Hostname: page-component-848d4c4894-ndmmz Total loading time: 0 Render date: 2024-06-11T16:57:56.690Z Has data issue: false hasContentIssue false

From regional to total geographic ranges: testing the relationship in Recent bivalves

Published online by Cambridge University Press:  08 April 2016

David Jablonski
Affiliation:
Department of the Geophysical Sciences, University of Chicago, 5734 S. Ellis Avenue, Chicago, Illinois 60637
James W. Valentine
Affiliation:
Department of Geological Sciences, University of California, Santa Barbara, California 93106

Abstract

Geographic range appears to be an important aspect of the biology of species, but ranges cannot be unambiguously determined from the fossil record: ancient species can rarely be traced to the full extent of their original geographic ranges, and some ancient provinces are far better characterized than others. Here we test the degree to which paleogeographic range data for species within ancient provinces can be used to estimate the relative magnitudes of total geographic ranges, focusing on the 212 living bivalve species of the Oregonian Province, a north-south (N-S) province of the northeastern Pacific shelf.

We break the total range of each species into its within-province and extraprovincial components. Nonparametric rank correlation tests and simple linear regressions yield highly significant correlations, both between within-province and total ranges (which are not independent variables) and between within-province and extraprovincial ranges. The strongest correlations are obtained when north-ranging species are excluded. These north-ranging species are thermally tolerant and have access to broad east-west (E-W) provinces, thus weakening the relation between their within-province and total ranges. On a coarser scale, we also found that the extraprovincial ranges of species with small within-province ranges (<650 km, ca. 25% of focal province) are significantly less than those for species with large within-province ranges (> 1950 km, ca. 75% of focal province). The significant correlations of within-province range with both total and extraprovincial ranges hold for species with within-province ranges >650 km, so caution is needed regarding species that extend into a focal province only a short distance from its edges. We conclude that within-province geographic range can generally be used to rank species by total geographic range and thus is an adequate proxy in most comparative studies of the paleobiologic consequences of range magnitudes.

