Hostname: page-component-76fb5796d-25wd4 Total loading time: 0 Render date: 2024-04-26T16:02:32.088Z Has data issue: false hasContentIssue false

Differences between evolution of mean form and evolution of new morphotypes: an example from Late Cretaceous planktonic foraminifera

Published online by Cambridge University Press:  08 February 2016

Michal Kucera
Affiliation:
Department of Marine Geology, Earth Sciences Centre, Göteborg University, S-413 81 Göteborg, Sweden. E-mail: michal@gvc.gu.se
Björn A. Malmgren
Affiliation:
Department of Marine Geology, Earth Sciences Centre, Göteborg University, S-413 81 Göteborg, Sweden. E-mail: michal@gvc.gu.se

Abstract

Morphological evolution in the Late Cretaceous (Maastrichtian) Contusotruncana lineage of planktonic foraminifera was studied at DSDP Sites 525 (South Atlantic) and 384 (North Atlantic). A multivariable approach was used to separate aspects of form controlled by geographical variation (size, spiral roundness of the test, percentage of kummerform specimens) from those due to changes that occurred simultaneously in geographically distant populations of the lineage (shell conicity, number of chambers in the last whorl).

A gradual increase in mean shell conicity was observed over the last 3 million years of the Cretaceous. It arose from the combination of a rapid development of highly conical shells after 68.5 Ma and a long-term trend of progressive disappearance of the ancestral morphotype. Therefore, despite the gradual change in “mean form,” the morphological evolution in the Contusotruncana lineage differs from the classical image of phyletic gradualism. The gradual increase in mean shell conicity in the lineage was accompanied by a remarkable decrease in its absolute abundance (shell accumulation rate), suggesting that the changes in shell morphology might not have been neutral with respect to natural selection. Apparently, gradual change in “mean form” of fossil lineages does not require an equally gradual development of morphological novelties. It may be caused by natural selection operating on a constant range of variation in populations living in environments without geographical barriers.

