Hostname: page-component-76fb5796d-skm99 Total loading time: 0 Render date: 2024-04-28T08:55:14.765Z Has data issue: false hasContentIssue false

Allometric heterochrony in the Pliocene-Pleistocene planktic foraminiferal clade Globoconella

Published online by Cambridge University Press:  08 February 2016

Kuo-Yen Wei*
Affiliation:
Department of Geology and Geophysics, Yale University, New Haven, Connecticut 06511-8130

Abstract

Allometric analysis of the size-shape relationships in the Pliocene-Pleistocene planktic foraminiferal Globorotalia (Globoconella) puncticulata-inflata plexus reveals several heterochronic modes underlying the morphological evolution of the clade. The ancestral lineage, G. puncticulata, is a peramorphocline, showing a pre-displacement mode of heterochrony between 3.5 Ma and 3.0 Ma and an acceleration mode from 3.0 to 2.7 Ma. A different peramorphosis process, isometric giantism (hypermorphosis), in the ontogeny of the ancestral stocks of Globoconella occurred at about 3.5 Ma and gave rise to the G. inflata lineage. The descendant lineage, G. inflata, appears to have adopted a paedomorphosis trend by delaying the onset of the neanic stage in ontogeny during the period of 3.5 to 2.35 Ma, resulting in a series of transposition allometries. During the interval of 2.4 to 1.73 Ma, the allometries shifted to the opposite direction, signifying a pre-displacement trend. Evolutionary stasis marks the evolution during 1.73 to 0.25 Ma. Neoteny concluded the final evolutionary stage of the G. inflata lineage during the latest Quaternary (0.26 to 0.05 Ma). The enormous plasticity and fluctuations in morphology of G. inflata are attributed to the highly positive allometric growth during the ontogeny and the wide-range transposing allometries in the phyletic history. The major changes in heterochronic mode coincide with paleoceanographic events, suggesting that the morphological evolution in the Globoconella clade has been modulated by changes in paleoceanographic conditions.

