Hostname: page-component-8448b6f56d-cfpbc Total loading time: 0 Render date: 2024-04-23T15:55:50.487Z Has data issue: false hasContentIssue false

POPULATION GENETICS OF THE SPRUCE BUDWORM, CHORISTONEURA FUMIFERANA (CLEM.) FREEMAN (LEPIDOPTERA: TORTRICIDAE), IN RELATION TO GEOGRAPHICAL AND POPULATION DENSITY DIFFERENCES

Published online by Cambridge University Press:  31 May 2012

G.T. Harvey
Affiliation:
Natural Resources Canada, Canadian Forest Service - Sault Ste. Marie, PO Box 490, Sault Ste. Marie, Ontario, Canada P6A 5M7

Abstract

Collections (68) of spruce budworm from 33 locations from Newfoundland to Alaska were analysed for isozyme frequencies using horizontal starch gels. Collections represented pre-, early-, mid-, late-, and post-outbreak stages of several populations in balsam fir, white spruce, and mixed host forests, as well as successive annual collections at several locations. Isozymes were measured at 11 loci in mature larvae and at six loci in pheromone-trapped males; frequencies were essentially the same in both stages, and from all host species. Three loci (IDH-2, LDH-1, and AAT-1) were found to be sex-linked, with no heterozygotes in females. Mean percentage heterozygosity ranged from 13.2 to 23.1; at individual locations it tended to decrease over successive years of outbreak and over successive collections in the same year. Contingency chi-square analysis indicated small differences related to location and outbreak history but all populations were generally homogeneous over the entire range. Nevertheless, one allozyme of AAT-1 exhibited a significant cline in frequency from the southeast to the northwest. Gene flow across the entire range appeared to be appreciable.

Résumé

Des récoltes (68) de Tordeuses des bourgeons de l’épinette en 33 localités, de Terre-Neuve à l’Alaska, ont été analysées par électrophorèse horizontale sur gel d’amidon quant à la fréquence des isozymes. Les échantillons représentaient divers stades épidémiques, pré-épidémie, début d’épidémie, milieu d’épidémie, fin d’épidémie, post-épidémie, de plusieurs populations, dans des forêts de sapins baumiers, d’épinettes blanches et dans des forêts mixtes, de même que des récoltes annuelles successives à différents endroits. Les isozymes ont été mesurés à 11 locus chez les larves à maturité et à six locus chez les mâles capturés dans des pièges à phéromones; la fréquence était essentiellement la même aux deux stades chez toutes les espèces d’hôtes. Trois locus (IDH-2, LDH-1 et AAT-1) se sont révélés liés aux chromosomes sexuels, mais il n’y avait pas d’hétérozygotes chez les femelles. Le pourcentage moyen d’hétérozygotie allait de 13,2 à 23,1; à certains endroits particuliers, ce pourcentage avait tendance à diminuer au cours des années successives d’épidémie et au cours des récoltes successives d’une même année. Un tableau de contingence (chi carré) a mis en lumière de petites différences reliées à la localité et au déroulement de l’épidémie, mais toutes les populations étaient généralement homogènes dans toute l’étendue de la répartition. Néanmoins, la fréquence d’un allozyme d’AAT-1 suivant une tendance significative du sud-est au nord-ouest. Le passage des gènes semble important dans toute d’étendue du territoire étudié.

[Traduit par la Rédaction]

