Hostname: page-component-848d4c4894-xfwgj Total loading time: 0 Render date: 2024-07-02T18:45:27.419Z Has data issue: false hasContentIssue false

From branch to bench: establishing wild spruce budworm populations into laboratory colonies for the exploration of local adaptation and plasticity

Published online by Cambridge University Press:  24 February 2021

K. Perrault
Affiliation:
Natural Resources Canada, Canadian Forest Service, Great Lakes Forestry Centre, 1219 Queen St. East, Sault Ste. Marie, Ontario, P6A 2E6, Canada
A.A. Wardlaw
Affiliation:
Natural Resources Canada, Canadian Forest Service, Great Lakes Forestry Centre, 1219 Queen St. East, Sault Ste. Marie, Ontario, P6A 2E6, Canada
J.N. Candau
Affiliation:
Natural Resources Canada, Canadian Forest Service, Great Lakes Forestry Centre, 1219 Queen St. East, Sault Ste. Marie, Ontario, P6A 2E6, Canada
C.L. Irwin
Affiliation:
Natural Resources Canada, Canadian Forest Service, Great Lakes Forestry Centre, 1219 Queen St. East, Sault Ste. Marie, Ontario, P6A 2E6, Canada
M. Demidovich
Affiliation:
Natural Resources Canada, Canadian Forest Service, Great Lakes Forestry Centre, 1219 Queen St. East, Sault Ste. Marie, Ontario, P6A 2E6, Canada
C.J.K. MacQuarrie
Affiliation:
Natural Resources Canada, Canadian Forest Service, Great Lakes Forestry Centre, 1219 Queen St. East, Sault Ste. Marie, Ontario, P6A 2E6, Canada
A.D. Roe*
Affiliation:
Natural Resources Canada, Canadian Forest Service, Great Lakes Forestry Centre, 1219 Queen St. East, Sault Ste. Marie, Ontario, P6A 2E6, Canada
*
*Corresponding author. Email: [email protected]

Abstract

Spruce budworm, Choristoneura fumiferana (Clemens) (Lepidoptera: Tortricidae), is a destructive defoliator found throughout the Nearctic boreal forest. This pest has a broad geographic range and shows regional variation in key life history traits. These population differences may represent important adaptations to local environmental conditions and reflect underlying genetic diversity. Existing laboratory colonies of spruce budworm do not capture this regional variation, so we established five new spruce budworm colonies from across its range to explore regional adaptations among spruce budworm populations within common garden experiments. We present methods for establishing new spruce budworm laboratory colonies from wild populations. We describe the process of flushing, rearing, and disease screening used on these new populations to produce healthy disease-free laboratory stocks.

Type
Research Papers
Copyright
© The authors, and Her Majesty, the Queen, in right of Canada, 2021. Published by Cambridge University Press on behalf of the Entomological Society of Canada

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.)

