Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-mp689 Total loading time: 0 Render date: 2024-04-20T16:35:19.103Z Has data issue: false hasContentIssue false

15 - Applications of molecular ecology to IPM: what impact?

Published online by Cambridge University Press:  04 August 2010

Marcos Kogan
Affiliation:
Oregon State University
Paul Jepson
Affiliation:
Oregon State University
Get access

Summary

Introduction

The use of molecular markers in biology, phylogeny and ecology has a long and distinguished history. Variable heritable protein markers (mostly allozymes) formed the basis of numerous studies in population biology and genetics (see Loxdale and den Hollander, 1989; Symondson and Liddell, 1996 for reviews). They enabled boundaries between functionally independent populations to be identified and allowed estimation of gene flow (and inferred migration) between them. Their use has been largely replaced by the use of DNA-based (often referred to as “molecular”) techniques and it is here where we focus our discussion. We use “molecular ecology” to mean the application of molecular biology to population ecology (see reviews in Avise, 1994; Schierwater et al., 1994; Moritz and Lavery, 1996; Carvalho, 1998; Sunnucks, 2000; see Loxdale and Lushai, 1999; MacDonald and Loxdale, 2004 for reviews on entomological applications).

Reasons for the shift to DNA-based methods for studying pests include the following.

  1. a. Technological factors, such as the fact that usable DNA can be obtained from very small specimens (e.g. tiny grape phylloxera Daktulosphaira vitifoliae, Downie, 2000; individual aphid eggs, Sloane et al., 2001) and preserved specimens such as pinned museum collections or ethanol-preserved suction-trapped insects.

  2. […]

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2007

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

Abdullahi, I., Winter, S., Atiri, G. I. and Thottappilly, G. (2003). Molecular characterization of whitefly, Bemisia tabaci (Hemiptera: Aleyrodidae) populations infesting cassava. Bulletin of Entomological Research, 93, 97–106.CrossRefGoogle ScholarPubMed
Abdullahi, I., Atiri, G. I., Thottappilly, G. and Winter, S. (2004). Discrimination of cassava-associated Bemisia tabaci in Africa from polyphagous populations, by PCR-RFLP of the internal transcribed spacer regions of ribosomal DNA. Journal of Applied Entomology, 128, 81–7.CrossRefGoogle Scholar
Aebi, A., Alvarez, N., Butcher, R. D. J.et al. (2004). Microsatellite markers in a complex of Horismenus sp. (Hymenoptera: Eulophidae), parasitoids of bruchid beetles. Molecular Ecology Notes, 4, 707–9.CrossRefGoogle Scholar
Agusti, N., Vincente, C. and Gabarra, R. (1999). Development of sequence amplified characterized region (SCAR) markers of Helicoverpa armigera: a new polymerase chain reaction-based technique for predator gut analysis. Molecular Ecology, 8, 1467–74.CrossRefGoogle ScholarPubMed
Agusti, N., Shayler, S. P., Harwood, J. D.et al. (2003). Collembola as alternative prey sustaining spiders in arable ecosystems: prey detection within predators using molecular markers. Molecular Ecology, 12, 3467–75.CrossRefGoogle ScholarPubMed
Aikhionbare, F. O. and Mayo, Z. B. (2000). Mitochondrial DNA sequences of greenbug (Homoptera: Aphididae) biotypes. Biomolecular Engineering, 16, 199–205.CrossRefGoogle ScholarPubMed
Alberti, A. C., Rodriguero, M. S., Cendra, P. G., Saidman, B. O. and Vilardi, J. C. (2003). Evidence indicating that argentine populations of Anastrepha fraterculus (Diptera: Tephritidae) belong to a single biological species. Annals of the Entomological Society of America, 95, 505–12.CrossRefGoogle Scholar
Aljanabi, S. M., Loiacono, M. S., Lourenco, R. T., Borges, M. and Tigano, M. S. (1998). RAPD analysis revealing polymorphism in egg parasitoids of soybean stink bugs (Hemiptera; Pentatomidae). Anais da Sociedade Entomologica do Brasil., 27, 413–20.CrossRefGoogle Scholar
Althoff, D. M. and Pellmyr, O. (2002). Examining genetic structure in a bogus yucca moth: a sequential approach to phylogeography. Evolution, 56, 1632–43.CrossRefGoogle Scholar
Andersen, L. W., Holm, L.-E., Siegismund, H. R.et al. (1997). A combined DNA-microsatellite and isozyme analysis of the population structure of the harbour porpoise in Danish waters and West Greenland. Heredity, 78, 270–6.CrossRefGoogle ScholarPubMed
Angers, B., Magnan, P., Plante, L. and Bernatchez, M. (1999). Canonical correspondence analysis for estimating spatial and environmental effects on microsatellite gene diversity in brook charr (Salvelinus fontinalis). Molecular Ecology, 8, 1043–55.CrossRefGoogle Scholar
Anstead, J. A., Burd, J. D. and Shufran, K. A. (2002). Mitochondrial DNA sequence divergence among Schizaphis graminum (Hemiptera: Aphididae) clones from cultivated and non-cultivated hosts: haplotype and host associations. Bulletin of Entomological Research, 92, 17–24.Google ScholarPubMed
Antolin, M. F., Bosio, C. F., Cotton, J.et al. (1996). Intensive linkage mapping in a wasp (Bracon hebetor) and a mosquito (Aedes aegypti) with single-strand conformation polymorphism analysis of Random Amplified Polymorphic DNA markers. Genetics, 143, 1727–38.Google Scholar
Avise, J. C., Lansman, R. A. and Shade, R. O. (1979). The use of restriction endonucleases to measure mitochondrial DNA sequence relatedness in natural populations. I. Population structure and evolution in the genus Peromyscus. Genetics, 92, 279–95.Google ScholarPubMed
Avise, J. C. (1994). Molecular Markers, Natural History and Evolution. London: Chapman and Hall.CrossRefGoogle Scholar
Baer, C. F., Tripp, D. W., Bjorksten, T. A. and Antolin, M. F. (2004). Phylogeography of a parasitoid wasp (Diaeretiella rapae): no evidence of host-associated lineages. Molecular Ecology, 13, 1859–69.CrossRefGoogle ScholarPubMed
Baker, D. A., Loxdale, H. D. and Edwards, O. R. (2004). Genetic variation and founder effects in the parasitoid wasp, Diaeretiella rapae (M'intosh) (Hymenoptera: Braconidae: Aphidiidae), affecting its potential as a biological control agent. Molecular Ecology, 12, 3303–11.CrossRefGoogle Scholar
Baldwin, I. T., Kessler, A. and Halitschke, R. (2002). Volatile signaling in plant–plant–herbivore interactions: what is real?Current Opinion in Plant Biology, 5, 351–4.CrossRefGoogle ScholarPubMed
Baliraine, F. N., Bonizzoni, M., Guglielmino, C. R.et al. (2004). Population genetics of the potentially invasive African fruit fly species, Ceratitis rosa and Ceratitis fasciventris (Diptera: Tephritidae). Molecular Ecology, 13, 683–95.CrossRefGoogle Scholar
Baliraine, F. N., Bonizzoni, M., Osir, E. O.et al. (2003). Comparative analysis of microsatellite loci in four fruit fly species of the genus Ceratitis (Diptera: Tephritidae). Bulletin of Entomological Research, 93, 1–10.CrossRefGoogle Scholar
Bang-Ce, Y., Peng, Z., Bincheng, Y. and Songyang, L. (2004). Estimation of relative allele frequencies of single-nucleotide polymorphisms in different populations by microarray hybridization of pooled DNA. Analytical Biochemistry, 333, 72–8.CrossRefGoogle ScholarPubMed
Barrette, R. J., Crease, T. J. and Hebert, P. D. N. (1994). Mitochondrial DNA diversity in the pea aphid Acyrthosiphon pisum. Genome, 37, 858–65.CrossRefGoogle Scholar
Beaumont, M. A. (2004). Recent developments in genetic data analysis: what can they tell us about human demographic history?Heredity, 92, 365–79.CrossRefGoogle ScholarPubMed
Beerli, P. and Felsenstein, T. J. (2001). Maximum likelihood estimation of a migration matrix and effective population sizes in n subpopulations by using a coalescent approach. Proceedings of the National Academy of Sciences USA, 98, 4563–8.CrossRefGoogle ScholarPubMed
Berry, S. D., Fondong, V. N., Rey, C., Rogan, D., Fauquet, C. M. and Brown, J. K. (2004 a). Molecular evidence for five distinct Bemisia tabaci (Homoptera: Aleyrodidae) geographic haplotypes associated with cassava plants in sub-Saharan Africa. Annals of the Entomological Society of America, 97, 852–9.CrossRefGoogle Scholar
Berry, O., Tocher, M. D. and Sarre, S. D. (2004 b). Can assignment tests measure dispersal?Molecular Ecology, 13, 551–61.CrossRefGoogle ScholarPubMed
Black, W. C. IV. (1993). PCR with arbitrary primers: approach with care. Insect Molecular Biology, 2, 1–6.CrossRefGoogle ScholarPubMed
Blackman, R. L., Watson, G. W. and Ready, P. D. (1995). The identity of the African pine woolly aphid: a multidisciplinary approach. Bulletin OEPP, 25, 337–41.CrossRefGoogle Scholar
Bodrossy, L., Stralis-Pavese, N., Murrell, J. C.et al. (2003). Development and validation of a diagnostic microbial microarray for methanotrophs. Environmental Microbiology, 5, 566–82.CrossRefGoogle ScholarPubMed
Bogdanowicz, S. M., Wallner, W. E., Bell, J., Odell, T. M. and Harrison, R. G. (1993). Asian gypsy moths (Lepidoptera: Lymantriidae) in North America: evidence from molecular data. Annals of the Entomological Society of America, 86, 710–15.CrossRefGoogle Scholar
Bogdanowicz, S. M., Mastro, V. C., Prasher, D. C. and Harrison, R. G. (1997). Microsatellite DNA variation among Asian and North American gypsy moths (Lepidoptera: Lymantriidae). Annals of the Entomological Society of America, 90, 768–75.CrossRefGoogle Scholar