Type
Articles
Copyright
Copyright © Entomological Society of Canada 1996

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

Anonymous. 1969. The Universal Transverse Mercator Grid. Department of Energy, Mines and Resources, Ottawa, Canada. 13 pp.Google Scholar
Ayala, F.J., Powell, J.R., Tracey, M.L., Mourão, C.A., and Pérez-Salas, S.. 1972. Enzyme variability in the Drosophila willistoni group. IV. Genic variation in natural populations of Drosophila willistoni. Genetics 70: 113139.CrossRefGoogle ScholarPubMed
Berryman, A.A. 1987. The theory and classification of outbreaks. pp. 3–29 in Barbosa, P., and Schultz, J.C. (Eds.), Insect Outbreaks. Academic Press, New York, NY. 578 pp.Google Scholar
Brussard, P.F., Ehrlich, P.R., Wilcox, D.D., and Wright, J.. 1985. Genetic distances and the taxonomy of checkerspot butterflies (Nymphalidae: Nymphalininae). Journal of the Kansas Entomological Society 58: 403412.Google Scholar
Carson, H. 1968. The population flush and its genetic consequences. pp. 123–137 in Lewontin, R.C. (Ed.), Population Biology and Evolution: Proceedings of the International Symposium, June 7–9, 1967. Syracuse University Press, Syracuse, NY. 205 pp.Google Scholar
Castrovillo, P.J., and Stock, M.W.. 1981. Electrophoretic techniques for detection of Glypta fumiferanae, an endoparasitoid of western spruce budworm. Entomologia experimentalis et applicata 30: 176180.CrossRefGoogle Scholar
David, J.R., Alonso-Moraga, A., Borai, F., Capy, P., Merçot, H., McEvey, S.F., Munoz-Serrano, A., and Tsakas, S.. 1989. Latitudinal variation of Adh gene frequencies in Drosophila melanogaster: A Mediterranean instability. Heredity 62: 1116.CrossRefGoogle ScholarPubMed
Feder, J.L., and Bush, G.L.. 1989. Gene frequency clines for host races of Rhagoletis pomonella in the midwestern United States. Heredity 63: 245266.CrossRefGoogle Scholar
Forest Insect and Disease Survey (Canada). 19701980. Annual Report of the Forest Insect and Disease Survey. Canadian Forestry Service, Ottawa, Ont.Google Scholar
Forest Insect and Disease Survey (Canada). 19811989. Forest Insect and Disease Conditions in Canada. Canadian Forestry Service, Ottawa, Ont.Google Scholar
Gooding, R.H., Rolseth, B.M., Byers, J.R., and Herle, C.E.. 1992. Electrophoretic comparisons of pheromotypes of the dingy cutworm, Feltia jaculifera (Gn.) (Lepidoptera: Noctuidae). Canadian Journal of Zoology 70: 7986.CrossRefGoogle Scholar
Greenbank, D.O., Schaefer, G.W., and Rainey, R.C.. 1980. Spruce Budworm Moth Flight and Dispersal: New Understanding from Canopy Observations, Radar, and Aircraft. Memoirs of the Entomological Society of Canada 110: 49 pp.Google Scholar
Grisdale, D. 1970. An improved laboratory method for rearing large numbers of spruce budworm, Choristoneura fumiferana (Lepidoptera: Tortricidae). The Canadian Entomologist 102: 11111117.CrossRefGoogle Scholar
Hardy, Y., Mainville, M., and Schmidt, D.M.. 1986. An Atlas of Spruce Budworm Defoliation in Eastern North America, 1938–80. Department of Agriculture, Forest Service, Cooperative State Research Service, Washington, DC. Miscellaneous Publication 1449: 52 pp.Google Scholar
Harvey, G.T. 1980. Sampling low density spruce budworm populations for isozyme analysis. The Canadian Entomologist 112: 969970.CrossRefGoogle Scholar
Harvey, G.T. 1985. The taxonomy of coniferophagous Choristoneura (Lepidoptera: Tortricidae): A review. pp. 16–59 in Sanders, C.J., Stark, R.W., Mullins, E.J., and Murphy, J.. (Eds.), Recent Advances in Spruce Budworms Research: Proceedings of the CANUSA Spruce Budworms Research Symposium, 16–20 September 1984, Bangor, Maine. Supply and Services, Ottawa, Canada. 527 pp.Google Scholar
Harvey, G.T., and Sohi, S.S.. 1985. Isozyme characterization of 28 cell lines from five insect species. Canadian Journal of Zoology 63: 22702276.CrossRefGoogle Scholar
Heywood, J.S. 1991. Spatial analysis of genetic variation in plant populations. Annual Review of Ecology and Systematics 22: 335355.CrossRefGoogle Scholar
Hudson, A., and Lefkovitch, L.P.. 1982. Allozyme variation in four Ontario populations of Xestia adela and Xestia dolosa and in a British population of Xestia c-nigrum (Lepidoptera: Noctuidae). Annals of the Entomological Society of America 75: 250256.CrossRefGoogle Scholar