Footnotes

Subject editor: Kevin Floate

References

Antonovics, J. 1971. The effects of a heterogeneous environment on the genetics of natural populations: the realization that environments differ has had a profound effect on our views of the origin and role of genetic variability in populations. American Scientist, 59: 593599.Google Scholar
Bauer, L.S. and Nordin, G.L. 1988. Pathogenicity of Nosema fumiferanae (Thomson) (Microsporida) in spruce budworm, Choristoneura fumiferana (Clemens), and implications of diapause conditions. The Canadian Entomologist, 120: 221229. https://doi.org/10.4039/Ent120221-3.CrossRefGoogle Scholar
Bauer, L.S. and Nordin, G.L. 1989. Effect of Nosema fumiferanae (Microsporidia) on fecundity, fertility, and progeny performance of Choristoneura fumiferana (Lepidoptera: Tortricidae). Environmental Entomology, 18: 261265.CrossRefGoogle Scholar
Berthier, K., Chapuis, M.P., Simpson, S.J., Ferenz, H.J., Kane, C.M.H., Kang, L., et al. 2010. Laboratory populations as a resource for understanding the relationship between genotypes and phenotypes. Advances in Insect Physiology, 39: 137.CrossRefGoogle Scholar
Bird, F.T. 1969. Infection and mortality of spruce budworm, Choristoneura fumiferana, and forest tent caterpillar, Malacosoma disstria, caused by nuclear and cytoplasmic polyhedrosis viruses. The Canadian Entomologist, 101: 12691285. https://doi.org/10.4039/Ent1011269-12.CrossRefGoogle Scholar
Blackburn, G.S., Brunet, B.M.T., Muirhead, K., Cusson, M., Beliveau, C., Levesque, R.C., et al. 2017. Distinct sources of gene flow produce contrasting population genetic dynamics at different range boundaries of a Choristoneura budworm. Molecular Ecology, 26: 66666684. https://doi.org/10.1111/mec.14386.CrossRefGoogle ScholarPubMed
Blais, J.R. 1983. Trends in the frequency, extent and severity of spruce budworm outbreaks in eastern Canada. Canadian Journal of Forest Research, 13: 539547. https://doi.org/10.1139/x83-079.CrossRefGoogle Scholar
Candau, J.-N., Fleming, R.A., and Hopkin, A. 1998. Spatiotemporal patterns of large-scale defoliation caused by the spruce budworm in Ontario since 1941. Canadian Journal of Forest Research, 28: 17331741. https://doi.org/10.1139/x98-164.CrossRefGoogle Scholar
Caprio, M.A. 2009. Genetic considerations and strategies for rearing high-quality insects. In Principles and procedures for rearing high-quality insects. Edited by J.C. Schneider. Mississippi State University, Starkville, Mississippi, United States of America. Pp. 8795.Google Scholar
Chevin, L.M., Lande, R., and Mace, G.M. 2010. Adaptation, plasticity, and extinction in a changing environment: towards a predictive theory. PLOS Biology, 8: e1000357. https://doi.org/10.1371/journal.pbio.1000357.CrossRefGoogle Scholar
de Villemereuil, P., Gaggiotti, O.E., Mouterde, M., and Till-Bottraud, I. 2016. Common garden experiments in the genomic era: new perspectives and opportunities. Heredity, 116: 249254. https:doi.org/10.1038/hdy.2015.93.CrossRefGoogle ScholarPubMed
Du, W.G., Warner, D.A., Langkilde, T., Robbins, T., and Shine, R. 2010. The physiological basis of geographic variation in rates of embryonic development within a widespread lizard species. The American Naturalist, 176: 522528.CrossRefGoogle ScholarPubMed
Ebling, P.M. and Dedes, J. 2015. Rearing diapause Choristoneura fumiferana. Insect Production Services Standard Operating Procedure IPS/003/003. Natural Resources Canada, Canadian Forest Service, Great Lakes Forestry Centre, Sault Ste. Marie, Ontario, Canada. 40 pp. Available from: https://cfs.nrcan.gc.ca/publications?id=35979 [accessed 13 March 2020].Google Scholar
Ebling, P.M. and Demidovich, M. 2015. Quality control for diapause Choristoneura fumiferana. Standard Operating Procedure IPS/006/005. Natural Resources Canada, Canadian Forest Service, Great Lakes Forestry Centre, Insect Production Services, Sault Ste. Marie, Ontario, Canada. 22 pp. Available from: https://cfs.nrcan.gc.ca/publications?id=35993 [accessed 13 March 2020].Google Scholar
Evans, H. and Shapiro, M. 1997. Viruses. In Manual of Techniques in Insect Pathology. Edited by L.A. Lacey. Academic Press, Cambridge, Massachusetts, United States of America. Pp. 1753.Google Scholar
