529535.Google Scholar
Burns, J.M., Janzen, D.H., Hajibabaei, M., Hallwachs, W. & Hebert, P.D.N. (2008) DNA barcodes and cryptic species of skipper butterflies in the genus Perichares in Area de Conservación Guanacaste, Costa Rica. Proceedings of the National Academy of Sciences of the United States of America 105, 63506355.CrossRefGoogle ScholarPubMed
Carlsson, J. (2008) Effects of microsatellite null alleles on assignment testing. Journal of Heredity 99, 616623.Google Scholar
Chapuis, M.P. & Estoup, A. (2007) Microsatellite null alleles and estimation of population differentiation. Molecular Biology and Evolution 24, 621631.CrossRefGoogle ScholarPubMed
Chinajariyawong, A. (1988) The Sap-Sucking Bugs Attacking Cotton: Biological Aspects and Economic Damage. The University of Queensland, Brisbane Australia.Google Scholar
Clarke, A. & Walter, G. (1995) ‘Strains’ and the classical biological control of insect pests. Canadian Journal of Zoology/Revue Canadienne de Zoologie 73, 17773567.Google Scholar
Coleman, R.J., Hereward, J.P., De Barro, P.J., Frohlich, D.J., Adamczyk, J.J., & Goolsby, J.A. (2008) Molecular comparison of Creontiades plant bugs from south Texas and Australia. Southwestern Entomologist 33, 111117.CrossRefGoogle Scholar
Dempster, A.P., Laird, N.M., & Rubin, D.B. (1977) Maximum likelihood from incomplete data via EM algorithm. Journal of the Royal Statistical Society Series B—Methodological 39, 138.Google Scholar
Foley, D.H. & Pyke, B.A. (1985) Developmental time of Creontiades dilutus (Stal) (Hemiptera, Miridae) in relation to temperature. Journal of the Australian Entomological Society 24, 125127.Google Scholar
Fournier, V., Hagler, J., Daane, K., de Leon, J. & Groves, R. (2008) Identifying the predator complex of Homalodisca vitripennis (Hemiptera: Cicadellidae): a comparative study of the efficacy of an ELISA and PCR gut content assay. Oecologia 157, 629640.Google Scholar
Gariepy, T.D., Kuhlmann, U., Gillott, C. & Erlandson, M. (2007) Parasitoids, predators and PCR: the use of diagnostic molecular markers in biological control of Arthropods. Journal of Applied Entomology 131, 225240.Google Scholar
Gross, K. & Rosenheim, J.A. (2011) Quantifying secondary pest outbreaks in cotton and their monetary cost with causal-inference statistics. Ecological Applications 21, 27702780.CrossRefGoogle ScholarPubMed
Hebert, P., Penton, E., Burns, J., Janzen, D. & Hallwachs, W. (2004) Ten species in one: DNA barcoding reveals cryptic species in the neotropical skipper butterfly Astraptes fulgerator. Proceedings of the National Academy of Sciences of the United States of America 101, 1481214817.Google Scholar
Hereward, J.P.H. & Walter, G.H. (2012) Molecular interrogation of the feeding behaviour of field captured individual insects for interpretation of multiple host plant use. PLoS ONE 7, e44435.Google Scholar
Hoogendoorn, M. & Heimpel, G.E. (2001) PCR-based gut content analysis of insect predators: using ribosomal ITS-1 fragments from prey to estimate predation frequency. Molecular Ecology 10, 20592067.Google Scholar
Hosseini, S.M., Asadi, H.B., Kamali, K., Shojaii, M. & Hadi, O. (2002) Study on bio-ecology of cotton shedder bug Creontiades pallidus Rambur (Het; Miridae) in cotton fields of Khorassan Iran. Journal of Agricultural Sciences – Islamic Azad University 8, Ar73.Google Scholar
Jakobsson, M. & Rosenberg, N.A. (2007) CLUMPP: a cluster matching and permutation program for dealing with label switching and multimodality in analysis of population structure. Bioinformatics 23, 18011806.CrossRefGoogle ScholarPubMed
Jurado-Rivera, J.A., Vogler, A.P., Reid, C.A.M., Petitpierre, E. & Gomez-Zurita, J. (2009) DNA barcoding insect–host plant associations. Proceedings of the Royal Society B–Biological Sciences 276, 639648.Google Scholar
Khan, M., Kelly, D., Hickman, M., Mensah, R., Brier, H. & Wilson, L. (2004) Mirid management in Australian cotton. Australian Cotton CRCReview 15. Australian Cotton Co-operative Research Centre, Technology Resource Centre Press, Narrabri, NSW, Australia.Google Scholar
Li, G., Feng, H., McNeil, J.N., Liu, B., Chen, P. & Qiu, F. (2011) Impacts of transgenic Bt cotton on a non-target pest, Apolygus lucorum (Meyer-Dur) (Hemiptera: Miridae), in northern China. Crop Protection 30, 15731578.Google Scholar
Lori, K., Michael, M., Susan, J.B., Margaret, B. & Megan, H.R. (2009) Development, characterization and transferability of microsatellite markers for Cullen australasicum (Leguminosae). Conservation Genetics 10, 18031805.Google Scholar