Jennings, M.J., and Philipp, D.P.. 1992. Genetic variation in the longear sunfish (Lepomis megalotis). Canadian Journal of Zoology 70: 16731680.CrossRefGoogle Scholar
Kettela, E. 1983. A Cartographic History of Spruce Budworm Defoliation 1967 to 1981 in Eastern North America. Canadian Forestry Service, Ottawa, Ontario, Information Report DPC–X–14: 8 pp.Google Scholar
Levene, H. 1949. On a matching problem arising in genetics. Annals of Mathamatical Statistics 20: 9194.CrossRefGoogle Scholar
Löfstedt, C. 1990. Population variation and genetic control of pheromone communication systems in moths. Entomologia experimentalis et applicata 54: 199218.CrossRefGoogle Scholar
May, B., Leonard, D.E., and Vadas, R.L.. 1977. Electophoretic variation and sex linkage in spruce budworm. Journal of Heredity 68: 355359.CrossRefGoogle Scholar
Mitter, C., and Schneider, J.C.. 1987. Genetic change and insect outbreaks. pp. 505–532 in Barbosa, P., and Schultz, J.C. (Eds.), Insect Outbreaks. Academic Press, New York, NY. 578 pp.Google Scholar
Morris, R.F. 1954. A sequential sampling technique for spruce budworm egg surveys. Canadian Journal of Zoology 32: 303313.CrossRefGoogle Scholar
Namkoong, G., Roberds, J.H., Nunnally, L.B., and Thomas, H.A.. 1979. Isozyme variations in populations of southern pine beetles. Forest Science 25: 197203.Google Scholar
Nei, M. 1978. Estimation of average heterozygosity and genetic distance from a small number of individuals. Genetics 89: 583590.CrossRefGoogle ScholarPubMed
Pashley, D.P., and Bush, G.L.. 1979. The use of allozymes in studying insect movement with special reference to the codling moth, Laspyresia pomonella (L.) (Olethreutidae). pp. 333–341 in Rabb, R.L., and Kennedy, G.G. (Eds.), Movement of Highly Mobile Insects: Concepts and Methodology in Research. North Carolina State University, Raleigh, NC. 546 pp.Google Scholar
Pashley, D.P., Johnson, S.J., and Sparks, A.N.. 1985. Genetic population structure of migratory moths: The fall armyworm (Lepidoptera: Noctuidae). Annals of the Entomological Society of America 78: 756762.CrossRefGoogle Scholar
Poulik, M.D. 1957. Starch gel electrophoresis in a discontinuous system of buffers. Nature (London) 180: 14771479.Google Scholar
Preziosi, R.F., and Fairbairn, D.J.. 1992. Genetic population structure and levels of gene flow in the stream dwelling waterstrider, Aquarius (=Gerris) remigis (Hemiptera: Gerridae). Evolution 46: 430444.Google ScholarPubMed
Régnière, J., Lysyk, T.J., and Auger, M.. 1989. Population density estimation of spruce budworm, Choristoneura fumiferana (Clem.) (Lepidoptera: Tortricidae) on balsam fir and white spruce from 45 - cm mid-crown branch tips. The Canadian Entomologist 121: 267281.CrossRefGoogle Scholar
Richardson, B.J., Baverstock, P.R., and Adams, M.. 1986. Allozyme Electrophoresis. A Handbook for Animal Systematics and Population Studies. Academic Press, Sydney, Australia. 410 pp.Google Scholar
Royama, T. 1984. Population dynamics of spruce budworm Choristoneura fumiferana. Ecological Monographs 54: 429462.CrossRefGoogle Scholar
Sanders, C.J. 1978. Evaluation of sex attractant traps for monitoring spruce budworm populations (Lepidoptera: Tortricidae). The Canadian Entomologist 110: 4350.CrossRefGoogle Scholar
Sanders, C.J. 1992. Sex Pheromone Traps for Monitoring Spruce Budworm Problems: Resolving Operational Problems. Forestry Canada, Ontario Region, Sault Ste. Marie, Ontario, Information Report O–X–425: 30 pp.Google Scholar
Sanders, C.J., Stark, R.W., Mullins, E.J., and Murphy, J. (Eds.). 1985. Recent Advances in Spruce Budworm Research: Proceedings of the CANUSA Spruce Budworms Research Symposium, 16–20 September 1984, Bangor, Maine. Supply and Services, Ottawa, Canada. 527 pp.Google Scholar
Shaw, C.R., and Prasad, R.. 1970. Starch gel electrophoresis of enzymes—a compilation of recipes. Biochemical Genetics 4: 297320.CrossRefGoogle ScholarPubMed