Eveleigh, E.S. and Johns, R.C. 2014. Intratree variation in the seasonal distribution and mortality of spruce budworm (Lepidoptera: Tortricidae) from the peak to collapse of an outbreak. Annals of the Entomological Society of America, 107: 435444. http://dx.doi.org/10.1603/AN13136.CrossRefGoogle Scholar
Fuxa, J.R., Sun, J.Z., Weidner, E.H., and LaMotte, L.R. 1999. Stressors and rearing diseases of Trichoplusia ni: evidence of vertical transmission of NPV and CPV. Journal of Invertebrate Pathology, 74: 149155. https://doi.org/10.1006/jipa.1999.4869.CrossRefGoogle ScholarPubMed
Godinho, D.P., Cruz, M.A., Charlery de la Masselière, M., Teodoro-Paulo, J., Eira, C., Fragata, I., et al. 2020. Creating outbred and inbred populations in haplodiploids to measure adaptive responses in the laboratory. Ecology and Evolution, 10: 116. https://doi.org/10.1002/ece3.6454.CrossRefGoogle ScholarPubMed
Han, E.N. and Bauce, E. 1998. Timing of diapause initiation, metabolic changes and overwintering survival of the spruce budworm, Choristoneura fumiferana . Ecological Entomology, 23: 160167. https://doi.org/10.1046/j.1365-2311.1998.00111.x.CrossRefGoogle Scholar
Hardy, Y., Mainville, M., and Schmitt, D.M. 1986. An atlas of spruce budworm defoliation in eastern North America, 1938–80. Miscellaneous Publication 1449. United States Department of Agriculture Forest Service, Beltsville, Maryland, United States of America. Available from: https://cfs.nrcan.gc.ca/publications?id=14608 [accessed 13 March 2020].Google Scholar
Harvey, G.T. 1983. A geographical cline in egg weights in Choristoneura fumiferana (Lepidoptera: Tortricidae) and its significance in population dynamics. Canadian Entomologist, 115: 11031108. https://doi.org/10.4039/Ent1151103-9.CrossRefGoogle Scholar
Harvey, G.T. 1985. The taxonomy of the coniferophagous Choristoneura (Lepidoptera: Tortricidae): a review. In Recent Advances in Spruce Budworms Research: Proceedings of the CANUSA Spruce Budworms Research Symposium, Bangor, Maine, September 16–20, 1984. Agriculture Canada, Ministry of State for Forestry, Headquarters, Ottawa, Ontario, Canada and United Stated Department of Agriculture Forest Service. 527 pp.Google Scholar
Inglis, G.D. and Sikorowski, P.P. 2009. Entomopathogens and insect rearing. In Principles and procedures for rearing high-quality insects. Edited by J.C. Schneider. Mississippi State University, Starkville, Mississippi, United States of America. Pp. 223288.Google Scholar
James, P.M.A., Cooke, B., Brunet, B.M.T., Lumley, L.M., Sperling, F.A.H., Fortin, M.J., et al. 2015. Life-stage differences in spatial genetic structure in an irruptive forest insect: implications for dispersal and spatial synchrony. Molecular Ecology, 24: 296309. https://doi.org/10.1111/mec.13025.CrossRefGoogle Scholar
Kawecki, T.J. and Ebert, D. 2004. Conceptual issues in local adaptation. Ecology Letters, 7: 12251241. https://doi.org/10.1111/j.1461-0248.2004.00684.x.CrossRefGoogle Scholar
Larroque, J., Legault, S., Johns, R., Lumley, L., Cusson, M., Renaut, S., et al. 2019. Temporal variation in spatial genetic structure during population outbreaks: distinguishing among different potential drivers of spatial synchrony. Evolutionary Applications, 12: 19311946. https://doi.org/10.1111/eva.12852.CrossRefGoogle ScholarPubMed
Lewis, L.C. and Lynch, R.E. 1970. Treatment of Ostrinia nubilalis larvae with Fumidil-B to control infections caused by Perezia pyraustae . Journal of Invertebrate Pathology, 15: 4348. https://doi.org/10.1016/0022-2011(70)90096-0.CrossRefGoogle Scholar
Lumley, L.M., Pouliot, E., Laroche, J., Boyle, B., Brunet, B.M.T., Levesque, R.C., et al. 2020. Continent-wide population genomic structure and phylogeography of North America’s most destructive conifer defoliator, the spruce budworm (Choristoneura fumiferana). Ecology and Evolution, 10: 914927. https://doi.org/10.1002/ece3.5950.CrossRefGoogle Scholar
Lysyk, T.J. 1989. Stochastic model of eastern spruce budworm (Lepidoptera: Tortricidae) phenology on white spruce and balsam fir. Journal of Economic Entomology, 82: 11611168. https://doi.org/10.1093/jee/82.4.1161.CrossRefGoogle Scholar