Bonizzoni, M., Katsoyannos, B. I., Marguerie, R.et al. (2003). Microsatellite analysis reveals remating by wild Mediterranean fruit fly females, Ceratitis capitata. Molecular Ecology, 11, 1915–21.CrossRefGoogle Scholar
Bonizzoni, M., Guglielmino, C. R., Smallridge, C. J.et al. (2004). On the origins of medfly invasion and expansion in Australia. Molecular Ecology, 13, 3845–55.CrossRefGoogle ScholarPubMed
Borghuis, A., Pinto, J. D., Platner, G. R. and Stouthamer, R. (2004). Partial cytochrome oxidase II sequences distinguish the sibling species Trichogramma minutum Riley and Trichogramma platneri Nagarkatti. Biological Control, 30, 90–4.CrossRefGoogle Scholar
Bournoville, R., Simon, J. C., Badenhausser, I.et al. (2000). Clones of pea aphid, Acyrthosiphon pisum (Hemiptera: Aphididae) distinguished using genetic markers, differ in their damaging effect on a resistant alfalfa cultivar. Bulletin of Entomological Research, 90, 33–9.CrossRefGoogle ScholarPubMed
Brodeur, J. and Rosenheim, J. A. (2000). Intraguild interactions in aphid parasitoids. Entomologia Experimentalis et Applicata, 97, 93–108.CrossRefGoogle Scholar
Brouat, C., Chevallier, H., Meusnier, S., Noblecourt, T. and Rasplus, J.-Y. (2004). Specialization and habitat: spatial and environmental effects on abundance and genetic diversity of forest generalist and specialist Carabus species. Molecular Ecology, 13, 1815–26.CrossRefGoogle ScholarPubMed
Brown, S., McLaughlin, W., Jerez, I. T. and Brown, J. K. (2002). Identification and distribution of Bemisia tabaci (Gennadius) (Homoptera: Aleyrodidae) haplotypes in Jamaica. Tropical Agriculture, 79, 140–9.Google Scholar
Brunner, P. C. and Frey, J. E. (2004). Isolation and characterization of six polymorphic microsatellite loci in the western flower thrips Frankliniella occidentalis (Insecta: Thysanoptera). Molecular Ecology Notes, 4, 599–601.CrossRefGoogle Scholar
Brunner, P. C., Chatzivassiliou, E. K., Katis, N. I. and Frey, J. E. (2004). Host-associated genetic differentiation in Thrips tabaci (Insecta: Thysanoptera), as determined from mtDNA sequence data. Heredity, 93, 364–70.CrossRefGoogle Scholar
Caillaud, M. C., Mondor-Genson, G., Levine-Wilkinson, S.et al. (2004). Microsatellite DNA markers for the pea aphid Acyrthosiphon pisum. Molecular Ecology Notes, 4, 446–8.CrossRefGoogle Scholar
Calvert, L. A., Cuervo, M., Arroyave, J. A.et al. (2001). Morphological and mitochondrial DNA marker analyses of whiteflies (Homoptera: Aleyrodidae) colonizing cassava and beans in Colombia. Annals of the Entomological Society of America, 94, 512–19.CrossRefGoogle Scholar
Carvalho, G. R., Maclean, N., Wratten, S. D., Carter, R. E. and Thurston, J. P. (1991). Differentiation of aphid clones using DNA fingerprints from individual aphids. Proceedings of the Royal Society London B series, 243, 109–14.CrossRefGoogle Scholar
Carvalho, G. R. (ed.) (1998). Advances in Molecular Ecology. IOS Press, Amsterdam.Google Scholar
Casey, D. G. and Burnell, A. M. (2001). The isolation of microsatellite loci in the Mediterranean fruitfly Ceratitis capitata (Diptera: Tephritidae) using a biotin/streptavidin enrichment technique. Molecular Ecology Notes, 1, 120–2.CrossRefGoogle Scholar
Castric, V. and Bernatchez, L. (2004). Individual assignment test reveals differential restriction to dispersal between two salmonids despite no increase of genetic differences with distance. Molecular Ecology, 13, 1299–312.CrossRefGoogle ScholarPubMed
Chakrabarti, S., Kambhampati, S., Grace, T. and Zurek, L. (2004). Characterization of microsatellite loci in the house fly, Musca domestica L. (Diptera: Muscidae). Molecular Ecology Notes, 4, 728–30.CrossRefGoogle Scholar
Chakraborty, R., Meagher, T. R. and Smouse, P. E. (1988). Parentage analysis with genetic markers in natural populations. I. The expected proportion of offspring with unambiguous paternity. Genetics, 52, 922–7.Google Scholar
Chang, S. C., Hu, N. T., Hsin, C. Y. and Sun, C. N. (2001). Characterization of differences between two Trichogramma wasps by molecular markers. Biological Control, 21, 75–8.CrossRefGoogle Scholar
Chang, W. X. Z., Tabashnik, B. E., Artelt, B.et al. (1997). Mitochondrial DNA sequence variation among geographic strains of diamondback moth (Lepidoptera: Plutellidae). Annals of the Entomological Society of America, 90, 590–5.CrossRefGoogle Scholar
Chapman, R. W. (2001). EcoGenomics – a consilience for comparative immunology?Developmental and Comparative Immunology, 25, 549–51.CrossRefGoogle ScholarPubMed
Chapuisat, M., Bocherens, S. and Rosset, H. (2004). Variable queen number in ant colonies: no impact on queen turnover, inbreeding, and population genetic differentiation in the ant Formica selysi. Evolution, 58, 1064–72.CrossRefGoogle ScholarPubMed
Cho, R. J., Mindrinos, M., Richards, D. R.et al. (1999). Genome-wide mapping with biallelic markers in Arabidopsis thaliana. Nature Genetics, 23, 203–7.CrossRefGoogle ScholarPubMed
Clark, T. L. and Hibbard, B. E. (2004). Comparison of nonmaize hosts to support western corn rootworm (Coleoptera: Chrysomelidae) larval biology. Environmental Entomology, 33, 681–9.CrossRefGoogle Scholar
Cliff, A. D. and Ord, J. K. (1981). Spatial Processes. Models and Applications. London: Pion Ltd.Google Scholar
Coates, B. S., Sumerford, D. V. and Hellmich, R. L. (2004). Geographic and voltinism differentiation among North American Ostrinia nubilalis (European corn borer) mitochondrial cytochrome c oxidase haplotypes, Journal of Insect Science, 4, 35.Google ScholarPubMed
Cognato, A. I., Harlin, A. D. and Fisher, M. L. (2003). Genetic structure among pinyon pine beetle populations (Scolytinae: Ips confusus). Environmental Entomology, 32, 1262–70.CrossRefGoogle Scholar
Contreras-Diaz, H. G., Moya, O., Oromi, P. and Juan, C. (2003). Phylogeography of the endangered darkling beetle species of Pimelia endemic to Gran Canaria (Canary Islands). Molecular Ecology, 12, 2131–43.CrossRefGoogle Scholar
Cornuet, J. M., Piry, S., Luikart, G., Estoup, A. and Solignac, M. (1999). New methods employing multilocus genotypes to select or exclude populations as origins of individuals. Genetics, 153, 1989–2000.Google ScholarPubMed
Corrie, A. M., Crozier, R. H., Heeswijck, R. and Hoffmann, A. A. (2002). Clonal reproduction and population genetic structure of grape phylloxera, Daktulosphaira vitifoliae, in Australia. Heredity, 88, 203–11.CrossRefGoogle ScholarPubMed
Corrie, A. M., Heeswijck, R. and Hoffmann, A. A. (2003). Evidence for host-associated clones of grape phylloxera Daktulosphaira vitifoliae (Hemiptera: Phylloxeridae) in Australia. Bulletin of Entomological Research, 93, 193–201.CrossRefGoogle Scholar
Corrie, A. M. and Hoffmann, A. A. (2004). Fine-scale genetic structure of grape phylloxera from the roots and leaves of Vitis. Heredity, 92, 118–27.CrossRefGoogle ScholarPubMed
Cuthbertson, A. G. S., Fleming, C. C. and Murchie, A. K. (2003). Detection of Rhopalosiphum insertum (apple-grass aphid) predation by the predatory mite Anystis baccarum using molecular gut analysis. Agricultural and Forest Entomology, 5, 219–25.CrossRefGoogle Scholar
D'Acier, A. C., Sembene, M., Audiot, P. and Rasplus, J. Y. (2004). Polymorphic microsatellites loci in the black Aphid, Aphis fabae Scopoli, 1763 (Hemiptera: Aphididae). Molecular Ecology Notes, 4, 306–8.CrossRefGoogle Scholar
Dai, S. M., Lin, C. C. and Chang, C. (2004). Polymorphic microsatellite DNA markers from the oriental fruit fly Bactrocera dorsalis (Hendel). Molecular Ecology Notes, 4, 629–31.CrossRefGoogle Scholar
Davies, N., Villablanca, F. X. and Roderick, G. K. (1999 a). Determining the source of individuals: multilocus genotyping in nonequilibrium population genetics. Trends in Ecology and Evolution, 14, 17–21.CrossRefGoogle ScholarPubMed
Davies, N., Villablanca, F. X. and Roderick, G. K. (1999 b). Bioinvasions of the medfly Ceratitis capitata: Source estimation using DNA sequences at multiple intron loci. Genetics, 153, 351–60.Google ScholarPubMed
Barro, P. J. and Driver, F. (1997). Use of RAPD PCR to distinguish the B biotype from other biotypes of Bemisia tabaci (Gennadius) (Hemiptera: Aleyrodidae). Australian Journal of Entomology, 36, 149–52.CrossRefGoogle Scholar
Barro, P. J., Driver, F., Naumann, I. D.et al. (2000 a). Descriptions of three species of Eretmocerus Haldeman (Hymenoptera: Aphelinidae) parasitising Bemisia tabaci (Gennadius) (Hemiptera: Aleyrodidae) and Trialeurodes vaporariorum (Westwood) (Hemiptera: Aleyrodidae) in Australia based on morphological and molecular data. Australian Journal of Entomology, 39, 259–69.CrossRefGoogle Scholar
Barro, P. J., Driver, F., Trueman, J. W. H. and Curran, J. (2000 b). Phylogenetic relationship of world populations of Bemisia tabaci (Gennadius) using ribosomal ITS1. Molecular Phylogeny and Evolution, 16, 29–36.CrossRefGoogle Scholar
Barro, P. J. and Hart, P. J. (2000). Mating interactions between two biotypes of the whitefly, Bemisia tabaci (Hemiptera: Aleyrodidae) in Australia. Bulletin of Entomological Research, 90, 103–12.CrossRefGoogle Scholar