Shepherd, R.F., Gray, T.G., and Harvey, G.T.. 1995. Geographical distribution of Choristoneura species (Lepidoptera: Tortricidae) feeding on Abies, Picea, and Pseudotsuga in western Canada and Alaska. The Canadian Entomologist 127: 813830.CrossRefGoogle Scholar
Shore, T.L., and Alfaro, R.I.. 1986. The spruce budworm, Choristoneura fumiferana (Lepidoptera: Tortricidae), in British Columbia. Journal of the Entomological Society of British Columbia 83: 3138.Google Scholar
Slatkin, M. 1985. Rare alleles as indicators of gene flow. Evolution 39: 5365.CrossRefGoogle ScholarPubMed
Slatkin, M., and Barton, N.H.. 1989. A comparison of three indirect methods for estimating average levels of gene flow. Evolution 43: 13491368.CrossRefGoogle ScholarPubMed
Snedecor, G.W. 1946. Statistical Methods. The Iowa State College Press, Ames, IA. 485 pp.Google ScholarPubMed
Stehr, G. 1959. Hemolymph polymorphism in a moth and the nature of sex-controlled inheritance. Evolution 13: 537560.CrossRefGoogle Scholar
Stock, M.W., and Castrovillo, P.J.. 1981. Genetic relationships among representative populations of five Choristoneura species: C. occidentalis, C. retiniana, C. biennis, C. lambertiana and C. fumiferana (Lepidoptera: Tortricidae). The Canadian Entomologist 113: 857865.CrossRefGoogle Scholar
Stock, M.W., and Robertson, J.L.. 1980. Inter- and intra-specific variation in selected Choristoneura species (Lepidoptera: Tortricidae): A toxicological and genetic survey. The Canadian Entomologist 112: 10191027.CrossRefGoogle Scholar
Sturgeon, K.B., and Mitton, J.B.. 1986. Allozyme and morphological differentiation of mountain pine beetles Dendroctonus ponderosae Hopkins (Coleoptera: Scolytidae) associated with host tree. Evolution 40: 290302.CrossRefGoogle ScholarPubMed
Swofford, D.L. 1981. On the utility of the Distance Wagner Procedure. pp. 25–43 in Funk, V.A., and Brooks, D.R. (Eds.), Advances in Cladistics. Proceedings of the First Meeting of the Willi Hennig Society. The New York Botanical Garden, Bronx, NY. 232 pp.Google Scholar
Swofford, D.L., and Selander, R.B.. 1981. BIOSYS-1: A Fortran program for the comprehensive analysis of electrophoretic data in population genetics and systematics. Journal of Heredity 72: 281283.CrossRefGoogle Scholar
Tumquist, R., and Ferris, R.. 1989. Forest Insect and Disease Conditions. Prince George Region 1988. Forestry Canada, Pacific Forestry Centre, Victoria, British Columbia, Report 89–4: 8 pp.Google Scholar
Volney, W.J.A. 1989. Biology and dynamics of North American coniferophagous Choristoneura populations. Agricultural Zoology Reviews 3: 113156.Google 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., Liebhold, A.M., and Waters, W.E.. 1983. Effects of temperature, sex and genetic background on coloration of Choristoneura spp. (Lepidoptera: Tortricidae) populations in south-central Oregon. The Canadian Entomologist 115: 15831596.CrossRefGoogle Scholar
Wallner, W.E. 1987. Factors affecting insect population dynamics: Differences between outbreak and non-outbreak species. Annual Review of Entomology 32: 317340.CrossRefGoogle Scholar
Wilhite, E.A. 1979. Genetics of Outbreaking Western Spruce Budworm, Choristoneura occidentalis Freeman (Lepidoptera: Tortricidae), Populations in Idaho and Montana. M.S. thesis, University of Idaho, Moscow, ID. 106 pp.Google Scholar
Wilhite, E.A., and Stock, M.W., 1983. Genetic variation among western spruce budworm (Choristoneura occidentalis) (Lepidoptera: Tortricidae) outbreaks in Idaho and Montana. The Canadian Entomologist 115: 4154.CrossRefGoogle Scholar
Wood, C.S., and Van Sickle, G.A.. 1989. Forest Insect and Disease Conditions. British Columbia and Yukon 1988. Forestry Canada, Pacific Forestry Centre, Victoria, British Columbia, Information Report BC–X–306: 33 pp.Google Scholar
Wood, C.S., Van Sickle, G.A., and Humble, L.M.. 1987. Forest Insect Disease Conditions. British Columbia and Yukon 1987. Canadian Forestry Service, Pacific Forestry Centre, Victoria, British Columbia, Information Report BC–X–296: 40 pp.Google Scholar
Wright, S. 1978. Evolution and the Genetics of Populations. Vol. 4. Variability Within and Among Natural Populations. University of Chicago Press, Chicago, IL. 580 pp.Google Scholar