Lu, Y., Wu, K., Jiang, Y., Xia, B., Li, P., Feng, H., Wyckhuys, K.A. & Guo, Y. (2010) Mirid bug outbreaks in multiple crops correlated with wide-scale adoption of Bt cotton in China. Science 328, 11511154.Google Scholar
Malipatil, M.B. & Cassis, G. (1997) Taxonomic review of Creontiades Distant in Australia (Hemiptera: Miridae: Mirinae). Australian Journal of Entomology 36, 113.Google Scholar
Mensah, R.K., & Khan, M. (1997) Use of Medicago sativa (L.) interplantings trap crops in the management of the green mirid, Creontiades dilutus (Stal) in commercial cotton in Australia. International Journal of Pest Management 43, 197202.Google Scholar
Miles, M.M. (1995) Identification, Pest Status, Ecology and Management of the Green Mirid, Creontiades dilutus (Stal) (Hemiptera: Miridae), a Pest of Cotton in Australia. The University of Queensland, Brisbane Australia.Google Scholar
Miller, S.A., Dykes, D.D. & Polesky, H.F. (1988) A simple salting out procedure for extracting DNA from human nucleated cells. Nucleic Acids Research 16, 12151215.Google Scholar
Muilenburg, V.L., Goggin, F.L., Hebert, S.L., Jia, L. & Stephen, F.M. (2008) Ant predation on red oak borer confirmed by field observation and molecular gut-content analysis. Agricultural and Forest Entomology 10, 205213.Google Scholar
Nevado, B., Koblmuller, S., Sturmbauer, C., Snoeks, J., Usano-Alemany, J. & Verheyen, E. (2009) Complete mitochondrial DNA replacement in a Lake Tanganyika cichlid fish. Molecular Ecology 18, 42404255.Google Scholar
Paterson, H.E.H. (1991) The recognition of cryptic species among economically important insects. pp. 110in Zalucki, M.P. (Ed.) Heliothis, Research Methods and Prospects. Springer-Verlag, New York.Google Scholar
Patil, S.B., Udikeri, S.S., Vandal, N.B., Modagi, S.S., Hirekurubar, R.B. & Guruprasad, G.S. (2010) Population dynamics of Creontiades biseratense (Distant) (Miridae: Hemiptera) on Bt cotton in Dharwad district. Karnataka Journal of Agricultural Sciences 23, 157158.Google Scholar
Petit, R. & Excoffier, L. (2009) Gene flow and species delimitation. Trends in Ecology and Evolution 24, 386479.Google Scholar
Powell, J.R. (1983) Interspecific Cytoplasmic gene flow in the absence of nuclear gene flow – evidence from Drosophila. Proceedings of the National Academy of Sciences of the United States of America–Biological Sciences 80, 492495.Google Scholar
Pritchard, J.K., Stephens, M. & Donnelly, P. (2000) Inference of population structure using multilocus genotype data. Genetics 155, 945959.Google Scholar
Rohini, R.S., Mallapur, C.P. & Udikeri, S.S. (2009) Incidence of mirid bug, Creontiades biseratense (Distant) on Bt cotton in Karnataka. Karnataka Journal of Agricultural Sciences 22, 680681.Google Scholar
Rosenberg, N.A. (2004) DISTRUCT: a program for the graphical display of population structure. Molecular Ecology Notes 4, 137138.Google Scholar
Schuelke, M. (2000) An economic method for the fluorescent labeling of pcr fragments. Nature Biotechnology 18, 233234.Google Scholar
Stam, P.A. (1987) Creontiades pallidus (Rambur) (Miridae, Hemiptera), a pest on cotton along the Euphrates river and its effect on yield and control action threshold in the Syrian-Arab-Republic. Tropical Pest Management 33, 273276.Google Scholar
Suriyagoda, L., Ryan, M., Renton, M. & Lambers, H. (2010) Multiple adaptive responses of Australian native perennial legumes with pasture potential to grow in phosphorus- and moisture-limited environments. Annals of Botany 105, 755767.Google Scholar
Taberlet, P., Gielly, L., Pautou, G. & Bouvet, J. (1991) Universal primers for amplification of 3 noncoding regions of chloroplast DNA. Plant Molecular Biology 17, 11051109.CrossRefGoogle Scholar
Torres, J.B. & Ruberson, J.R. (2006) Interactions of Bt-cotton and the omnivorous big-eyed bug Geocoris punctipes (Say), a key predator in cotton fields. Biological Control 39, 4757.Google Scholar
Torres, J.B. & Ruberson, J.R. (2008) Interactions of Bacillus thuringiensis Cry1Ac toxin in genetically engineered cotton with predatory heteropterans. Transgenic Research 17, 345354.Google Scholar
Walter, G.H. (2003) Insect Pest Management and Ecological Research. Cambridge University Press, USA.Google Scholar
Weir, B.S. (1996) Genetic Data Analysis II. Sinauer Associates, Sunderland, MA.Google Scholar
Whitehouse, M.E.A. (2011) IPM of mirids in Australian cotton: why and when pest managers spray for mirids. Agricultural Systems 104, 3041.Google Scholar
Whitehouse, M.E.A., Wilson, L.J. & Fitt, G.P. (2005) A comparison of arthropod communities in transgenic Bt and conventional cotton in Australia. Environmental Entomology 34, 12241241.CrossRefGoogle Scholar