Type
Research Article
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

Alberch, P., Gould, S. J., Oster, G. F., and Wake, D. B. 1979. Size and shape in ontogeny and phylogeny. Paleobiology 5:296317.CrossRefGoogle Scholar
Albercht, G. H. 1987. The simple allometry equation reconsidered: assumptions, problems and alternative solutions. American Journal Physical Anthropology 74:174.Google Scholar
Barton, C. E., and Bloemendal, J. 1986. Paleomagnetism of sediments collected during Leg 90, southwest Pacific. Pp. 12731316in Kennett, and von der Borch, et al. 1986.Google Scholar
Berger, W. H., and Piper, D. J. W. 1972. Planktonic foraminifera: differential settling, dissolution and redeposition. Limnology and Oceanography 17:275287.CrossRefGoogle Scholar
Berggren, W. A., Kent, D. V., and Van Couvering, J. A. 1985. Neogene geochronology and chronostratigraphy. Pp. 211260in Snelling, N. J., ed. The chronology of the geological record. Geological Society of London.Google Scholar
Blackstone, N. W. 1987. Size and time. Systematic Zoology 36:7678.CrossRefGoogle Scholar
Brummer, G. J. A., Hemleben, C., and Spindler, M. 1986. Planktonic foraminiferal ontogeny and new perspectives for micropaleontology. Nature (London) 319:5052.CrossRefGoogle Scholar
Brummer, G. J. A., Hemleben, C., and Spindler, M. 1987. Ontogeny of extant globigerinid planktonic foraminifera; a concept exemplified by Globigerinoides sacculifer (Brady) and G. ruber (d'Orbigny). Marine Micropaleontology 12:357381.CrossRefGoogle Scholar
Druazzi, J. T. 1981. Stable-isotope studies of planktonic foraminifera in North Atlantic core tope. Palaeogeography, Palaeoclimatology, Palaeoecology 33:157172.CrossRefGoogle Scholar
Dudley, W. C., and Nelson, C. S. 1989 Quaternary surface-water stable isotope signal from calcareous nannofossils at DSDP Site 593, southern Tasman Sea. Marine Micropaleontology 13:353373.CrossRefGoogle Scholar
Edwards, A. R. 1987. An integrated biostratigraphy, magnetostratigraphy and oxygen isotope stratigraphy for the late Neogene of New Zealand. New Zealand Geological Survey Record 23.Google Scholar
Elmstrom, K. M., and Kennett, J. P. 1986. Late Neogene paleoceanographic evolution of Site 590: southwest Pacific. Pp. 13611381in Kennett, and von der Borch, et al. 1986.Google Scholar
Emiliani, C. 1954. Depth habitats of some species of pelagic foraminifera as indicated by oxygen isotope ratios. American Journal of Science 252:149158.CrossRefGoogle Scholar
Fairbanks, R. G., and Wiebe, P. H. 1980. Foraminifera and chlorophyll maximum: vertical distribution, seasonal succession, and paleoceanographic significance. Science 209:15241526.CrossRefGoogle ScholarPubMed
Fok-Pun, L., and Komar, P. D. 1983. Settling velocities of planktonic foraminifera: density variations and shape effects. Journal of Foraminiferal Research 13:6068.CrossRefGoogle Scholar
Gould, S. J. 1966. Allometry and size in ontogeny and phylogeny. Biological Review 41:587640.CrossRefGoogle ScholarPubMed
Gould, S. J. 1972. Allometric fallacies and the evolution of Gryphaea: a new interpretation based on White's criterion of geometric similarity. Evolutionary Biology 6:91119.Google Scholar
Gould, S. J. 1977. Ontogeny and phylogeny. Belknap Press, Cambridge, Mass.Google Scholar
Healy-Williams, N., Ehrlich, R., and Williams, D. F. 1985. Morphometric and stable isotopic evidence for subpopulations of Globorotalia truncatulinoides. Journal of Foraminiferal Research 15:242253.CrossRefGoogle Scholar
Hemleben, C., Spindler, M., and Anderson, O. R. 1988. Modern planktonic foraminifera. Springer, New York.Google Scholar
Hodell, D. A., and Kennett, J. P. 1986. Late Miocene-early Pliocene stratigraphy and paleoceanography of the south Atlantic and southwest Pacific oceans: a synthesis. Paleoceanography 1:285311.CrossRefGoogle Scholar
Hornibrook, N. de B. 1980. Correlation of Pliocene biostratigraphy, magnetostratigraphy and δ18O fluctuations in New Zealand and DSDP Site 284. Newsletter of Stratigraphy 9:114120.CrossRefGoogle Scholar
Hornibrook, N. de B. 1982. Late Miocene to Pleistocene Globorotalia (Foraminiferida) from Deep Sea Drilling Project Leg 29, Site 284, Southwest Pacific. New Zealand Journal of Geology and Geophysics 25:8399.CrossRefGoogle Scholar
Hoskins, R. H. 1990. Planktic Foraminiferal Correlation of the Late Pliocene to Early Pleistocene of DSDP Sites 284 and 593 (Challenger Plateau) with New Zealand Stages. New Zealand Geological Survey Report PAL 149.Google Scholar