Barro, P. J., Scott, K. D., Graham, G. C., Lange, C. L. and Schutze, M. K. (2003). Isolation and characterization of microsatellite loci in Bemisia tabaci. Molecular Ecology Notes, 3, 40–3.CrossRefGoogle Scholar
Barro, P. J., Sherratt, T. N., Brookes, C. P., Markovic, O. and Maclean, N. (1995 c). Spatial and temporal genetic variation in the population structure of the grain aphid, Sitobion avenae (F.) (Hemiptera: Aphididae) studied using RAPD-PCR. Proceedings of the Royal Society, Series B, London, 262, 321–7.CrossRefGoogle Scholar
Barro, P. J., Sherratt, T. N., Carvalho, G. R.et al. (1994). An analysis of secondary spread by putative clones of Sitobion avenae within a Hampshire wheat field using the multilocus (GATA)4 probe. Insect Molecular Biology, 3, 253–60.CrossRefGoogle ScholarPubMed
Barro, P. J., Sherratt, T. N., Carvalho, G. R.et al. (1995 a). Geographic and microgeographic genetic differentiation in two aphid species over southern England using the multilocus (GATA)4 probe. Molecular Ecology, 4, 375–82.CrossRefGoogle ScholarPubMed
Barro, P. J., Sherratt, T. N., Markovic, O. and Maclean, N. (1995 b). An investigation of the differential performance of clones of the cereal aphid, Sitobion avenae on two host species. Oecologia, 104, 375–85.CrossRefGoogle Scholar
Barro, P. J., Trueman, J. W. H. and Frohlich, D. R. (2005). Bemisia argentifolii is a race of B. tabaci: the molecular genetic differentiation of B. tabaci populations around the world. Bulletin of Entomological Research, 14, 3695–718.Google Scholar
Brito, R. A., Manfrin, M. H. and Sene, F. M. (2002). Nested cladistic analysis of Brazilian populations of Drosophila serido. Molecular Phylogenetics and Evolution, 22, 131–43.CrossRefGoogle ScholarPubMed
Leon, J. H. and Jones, N. A. (2004). Detection of DNA Polymorphisms in Homalodisca coagulata (Homoptera: Cicadellidae) by polymerase chain reaction-based DNA fingerprinting methods. Annals of the Entomological Society of America, 97, 574–85.CrossRefGoogle Scholar
Meeus, T., Beati, L., Delaye, C., Aeschlimann, A. and Renaud, F. (2002). Sex-biased genetic structure in the vector of Lyme disease, Ixodes ricinus. Evolution, 56, 1802–7.CrossRefGoogle ScholarPubMed
Delmotte, F., Leterme, N., Gauthier, J. P., Rispe, C. and Simon, J. C. (2002). Genetic architecture of sexual and asexual populations of the aphid Rhopalosiphum padi based on allozyme and microsatellite markers. Molecular Ecology, 11, 711–23.CrossRefGoogle ScholarPubMed
Dennis, P., Edwards, E. A., Liss, S. N. and Fulthorpe, R. (2002). Monitoring gene expression in mixed microbial communities by using DNA microarrays. Applied and Environmental Microbiology, 69, 769–78.CrossRefGoogle Scholar
Donnelly, P. and Tavare, S. (1995). Coalescents and genealogical structure under neutrality. Annual Review of Genetics, 29, 401–21.CrossRefGoogle ScholarPubMed
Dopman, E. B., Bogdanowicz, S. M. and Harrison, R. G. (2004). Genetic mapping of sexual isolation between E and Z pheromone strains of the European corn borer (Oshinia nubilalis). Genetics, 167, 301–9.CrossRefGoogle Scholar
Downie, D. A. (2000). Patterns of genetic variation in the native grape phylloxera on two sympatric host species. Molecular Ecology, 9, 505–14.CrossRefGoogle ScholarPubMed
Downie, D. A., Fisher, J. R. and Granett, J. (2001). Grapes, galls, and geography: the distribution of nuclear and mitochondrial DNA variation across host-plant species and regions in a specialist herbivore. Evolution, 55, 1345–62.CrossRefGoogle Scholar
Eber, S. and Brandl, R. (1994). Ecological and genetic spatial patterns of Urophora cardui (Diptera: Tephritidae) as evidence for population structure and biogeographical processes. Journal of Animal Ecology, 63, 187–99.CrossRefGoogle Scholar
Edwards, O. R. and Hopper, K. R. (1999). Using superparasitism by a stem borer parasitoid to infer a host refuge. Ecological Entomology, 24, 7–12.CrossRefGoogle Scholar
Edwards, O. R. and Hoy, M. A. (1995 a). Monitoring laboratory and field biotypes of the walnut aphid parasite, Trioxys pallidus, in population cages using RAPD-PCR. Biocontrol Science and Technology, 5, 313–27.CrossRefGoogle Scholar
Edwards, O. R. and Hoy, M. A. (1995 b). Random amplified polymorphic DNA markers to monitor laboratory-selected, pesticide-resistant Trioxys pallidus (Hymenoptera: Aphidiidae) after release into three California walnut orchards. Environmental Entomology, 24, 487–96.CrossRefGoogle Scholar
Edwards, S. V. and Beerli, P. (2000). Perspective: gene divergence, population divergence, and the variance in coalescence time in phylogeographic studies. Evolution, 54, 1839–54.Google ScholarPubMed
Ehtesham, N. Z., Bentur, J. S. and Bennett, J. (1995). Highly repetitive DNA sequence elements from Orseolia oryzae (Wood–Mason) discriminate between the Indian isolates of the Asian rice gall midge and the paspalum midge. Electrophoresis, 16, 1762–5.CrossRefGoogle ScholarPubMed
Epperson, B. K. (1993). Spatial and space–time correlations in systems of subpopulations with genetic drift and migration. Genetics, 133, 711–27.Google ScholarPubMed
Epperson, B. K. (1995 a). Spatial structure of two-locus genotypes under isolation by distance. Genetics, 140, 365–75.Google ScholarPubMed
Epperson, B. K. (1995 b). Fine-scale spatial structure – correlations for individual genotypes differ from those for local gene frequencies. Evolution, 49, 1022–6.CrossRefGoogle ScholarPubMed
Epperson, B. K. and Li, T. (1996). Measurement of genetic structure within populations using Moran's spatial autocorrelation statistics. Proceeding of the National Academy of Sciences USA, 93, 10528–32.CrossRefGoogle ScholarPubMed
Excoffier, L., Smouse, P. E. and Quattro, J. (1992). Analysis of molecular variance from metric distance among DNA haplotypes: application to human mitochondrial DNA restriction data. Genetics, 131, 479–91.Google Scholar
Evans, J. D., Pettis, J. S., Hood, W. M. and Shimanuki, H. (2003). Tracking an invasive honey bee pest: mitochondrial DNA variation in North American small hive beetles. Apidologie, 34, 103–9.CrossRefGoogle Scholar
Fairley, T. L., Povoa, M. M. and Conn, J. E. (2002). Evaluation of the Amazon River delta as a barrier to gene flow for the regional malaria vector, Anopheles aquasalis (Diptera: Culicidae) in northeastern Brazil. Journal of Medical Entomology, 39, 861–9.CrossRefGoogle Scholar
Favre, L., Balloux, F., Goudet, J. and Perrin, N. (1997). Female-biased dispersal in themonogamous mammal Crocidura russula: evidence from field data and microsatellite patterns. Proceedings of the Royal Society of London Series B, 264, 127–32.CrossRefGoogle ScholarPubMed
Fenton, B., Malloch, G., Jones, A. T.et al. (1995). Species identification of Cecidophyopsis mites (Acari: Eriophyidae) from different Ribes species and countries using molecular genetics. Molecular Ecology, 4, 383–7.CrossRefGoogle ScholarPubMed
Fenton, B., Jones, A. T., Malloch, G. and Thomas, W. P. (1996). Molecular Ecology of some Cecidophyopsis mites (Acari: Eriophyidae) on Ribes species and evidence for their natural cross colonization of blackcurrant (R. nigrum). Annals of Applied Biology, 128, 405–14.CrossRefGoogle Scholar
Fenton, B., Malloch, G., Navajas, M., Hillier, J. and Birch, A. N. E. (2003). Clonal composition of the peach-potato aphid Myzus persicae (Homoptera: Aphididae) in France and Scotland: comparative analysis with IGS fingerprinting and microsatellite markers. Annals of Applied Biology, 142, 255–67.CrossRefGoogle Scholar
Ferry, N., Edwards, M. G., Gatehouse, J. A. and Gatehouse, A. M. R. (2004). Plant–insect interactions: molecular approaches to insect resistance. Current Opinions in Biotechnology, 15, 155–61.CrossRefGoogle ScholarPubMed
Field, L. M., Crick, S. E. and Devonshire, A. L. (1996). Polymerase chain reaction-based identification of insecticide resistance genes and DNA methylation in the aphid Myzus persicae (Sulzer). Insect Molecular Biology, 5, 197–202.CrossRefGoogle Scholar
Figueroa, C. C., Simon, J.-C., Gallic, J.-F.et al. (2005). Genetic structure and clonal diversity of an introduced pest in Chile, the cereal aphid Sitobion avenae. Heredity, 95, 24–33.CrossRefGoogle ScholarPubMed
Foley, D. H., Russell, R. C. and Bryan, J. H. (2004). Population structure of the peridomestic mosquito Ochlerotatus notoscriptus in Australia. Medical and Veterinary Entomology, 18, 180–90.CrossRefGoogle ScholarPubMed
Fordyce, J. A. and Nice, C. C. (2003). Contemporary patterns in a historical context: phylogeographic history of the pipevine swallowtail, Battus philenor (Papilionidae). Evolution, 57, 1089–99.CrossRefGoogle Scholar
Fritz, A. H. and Schable, N. (2004). Microsatellite loci from the Caribbean fruit fly, Anastrepha suspensa (Diptera: Tephritidae). Molecular Ecology Notes, 4, 443–5.CrossRefGoogle Scholar
Frohlich, D. R., Torres-Jerez, I., Bedford, I. D., Markham, P. G. and Brown, J. K. (1999). A phylogeographical analysis of the Bemisia tabaci species complex based on mitochondrial DNA markers. Molecular Ecology, 8, 1867–77.CrossRefGoogle ScholarPubMed
