Hostname: page-component-848d4c4894-4rdrl Total loading time: 0 Render date: 2024-06-16T15:12:02.918Z Has data issue: false hasContentIssue false

Clinal geographic variation in mammals: implications for the study of chronoclines

Published online by Cambridge University Press:  08 April 2016

Paul L. Koch*
Affiliation:
Museum of Paleontology and Department of Geological Sciences, University of Michigan, Ann Arbor, Michigan 48109

Abstract

Mammalian species often exhibit clinal geographic variation in body size: individuals tend to be larger in areas with lower mean annual temperature. Climatic change involving increasing or decreasing mean annual temperature may cause clines to shift geographically, resulting in a phenotypic shift at all affected locales within a species' range. I assess the potential of shifting geographic clines to produce morphological trends in the fossil record. Five extant North American mammalian species (Didelphis virginiana, Mephitis mephitis, Odocoileus virginianus, Scalopus aquaticus, and Sciurus carolinensis) are examined to quantify size change along latitudinal clines and to estimate the geographic range and temperature difference commonly associated with a given difference in body size. Relative to body size, the observed size range of skeletal characters within each of these five species is comparable to that seen in a much larger sample of North American mammals. Thus patterns of variation documented for the five species may be used to assess the likelihood of dine translocation as an explanation of size change in the mammalian fossil record. As a case study, I examine three lineages from the Early Eocene of the Bighorn Basin, Wyoming. I determine that size change in these chronoclines represents evolutionary change and is not merely the result of shifting geographic clines.