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

Baxter, R. 1989. Mollusks of Alaska. Alaska Department of Fish and Game; Bethel, Alaska[19 February 1989 edition].Google Scholar
Bernard, F. R. 1983. Catalogue of the living Bivalvia of the eastern Pacific Ocean: Bering Strait to Cape Horn. Canadian Special Publication of Fisheries and Aquatic Sciences 61.Google Scholar
Bernard, F. R. MS.Bivalvia of the Northeastern Pacific Ocean. In preparation, to be published by Stanford University Press, Stanford, California.Google Scholar
Briggs, J. C. 1974. Marine Zoogeography. McGraw-Hill; New York.Google Scholar
Brusca, R. C., and Wallerstein, B. R. 1979. Zoogeographic patterns of idoteid isopods in the northeast Pacific, with a review of shallow water zoogeography of the area. Bulletin of the Biological Society of Washington 3:67105.Google Scholar
Carlton, J. T. 1975. Introduced intertidal invertebrates. Pp. 1725. In Smith, R. I., and Carlton, J. T. (eds.), Light's Manual: Intertidal Invertebrates of the Central California Coast. Third Edition. University of California Press; Berkeley, California.Google Scholar
Carlton, J. T. 1987. Patterns of transoceanic marine biological invasions in the Pacific Ocean. Bulletin of Marine Science 41:452465.Google Scholar
Conover, W. J. 1980. Practical Nonparametric Statistics. Second Edition. Wiley; New York.Google Scholar
Cronin, T. M., and Ikeya, N. 1988. The Omma-Manganji ostracod fauna (Plio-Pleistocene) of Japan and the zoogeography of circumpolar species. Journal of Micropalaeontology 6:6588.Google Scholar
Doyle, R. F. 1985. Biogeographical Studies of the Point Conception Region, California. Ph.D. dissertation, University of California, Santa Barbara.Google Scholar
Durham, J. W., and Macneil, F. S. 1967. Cenozoic migrations of marine invertebrates through the Bering Strait region. Pp. 326349. In Hopkins, D. M. (ed.), The Bering Land Bridge. Stanford University Press; Stanford, California.Google Scholar
Ekman, S. 1953. Zoogeography of the Sea. Sidgwick and Jackson; London.Google Scholar
Franz, D. R., and Merrill, A. S. 1980a. Molluscan distribution patterns on the continental shelf of the Middle Atlantic Bight (northwest Atlantic). Malacologia 19:209225.Google Scholar
Franz, D. R. and Merrill, A. S. 1980b. The origins and determinants of distribution of molluscan faunal groups on the shallow continental shelf of the northwest Atlantic. Malacologia 19:227248.Google Scholar
Hall, C. A. Jr. 1964. Shallow-water marine climates and molluscan provinces. Ecology 45:226234.Google Scholar
Hanna, G. D. 1966. Introduced mollusks of western North America. California Academy of Sciences Occasional Paper 48.Google Scholar
Hansen, T. A. 1980. Influence of larval dispersal and geographic distribution on species longevity in neogastropods. Paleobiology 6:193207.CrossRefGoogle Scholar
Hayden, B. P., and Dolan, R. 1976. Coastal marine fauna and marine climates of the Americas. Journal of Biogeography 3:7181.Google Scholar
Hoagland, K. E., and Turner, R. D. 1980. Range extensions of teredinids (shipworms) and polychaetes in the vicinity of a temperate-zone nuclear generating station. Marine Biology 58:5564.Google Scholar
Hoffman, A., and Szubdza-Studencka, B. 1982. Bivalve species duration and ecological characteristics in the Badenian (Miocene) marine sandy facies of Poland. Neues Jahrbuch für Geologie und Paläontologie Abhandlungen 163:122135.Google Scholar
Horn, M. H., and Allen, L. G. 1978. A distributional analysis of California coastal marine fishes. Journal of Biogeography 5:2342.Google Scholar
Jablonski, D. 1980. Apparent versus real effects of transgressions and regressions. Paleobiology 6:397407.Google Scholar
Jablonski, D. 1986a. Background and mass extinction: The alternation of macroevolutionary regimes. Science 231:129133.CrossRefGoogle ScholarPubMed
Jablonski, D. 1986b. Larval ecology and macroevolution in marine invertebrates. Bulletin of Marine Science 39:565587.Google Scholar
Jablonski, D. 1986c. Causes and consequences of mass extinctions: A comparative approach. Pp. 183229. In Elliott, D. K. (ed.), Dynamics of Extinction. Wiley; New York.Google Scholar
Jablonski, D. 1986d. Evolutionary consequences of mass extinctions. Pp. 313329. In Raup, D. M., and Jablonski, D. (eds.), Patterns and Processes in the History of Life. Springer-Verlag; Berlin.CrossRefGoogle Scholar
Jablonski, D. 1987. Heritability at the species level: Analysis of geographic ranges of Cretaceous mollusks. Science 238:360363.CrossRefGoogle ScholarPubMed
Jablonski, D. 1988. Estimates of species duration. Science 240:969.CrossRefGoogle ScholarPubMed
Jablonski, D. 1989. The biology of mass extinction: A paleontological view. Philosophical Transactions of the Royal Society of London 325:357368.Google Scholar
Jablonski, D., Flessa, K. W., and Valentine, J. W. 1985. Biogeography and paleobiology. Paleobiology 11:7590.CrossRefGoogle Scholar
Jablonski, D., and Valentine, J. W. 1981. Onshore-offshore gradients in Recent eastern Pacific shelf faunas and their paleobiogeographic significance. Pp. 441453. In Scudder, G. G. E., and Reveal, J. L. (eds.), Evolution Today. Carnegie-Mellon University; Pittsburgh, Pennsylvania.Google Scholar
Jackson, J. B. C. 1974. Biogeographic consequences of eurytopy and stenotopy among marine bivalves and their evolutionary significance. American Naturalist 108:541560.CrossRefGoogle Scholar
Keen, A. M. 1971. Sea Shells of Tropical West America. Second Edition. Stanford University Press; Stanford, California.Google Scholar