Fuentes-Contreras, E., Figueroa, C. C., Reyes, M., Briones, L. M. and Niemeyer, H. M. (2004). Genetic diversity and insecticide resistance of Myzus persicae (Hemiptera: Aphididae) populations from tobacco in Chile: evidence for the existence of a single predominant clone. Bulletin of Entomological Research, 94, 11–18.CrossRefGoogle ScholarPubMed
Fuller, S. J., Chavigny, P., Lapchin, L. and Vanlerberghe-Masutti, F. (1999). Variation in clonal diversity in glasshouse infestations of the aphid, Aphis gossypii Glover in southern France. Molecular Ecology, 8, 1867–77.CrossRefGoogle ScholarPubMed
Garner, K. J. and Slavicek, J. M. (1996). Identification and characterization of a RAPD-PCR marker for distinguishing Asian and North American gypsy moths. Insect Molecular Biology, 5, 81–91.CrossRefGoogle ScholarPubMed
Gaublomme, E., Dhuyvetter, H., Verdyck, P.et al. (2003). Isolation and characterization of microsatellite loci in the ground beetle Carabus problematicus (Coleoptera: Carabidae). Molecular Ecology Notes, 3, 341–3.CrossRefGoogle Scholar
Gauthier, N. and Rasplus, J. Y. (2004). Polymorphic microsatellite loci in the coffee berry borer, Hypothenemus hampei (Coleoptera, scolytidae). Molecular Ecology Notes, 4, 294–6.CrossRefGoogle Scholar
Gawel, N. J. and Bartlett, A. C. (1993). Characterization of differences between whiteflies using RAPD-PCR. Insect Molecular Biology, 2, 33–8.CrossRefGoogle ScholarPubMed
Giblin-Davis, R. M., Gries, R., Crespi, B.et al. (2000). Aggregation pheromones of two geographical isolates of the New Guinea sugarcane weevil, Rhabdoscelus obscurus. Journal of Chemical Ecology, 26, 2763–80.CrossRefGoogle Scholar
Gilchrist, A. S., Sved, J. A. and Meats, A. (2004). Genetic relations between outbreaks of the Queensland fruit fly, Bactrocera tryoni (Froggatt) (Diptera: Tephritidae), in Adelaide in 2000 and 2002. Australian Journal of Entomology, 43, 157–63.CrossRefGoogle Scholar
Gilles, J., Litrico, I., Sourrouille, P. and Duvallet, G. (2004). Microsatellite DNA markers for the stable fly, Stomoxys calcitrans (Diptera: Muscidae). Molecular Ecology Notes, 4, 635–7.CrossRefGoogle Scholar
Goldson, S. L., Phillips, C. B., McNeill, M. R. and Barlow, N. D. (1997). The potential of parasitoid strains in biological control: observations to date on Microctonus spp. intraspecific variation in New Zealand. Agriculture, Ecosystems Environment, 64, 115–24.CrossRefGoogle Scholar
Gomulski, L. M., Bourtzis, K., Brogna, S.et al. (1998). Intron size polymorphism of the Adh1 gene parallels the worldwide colonization history of the Mediterranean fruit fly, Ceratitis capitata. Molecular Ecology, 7, 1729–41.CrossRefGoogle Scholar
Graur, D. (1985). Gene diversity in Hymenoptera. Evolution, 39, 190–9.CrossRefGoogle ScholarPubMed
Greene, E. A. and Voordouw, G. (2003). Analysis of environmental microbial communities by reverse sample genome probing. Journal of Microbiological Methods, 53, 211–19.CrossRefGoogle ScholarPubMed
Griffin, T. J., Hall, J. G., Prudent, J. R. and Smith, L. M. (1999). Direct genetic analysis by matrix-assisted laser desorption/ionization mass spectrometry. Proceedings of the National Academy of Sciences USA, 96, 6301–6.CrossRefGoogle ScholarPubMed
Guillemaud, T., Mieuzet, L. and Simon, J. C. (2003). Spatial and temporal genetic variability in French populations of the peach-potato aphid, Myzus persicae. Heredity, 91, 143–52.CrossRefGoogle ScholarPubMed
Haack, L., Gauthierm, J.-P., Plantegenest, M. and Dedryver, C.-A. (2000). Predominance of generalist clones in a cyclically parthenogenetic organism evidenced by combined demographic and genetic analyses. Molecular Ecology, 9, 2055–66.CrossRefGoogle Scholar
Haff, L. A. and Smirnov, I. P. (1997). Single-nucleotide polymorphism identification assays using a thermostable DNA polymerase and delayed extraction MALDI-TOF mass spectrometry. Genome Research, 7, 378–88.CrossRefGoogle ScholarPubMed
Haig, S. M., Gratto-Trevor, C. L., Mullins, T. D. and Colwell, M. A. (1997). Population identification of western hemisphere shorebirds throughout the annual cycle. Molecular Ecology, 6, 412–27.CrossRefGoogle Scholar
Hales, D. F., Sloane, M. A., Wilson, A. C. C. and Sunnucks, P. (2002 a). Segregation of autosomes during spermatogenesis in the peach-potato aphid (Myzus persicae) (Sulzer), (Hemiptera: Aphididae). Genetical Research, 79, 119–27.CrossRefGoogle Scholar
Hales, D. F., Sloane, M. A., Wilson, A. C. C.et al. (2002 b). A lack of detectable genetic recombination on the X chromosome during the parthenogenetic production of female and male aphids. Genetical Research, 79, 203–9.CrossRefGoogle ScholarPubMed
Hardy, O. J. and Vekemans, X. (1999). Isolation by distance in a continuous population: reconciliation between spatial autocorrelation analysis and population genetics models. Heredity, 83, 145–54.CrossRefGoogle Scholar
Harper, G. L., Maclean, N. and Goulson, D. (2003). Microsatellite markers to assess the influence of population size, isolation and demographic change on the genetic structure of the UK butterfly Polyommatus bellargus. Molecular Ecology, 12, 3349–57.CrossRefGoogle ScholarPubMed
Haymer, D. S., McInnis, D. O. and Arcangeli, L. (1992). Genetic variation between strains of the Mediterranean fruit fly, Ceratitis capitata, detected by DNA fingerprinting. Genome, 35, 528–33.CrossRefGoogle ScholarPubMed
Haymer, D. S., He, M. and McInnis, D. O. (1997). Genetic marker analysis of spatial and temporal relationships among existing populations and new infestations of the Mediterranean fruit fly (Ceratitis capitata). Heredity, 97, 302–9.CrossRefGoogle Scholar
He, M. and Haymer, D. S. (1999). Genetic relationships of populations and the origins of new infestations of the Mediterranean fruit fly. Molecular Ecology, 8, 1247–57.CrossRefGoogle ScholarPubMed
Heckel, D. G., Gahan, L. J., Tabashnik, B. E. and Johnson, M. W. (1995). Randomly amplified polymorphic DNA differences between strains of diamondback moth (Lepidoptera: Plutellidae) susceptible or resistant to Bacillus thuringiensis. Annals of the Entomological Society of America, 88, 531–7.CrossRefGoogle Scholar
Heywood, . (1991). Spatial analysis of genetic variation in plant populations. Annual Review of Ecology and Systematics, 22, 335–55.CrossRefGoogle Scholar
Hibbard, B. E., Higdon, M. L., Duran, D. P., Chweikert, Y. M. and Ellersieck, M. R. (2004). Role of egg density on establishment and plant-to-plant movement by western corn rootworm larvae (Coleoptera: Chrysomelidae). Journal of Economic Entomology, 97, 871–82.CrossRefGoogle Scholar
Hillis, D. M., Moritz, C. and Mable, B. K. (1996). Molecular Systematics. Sinauer Associates, Sunderland MA.Google Scholar
Hinomoto, N., Muraji, M., Noda, T., Shimizu, T. and Kawasaki, K. (2004). Identification of five Orius species in Japan by multiplex polymerase chain reaction. Biological Control, 31, 276–9.CrossRefGoogle Scholar
Hoogendoorn, M. and 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, 2059–67.CrossRefGoogle ScholarPubMed
Hufbauer, R. A., Bogdanowicz, S. M., Perez, L. and Harrison, R. G. (2001). Isolation and characterization of microsatellites in Aphidius ervi (Hymenoptera: Braconidae) and their applicability to related species. Molecular Ecology Notes, 1, 197–9.CrossRefGoogle Scholar
Hufbauer, R. A., Bogdanowicz, S. M. and Harrison, R. G. (2004). The population genetics of a biological control introduction: mitochondrial DNA and microsatellie variation in native and introduced populations of Aphidus ervi, a parisitoid wasp. Molecular Ecology, 13, 337–48.CrossRefGoogle ScholarPubMed
Hurme, P. and Savolainen, O. (1999). Comparison of homology and linkage of random amplified polymorphic DNA (RAPD) markers between individual trees of Scots pine (Pinus sylvestris L.). Molecular Ecology, 8, 15–22.CrossRefGoogle Scholar
Ibrahim, K. M., Yassin, Y. and Elguzouli, A. (2004). Polymerase chain reaction primers for polymorphic microsatellite loci in the African armyworm, Spodoptera exempta (Lepidoptera: Noctuidae). Molecular Ecology Notes, 4, 653–5.CrossRefGoogle Scholar
Isabel, N., Beaulieu, J., Theriault, P. and Bousquet, J. (1999). Direct evidence for biased gene diversity estimates from dominant random amplified polymorphic DNA (RAPD) fingerprints. Molecular Ecology, 8, 477–83.CrossRefGoogle Scholar
Jeffreys, A. J., Wilson, V. and Thein, S. L. (1985). Hypervariable ‘minisatellite’ regions in human DNA. Nature, 314, 67–73.CrossRefGoogle ScholarPubMed
Johnson, K. P., Williams, B. L., Drown, D. M., Adams, R. J. and Clayton, D. H. (2002). The population genetics of host specificity: genetic differentiation in dove lice (Insecta: Phthiraptera). Molecular Ecology, 11, 25–38.CrossRefGoogle Scholar
Kauwe, J. S. K., Shizawa, D. K. and Evans, R. P. (2004). Phylogeographic and nested clade analysis of the stonefly Pteronarcys californica (Plecoptera: Pteronarcyidae) in the western USA. Journal of the North American Benthological Society, 23, 824–38.2.0.CO;2>CrossRefGoogle Scholar
Keller, I. and Largiaderm, C. R. (2003). Five microsatellite DNA markers for the ground beetle Abax parallelepipedus (Coleoptera: Carabidae). Molecular Ecology Notes, 3, 113–14.CrossRefGoogle Scholar
