Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-24T12:18:27.784Z Has data issue: false hasContentIssue false

Evolution and speciation in the Eocene planktonic foraminifer Turborotalia

Published online by Cambridge University Press:  08 April 2016

Paul N. Pearson
Affiliation:
School of Earth and Ocean Sciences, Cardiff University, Main Building, Park Place, Cardiff CF10 3YE, U.K. E-mail: PearsonP@cardiff.ac.uk
Thomas H. G. Ezard
Affiliation:
Centre for Biological Sciences, University of Southampton, Highfield Campus, Southampton SO17 1BJ, U.K.

Abstract

Marine planktonic microfossils have provided some of the best examples of evolutionary rates and patterns on multi-million-year time scales, including many instances of gradual evolution. Lineage splitting as a result of speciation has also been claimed, but all such studies have used subjective visual species discrimination, and interpretation has often been complicated by relatively small sample sizes and oceanographic complexity at the study sites. Here we analyze measurements on a collection of 10,200 individual tests of the Eocene planktonic foraminifer Turborotalia in 51 stratigraphically ordered samples from a site within the oceanographically stable tropical North Pacific gyre. We use novel multivariate statistical clustering methods to test the hypothesis that a single evolutionary species was present from 45 Ma to its extinction ca. 34 Ma. After identification of a set of biologically relevant traits, the protocol we apply does not require a prior assignment of individuals to species. We find that for most of the record, contemporaneous specimens form one morphological cluster, which we interpret as an evolving species that shows quasi-continuous but non-directional gradual evolutionary change (anagenesis). However, in the upper Eocene from ca. 36 to ca. 34 Ma there are two clusters that persistently occupy distinct areas of morphospace, from which we infer that speciation (cladogenesis) must have occurred.

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

Alizon, S., Kucera, M., and Jansen, V. A. A. 2008. Competition between cryptic species explains variations in rates of lineage evolution. Proceedings of the National Academy of Sciences USA 105:12,38212,386.CrossRefGoogle ScholarPubMed
André, A., Weiner, A., Quillévéré, F., Aurahs, R., Morard, R., Douady, C. J., de Garidel-Thoron, T., Escarguel, G., de Vargas, C., and Kucera, M. 2013. The cryptic and the apparent reversed: lack of genetic differentiation within the morphologically diverse plexus of the planktonic foraminifer Globigerinoides sacculifer. Paleobiology 39:2139.Google Scholar
Arnold, A. J. 1983. Phyletic evolution in the Globorotalia crassaformis (Galloway and Wissler) lineage: a preliminary report. Paleobiology 9:390398.Google Scholar
Arnold, S. J., Pfrender, M. E., and Jones, A. G. 2001. The adaptive landscape as a conceptual bridge between micro- and macroevolution. Genetica 112–113:932.CrossRefGoogle Scholar
Aze, T., Ezard, T. H. G., Purvis, A., Coxall, H. K., Stewart, D. R. M., Wade, B. S., and Pearson, P. N. 2011. A phylogeny of macroperforate planktonic foraminifera from fossil data. Biological Review of the Cambridge Philosophical Society 86:900927.Google Scholar
Aze, T., Ezard, T. H. G., Purvis, A., Coxall, H. K., Stewart, D. R. M., Wade, B. S., and Pearson, P. N. 2013. Identifying anagenesis and cladogenesis in the fossil record. Proceedings of the National Academy of Sciences USA 110. Available online: doi: 10.1073/pnas.1307562110.CrossRefGoogle ScholarPubMed
Backman, J., and Hermelin, J. O. R. 1986. Morphometry of the Eocene nannofossil Reticulofenestra umbilicus lineage and its biochronological consequences. Palaeogeography, Palaeoclimatology, Palaeoecology 57:103116.Google Scholar