Koch, C. F. 1987. Prediction of sample size effects on the measured temporal and geographic distribution patterns of species. Paleobiology 13:100107.Google Scholar
Koch, C. F., and Morgan, J. P. 1988. On the expected distribution of species' ranges. Paleobiology 14:126138.Google Scholar
Lankford, R. R., and Phleger, F. B. 1973. Foraminifera from the nearshore turbulent zone, western North America. Journal of Foraminiferal Research 3:101132.Google Scholar
Macneil, F. S. 1965. Evolution and distribution of the genus Mya, and Tertiary migrations of Mollusca. U.S. Geological Survey Professional Paper 483-G.Google Scholar
Maluf, L. Y. 1988. Biogeography of the central eastern Pacific shelf echinoderms. Pp. 389398. In Burke, R. D., Mladanov, P. V., Lambert, P., and Parsley, R. L. (eds.), Echinoderm Biology. Proceedings of the 6th International Echinoderm Conference. A.A. Balkema; Rotterdam.Google Scholar
Murray, S. N., and Littler, M. M. 1981. Biogeographical analysis of intertidal macrophyte floras of southern California. Journal of Biogeography 8:339351.Google Scholar
Newman, W. A. 1979. Californian Transition Zone: Significance of shoft-range endemics. Pp. 399416. In Gray, J., and Boucot, A. J. (eds.), Historical Biogeography, Plate Tectonics, and the Changing Environment. Oregon State University Press; Corvallis, Oregon.Google Scholar
Owen, R. W. 1980. Eddies of the California Current System: Physical and ecological characteristics. Pp. 217263. In Power, D. M. (ed.), The California Islands: Proceedings of a Multidisciplinary Symposium. Santa Barbara Museum of Natural History; Santa Barbara, California.Google Scholar
Pease, C. M. 1988. On comparing the geologic durations of easily versus poorly fossilized taxa. Journal of Theoretical Biology 133:255257.CrossRefGoogle Scholar
Peden, A. E., and Wilson, D. E. 1976. Distribution of intertidal and subtidal fishes of northern British Columbia and southeastern Alaska. Syesis 9:221248.Google Scholar
Raffi, S., Stanley, S. M., and Marasti, R. 1986. Biogeographic patterns and Plio-Pleistocene extinction of Bivalvia in the Mediterranean and southern North Sea. Paleobiology 11:368388.Google Scholar
Reid, J. L., Roden, G. I., and Wyllie, J. G. 1958. Studies of the California Current System. California Cooperative Fisheries Investigation Progress Report, 1 July 1956 to 1 January 1958:2756.Google Scholar
Ricklefs, R. E. 1989. Speciation and diversity: The integration of local and regional processes. Pp. 599622. In Otte, D., and Endler, J. A. (eds.), Speciation and its Consequences. Sinauer; Sunderland, Massachusetts.Google Scholar
Rosen, B. R. 1984. Reef coral biogeography and climate through the Late Cainozoic: Just islands in the sun or a critical pattern of islands? Pp. 201262. In Brenchley, P. J. (ed.), Fossils and Climate. Wiley; Chichester.Google Scholar
Russell, M. P., and Lindberg, D. R. 1988. Real and random patterns associated with molluscan spatial and temporal distributions. Paleobiology 14:322330.CrossRefGoogle Scholar
Scarlto, O. A. 1981. Bivalve molluscs of the temperate latitudes of the western part of the Pacific Ocean. Opredeliteli po Faune SSSR 125. [In Russian]Google Scholar
Snedecor, G. W., and Cochran, W. G. 1980. Statistical Methods. Seventh Edition. Iowa State University Press; Ames, Iowa.Google Scholar
Sokal, R. R., and Rohlf, F. J. 1981. Biometry. Second Edition. Freeman; San Francisco.Google Scholar
Stanley, S. M. 1986a. Population size, extinction, and speciation: The fission effect in Neogene Bivalvia. Paleobiology 12:89110.Google Scholar
Stanley, S. M. 1986b. Anatomy of a regional mass extinction: Plio-Pleistocene decimation of the Western Atlantic bivalve fauna. Palaios 1:1736.Google Scholar
Stanley, S. M. 1988. Paleozoic mass extinctions: Shared patterns suggest global cooling as a common cause. American Journal of Science 288:334352.CrossRefGoogle Scholar
Strauch, F. 1972. Phylogenese, Adaptation und Migration einiger nordischer mariner Molluskengenera (Neptunea, Panomya, Cyrtodaria und Mya). Abhandlungen der Senckenbergischen Naturforschenden Gesellschaft 531.Google Scholar
Sverdrup, H. U., Johnson, M. W., and Fleming, R. H. 1942. The Oceans. Prentice-Hall; Englewood Cliffs, New Jersey.Google Scholar
Turner, R. D. 1966. A survey and illustrated catalogue of the Teredinidae. Museum of Comparative Zoology, Harvard University; Cambridge, Massachusetts.Google Scholar
Valentine, J. W. 1966. Numerical analysis of marine molluscan ranges on the extratropical northeastern Pacific shelf. Limnology and Oceanography 11:198211.Google Scholar
Valentine, J. W. 1984. Neogene marine climate trends: Implications for biogeography and evolution of the shallow-sea biota. Geology 12:647650.Google Scholar
Valentine, J. W., and Jablonski, D. 1983. Speciation in the shallow seas: General patterns and biogeographic controls. Pp. 201226. In Sims, R. W., Price, J. H., and Whalley, P. E. S. (eds.), Evolution, Time and Space: The Emergence of the Biosphere. Systematics Association Special Volume 23. Academic Press; London.Google Scholar
Valentine, P. C. 1976. Zoogeography of Holocene Ostracoda off western North America and paleoclimatic implications. U.S. Geological Survey Professional Paper 916.Google Scholar
Vermeij, G. J. 1978. Biogeography and Adaptation. Harvard University Press, Cambridge, Mass.Google Scholar
Vermeij, G. J. 1989. Invasion and extinction: The last three million years of North Sea pelecypod history. Conservation Biology 3:274281.Google Scholar
Zenkevitch, L. 1963. Biology of the Seas of the U.S.S.R. Interscience; New York.Google Scholar