Kerdelhue, C., Mondor-Genson, G., Rasplus, J. Y., Robert, A. and Lieutier, F. (2003). Characterization of five microsatellite loci in the pine shoot beetle Tomicus piniperda (Coleoptera: Scolytidae). Molecular Ecology Notes, 3, 100–1.CrossRefGoogle Scholar
Kessler, A. and Baldwin, I. T. (2001). Defensive function of herbivore-induced plant volatile emissions in nature. Science, 291, 2141–4.CrossRefGoogle ScholarPubMed
Kessler, A. and Baldwin, I. T. (2004). Herbivore-induced plant vaccination. Part I. The orchestration of plant defenses in nature and their fitness consequences in the wild tobacco Nicotiana attenuate. The Plant Journal, 38, 639–49.CrossRefGoogle Scholar
Kessler, A., Halitschke, R. and Baldwin, I. T. (2004). Silencing the jasmonate cascade: induced plant defenses and insect populations. Science, 305, 665–8.CrossRefGoogle ScholarPubMed
Kim, K. S. and Sappington, T. W. (2004 a). Isolation and characterization of polymorphic microsatellite loci in the boll weevil/Anthonomus grandis Boheman (Coleoptera: Curculionidae). Molecular Ecology Notes, 4, 701–3.CrossRefGoogle Scholar
Kim, K. S. and Sappington, T. W. (2004 b). Boll weevil (Anthonomus grandis Boheman) (Coleoptera: Curculionidae) dispersal in the southern United States: Evidence from mitochondrial DNA variation. Environmental Entomology, 33, 457–70.CrossRefGoogle Scholar
Kinnear, M. W., Bariana, H. S., Sved, J. A. and Frommer, M. (1998). Polymorphic microsatellite markers for population analysis of a tephritid pest species, Bactrocera tryoni. Molecular Ecology, 7, 1489–95.CrossRefGoogle ScholarPubMed
Kirk, A. A., Lacey, L. A., Brown, J. K.et al. (2000). Variation in the Bemisia tabaci s. l. species complex (Hemiptera: Aleyrodidae) and its natural enemies leading to successful biological control of Bemisia biotype B in the USA. Bulletin of Entomological Research, 90, 317–27.CrossRefGoogle Scholar
Kjaer, E. D., Siegismund, H. R. and Suangtho, V. (1996). A multivariate study on genetic variation in teak (Tectona grandis L.). Silvae Genetica, 45, 361–8.Google Scholar
Knowles, L. L. (2004). The burgeoning field of statistical phylogeography. Journal of Evolutionary Biology, 17, 1–10.CrossRefGoogle ScholarPubMed
Koshio, C., Tomishima, M., Shimizu, K., Kim, H. S. and Takenaka, O. (2002). Microsatellites in the gypsy moth, Lymantria dispar L. (Lepidoptera: Lymantriidae). Applied Entomology and Zoology, 37, 309–12.CrossRefGoogle Scholar
Kreiger, M. J. B. and Ross, K. (2002). Identification of a major gene regulating complex social behavior. Science, 295, 328–32.CrossRefGoogle Scholar
Kuhner, M. K., Yamato, J. and Felsenstein, J. (1998). Maximum likelihood estimation of population growth rates based on the coalescent. Genetics, 149, 429–34.Google ScholarPubMed
Kurokawa, T., Yao, I., Akimoto, S. I. and Hasegawa, E. (2004). Isolation of six microsatellite markers from the pea aphid, Acyrthosiphon pisum (Homoptera: Aphididae). Molecular Ecology Notes, 4, 523–4.CrossRefGoogle Scholar
Laffin, R. D., Langor, D. W. and Sperling, F. A. H. (2004). Population structure and gene flow in the white pine weevil Pissodes strobe (Coleoptera: Curculionidae). Annals of the Entomological Society of America, 97, 949–56.CrossRefGoogle Scholar
Landry, B. S., Dextraze, L. and Boivin, G. (1993). Random amplified polymorphic DNA markers for DNA fingerprinting and genetic variability assessment of minute parasitic wasp species (Hymenoptera: Mymaridae and Trichogrammatidae) used in biological control programs of phytophagous insects. Genome, 36, 580–7.CrossRefGoogle ScholarPubMed
Lee, M. L. and Lee, M. H. (1997). Amplified mitochondrial DNA identify four species of Tetranychus mites (Acarina: Tetranychidae) in Korea. Korean Journal of Applied Entomology, 36, 30–6.Google Scholar
Legaspi, J. C., Legaspi, B. C., Carruthers, R. I.et al. (1996). Foreign exploration for natural enemies of Bemisia tabaci from Southeast Asia. Subtropical Plant Science, 48, 43–8.Google Scholar
Legg, J. P., French, R., Rogan, D., Okao-Okuja, G. and Brown, J. K. (2002). A distinct Bemisia tabaci (Gennadius) (Hemiptera: Sternorrhyncha: Aleyrodidae) genotype cluster is associated with the epidemic of severe cassava mosaic virus disease in Uganda. Molecular Ecology, 11, 1219–29.CrossRefGoogle ScholarPubMed
Lessa, E. P. (1990). Multidimensional analysis of geographic genetic structure. Systematic Zoology, 39, 242–52.CrossRefGoogle Scholar
Li, Z. X., Zheng, L. and Shen, Z. R. (2004). Using internally transcribed spacer. 2 sequences to re-examine the taxonomic status of several cryptic species of Trichogramma (Hymenoptera: Trichogrammatidae). European Journal of Entomology, 101, 347–58.CrossRefGoogle Scholar
Llewellyn, K. S., Loxdale, H. D., Harrington, R.et al. (2003). Migration and genetic structure of the grain aphid (Sitobion avenae) in Britain related to climate and clonal fluctuation as revealed using microsatellites. Molecular Ecology, 12, 21–34.CrossRefGoogle ScholarPubMed
Llewellyn, K. S., Loxdale, H. D., Harrington, R., Clark, S. J. and Sunnucks, P. (2004). Evidence for gene flow and local clonal selection in field populations of the grain aphid (Sitobion avenae) in Britain revealed using microsatellites. Heredity, 93, 143–53.CrossRefGoogle ScholarPubMed
Loxdale, H. D., Brookes, C. P., Wynne, I. R. and Clark, S. J. (1998). Genetic variability within and between English populations of the damson-hop aphid, Phorodon humuli (Hemiptera: Aphididae), with special reference to esterases associated with insecticide resistance. Bulletin of Entomological Research, 88, 513–26.CrossRefGoogle Scholar
Loxdale, H. D. and Hollander, J. (eds.) (1989). Electrophoretic Studies on Agricultural Pests. Systematics Association Special Volume No. 39. Oxford: Clarendon Press.Google Scholar
Loxdale, H. D. and Lushai, G. (1998). Molecular markers in entomology. Bulletin of Entomological Research, 88, 577–600.CrossRefGoogle Scholar
Loxdale, H. D. and Lushai, G. (1999). Slaves of the environment: the movement of herbivorous insects in relation to their ecology and genotype. Philosophical Transactions of the Royal Society London, Series B, 354, 1479–95.CrossRefGoogle Scholar
Loxdale, H. D., Tarr, I. J., Weber, C. P.et al. (1985). Electrophoretic study of enzymes from cereal aphid populations. III. Spatial and temporal genetic variation of populations of Sitobion avenae (F.) (Hemiptera: Aphididae). Bulletin of Entomological Research, 75, 121–41.CrossRefGoogle Scholar
Lu, Y. and Adang, M. J. (1996). Distinguishing fall armyworm (Lepidoptera: Noctuidae) strains using a diagnostic mitochondrial DNA marker. Florida Entomologist, 79, 48–55.CrossRefGoogle Scholar
Lu, Y. J., Adang, M. J., Isenhour, D. J. and Kochert, G. D. (1992). RFLP analysis of genetic variation in North American populations of the fall armyworm moth Spodoptera frugiperda (Lepidoptera: Noctuidae). Molecular Ecology, 1, 199–208.CrossRefGoogle Scholar
Lu, Y., Kochert, G. D., Isenhour, D. and Adang, M. J. (1994). Molecular characterization of a strain-specific repeated DNA sequence in the fall armyworm Spodoptera frugiperda (Lepidoptera: Noctuidae). Insect Molecular Biology, 3, 123–30.CrossRefGoogle Scholar
Luikart, G. and England, P. E. (1999). Statistical analysis of microsatellite data. Trends in Ecology and Evolution, 14, 253–6.CrossRefGoogle Scholar
MacDonald, C., Brookes, P., Edwards, K. J.et al. (2003). Microsatellite isolation and characterization in the beneficial parasitoid wasp Diaeretiella rapae (M'Intosh) (Hymenoptera: Braconidae: Aphidiinae). Molecular Ecology Notes, 4, 601–3.CrossRefGoogle Scholar
Macdonald, C. and Loxdale, H. D. (2004). Molecular markers to study population structure and dynamics in beneficial insects (predators and parasitoids). International Journal of Pest Management, 50, 215–24.CrossRefGoogle Scholar
Mantel, N. (1967). The detection of disease clustering and a generalised regression approach. Cancer Research, 27, 209–20.Google Scholar
Marshall, T. C., Slate, J., Kruuk, L. E. B. and Pemberton, J. M. (1998). Statistical confidence for likelihood-based paternity inference in natural populations. Molecular Ecology, 7, 639–55.CrossRefGoogle ScholarPubMed
Maruthi, M. N., Colvin, J., Thwaites, R. M.et al. (2004). Reproductive incompatibility and cytochrome oxidase I gene sequence variability amongst host-adapted and geographically separate Bemisia tabaci populations (Hemiptera: Aleyrodidae). Systematic Entomology, 29, 560–8.CrossRefGoogle Scholar
McKenzie, J. A. and Batterham, P. (1998). Predicting insecticide resistance: mutagenesis, selection and response. Philosophical Transactions of the Royal Society London, Series B, 353, 1729–34.CrossRefGoogle Scholar
McKenzie, L., Dransfield, L., Driver, F. and Curran, J. (1999). Molecular Protocols for Species Identification of Tephritid Fruit Flies. Report to the Australian Quarantine Inspection Service Canberra, CSIRO Entomology.Google Scholar
McMichael, M. and Prowell, D. P. (1999). Differences in amplified fragment-length polymorphisms in fall armyworm (Lepidoptera: Noctuidae) host strains. Annals of the Entomological Society of America, 92, 175–81.CrossRefGoogle Scholar