Biolzi, M. 1991. Morphometric analysis of the Late Neogene planktonic foraminiferal lineage Neogloboquadrina dutertrei. Marine Micropaleontology 18:129142.Google Scholar
Blow, W. H. 1979. The Cainozoic Globigerinida: a study of the morphology, taxonomy, evolutionary relationships and the stratigraphical distribution of some of the Globigerinida (Mainly Globigerinaceae). E. J. Brill, Leiden.Google Scholar
Blow, W. H., and Banner, F. T. 1962. The mid-Tertiary (upper Eocene to Aquitanian) Globigerinaceae. Pp. 61151inEames, F. E., Banner, F. T., Blow, W. H., and Clarke, W. J., eds. Fundamentals of mid-Tertiary stratigraphic correlation. Cambridge University Press, Cambridge.Google Scholar
Boersma, A., Premoli Silva, I., and Shackleton, N. J. 1987. Atlantic Eocene planktonic foraminiferal paleohydrographic indicators and stable isotope paleoceanography. Paleoceanography 2:287331.CrossRefGoogle Scholar
Bolli, H. M. 1957. Planktonic foraminifera from the Eocene Navet and San Fernando formations of Trinidad, B.W.I. In Loeblich, A. R. Jr. and collaborators, eds. Studies in Foraminifera. U.S. National Museum Bulletin 215:155172.Google Scholar
Bralower, T. J., and Mutterlose, J. 1995. Calcareous nannofossil biostratigraphy of Site 865, Allison Guyot, Central Pacific Ocean: a tropical Paleogene reference section. InWinterer, E. L., Sager, W. W., Firth, J. V., and Sinton, J. M., eds. Proceedings of the Ocean Drilling Program, Scientific Results 143:3174. Ocean Drilling Program, College Station, Tex.Google Scholar
Bralower, T. J., Zachos, J. C., Thomas, E., Parrow, M., Paull, C. K., Kelly, D. C., Premoli Silva, I., Sliter, V. W., and Lohmann, K. C., 1995. Late Paleocene to Eocene paleoceanography of the equatorial Pacific Ocean: stable isotopes recorded at Ocean Drilling Program Site 865, Allison Guyot. Paleoceanography 10:841865.Google Scholar
Burnham, K. P., and Anderson, D. R., 2002. Model selection and multimodal inference: a practical information-theoretical approach. Springer, New York.Google Scholar
Cole, W. S. 1928. A foraminiferal fauna from the Chapapote Formation in Mexico. Bulletin of American Paleontology 14:132.Google Scholar
Coxall, H. K., Pearson, P. N., Shackleton, N. J., and Hall, M. A. 2000. Hantkeninid depth adaptation: an evolving life strategy in a changing ocean. Geology 28:8790.Google Scholar
Croux, C., and Ruiz-Gazen, A. 2005. High breakdown estimators for principal components: the projection-pursuit approach revisited. Journal of Multivariate Analysis 95:206226.Google Scholar
Cushman, J. A. 1928. Additional foraminifera from the Upper Eocene of Alabama. Contributions from the Cushman Laboratory for Foraminiferal Research 4:7379.Google Scholar
Darling, K. F., and Wade, C. M. 2008. The genetic diversity of planktonic foraminifera and the global distribution of ribosomal genotypes. Marine Micropaleontology 67:216238.CrossRefGoogle Scholar
Darwin, C. 1859. On the origin of species by means of natural selection, or the preservation of favoured races in the struggle for life. J. Murray, London.Google Scholar
de Vargas, C., Norris, R. D., Zaninetti, L., Gibb, S. W., and Pawlowski, J. 1999. Molecular evidence of cryptic speciation in planktonic foraminifers and their relation to oceanic provinces. Proceedings of the National Academy of Sciences USA, 96:28642868.CrossRefGoogle ScholarPubMed
Dinno, A. 2009. Implementing Horn's parallel analysis for principal components analysis and factor analysis. Stata Journal 9:291298.Google Scholar
Ezard, T. H. G., Pearson, P. N., and Purvis, A. A., 2010. Algorithmic approaches to aid species' delimitation in multidimensional morphospace. BMC Evolutionary Biology 10:175.Google Scholar