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

Arnold, A. J. 1983. Phyletic evolution in the Globorotalia crassaformis (Galloway and Wissler) lineage: a preliminary report. Paleobiology 9:390398.CrossRefGoogle Scholar
Berger, W. H. 1969. Planktonic foraminifera: basic morphology and ecologic implications. Journal of Paleontology 43:13691383.Google Scholar
Berggren, W. A., Kent, D. V., Swisher, C. C. III, and Aubry, M. P. 1995. A revised Cenozoic geochronology and chronostratigraphy. Society of Economic Paleontologists and Mineralogists Special Publication 54:129212.Google Scholar
Boersma, A., and Shackleton, N. J. 1981. Oxygen- and carbon-isotope variations and planktonic-foraminifer depth habitats, Late Cretaceous to Paleocene, central Pacific, Deep Sea Drilling Project Sites 463 and 465. Initial Reports of the Deep Sea Drilling Project 62:513526Google Scholar
Bookstein, F. L. 1987. Random walk and the existence of evolutionary rates. Paleobiology 13:446464.CrossRefGoogle Scholar
Bookstein, F. L. 1991. Morphometric tools for landmark data: geometry and biology. Cambridge University Press, New York.Google Scholar
Caron, M. 1985. Cretaceous planktic foraminifera. Pp. 1786in Bolli, H. M., Saunders, J. B., and Perch-Nielsen, K., eds. Plankton stratigraphy. Cambridge University Press, Cambridge.Google Scholar
Charlesworth, B. 1984. The cost of phenotypic evolution. Paleobiology 10:319327.CrossRefGoogle Scholar
Chave, A. D. 1984. Lower Paleocene–Upper Cretaceous magnetostratigraphy, Sites 525, 527, 528, and 529, Deep Sea Drilling Project. Initial Reports of the Deep Sea Drilling Project 74:525531.Google Scholar
D'Hondt, S., and Arthur, M. A. 1995. Interspecies variation in stable isotopic signals of Maastrichtian planktonic foraminifera. Paleoceanography 10:123135.CrossRefGoogle Scholar
D'Hondt, S., and Lindinger, M. 1994. A stable isotope record of the Maastrichtian ocean-climate system: South Atlantic DSDP site 528. Palaeogeography, Palaeoclimatology, Palaeoecology 112:363378.CrossRefGoogle Scholar
Eldredge, N., and Gould, S. J. 1972. Punctuated equilibria: an alternative to phyletic gradualism. Pp. 82115in Schopf, T. J. M., ed. Models in paleobiology. W. H. Freeman, San Francisco.Google Scholar
Eldredge, N., and Gould, S. J. 1988. Punctuated equilibrium prevails. Nature 332:211212.CrossRefGoogle Scholar
El-Naggar, Z. R. 1966. Stratigraphy and planktonic foraminifera of the Upper Cretaceous–Lower Tertiary succession in the Esna-Idfu region, Nile Valley, Egypt, U.A.R. Bulletin of the British Museum of Natural History, Supplements in Geology 2:1291.CrossRefGoogle Scholar
El-Naggar, Z. R., and Haynes, J. 1967. Globotruncana caliciformis in the Maestrichtian Sharawna Shale of Egypt. Cushman Foundation for Foraminiferal Research, Contribution 18:113.Google Scholar
Fortey, R. A. 1988. Seeing is believing: gradualism and punctuated equilibria in the fossil record. Science Progress, Oxford 72:119.Google ScholarPubMed
Gandolfi, R. 1955. The genus Globotruncana in northeastern Colombia. Bulletin of American Paleontology 36:1118.Google Scholar
Gould, S. J. 1990. Speciation and sorting as the source of evolutionary trends, or “Things are seldom what they seem.” Pp. 327in McNamara, K. J., ed. Evolutionary trends. Belhaven, London.Google Scholar
Gould, S. J., and Eldredge, N. 1993. Punctuated equilibrium comes of age. Nature 366:223227.CrossRefGoogle ScholarPubMed
Hemleben, Ch., Spindler, M., and Anderson, R. O. 1989. Modern planktonic foraminifera. Springer, New York.CrossRefGoogle Scholar
Herm, D. 1962. Stratigraphische und mikropaläontologische Untersuchungen der Oberkreide im attengebirge und Nierental. Abhandlungen Bayerisches Akademie der Wissenschaften, Mathematisch-physikalische Klasse, Neue Folge 104:1119.Google Scholar
Huber, B. T., Hodell, D. A., and Hamilton, C. P. 1995. Middle-Late Cretaceous climate of the southern high latitudes: stable isotopic evidence for minimal equator-to-pole thermal gradients. Geological Society of America Bulletin 107:11641191.2.3.CO;2>CrossRefGoogle Scholar
Hunter, R. S. T., Arnold, A. J., and Parker, W. C. 1988. Evolution and homeomorphy in the development of the Paleocene Planorotalites pseudomenardii and the Miocene Globorotalia (Globorotalia) margaritae lineages. Micropaleontology 34:181192.CrossRefGoogle Scholar
Johnson, J. G. 1982. Occurrence of phyletic gradualism and punctuated equilibria through geologic time. Journal of Paleontology 56:13291331.Google Scholar
Kellogg, D. E. 1975. The role of phyletic change in the evolution of Pseudocubus vema (Radiolaria). Paleobiology 1:359370.CrossRefGoogle Scholar
Kellogg, D. E. 1983. Phenology of morphologic change in radiolarian lineages from deep-sea cores: implications for macroevolution. Paleobiology 9:355362.CrossRefGoogle Scholar