Meglécz, E., Petenian, F., Danchin, E.et al. (2004). High similarity between flanking regions of different microsatellites detected within each of two species of Lepidoptera: Parnassius apollo and Euphydryas aurinia. Molecular Ecology, 13, 1693–700.CrossRefGoogle ScholarPubMed
Meixner, M. D., McPheron, B. A., Silva, J. G., Gasparich, G. E. and Sheppard, W. S. (2002). The Mediterranean fruit fly in California: evidence for multiple introductions and persistent populations based on microsatellite and mitochondrial DNA variability. Molecular Ecology, 11, 891–9.CrossRefGoogle ScholarPubMed
Mendel, Z., Nestle, D. and Gafny, R. (1994). Examination of the origin of the Israeli population of Matsucoccus josephi (Homoptera: Matsucoccidae) using random amplified polymorphic DNA-polymerase chain reaction method. Annals of the Entomological Society of America, 87, 165–9.CrossRefGoogle Scholar
Miller, L. J., Allsopp, P. G., Graham, G. C. and Yeates, D. K. (1999). Identification of morphologically similar canegrubs (Coleoptera: Scarabaeidae: Melolonthini) using a molecular diagnostic technique. Australian Journal of Entomology, 38, 189–96.CrossRefGoogle Scholar
Miller, N. J., Birley, A. J., Overall, A. D. J. and Tatchell, G. M. (2003). Population genetic structure of the lettuce root aphid, Pemphigus bursarius (L.), in relation to geographic distance, gene flow and host plant usage. Heredity, 91, 217–23.CrossRefGoogle Scholar
Miller, N. J., Birley, A. J. and Tatchell, G. M. (2000). Polymorphic microsatellite loci from the lettuce root aphid, Pemphigus bursarius. Molecular Ecology, 9, 1951–2.CrossRefGoogle ScholarPubMed
Monteiro, F. A., Donnelly, M. J., Beard, C. B. and Costa, J. (2004). Nested clade and phylogeographic analyses of the Chagas disease vector Triatoma brasiliensis in Northeast Brazil. Molecular Phylogenetics and Evolution, 32, 46–56.CrossRefGoogle ScholarPubMed
Morin, P. A., Luikart, G. and Wayne, R. K. (2004). SNPs in ecology, evolution, and conservation. Trends in Ecology and Evolution, 19, 209–16.CrossRefGoogle Scholar
Moritz, C. and Lavery, S. (1996). Molecular ecology: contributions from molecular genetics to population ecology. In Floyd, R. B., Sheppard, A. W. and Barro, P. J. (eds.), Frontiers of Population Ecology. Melbourne: CSIRO Publishing. pp. 433–50.Google Scholar
Morris, D. C. and Mound, L. A. (2004). Molecular relationships between populations of South African citrus thrips (Scirtothrips aurantii Faure) in South Africa and Queensland, Australia. Australian Journal of Entomology, 43, 353–8.CrossRefGoogle Scholar
Moya, O., Contreras-Diaz, H. G., Oromi, P. and Juan, C. (2004). Genetic structure, phylogeography and demography of two ground-beetle species endemic to the Tenerife laurel forest (Canary Islands). Molecular Ecology, 13, 3153–67.CrossRefGoogle Scholar
Mueller, U. G. and Wolfenbarger, L. L. (1999). AFLP genotyping and fingerprinting. Trends in Ecology and Evolution, 14, 389–94.CrossRefGoogle ScholarPubMed
Mun, J., Bohonak, A. J. and Roderick, G. K. (2003). Population structure of the pumpkin fruit fly Bactrocera depressa (Tephritidae) in Korea and Japan: Pliocene allopatry or recent invastion?Molecular Ecology, 12, 2941–51.CrossRefGoogle ScholarPubMed
Muraji, M. and Nakahara, S. (2002). Discrimination among pest species of Bactrocera (Diptera: Tephritidae) based on PCR-RFLP of the mitochondrial DNA. Applied Entomology and Zoology, 37, 437–46.CrossRefGoogle Scholar
Muraji, M., Kawasaki, K., Shimizu, T. and Noda, T. (2004). Discrimination among Japanese species of the Orius flower bugs (Heteroptera: Anthocoridae) based on PCR-RFLP of the nuclear and mitochondrial DNAs. Japan Agricultural Research Quarterly, 38, 91–5.CrossRefGoogle Scholar
O'Hanlon, P. C., Peakall, R. and Briese, D. T. (1999). AFLP reveals introgression in weedy Onopordum thistles: hybridization and invasion. Molecular Ecology, 8, 1239–46.CrossRefGoogle ScholarPubMed
Paetkau, D., Calvert, W., Stirling, I. and Strobeck, C. (1995). Microsatellite analysis of population structure in Canadian polar bears. Molecular Ecology, 4, 347–54.CrossRefGoogle ScholarPubMed
Papura, D., Simon, J. C., Halkett, F.et al. (2003). Predominance of sexual reproduction in Romanian populations of the aphid Sitobion avenae inferred from phenotypic and genetic structure. Heredity, 90, 397–404.CrossRefGoogle ScholarPubMed
Parsons, Y. M. and Shaw, K. L. (2002). Mapping unexplored genomes: a genetic linkage map of the Hawaiian cricket Laupala. Genetics, 162, 1275–82.Google ScholarPubMed
Pashley, D. P., McMichael, M. and Silvain, J.-F. (2004). Multilocus genetic analysis of host use, introgression, and speciation in host strains of fall armyworm (Lepidoptera: Noctuidae). Annals of the Entomological Society of America, 97, 1034–44.Google Scholar
Peakall, R. and Smouse, P. E. (2001). GenAlEx V5: Genetic Analysis in Excel. Population Genetic Software for Teaching and Research. Canberra, Australia: Australian National University.Google Scholar
Pérez, T., Albornoz, J. and Domínguez, A. (1998). An evaluation of RAPD reproducibility and nature. Molecular Ecology, 7, 1347–58.CrossRefGoogle ScholarPubMed
Perring, T. M., Cooper, A. D., Rodriguez, R. J., Farrar, C. A. and Bellows, T. S. (1993). Identification of a whitefly species by genomic and behavioral studies. Science, 259, 74–7.CrossRefGoogle ScholarPubMed
Pfeifer, T. A., Humble, L. M., Ring, M. and Grigliatti, T. A. (1995). Characterization of gypsy moth populations and related species using a nuclear DNA marker. Canadian Entomologist, 127, 49–58.CrossRefGoogle Scholar
Podani, J. (1995). SYN-TAX 5.02.Mac: Computer Programs for Multivariate Data Analysis on the Macintosh System. Budapest: Scientia Publishing.Google Scholar
Polaszek, A., Manzari, S. and Quicke, D. L. J. (2004). Morphological and molecular taxonomic analysis of the Encarsia meritoria species-complex (Hymenoptera: Aphelinidae), parasitoids of whiteflies (Hemiptera: Aleyrodidae) of economic importance. Zoologica Scripta, 33, 403–21.CrossRefGoogle Scholar
Posada, D., Crandall, K. A. and Templeton, A. R. (2000). GeoDis: a program for the cladistic nested analysis of the geographical distribution of genetic haplotypes. Molecular Ecology, 9, 487–8.CrossRefGoogle ScholarPubMed
Powers, T. O., Jensen, S. G., Kindler, S. D., Stryker, C. J. and Sandall, L. J. (1989). Mitochondrial DNA divergence among Greenbug (Homoptera: Aphididae) biotypes. Annals of the Entomological Society of America, 82, 298–302.CrossRefGoogle Scholar
Pritchard, J. K., Stephens, M. and Donnelly, P. (2000). Inference of population structure using multilocus genotype data. Genetics, 155, 945–59.Google ScholarPubMed
Prowell, D. P., McMichael, M. and Silvain, J. F. (2004). Multilocus genetic analysis of host use, introgression, and speciation in host strains of all armyworm (Lepidoptera: Noctuidae). Annals of the Entomological Society of America, 97, 1034–44.CrossRefGoogle Scholar
Rabouam, C., Comes, A. M., Bretagnolle, V.et al. (1999). Features of DNA fragments obtained by random amplified polymorphic DNA (RAPD) assays. Molecular Ecology, 8, 493–504.CrossRefGoogle ScholarPubMed
Rannala, B. and Mountain, J. L. (1997). Detecting immigration by using multilocus genotypes. Proceedings of the National Academy of Sciences USA, 94, 9197–201.CrossRefGoogle ScholarPubMed
Reyes, A. and Ochando, M. D. (2004). Mitochondrial DNA variation in Spanish populations of Ceratitis capitata (Wiedemann) (Tephritidae) and the colonization process. Journal of Applied Entomology, 128, 358–64.CrossRefGoogle Scholar
Robinson, M. T., Weeks, A. R. and Hoffmann, A. A. (2002). Geographic patterns of clonal diversity in the earth mite species Penthaleus major with particular emphasis on species margins. Evolution, 56, 1160–7.CrossRefGoogle ScholarPubMed
Roda, A., Halitschke, R., Anke, S. and Baldwin, I. T. (2004). Individual variability in herbivore-specific elicitors from the plant's perspective. Molecular Ecology, 13, 2421–33.CrossRefGoogle ScholarPubMed
Ross, P., Hall, L., Smirnov, I. and Haff, L. (1998). High level multiplex genotyping by MALDI-TOF mass spectrometry. Nature Biotechnology, 16, 1347–551.CrossRefGoogle ScholarPubMed
Roy, A. S., McNamara, D. G. and Smith, I. M. (1995). Situation of Lymantria dispar in Europe. Bulletin OEPP, 25, 611–16.CrossRefGoogle Scholar
Saiki, R. K., Gelfand, D. H., Stoffel, S.et al. (1988). Science, 219, 487.CrossRef
Salle, A., Kerdelhue, C., Breton, M. and Lieutier, F. (2003). Characterization of microsatellite loci in the spruce bark beetle Ips typographus (Coleoptera: Scolytinae). Molecular Ecology Notes, 3, 336–7.CrossRefGoogle Scholar
Sandstrom, J. P., Russell, J. A., White, J. P. and Moran, N. A. (2001). Independent origins and horizontal transfer of bacterial symbionts of aphids. Molecular Ecology, 10, 217–28.CrossRefGoogle ScholarPubMed
Sapolsky, R. J., Hsie, L., Berno, A.et al. (1999). High-throughput polymorphism screening and genotyping with high-density oligonucleotide arrays. Genetic Analysis: Biomolecular Engineering, 14, 187–92.CrossRefGoogle ScholarPubMed
Scheffer, S. J. and Lewis, M. L. (2001). Two nuclear genes confirm mitochondrial evidence of cryptic species within Liriomyza huidobrensis (Diptera: Agromyzidae). Annals of the Entomological Society of America, 94, 648–53.CrossRefGoogle Scholar