Ezard, T. H. G., Pearson, P. N., Aze, T., and Purvis, A. 2012. The meaning of birth and death (in macroevolutionary birth-death models). Biology Letters 8:139142.Google Scholar
Fisher, R. A. 1930. The genetical theory of natural selection. Clarendon, Oxford.Google Scholar
Filzmoser, P., Garrett, R. G., and Reimann, C. 2005. Multivariate outlier detection in exploration geochemistry. Computers and Geosciences 31:579587.Google Scholar
Filzmoser, P., Maronna, R., and Werner, M. 2008. Outlier identification in high dimensions. Computational Statistics and Data Analysis 52:16941711.CrossRefGoogle Scholar
Fraley, C., and Raftery, A. E. 2002. Model-based clustering, discriminant analysis, and density estimation. Journal of the American Statistical Association 97:611631.Google Scholar
Gee, H. 2000. Deep time: cladistics, the revolution in evolution. Fourth Estate, London.Google Scholar
Gingerich, P. D. 1990. Stratophenetics. Pp. 437442inBriggs, D. E. G. and Crowther, P. R., eds. Palaeobiology: a synthesis. Blackwell Science, Oxford.Google Scholar
Gould, S. J. 1978. Sociobiology: the art of storytelling. New Scientist, 16 November 1978:530533.Google Scholar
Gould, S. J. 2002. The structure of evolutionary theory. Harvard University Press, Cambridge.Google Scholar
Gould, S. J., and Lewontin, R. C. 1979. The spandrels of San Marco and the Panglossian paradigm: a critique of the adaptationist programme. Proceedings of the Royal Society of London B 205:581598Google Scholar
Hays, J. D. 1970. Stratigraphy and evolutionary trends of Radiolaria in North Pacific deep-sea sediments. InHays, J. D., ed. Geological investigations of the North Pacific. Geological Society of America Memoir 126:185218.Google Scholar
Hemleben, C., Spindler, M., and Anderson, O. R. 1989. Modern planktonic foraminifera. Springer, New York.Google Scholar
Huber, B. T., Bijma, J., and Darling, K. 1997. Cryptic speciation in the living planktonic foraminifer Globigerinella siphonifera (d'Orbigny). Paleobiology 23:3362.Google Scholar
Hull, P. M., and Norris, R. D. 2009. Evidence for abrupt speciation in a classic case of gradual evolution. Proceedings of the National Academy of Sciences USA 106:21,22421,229.Google Scholar
Hunt, G. 2006. Fitting and comparing models of phyletic evolution: random walks and beyond. Paleobiology 32:578601.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
Kellogg, D. E. 1975. The role of phyletic change in the evolution of Pseudocubus vema (Radiolaria). Paleobiology 1:359370.Google Scholar
Kellogg, D. E. 1976. Character displacement in the radiolarian genus Eucyrtidium. Evolution 29:736749.Google Scholar
Knappertsbusch, M. 2000. Morphologic evolution of the coccolithophorid Calcidiscus leptoporus from the Early Miocene to Recent. Journal of Paleontology, 74:712730.Google Scholar
Knappertsbusch, M. 2007. Morphological variability of Globorotalia menardii (planktonic Foraminifera) in two DSDP cores from the Caribbean Sea and the Eastern Equatorial Pacific. Carnets de Géologie / Notebooks on Geology, Brest, Article 2007/04 (CG2007_AO4).Google Scholar
Kucera, M., and Malmgren, B. A. 1998. Differences between evolution of mean form and evolution of new morphotypes: an example from Late Cretaceous planktonic foraminifera. Paleobiology 24:4963.Google Scholar
Lazarus, D. 1986. Tempo and mode of morphologic evolution near the origin of the radiolarian lineage Pterocanium prismatium. Paleobiology 12:175189.Google Scholar
Lazarus, D. 2011. The deep-sea microfossil record of macroevolutionary change in plankton and its study. Geological of London Special Publication 358:141166.Google Scholar