Type
Articles
Copyright
Copyright © The Paleontological Society 

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

Literature Cited

Barnett, R.J. 1977. Bergmann's Rule and variation in structures related to feeding in the gray squirrel. Evolution. 31:538545.Google Scholar
Bell, M. A. and Haglund, T. R. 1982. Fine-scale temporal variation of the Miocene stickleback Gasterosteus doryssus. Paleobiology. 8:282292.Google Scholar
Boucot, A. J. 1982. Ecophenotypic or genotypic? Nature. 296:609.CrossRefGoogle Scholar
Boyce, M. S. 1978. Climatic variability and body size variation in the muskrat (Ondatra zibethicus) of North America. Oecologia. 36:119.Google Scholar
Brown, J. H. and Lee, A. K. 1969. Bergmann's Rule and climatic adaptation in woodrats (Neotoma). Evolution. 23:329338.Google Scholar
Buchardt, B. 1978. Oxygen isotope palaeotemperatures from the Tertiary of the North Sea area. Nature. 275:121123.CrossRefGoogle Scholar
Davis, S. J. M. 1981. The effects of temperature change and domestication on the body size of late Pleistocene to Holocene mammals of Israel. Paleobiology. 7:101114.CrossRefGoogle Scholar
Dorf, E. 1964. The use of plants in paleoclimatic interpretations. Pp. 1331. In: Nairn, A. E. M., ed. Problems in Paleoclimatology. Interscience; London.Google Scholar
Gauthreaux, S. A. Jr. 1980. The influence of long-term and short-term climatic changes on the dispersal and migrations of organisms. Pp. 103174. In: Gauthreaux, S. A. Jr., ed. Animal Migration, Orientation and Navigation. Academic Press; New York.Google Scholar
Gazin, C. L. 1968. A study of the Eocene condylarthran mammal Hyopsodus. Smithsonian Misc. Coll. 153:190.Google Scholar
Gingerich, P. D. 1974. Size variability of the teeth in living mammals and diagnosis of closely related sympatric fossil species. J. Paleontol. 48:895903.Google Scholar
Gingerich, P. D. 1976. Paleontology and phylogeny: patterns of evolution at the species level in early Tertiary mammals. Am. J. Sci. 276:128.Google Scholar
Gingerich, P. D. 1979. The stratophenetic approach to phylogeny reconstruction in vertebrate paleontology. Pp. 4177. In: Cracraft, J. and Eldredge, N., eds. Phylogenetic Analysis and Paleontology. Columbia Univ. Press; New York.Google Scholar
Gingerich, P. D. 1980. Evolutionary pattern in early Cenozoic mammals. Ann. Rev. Ear. Planet. Sci. 8:407424.Google Scholar
Gingerich, P. D. and Schoeninger, M. J. 1979. Patterns of tooth size variability in the dentition of primates. Am. J. Phys. Anthropol. 51:457466.Google Scholar
Gingerich, P. D. and Simons, E. L. 1977. Systematics, phylogeny, and evolution of Early Eocene Adapidae (Mammalia, Primates) in North America. Contr. Mus. Paleontol. Univ. Mich. 24:245279.Google Scholar
Gingerich, P. D., Smith, B. H., and Rosenberg, K. 1980. Patterns of allometric scaling in the primate dentition and prediction of body size from tooth size. Am. J. Phys. Anthropol. 52:321322.Google Scholar
Gingerich, P. D. and Winkler, D. A. 1979. Patterns of variation and correlation in the dentition of the red fox Vulpes vulpes. J. Mamm. 60:691704.CrossRefGoogle Scholar
Godinot, M. 1981. Les mammifères de Rians (Eocene inferieur, Provence). Palaeovertebrata. 10:43126.Google Scholar
Gould, S. J. 1975. On the scaling of tooth sizes in mammals. Am. Zool. 15:351362.Google Scholar
Gould, S. J. and Eldredge, N. 1977. Punctuated equilibria: the tempo and mode of evolution reconsidered. Paleobiology. 3:115151.CrossRefGoogle Scholar
Guilday, J. E. 1967. Differential extinction during the late Pleistocene and Recent times. Pp. 121140. In: Martin, P. S. and Wright, H. E. Jr., eds. Pleistocene Extinctions: The Search for a Cause. Yale Univ. Press; New Haven.Google Scholar
Guilday, J. E., Martin, P. S., and McCrady, A. D. 1964. New Paris No. 4: a Pleistocene cave deposit in Bedford County, Pennsylvania. Bull. Nat. Speleol. Soc. 26:122194.Google Scholar
Hall, R. E. 1981. The Mammals of North America. Vol. 1 and 2. Wiley; New York.Google Scholar
Hambridge, G., ed. 1941. Climate and Man: Yearbook of Agriculture 1941. U.S. Gov. Printing Office; Washington, D.C.Google Scholar
Herreid, C. F. and Kessel, B. 1967. Thermal conductance in birds and mammals. Comp. Biochem. Physiol. 21:405414.Google Scholar
Hickey, L. J. 1977. Stratigraphy and paleobotany of the Golden Valley Formation (early Tertiary) of western North Dakota. Geol. Soc. Am. Mem. 150:1183.Google Scholar
Hickey, L. J. 1980. Clark's Fork Basin flora. Univ. Mich. Papers Paleontol. 24:3349.Google Scholar
Hooijer, D. A. 1967. Indo-Australian insular elephants. Genetica. 38:143162.CrossRefGoogle ScholarPubMed
Hutchinson, G. E. 1959. Homage to Santa Rosalia, or Why are there so many kinds of animals? Am. Nat. 93:145159.Google Scholar
Johnson, D. L. 1978. The origin of island mammoths and the Quaternary land bridge history of the North Channel Islands, California. Quat. Res. 10:204225.Google Scholar