Huang, C. Y. 1981. Observations on the interior of some late Neogene planktonic foraminifera. Journal of Foraminiferal Research 11:173190.CrossRefGoogle Scholar
Huber, B.In press. Ontogenetic morphometrics of some Late Cretaceous trochospiral planktonic foraminifera from the austral realm. Smithsonian Contributions to Paleobiology.Google Scholar
Humpheries, J. M., Bookstein, F. L., Chernoff, B., Smith, G. R., Elder, R. L., and Poss, S. G. 1981. Multivariate discrimination by shape in relation to size. Systematic Zoology 30:291308.CrossRefGoogle Scholar
Huxley, J. 1932. Problems of relative growth. Cambridge University Press, Cambridge, Mass.Google Scholar
Jansen, E., and Sejrup, H. P. 1986. Stable isotope stratigraphy and amino-acid epimerization for the last 2.4 M.Y. at Site 610, Holes 610 and 610A. Pp. 879888in Ruddiman, W. F., Kidd, R. B., and Thomas, E., eds. Initial Reports of the Deep Sea Drilling Project, Vol. 94. U.S. Government Printing Office, Washington, D.C.Google Scholar
Johnson, R. G., and Andrews, J. T. 1986. Glacial terminations in the oxygen isotope record of deep sea cores: hypothesis of massive antarctic ice-shelf destruction. Palaeogeography, Palaeoclimatology, Palaeoecology 53:107138.CrossRefGoogle Scholar
Jolicouer, P. 1963. The multivariate generalization of the allometry equation. Biometrics 19:497499.CrossRefGoogle Scholar
Kennett, J. P., and von der Borch, C. C. 1986. Southwest Pacific Cenozoic paleoceanography. Pp. 14931517in Kennett, and von der Borch, et al. 1986.Google Scholar
Kennett, J. P., and von der Borch, C. C. et al., eds. 1986. Initial Reports of the Deep Sea Drilling Project, Vol. 90. U.S. Government Printing Office, Washington, D.C.Google Scholar
Kuhry, B., and Marcus, L. F. 1977. Bivariate linear models in biometry. Systematic Zoology 26:201209.CrossRefGoogle Scholar
Lipps, J. H. 1979. Ecology and paleoecology of planktic Foraminifera. Pp. 62104in Lipps, J. H. and Berger, W. H., eds. Foraminiferal ecology and paleoecology. Society of Economic Paleontologists and Mineralogists, Houston, Texas.CrossRefGoogle Scholar
Lohmann, G. P., and Schweitzer, P. N. 1990. Globorotalia truncatulinoides' growth and chemistry as probes of the past thermocline. I. Shell size. Paleoceanography 5:5575.CrossRefGoogle Scholar
Lohmann, G. P. 1992. Accelerated development in planktonic foraminifera: adaptive response to reduced ocean mixing. Fifth North American Paleontological Convention—Abstracts and Program, p. 188.Google Scholar
Loubere, P., and Jackiel, R. 1984. A sedimentological, faunal, and isotopic record of the middle-to-late Pliocene transition in the northeastern Atlantic, Deep Sea Drilling Project Site 548. Pp. 473488in Graciansky, P. C. and Poag, C. W. et al., eds. Initial Reports of the Deep Sea Drilling Project, Vol. 80. U.S. Government Printing Office, Washington, D.C.Google Scholar
Malmgren, B. A., and Kennett, J. P. 1981. Phyletic gradualism in Late Cenozoic planktonic foraminiferal lineage: DSDP Site 284, southwest Pacific. Paleobiology 7:230240.CrossRefGoogle Scholar
Malmgren, B. A., and Kennett, J. P. 1982. The potential of morphometrically based phylozonation: application of a late Cenozoic foraminiferal lineage. Marine Micropaleontology 7:285296.CrossRefGoogle Scholar
McKinney, M. L. 1986. Ecological causation of heterochrony: a test and implications for evolutionary theory. Paleobiology 12:282289.CrossRefGoogle Scholar
McKinney, M. L., and McNamara, K. J. 1991. Heterochrony: the evolution of ontogeny. Plenum, New York.CrossRefGoogle Scholar
McNamara, K. J. 1986. A guide to the nomenclature of heterochrony. Journal of Paleontology 60:413.CrossRefGoogle Scholar
McNown, J. S., and Malaika, J. 1950. Effects of particle shape on settling velocity at low Reynolds numbers. Transactions, American Geophysical Union 31:7482.Google Scholar
Meunier, , 1959a. Die Allomtriedes Vogelflugels. Zeitschrift fuer Wissenschaftiche Zoologie 161:444482.Google Scholar
Meunier, . 1959b. Die Grossenabhangigkeit der Korperform bei Vogeln. Zeitschrift fuer Wissenschaftiche Zoology 162:328355.Google Scholar
Nelson, C. S., Hendy, C. H., and Dudley, W. C. 1986. Quaternary isotope stratigraphy of Hole 593, Challenger Plateau, South Tasman Sea: preliminary observations based on foraminiferas and calcareous nannofossils. Pp. 14711491in Kennett, and von der Borch, et al. 1986.Google Scholar