Scheffer, S. J., Wijesekara, A., Visser, D. and Hallett, R. H. (2001). Polymerase chain reaction-restriction fragment-length polymorphism method to distinguish Liriomyza huidobrensis from L-langei (Diptera: Agromyzidae) applied to three recent leafminer invasions. Journal of Economic Entomology, 94, 1177–82.CrossRefGoogle ScholarPubMed
Schierwater, B., Streit, B., Wagner, G. P. and DeSalle, R. (eds.) (1994). Molecular Ecology and Evolution: Approaches and Applications. Birkhäuser Verlag, Basel.CrossRefGoogle Scholar
Schmidt, S., Naumann, I. D. and Barro, P. J. (2001). Encarsia species (Hymenoptera: Aphelinidae) of Australia and the Pacific Islands attacking Bemisia tabaci and Trialeurodes vaporariorum (Hemiptera: Aleyrodidae) – a pictorial key and descriptions of four new species. Bulletin of Entomological Research, 91, 369–87.CrossRefGoogle ScholarPubMed
Schneider, C. C., Kueffer, J. M., Roessli, D. and Excoffier, L. (1997). ARLEQUIN, version 1.1: A Software for Population Genetic Data Analysis. Genera: Genetics and Biometry Laboratory, University of Geneva.Google Scholar
Schultheis, A. S., Weigt, L. A. and Hendricks, A. C. (2002). Gene flow, dispersal, and nested clade analysis among populations of the stonefly Peltoperla tarteri in the southern Appalachians. Molecular Ecology, 11, 317–27.CrossRefGoogle ScholarPubMed
Scott, K. D., Lange, C. L., Scott, L. J. and Graham, G. C. (2004). Isolation and characterization of microsatellite loci from Helicoverpa armigera Hubner (Lepidoptera: Noctuidae). Molecular Ecology Notes, 4, 204–5.CrossRefGoogle Scholar
Segraves, K. A. and Pellmyr, O. (2004). Testing the out-of-Florida hypothesis on the origin of cheating in the yucca-yucca moth mutualism. Evolution, 58, 2266–79.Google ScholarPubMed
Sembene, M., Vautrin, D., Silvain, J. F., Rasplus, J. Y. and Delobel, A. (2003). Isolation and characterization of polymorphic microsatellites in the groundnut seed beetle, Caryedon serratus (Coleoptera: Bruchidae)Molecular Ecology Notes, 3, 299–301.CrossRefGoogle Scholar
Service, R. F. (1998). Microchip arrays put DNA on the spot. Science, 282, 396–9.CrossRefGoogle ScholarPubMed
Sheppard, W. S., Steck, G. J. and McPheron, B. A. (1992). Geographic populations of the medfly may be differentiated by mitochondrial DNA variation. Experientia, 48, 1010–13.CrossRefGoogle Scholar
Shoemaker, J. S., Painter, I. S. and Weir, B. S. (1999). Bayesian statistics in genetics: a guide for the uninitiated. Trends in Genetics, 15, 354–8.CrossRefGoogle ScholarPubMed
Shufran, K. A., Black, W. C. and Margolies, D. C. (1991). DNA fingerprinting to study spatial and temporal distributions of an aphid, Schizaphis graminum (Homoptera: Aphididae). Bulletin of Entomological Research, 81, 303–13.CrossRefGoogle Scholar
Shufran, K. A., Margolies, D. C. and Black, W. C. IV. (1992). Variation between biotype E clones of Schizaphis graminum (Homoptera: Aphididae). Bulletin of Entomological Research, 82, 407–16.CrossRefGoogle Scholar
Shufran, K. A. and Wilde, G. E. (1994). Clonal diversity in overwintering populations of Schizaphis graminum (Homoptera: Aphididae). Bulletin of Entomological Research, 84, 105–14.CrossRefGoogle Scholar
Shufran, K. A., Burd, J. D., Anstead, J. A. and Lushai, G. (2000). Mitochondrial DNA sequence divergence among greenbug (Homoptera: Aphididae) biotypes: evidence for host-adapted races. Insect Molecular Biology, 9, 179–84.CrossRefGoogle ScholarPubMed
Shufran, K. A., Weathersbee, A. A., Jones, D. B. and Elliott, N. C. (2004). Genetic similarities among geographic isolates of Lysiphlebus testaceipes (Hymenoptera: Aphidiidae) differing in cold temperature tolerances. Environmental Entomology, 33, 776–8.CrossRefGoogle Scholar
Silva, J. G., Meixner, M. D., McPheron, B. A., Steck, G. J. and Sheppard, W. S. (2003). Recent Mediterranean fruit fly (Diptera: Tephritidae) infestations in Florida – A genetic perspective. Journal of Economic Entomology, 96, 1711–18.CrossRefGoogle ScholarPubMed
Simon, J. C., Baumann, S., Sunnucks, P.et al. (1999 a). Reproductive mode and population genetic structure of the cereal aphid Sitobion avenae studied using phenotypic and microsatellite markers. Molecular Ecology, 8, 531–45.CrossRefGoogle ScholarPubMed
Simon, J. C., Carre, S., Boutin, M.et al. (2003). Host-based divergence in populations of the pea aphid: insights from nuclear markers and the prevalence of facultative symbionts. Proceedings of the Royal Society of London, Series B – Biological Science, 270, 1703–12.CrossRefGoogle ScholarPubMed
Simon, C., Frati, F., Beckenbach, A.et al. (1994). Evolution, weighting, and phylogenetic utility of mitochondrial gene sequences and a compilation of conserved polymerase chain reaction primers. Annals of the Entomological Society of America, 87, 651–701.CrossRefGoogle Scholar
Simon, J. C., Leterme, N., Delmotte, F., Martin, O. and Estoup, A. (2001). Isolation and characterization of microsatellite loci in the aphid species, Rhopalosiphum padi. Molecular Ecology Notes, 1, 4–5.CrossRefGoogle Scholar
Simon, J. C., Leterme, N. and Latorre, A. (1999 b). Molecular markers linked to breeding system differences in segregating and natural populations of the cereal aphid Rhopalosiphum padi L. Molecular Ecology, 8, 965–73.CrossRefGoogle ScholarPubMed
Simon, C., McIntosh, C. and Deniega, J. (1993). Standard restriction fragment length analysis of the mitochondrial genome is not sensitive enough for phylogenetic analysis or identification of 17-year periodical cicada broods (Hemiptera: Cicadidae): The potential for a new technique. Annals of the Entomological Society of America, 86, 228–38.CrossRefGoogle Scholar
Sin, F. Y. T., Suckling, D. M. and Marshall, J. W. (1995). Differentiation of the endemic New Zealand greenheaded and brownheaded leafroller moths by restriction fragment length variation in the ribosomal gene complex. Molecular Ecology, 4, 253–6.CrossRefGoogle Scholar
Slatkin, M. (1985). Gene flow in natural populations. Annual Review of Ecology and Systematics, 16, 393–430.CrossRefGoogle Scholar
Slatkin, M. and Maddison, W. P. (1989). A cladistic measure of gene flow inferred from the phylogenies of alleles. Genetics, 123, 603–13.Google Scholar
Sloane, M. A., Sunnucks, P., Wilson, A. C. C., Hales, D. F. and Sunnucks, P. (2001). Microsatellite isolation, linkage group identification and determination of recombination frequency in the peach-potato aphid, Myzus persicae (Sulzer) (Hemiptera: Aphididae). Genetical Research, 77, 251–60.CrossRefGoogle Scholar
Smouse, P. E. (1998). To tree or not to tree. Molecular Ecology, 7, 399–412.CrossRefGoogle Scholar
Smouse, P. E. and Peakall, R. (1999). Spatial autocorrelation analysis of individual multiallele and multilocus genetic structure. Heredity, 82, 561–73.CrossRefGoogle ScholarPubMed
Sokal, R. R. and Oden, N. L. (1978). Spatial autocorrelation in biology. 1. Methodology. Biological Journal of the Linnaean Society, 10, 199–228.CrossRefGoogle Scholar
Spackman, M. E. and McKechnie, S. W. (1995). Assessing the value of mitochondrial DNA variation for detecting population subdivision in the cotton bollworm, Helicoverpa armigera (Lepidoptera: Noctuidae) in Australia. Proceedings Beltwide Cotton Conference, 2, 811–13.Google Scholar
Sperling, F. A. H., Landry, J. F. and Hickey, D. A. (1995). DNA-based identification of introduced ermine moth species in North America (Lepidoptera: Yponomeutidae). Annals of the Entomological Society of America, 88, 155–62.CrossRefGoogle Scholar
Steiner, W. W. M., Voegtlin, D. J. and Irwin, M. E. (1985 a). Genetic differentiation and its bearing on migration in North American populations of the corn leaf aphid, Rhopalosiphum maidis (Fitch) (Homoptera: Aphididae). Annals of the Entomological Society of America, 78, 518–25.CrossRefGoogle Scholar
Steiner, W. W. M., Voegtlin, D. J., Irwin, M. E. and Kampmeier, G. (1985 b). Electrophoretic comparison of aphid species: detecting differences based on taxonomic status and host plant. Journal of Comparative Biochemistry and Physiology B, 81, 295–9.CrossRefGoogle Scholar
Stern, D. L., Aoki, S. and Kurosu, U. (1997). Determining aphid taxonomic affinities and life cycles with molecular data: a case study of the tribe Cerataphidini (Hormaphididae: Aphidoidea: Hemiptera). Systematic Entomology, 22, 81–96.CrossRefGoogle Scholar
Stine, O. C., Carnahan, A., Singh, R.et al. (2003). Characterization of microbial communities from coastal waters using microarrays. Environmental Monitoring and Assessment, 81, 327–36.CrossRefGoogle Scholar
Stipp, D. (1997). Gene chip breakthrough. Fortune, 135, 56–66.Google Scholar
Stone, G. N. and Sunnucks, P. (1993). Genetic consequences of an invasion through a patchy environment – the cynipid gallwasp Andricus quercuscalicis (Hymenoptera: Cynipidae). Molecular Ecology, 2, 251–68.CrossRefGoogle Scholar
Sunnucks, P., England, P. R., Taylor, A. C. and Hales, D. F. (1996). Microsatellite and chromosome evolution of parthenogenetic Sitobion aphids in Australia. Genetics, 144, 747–56.Google ScholarPubMed