Kay, R. F. 1975. The functional adaptations of primate molar teeth. Am. J. Phys. Anthropol. 43:195215.Google Scholar
Kendeigh, S. C. 1969. Tolerance to cold and Bergmann's Rule. Auk. 86:1325.CrossRefGoogle Scholar
King, J. F. and Saunders, J. J. 1984. Environmental insularity and the extinction of the American mastodont. Pp. 315339. In: Martin, P. S. and Klein, R. G., eds. Quaternary Extinctions: A Prehistoric Revolution. Univ. Ariz. Press; Tucson.Google Scholar
Kowalski, K. 1970. Variation and speciation in fossil voles. Symp. Zool. Soc. Lond. 26:149161.Google Scholar
Kurtén, B. 1960. Chronology and faunal evolution of the earlier European glaciations. Comm. Biol. Soc. Sci. Fennica. 21:162.Google Scholar
Kurtén, B. 1968. Pleistocene Mammals of Europe. Weidenfield Nicolson; London.Google Scholar
Lindsey, C. C. 1966. Body sizes of poikilotherm vertebrates at different latitudes. Evolution. 20:456465.Google Scholar
Lohmann, G. P. and Malmgren, B. A. 1983. Equatorward migration of Globorotalia truncatulinoides ecophenotypes through the late Pleistocene: gradual evolution or ocean change? Paleobiology. 9:414421.Google Scholar
MacGinitie, H. D. 1974. An early Middle Eocene flora from the Yellowstone-Absaroka Volcanic Province northwestern Wind River Basin, Wyoming. Cal. Univ. Pubs. Geo. Sci. 108:1103.Google Scholar
Marshall, L. G. and Corruccini, R. S. 1978. Variability, evolutionary rates, and allometry in dwarfing lineages. Paleobiology. 4:101118.Google Scholar
Mayr, E. 1956. Geographic character gradients and climatic adaptation. Evolution. 10:105108.Google Scholar
Mayr, E. 1982. Questions concerning speciation. Nature. 296:609.Google Scholar
McDonald, J. N. 1984. The recorded North American selection regime and late Quaternary megafaunal extinctions. Pp. 404439. In: Martin, P. S. and Klein, R. G., eds. Quaternary Extinctions: A Prehistoric Revolution. Univ. Ariz. Press; Tucson.Google Scholar
McNab, B. K. 1971. On the ecological significance of Bergmann's Rule. Ecology. 52:845854.Google Scholar
Murphy, E. C. 1985. Bergmann's Rule, seasonality, and geographic variation in body size of house sparrows. Evolution. 39:13271334.Google Scholar
Purdue, J. R. 1980. Clinal variation of some mammals during the Holocene in Missouri. Quat. Res. 13:242258.CrossRefGoogle Scholar
Ray, C. 1960. The application of Bergmann's and Allen's Rule to the poikilotherms. J. Morphol. 106:85108.Google Scholar
Rees, J. W. 1969. Morphologic variation in the mandible of the white-tailed deer (Odocoileus virginanus): a study of populational skeletal variation by principal component and canonical analyses. J. Morphol. 128:113130.Google Scholar
Rensch, B. 1936. Studien über klimatische Parallelität der Merkmalsausprägung bei Vögeln und Säugern. Arch. Naturgesch. (N. F.). 5:317363.Google Scholar
Rose, K. D. 1981. The Clarkforkian land-mammal age and mammalian faunal composition across the Paleocene-Eocene boundary. Univ. Mich. Papers Paleontol. 26:1196.Google Scholar
Rosenzweig, M. L. 1968. The strategy of body size in mammalian carnivores. Am. Midl. Nat. 80:299315.Google Scholar
Russell, D. E. 1975. Paleoecology of the Paleocene-Eocene transition in Europe. Pp. 2861. In: Szalay, F. S., ed. Approaches to Primate Paleobiology, Contributions to Primatology, vol. 5. S. Krager; New York.Google Scholar
Savin, S. M., Douglas, R. G., and Stehli, F. G. 1975. Tertiary marine paleotemperatures. Geol. Soc. Am. Bull. 86:14991510.Google Scholar
Simons, E. L. 1962. A new Eocene primate genus, Cantius, and a revision of some allied European lemuroids. Bull. Brit. Mus. Nat. Hist. Geol. 7:136.Google Scholar
Sokal, R. R. and Rohlf, F. J. 1981. Biometry. 2d ed.W. H. Freeman; San Francisco.Google Scholar
Tanai, T. and Huzioka, K. 1967. Climatic implications of Tertiary floras in Japan. Pp. 7787. In: Tertiary Correlation and Climatic Changes in the Pacific. 11th Pac. Sci. Congr., Tokyo 1966, Symp. 25.Google Scholar
Williamson, P. G. 1981. Palaeontological documentation of speciation in Cenozoic molluscs from Turkana Basin. Nature. 293:437443.Google Scholar
Williamson, P. G. 1982. Williamson replies. Nature. 296:611612.CrossRefGoogle Scholar
Wolfe, J. A. 1978. A paleobotanical interpretation of Tertiary climates in the Northern Hemisphere. Am. Sci. 66:694703.Google Scholar
Wolfe, J. A. 1980. Tertiary climates and floristic relationships at high latitudes in the Northern Hemisphere. Palaeogeogr. Palaeoclimatol. Palaeoecol. 30:313323.Google Scholar
Wolfe, J. A. and Hopkins, D. M. 1967. Climatic changes recorded by Tertiary land floras in northwestern North America. Pp. 6776. In: Tertiary Correlations and Climatic Changes in the Pacific. 11th Pac. Sci. Congr., Tokyo 1966, Symp. 25.Google Scholar
World Weather Records. 1961–1970. V. 1–3. U.S. Dept. of Commerce, Environmental Science Services Administration. Environmental Data Service; Washington, D.C.Google Scholar