Lazarus, D., Scherer, R., and Prothero, D. 1985. Evolution of the radiolarian species complex Pterocanium: a preliminary survey. Journal of Paleontology 59:183220.Google Scholar
Lazarus, D., Hilbrecht, H., Spencer-Cervato, C., and Thierstein, H. 1995. Sympatric speciation and phyletic change in Globorotalia truncatulinoides. Paleobiology 21:2851.Google Scholar
Li, G., and Chen, Z. 1985. Projection-pursuit approach to robust dispersion matrices and principal components: primary theory and Carlo, Monte. Journal of the American Statistical Association 80:759766.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
Longhurst, A. R. 2007. Ecological geography of the sea, 2nd ed. Academic Press, London.Google Scholar
Malmgren, B. A., and Kennett, J. P. 1981. Phyletic gradualism of a late Cenozoic planktonic foraminiferal lineage, D.S.D.P. Site 284, Southwest Pacific. Paleobiology 7:230240.Google Scholar
Malmgren, B. A., Berggren, W. A., and Lohmann, G. P. 1983. Evidence for punctuated gradualism in the late Neogene Globorotalia tumida lineage of planktonic foraminifera. Paleobiology 9:377389.Google Scholar
Malmgren, B. A., Kucera, M., and Ekman, G. 1996. Evolutionary changes in supplementary apertural characteristics of the Late Neogene Sphaeroidinella dehiscens lineage. Palaios 11:192206.Google Scholar
Norris, R. D. 2000. Pelagic species diversity, biogeography, and evolution. Paleobiology 26:236258.Google Scholar
Norris, R. D., Corfield, R. M., and Cartlidge, J. 1996. What is gradualism? Cryptic speciation in globorotaliid foraminifera. Paleobiology 22:386405.Google Scholar
Pearson, P. N. 1993. A lineage phylogeny for the Paleogene planktonic foraminifera. Micropaleontology 39:193232.Google Scholar
Pearson, P. N. 1998. Evolutionary concepts in biostratigraphy. Pp. 123144inDoyle, P. and Bennett, M. R., eds. Unlocking the stratigraphic record. Wiley, New York.Google Scholar
Pearson, P. N., Shackleton, N. J., and Hall, M. A. 1997. Stable isotopic evidence for the sympatric divergence of Globigerinoides trilobus and Orbulina universa (planktonic foraminifera). Journal of the Geological Society, London 154:295302.Google Scholar
Pearson, P. N., Ditchfield, P. W., Singano, J., Harcourt-Brown, K. G., Nicholas, C. J., Olsson, R. K., Shackleton, N. J., and Hall, M. A. 2001. Warm tropical sea surface temperatures in the Late Cretaceous and Eocene epochs. Nature 413:481487.Google Scholar
Pearson, P. N., Premec-Fucek, V., and Premoli Silva, I. 2006. Taxonomy, biostratigraphy, and phylogeny of Eocene Turborotalia. InPearson, P. N., Olsson, R. K., Huber, B. T., Hemleben, C., and Berggren, W. A., eds. Atlas of Eocene planktonic foraminifera. Cushman Foundation Special Publication 41:433460.Google Scholar
Peres-Neto, P. R., Jackson, D. A., and Somers, K. M. 2005. How many principal components? Stopping rules for determining the number of non-trivial axes revisited. Computational Statistics and Data Analysis 49:974997.Google Scholar
Poore, R. Z., and Matthews, R. K. 1984. Oxygen isotope ranking of Late Eocene and Oligocene planktonic foraminifers: implications for Oligocene sea-surface temperatures and global ice volume. Marine Micropaleontology 9:111134.Google Scholar
Premoli Silva, I., and Boersma, A. 1988. Atlantic Paleogene planktonic foraminiferal bioprovincial indices. Marine Micropaleontology 14:357371.Google Scholar
R Development Core Team. 2010. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna. http://www.R-project.org/.Google Scholar
Scott, G. H. 1982. Tempo and stratigraphic record of speciation in Globorotalia puncticulata. Journal of Foraminiferal Research 12:112.Google Scholar