Oba, T. 1990. Paleoceanographic information obtained by the isotopic measurement of individual foraminiferal specimens. Pp. 169180in Wang, P., Lao, Q., and He, Q., eds. Proceedings of the First International Conference on Asian Marine Geology, China Ocean Press, Beijing.Google Scholar
Raymo, M. E., Ruddiman, W. F., Backman, J., Clement, B. M., and Martinson, D. G. 1989. Late Pliocene variation in North Hemisphere ice sheets and North Atlantic Deep Water circulation. Paleoceanography 4:413446.CrossRefGoogle Scholar
Rohlf, F. J., and Bookstein, F. L. 1987. A comment on shearing as a method for “size correction.” Systematic Zoology 36:356367.CrossRefGoogle Scholar
Rohlf, F. J., and Bookstein, F. L. 1990. Proceedings of the Michigan Morphometrics Workshop. Special Publication No. 2. The University of Michigan Museum of Zoology, Ann Arbor.Google Scholar
Ruddiman, W. F., and Raymo, M. E. 1988. Northern hemisphere climate regimes during the last 3 Ma: possible tectonic connections. Philosophical Transactions of the Royal Society of London, B 318:411430.Google Scholar
Schneider, C. E., and Kennett, J. P. 1991. Oxygen isotopic evidence for seasonal or vertical segregation of ancestor and descendant species within the Neogene planktonic foraminiferal lineage Globorotalia (Globoconella). Geological Society of America Abstract with Programs A3435.Google Scholar
Shackleton, N. J., Backman, J., Zimmerman, H. et al. 1984. Oxygen isotope calibration of the onset of ice-rafting and history of glaciation in the North Atlantic region. Nature (London) 307:620623.CrossRefGoogle Scholar
Shackleton, N. J., Imbrie, J., and Pisias, N. G. 1988. The evolution of oceanic oxygen-isotope variability in the North Atlantic over the past three million years. Philosophical Transactions of Royal Society of London B 318:679688.Google Scholar
Signes, M., Bijma, J., Hemleben, C., and Ott, R. 1993. A model for planktic foraminiferal shell growth. Paleobiology 19:7191.CrossRefGoogle Scholar
Sokal, R. R., and Rohlf, F. J. 1981. Biometry. 2d ed.W. H. Freeman, New York.Google Scholar
Stein, R. 1986. Late Neogene evolution of paleoclimate and paleoceanic circulation in the Northern and Southern Hemisphere—a comparison. Geologische Rundschau 75:125138.CrossRefGoogle Scholar
Stein, R., and Roberts, C. 1986. Siliciclastic sediments at sites 588, 590, and 591: Neogene and Paleogene evolution in the southwest pacific and Australian climate. Pp. 14371455in Kennett, and von der Borch, et al. 1986.Google Scholar
Strauss, R. E. 1987. On allometry and relative growth. Systematic Zoology 36:7275.CrossRefGoogle Scholar
Takahashi, K., and , A. W. H. 1984. Planktonic foraminifera: factors controlling sinking speeds. Deep-Sea Research 12:14771500.CrossRefGoogle Scholar
Vella, P. P., and Collen, J. D. 1984. Four rhyolitic tuff marker beds, Lower Pliocene, Wairarapa. New Zealand Journal of Royal Society of New Zealand 14:133138.CrossRefGoogle Scholar
Weaver, P. P. E., and Clement, B. M. 1986. Synchroneity of Pliocene planktonic foraminiferal datums in the North Atlantic. Marine Micropaleontology 10:295307.CrossRefGoogle Scholar
Wei, K.-Y. 1987a. Multivariate morphometric differentiation of chronospeices in the late Neogene planktonic foraminiferal lineage Globoconella. Marine Micropaleontology 12:183202.CrossRefGoogle Scholar
Wei, K.-Y. 1987b. Tempo and mode of evolution in Neogene planktonic foraminifera: taxonomic and morphometric evidence. Ph.D. dissertation. University of Rhode Island, Kingston.Google Scholar
Wei, K.-Y. 1994. Stratophenetic tracing of phylogeny using SIMCA pattern recognition technique: a case study of the Late Neogene planktic foraminifera Globoconella plexus. Paleobiology 20:5265.CrossRefGoogle Scholar
Wei, K.-YIn press. Statistical pattern recognition in paleontology using SIMCA-MACUP. Journal of Paleontology.Google Scholar
Wei, K.-Y., and Kennett, J. P. 1988. Phyletic gradualism and punctuated equilibrium in the late Neogene planktic foraminiferal clade Globoconella. Paleobiology 14:345363.CrossRefGoogle Scholar
Wei, K.-Y., Zhang, Z.-W., and Wray, C. 1992. Shell ontogeny of Globorotalia inflata (I): growth dynamics and ontogenetic stages. Journal of Foraminiferal Research 22:318327.CrossRefGoogle Scholar