MacLean, D.A. 1980. Vulnerability of fir spruce stands during uncontrolled spruce budworm outbreaks: a review and discussion. The Forestry Chronicle, 56: 213221. https://doi.org/10.5558/tfc56213-5.CrossRefGoogle Scholar
Marshall, K.E. and Sinclair, B.J., 2015. The relative importance of number, duration and intensity of cold stress events in determining survival and energetics of an overwintering insect. Functional Ecology, 29: 357366. https://doi.org/10.1111/1365-2435.12328.CrossRefGoogle Scholar
Miller, C.A. 1957. A technique to assess second-instar spruce budworm populations with notes on overwintering mortality. Interim Report 1956–3. Canadian Forest Service, Victoria, British Columbia, Canada. 32 pp.Google Scholar
Miller, C.A. 1958. The measurement of spruce budworm populations and mortality during the first and second larval instars. Canadian Journal of Zoology, 36: 409422.CrossRefGoogle Scholar
McMorran, A. 1965. A synthetic diet for the spruce budworm, Choristoneura fumiferana (Clem.) (Lepidoptera: Tortricidae). The Canadian Entomologist, 97: 5862.CrossRefGoogle Scholar
McMorran, A. 1973. Effects of pre-storage treatment on survival of diapausing larvae of the spruce budworm, Choristoneura fumiferana (Lepidoptera: Tortricidae). The Canadian Entomologist, 105: 10051009.CrossRefGoogle Scholar
Mori, H. and Metcalf, P. 2010. Cypoviruses. In Insect Virology. Edited by S. Asgari and K.N. Johnson. Caister Academic Press, Norfolk, United Kingdom. Pp. 307323.Google Scholar
Palli, S.R., Ladd, T.R., Tomkins, W.L., Shu, S., Ramaswamy, S.B., Tanaka, Y., et al. 2000. Choristoneura fumiferana entomopoxvirus prevents metamorphosis and modulates juvenile hormone and ecdysteroid titers. Insect Biochemistry and Molecular Biology, 30: 869876.CrossRefGoogle ScholarPubMed
Podgwaite, J.D., Shields, K.S., Zerillo, R.T., and Bruen, R.B. 1979. Environmental persistence of the nucleopolyhedrosis virus of the gypsy moth, Lymantria dispar . Environmental Entomology, 8: 528536.CrossRefGoogle Scholar
Roe, A.D., Demidovich, M., and Dedes, J. 2018. Origins and history of laboratory insect stocks in a multispecies insect production facility, with the proposal of standardized nomenclature and designation of formal standard names. Journal of Insect Science, 18: 1.CrossRefGoogle Scholar
Sanders, C.J. 1980. Summary of current techniques used for sampling spruce budworm populations and estimating defoliation in Eastern Canada. Information Report O-X-306. Canadian Forestry Service, Great Lakes Forest Research Centre, Sault Ste. Marie, Ontario, Canada. 33 pp. Available from: https://cfs.nrcan.gc.ca/publications?id=8982 [accessed 13 March 2020].Google Scholar
Weber, J.D., Volney, W.J.A., and Spence, J.R. 1999. Intrinsic development rate of spruce budworm (Lepidoptera: Tortricidae) across a gradient of latitude. Environmental Entomology, 28: 224232.CrossRefGoogle Scholar
van Frankenhuyzen, K., Ebling, P., McCron, B., Ladd, T., Gauthier, D., and Vossbrinck, C. 2004. Occurrence of Cystosporogenes sp. (Protozoa, Microsporidia) in a multi-species insect production facility and its elimination from a colony of the eastern spruce budworm, Choristoneura fumiferana (Clem.) (Lepidoptera: Tortricidae). Journal of Invertebrate Pathology, 87: 1628.CrossRefGoogle Scholar
van Frankenhuyzen, K., Nystrom, C.W., and Tabashnik, B.E. 1995. Variation in tolerance to Bacillus thuringiensis among and within populations of the spruce budworm (Lepidoptera: Tortricidae) in Ontario. Journal of Economic Entomology, 88: 97105.CrossRefGoogle Scholar
Volney, W.J.A. and Cerezke, H.F. 1992. The phenology of white spruce and the spruce budworm in northern Alberta. Canadian Journal of Forest Research, 22: 198205.CrossRefGoogle Scholar
Volney, W.J.A. and Fleming, R.A. 2007. Spruce budworm (Choristoneura spp.) biotype reactions to forest and climate characteristics. Global Change Biology, 13: 16301643.CrossRefGoogle Scholar
Whitlock, V.H. 1974. Symptomatology of two viruses infecting Heliothis armigera . Journal of Invertebrate Pathology, 23: 7075.CrossRefGoogle Scholar