Scott, G. H., Kennett, J. P., Wilson, K. J., and Hayward, B. W. 2007. Globorotalia puncticulata: population divergence, dispersal and extinction related to Pliocene-Quaternary water masses. Marine Micropaleontology 62:235253.Google Scholar
Shipboard Scientific Party. 1993. Site 865. InWinterer, E. L., Firth, J. V., et al., eds. Proceedings of the Ocean Drilling Program, Initial Results 143:111180. Ocean Drilling Program, College Station, Tex.Google Scholar
Simpson, G. G. 1944. Tempo and mode in evolution. Columbia University Press, New York.Google Scholar
Simpson, G. G. 1961. Principles of animal taxonomy. Columbia University Press, New York.Google Scholar
Sorhannus, U., Fenster, E. J., Burckle, L. H., and Hoffman, A. 1988. Cladogenetic and anagenetic changes in the morphology of Rhizosolenia praebergonii Mukhina. Historical Biology 1:185205.Google Scholar
Strotz, L. C., and Allen, A. P. 2013. Assessing the role of cladogenesis in macroevolution by integrating fossil and molecular evidence. Proceedings of the National Academy of Sciences USA 110:29042909.Google Scholar
Toumarkine, M. 1975. Middle and Late Eocene planktonic foraminifera from the northwestern Pacific Ocean, Leg 32 of the Deep Sea Drilling Project. InLarson, R. L., Moberly, R., et al., eds. Initial Reports of the Deep Sea Drilling Project 32:735751. U.S. Government. Printing Office, Washington, D.C.Google Scholar
Toumarkine, M. 1978. Planktonic foraminiferal biostratigraphy of the Paleogene of Sites 360 to 364 and the Neogene of Sites 362A, 363 and 364 Leg 40. InBolli, H. M., Ryan, W. B. F., et al., eds. Initial Reports of the Deep Sea Drilling Project 40:679721. U.S. Government. Printing Office, Washington, D.C.Google Scholar
Toumarkine, M., and Bolli, H. M. 1970. Evolution de Globorotalia cerroazulensis (Cole) dans l'Éocène Moyen et Supérieur de Possagno (Italie). Revue de Micropaléontologie 13:131145.Google Scholar
Toumarkine, M., and Luterbacher, H.-P. 1985. Paleocene and Eocene planktic foraminifera. Pp. 87154inBolli, H. M., Saunders, J. B., and Perch-Nielsen, K., eds. Plankton stratigraphy. Cambridge University Press, Cambridge.Google Scholar
Tripati, A. K., and Elderfield, H. 2004. Abrupt hydrographic changes in the equatorial Pacific and subtropical Atlantic from foraminiferal Mg/Ca indicate greenhouse origin for the thermal maximum at the Paleocene-Eocene boundary. Geochemistry, Geophysics, Geosystems 5: Q02006. doi:10.1029/2003GC000631.Google Scholar
Trueman, A. E. 1930. Results of some recent statistical investigations of invertebrate fossils. Biological Reviews 5:296308.Google Scholar
Wade, B. S., and Pearson, P. N. 2008. Planktonic foraminiferal turnover, diversity fluctuations and geochemical signals across the Eocene/Oligocene boundary in Tanzania. Marine Micropaleontology 68:244255.Google Scholar
Wade, B. S., Pearson, P. N., Berggren, W. A., and Pälike, H. 2011. Review and revision of Cenozoic tropical planktonic foraminiferal biostratigraphy and calibration to the geomagnetic polarity and astronomical time scale. Earth Science Reviews 104:111142.Google Scholar
Wei, K.-Y. 1994. Allometric heterochrony in the Pliocene-Pleistocene planktic foraminiferal clade Globoconella. Paleobiology 20:6684.Google Scholar
Wei, K.-Y., and Kennett, J. P. 1988. Phyletic gradualism and punctuated equilibrium in the late Neogene planktonic foraminiferal clade Globoconella. Paleobiology 14:345363.Google Scholar
Wright, S. 1931. Evolution in Mendelian populations. Genetics 16:97159.Google Scholar
Young, J. 1990. Size variation of Neogene Reticulofenestra coccoliths from Indian Ocean DSDP cores. Journal of Micropalaeontology 9:7185.Google Scholar
Zachos, J. C., Dickens, G. R., and Zeebe, R. 2008. An early Cenozoic perspective on greenhouse warming and carbon-cycle dynamics. Nature 451:279283.Google Scholar