Sunnucks, P., Barro, P. J., Lushai, G., Maclean, N. and Hales, D. (1997 a). Genetic structure of an aphid studies using microsatellites: cyclic parthenogenesis, differentiated lineages, and host specialisation. Molecular Ecology, 6, 1059–73.CrossRefGoogle Scholar
Sunnucks, P., Driver, F., Brown, W. V.et al. (1997 b). Biological and genetic characterization of morphologically similar Therioaphis trifolii (Hemiptera: Aphididae) with different host utilization. Bulletin of Entomological Research, 87, 152–62.CrossRefGoogle Scholar
Sunnucks, P. (2000). Efficient genetic markers for population biology. Trends in Ecology and Evolution, 15, 199–203.CrossRefGoogle ScholarPubMed
Sunnucks, P., Wilson, A. C. C., Beheregaray, L. B.et al. (2000). SSCP is not so difficult: the application and utility of single-stranded conformation polymorphism in evolutionary biology and molecular ecology. Molecular Ecology, 9, 1699–710.CrossRefGoogle Scholar
Symondson, W. O. C. and Liddell, J. E. (eds.) (1996). The Ecology of Agricultural Pests: Biochemical Approaches. London: Chapman and Hall.Google Scholar
Takami, Y. and Katada, S. (2001). Microsatellite DNA markers for the ground beetle Carabus insulicola. Molecular Ecology Notes, 1, 128–30.CrossRefGoogle Scholar
Templeton, A. R. (1998). Nested clade analyses of phylogeographic data: testing hypotheses about gene flow and population history. Molecular Ecology, 7, 381–97.CrossRefGoogle ScholarPubMed
Terradot, L., Simon, J. C., Leterme, N.et al. (1999). Molecular characterization of clones of the Myzus persicae complex (Hemiptera: Aphididae) differing in their ability to transmit the potato leafroll luteovirus (PLRV). Oecologia, 108, 121–9.Google Scholar
Toda, S. and Komazaki, S. (2002). Identification of thrips species (Thysanoptera: Thripidae) on Japanese fruit trees by polymerase chain reaction and restriction fragment length polymorphism of the ribosomal ITS2 region. Bulletin of Entomological Research, 92, 359–63.CrossRefGoogle ScholarPubMed
Torsvik, V. and Øvreås, L. (2002). Microbial diversity and function in soil: from genes to ecosystems. Current Opinion in Microbiology, 5, 240–5.CrossRefGoogle ScholarPubMed
Tsagkarakou, A. and Roditakis, N. (2003). Isolation and characterization of microsatellite loci in Bemisia tabaci (Hemiptera: Aleyrodidae). Molecular Ecology Notes, 3, 196–8.CrossRefGoogle Scholar
Turgeon, J. and McPeek, M. A. (2002). Phylogeographic analysis of a recent radiation of Enallagma damselflies (Odonata: Coenagrionidae). Molecular Ecology, 11, 1989–2001.CrossRefGoogle Scholar
Via, S. (1999). Reproductive isolation between sympatric races of pea aphids. I. Gene flow restriction and habitat choice. Evolution, 53, 1446–57.CrossRefGoogle ScholarPubMed
Villablanca, F. X., Roderick, G. K. and Palumbi, S. R. (1998). Invasion genetics of the Mediterranean fruit fly: variation in multiple nuclear introns. Molecular Ecology, 7, 547–60.CrossRefGoogle ScholarPubMed
Vink, C. J., Phillips, C. B., Mitchell, A. D., Winder, L. M. and Cane, R. P. (2003). Genetic variation in Microctonus aethiopoides (Hymenoptera: Braconidae). Biological Control, 28, 251–64.CrossRefGoogle Scholar
Viscarret, M. M., Torres-Jerez, I., Manero, E. A.et al. (2003). Mitochondrial DNA evidence for a distinct new world group of Bemisia tabaci (Gennadius) (Hemiptera: Aleyrodidae) indigenous to Argentina and Bolivia, and presence of the Old World B biotype in Argentina. Annals of the Entomological Society of America, 96, 65–72.CrossRefGoogle Scholar
Voelckel, C. and Baldwin, I. T. (2004 a). Herbivore-induced plant vaccination. Part II. Array-studies reveal the transiene of herbivore-specific transcriptional imprints and a distinct imprint from stress combinations. The Plant Journal, 38, 650–63.CrossRefGoogle Scholar
Voelckel, C. and Baldwin, I. T. (2004 b). Generalist and specialist lepidopteran larvae elicit different transcriptional responses in Nicotiana attenuata, which correlate with larval FAC profiles. Ecology Letters, 7, 770–5.CrossRefGoogle Scholar
Voelckel, C., Weisser, W. W. and Baldwin, I. T. (2004). An analysis of plant–aphid interactions by different microarray hybridization strategies. Molecular Ecology, 13, 3187–95.CrossRefGoogle ScholarPubMed
Vorburger, C., Lancaster, M. and Sunnucks, P. (2003). Environmentally related patterns of reproductive modes in the aphid Myzus persicae and the predominance of two ‘superclones’ in Victoria, Australia. Molecular Ecology, 12, 3493–504.CrossRefGoogle ScholarPubMed
Vorburger, C. (2005). Positive genetic correlations among major life-history traits relative to ecological success in the aphid Myzus persicae. Evolution, 59, 3493–3504.Google Scholar
Wagener, B., Reineke, A., Lohr, B. and Zebitz, C. P. W. (2004). A PCR-based approach to distinguish important Diadegma species (Hymenoptera: Ichneumonidae) associated with diamondback moth, Plutella xylostella (Lepidoptera: Plutellidae). Bulletin of Entomological Research, 94, 465–71.CrossRefGoogle Scholar
Wang, D., Fan, J., Siao, C.et al. (1998). Large-scale identification, mapping, and genotyping of single-nucleotide polymorphisms in the human genome. Science, 280, 1077–82.CrossRefGoogle ScholarPubMed
Waser, P. M. and Strobeck, C. (1998). Genetic signatures of interpopulation dispersal. Trends in Ecology and Evolution, 13, 43–4.CrossRefGoogle ScholarPubMed
Weeks, A. R., Marec, F. and Breeuwer, A. J. (2001). A mite species that consists entirely of haploid females. Science, 292, 2479–82.CrossRefGoogle ScholarPubMed
Wells, M. M. (1994). Small genetic distances among populations of green lacewings of the genus Chrysoperla (Neuroptera: Chrysopidae). Annals of the Entomological Society of America, 87, 737–44.CrossRefGoogle Scholar
Williams, C. L., Goldson, S. L., Baird, D. B. and Bullock, D. W. (1994). Geographical origin of an introduced insect pest, Listronotus bonariensis (Kuschel), determined by RAPD analysis. Heredity, 72, 412–19.CrossRefGoogle Scholar
Wilson, A. C. C., Sunnucks, P. and Hales, D. F. (1999). Microevolution, low clonal diversity and genetic affinities of parthenogenetic Sitobion aphids in New Zealand. Molecular Ecology, 8, 1655–66.CrossRefGoogle ScholarPubMed
Wilson, A. C. C., Sunnucks, P., Blackman, R. L. and Hales, D. F. (2002). Microsatellite variation in cyclically parthenogenetic populations of Myzus persicae in south-eastern Australia. Heredity, 88, 258–66.CrossRefGoogle ScholarPubMed
Wilson, A. C. C., Sunnucks, P. and Hales, D. F. (2003). Heritable genetic variation and potential for adaptive evolution in asexual aphids (Aphidoidea). Biological Journal of the Linnean Society, 79, 115–35.CrossRefGoogle Scholar
Wilson, A. C. C., Massonnet, B., Simon, J. C.et al. (2004). Cross-species amplification of microsatellite loci in aphids: assessment and application. Molecular Ecology Notes, 4, 104–9.CrossRefGoogle Scholar
Wright, S. (1951). The genetical structure of populations. Annals of Eugenics, 15, 323–54.CrossRefGoogle ScholarPubMed
Wu, L., Thompson, D. K., Li, G.et al. (2001). Development and evaluation of functional gene arrays for detection of selected genes in the environment. Applied and Environmental Microbiology, 67, 5780–90.CrossRefGoogle ScholarPubMed
Wyman, A. R. and White, R. (1980). A highly polymorphic locus in human DNA. Proceedings of the National Academy of Science, 77, 6754–8.CrossRefGoogle ScholarPubMed
Yu, H., Frommer, M., Robson, M. K.et al. (2001). Microsatellite analysis of the Queensland fruit fly Bactrocera tryoni (Diptera: Tephritidae) indicates spatial structuring: implications for population control. Bulletin of Entomological Research, 91, 139–47.Google ScholarPubMed
Yulin, A., Caihua, D., Hongbing, Z. and Guoyao, J. (1998). RAPD assessment of three sibling species of Monochamus guer (Coleoptera: Cerambycidae). Journal of Nanjing Forestry University, 22, 35–8.Google Scholar
Zanic, K., Cenis, J. L., Kacic, S. and Katalinic, M. (2005). Current status of Bemisia tabaci in coastal Croatia. Phytoparasitica, 33, 60–4.CrossRefGoogle Scholar
Zhao, J. T., Frommer, M., Sved, J. A. and Gillies, C. B. (2003). Genetic and molecular markers of the Queensland fruit fly, Bactrocera tryoni. Journal of Heredity, 94, 416–20.CrossRefGoogle ScholarPubMed
Zhang, A., Dunn, J. B. and Clark, J. M. (1999). An efficient strategy for validation of a point mutation associated with acetylcholinesterase sensitivity to azinphosmethyl in Colorado potato beetle. Pesticide Biochemistry and Physiology, 65, 25–35.CrossRefGoogle Scholar
Zhang, D.-X. (2004). Lepidopteran microsatellite DNA: redundant but promising. Trends in Ecology and Evolution, 19, 507–9.CrossRefGoogle ScholarPubMed
Zhou, J. (2003). Microarrays for bacterial detection and microbial community analysis. Current Opinion in Microbiology, 6, 288–94.CrossRefGoogle ScholarPubMed
Zhu, Y. C., Burd, J. D., Elliott, N. C. and Greenstone, M. H. (2000). Specific ribosomal DNA marker for early polymerase chain reaction detection of Aphelinus hordei (Hymenoptera: Aphelinidae) and Aphidius colemani (Hymenoptera: Aphidiidae) from Diuraphis noxia (Homoptera: Aphididae). Annals of the Entomological Society of America, 93, 486–91.CrossRefGoogle 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
×