Kucera, M., and Malmgren, B. A. 1996. Latitudinal variation in the planktic foraminifer Contusotruncana contusa in the terminal Cretaceous ocean. Marine Micropaleontology 28:3152.CrossRefGoogle Scholar
Lande, R. 1976. Natural selection and random genetic drift in phenotypic evolution. Evolution 30:314334.CrossRefGoogle ScholarPubMed
Larson, P. A., and Opdyke, N. D. 1979. Paleomagnetic results from Early Tertiary/Late Cretaceous sediments of Site 384. Initial Reports of the Deep Sea Drilling Project 43:785788.Google Scholar
Lazarus, D. 1986. Tempo and mode of morphologic evolution near the origin of the rediolarian lineage Pterocanium prismatium. Paleobiology 12:175189.CrossRefGoogle Scholar
Lazarus, D., Hilbrecht, H., Spencer-Cervato, C., and Thierstein, H. 1995. Sympatric speciation and phyletic change in Globorotalia truncatulinoides. Paleobiology 21:2851.CrossRefGoogle Scholar
Lynch, J. D. 1989. The gauge of speciation: on frequencies of models of speciation. Pp. 527553in Otte, D. and Endler, J. A., eds. Speciation and its consequences. Sinauer, Sunderland, Mass.Google Scholar
MacLeod, K. G., and Huber, B. T. 1996. Reorganization of deep ocean circulation accompanying a Late Cretaceous extinction event. Nature 380:422425.CrossRefGoogle Scholar
Malmgren, B. A. 1987. Differential dissolution of Upper Cretaceous planktonic foraminifera from a temperate region of the South Atlantic Ocean. Marine Micropaleontology 11:51271.CrossRefGoogle Scholar
Malmgren, B. A. 1989. Coiling patterns in terminal Cretaceous planktonic foraminifera. Journal of Foraminiferal Research 19:311323.CrossRefGoogle Scholar
Malmgren, B. A. 1991. Biogeographic patterns in terminal Cretaceous planktonic foraminifera from Tethyan and warm Transitional waters. Marine Micropaleontology 18:7399.CrossRefGoogle Scholar
Malmgren, B. A., and Kennett, J. P. 1981. Phyletic gradualism in a Late Cenozoic planktonic foraminiferal lineage; DSDP Site 284, southwest Pacific. Paleobiology 7:230240.CrossRefGoogle Scholar
Manivit, H. 1984. Paleogene and upper Cretaceous nannofossils. Initial Reports of the Deep Sea Drilling Project 74:475499.Google Scholar
Masters, B. A. 1977. Mesozoic planktonic foraminifera. Pp. 301731in Ramsay, A. T. S., ed. Oceanic micropaleontology, Vol. 1. Academic Press, London.Google Scholar
Moore, T. C. Jr. et al. 1984. Site 525. Initial Reports of the Deep Sea Drilling Project 74:41160.Google Scholar
Mulder, E., and Marks, P. 1983. Umbilical structures of Globotruncana fornicata (Cushman) in exceptionally well-preserved material from Blake Escarpment (D.S.D.P. Leg 44, Site 390A). Cretaceous Research 4:211214.CrossRefGoogle Scholar
Norris, R. D., Corfield, R. M., and Cartlidge, J. 1996. What is gradualism? Cryptic speciation in globorotalid foraminifera. Paleobiology 22:386405.CrossRefGoogle Scholar
Pessagno, E. A. Jr. 1967. Upper Cretaceous planktonic foraminifera from the western Gulf Coastal Plain. Paleontographica Americana 5:245445.Google Scholar
Scott, G. H. 1982. Tempo and stratigraphic record of speciation in Globorotalia puncticulata. Journal of Foraminiferal Research 12:112.CrossRefGoogle Scholar
Sheldon, P. R. 1990. Shaking up evolutionary patterns. Nature 345:772.CrossRefGoogle Scholar
Sheldon, P. R. 1996. Plus ça change—a model for stasis and evolution in different environments. Palaeogeography, Palaeoclimatology, Palaeoecology 127:209227.CrossRefGoogle Scholar
Signes, M., Bijma, J., Hemleben, Ch., and Ott, R. 1993. A model for planktic foraminiferal shell growth. Paleobiology 19:7191.CrossRefGoogle Scholar
Sugarman, P. J., Miller, K. G., Bukry, D., and Feigenson, M. D. 1995. Uppermost Campanian–Maestrichtian strontium isotopic, biostratigraphic, and sequence stratigraphic framework of the New Jersey Coastal Plain. Geological Society of America Bulletin 107:1937.2.3.CO;2>CrossRefGoogle Scholar
Thierstein, H. R. 1981. Late Cretaceous nannoplankton and the change at the Cretaceous-Tertiary boundary. The Deep Sea Drilling Project: a decade of progress. Society of Economic Paleontologists and Mineralogists Special Publication 32:355394.CrossRefGoogle Scholar
Thierstein, H. R., and Okada, H. 1979. The Cretaceous/Tertiary boundary event in the North Atlantic. Initial Reports of the Deep Sea Drilling Project 43:601616.Google Scholar
Todd, R. 1970. Maestrichtian (Late Cretaceous) foraminifera from a deep-sea core off southwestern Africa. Revista Española de Micropaleontología 2:131154.Google Scholar
Tucholke, B. E. et al. 1979. Site 384: The Cretaceous/Tertiary boundary, Aptian reefs, and the J- Anomaly Ridge. Initial Reports of the Deep Sea Drilling Project 43:107154.Google Scholar